You are on page 1of 5

Earth and Planetary Science Letters 259 (2007) 625 – 629

www.elsevier.com/locate/epsl

Discussion
A comment on “Rapid late Miocene rise of the Bolivian Altiplano:
Evidence for removal of mantle lithosphere” by C.N. Garzione et al.
[Earth Planet. Sci. Lett. 241 (2006) 543–556]
A.J. Hartley a,⁎, T. Sempere b , G. Wörner c
a
Department of Geology & Petroleum Geology, School of Geoscience, University of Aberdeen, Aberdeen AB24 3UE, UK
b
LMTG, Université de Toulouse, CNRS, IRD, OMP, 14 avenue Edouard Belin, 31400 Toulouse, France
c
Geowissenschaftliches Zentrum Göttingen, Abt. Geochemie, Goldschmidtstr. 1, Universität Göttingen, 37077 Göttingen, Germany

Received 2 August 2006; received in revised form 21 March 2007; accepted 2 April 2007
Available online 11 April 2007
Editor: H. Elderfield

1. Introduction samples spread over a 11.54–5.46 Ma stratigraphic


interval. They assumed that the measured δ18O values
Garzione et al. (2006) measured oxygen isotope from pedogenic, palustrine and lacustrine limestones
compositions of Miocene paleosol carbonate nodules accurately represent the composition of the paleo–
and lacustrine and palustrine limestones deposited in the meteoric water from which they precipitated based on
northern Bolivian Altiplano. They suggested that these the relationship between δ18O values of meteoric water
compositions record an increase in elevation which and altitude (Gonfiantini et al., 2001). They observed a
indicate that the Altiplano surface was uplifted by 2.5– 3–4‰ change to increasingly negative δ18O values
3.5 km between ∼10.3 and ∼6.8 Ma. They further between 10.3 and 6.8 Ma interpreted to represent a
suggested that the rapidity and magnitude of this uplift 2500–3500 m increase in altitude during this time
could only be accounted for by crustal delamination. interval. However, seasonal variations in δ18O values
There are, however, alternative ways of interpreting these between 10 to 15‰ occur in modern meteoric waters on
data which do not require such rapid uplift and are not at the Altiplano and evaporative enrichment can account
odds with other geological data from the Central Andes. for a shift in δ18O values of up to 2‰ (Ghosh et al.,
2006). The possibility therefore that the δ18O values
2. Oxygen isotope data and paleoelevation estimates presented by Garzione et al. do not accurately record
paleo–meteoric water should be considered, particularly
Garzione et al. (2006) sampled the upper part of a as the likely impact of fractionation on the older samples
≥ 8 km thick section of continental sediments present in is to significantly underestimate paleoelevation.
the Corque Syncline of the northern Altiplano of Bolivia. The δ18O values for specific time intervals reported
They presented oxygen isotope data for 79 carbonate by Garzione et al. (2006) are − 9.5‰ to − 13.3‰
between 11.5 and 10.3 Ma, − 8.3‰ to −14.1‰ between
7.6 and 6.8 Ma, and −13.8‰ to − 15.3‰ between 6.8
⁎ Corresponding author. Tel.: +44 1224 273712; fax: +44 1224 and 5.3 Ma (excluding outliers). The δ18O values for the
272785. 11.5–10.3 Ma and 7.6–6.8 Ma time intervals show
E-mail address: a.hartley@abdn.ac.uk (A.J. Hartley). complete overlap (Fig. 1) indicating no significant shift
0012-821X/$ - see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.epsl.2007.04.012
626 A.J. Hartley et al. / Earth and Planetary Science Letters 259 (2007) 625–629

affected paleosurfaces in the Eastern Cordillera beginning


at 6.5 Ma. The geomorphological evidence suggests that
2.0–2.5 km of uplift relative to the foreland has affected the
Eastern Cordillera since 10 Ma, but only 800 m of
dissection occurred between 10 and 3 Ma. On the western
slope of the Altiplano, the timing of valley incision appears
to be controlled by (1) the eventual “breaking” of the sealed
ignimbrite surfaces at some critical time (Thouret et al.,
2003), (2) increased runoff around 7 Ma, a period when
climatic conditions were distinctly less arid than today
(Gaupp et al., 1999), and (3) the onset of glaciation and
increased runoff from about 3 to 4 Ma (Clapperton, 1993).
In addition, geomorphic evidence suggests that the volcanic
highlands on the western flank of the Altiplano were above
2000 m between 20–17 Ma (Sébrier et al., 1998). In
Fig. 1. Redrawn Fig. 4 of Garzione et al. (2006) showing δ18O values summary, geomorphological evidence for post 6.5 Ma
(taken from their Table 1) of carbonate versus stratigraphic level. Open incision generated by widespread regional surface uplift at
diamonds are palaesol carbonates, closed diamonds are palustrine
this time across the Central Andes is at best ambiguous.
carbonates. Data from the lacustrine interval (10.33 to 7.6 Ma) and
sandstone cements have been excluded in order to simplify the diagram. It is clear from structural studies that a significant
These data were interpreted by Garzione et al. (2006) to indicate a 3– amount of shortening took place before 10 Ma along the
4‰ shift to more negative δ18O values between 10.3 and 6.8 Ma. western flank of the Altiplano (Victor et al., 2004), within
the Altiplano–Puna plateau (Elger et al., 2005), and in the
to more negative δ18O values and therefore no change in Eastern Cordillera (McQuarrie et al., 2005). Reconstruction
paleoelevation. Indeed, the abrupt shift to more negative of the development of the western Altiplano (Victor et al.,
values at ∼ 6.8 Ma would then suggest that almost 2004) showed that 2600 m of structural uplift took place
instantaneous uplift occurred at this time. between 30 and 7 Ma with maximum shortening between
Garzione et al. (2006) estimated paleoelevation from 17 and 10 Ma, and elevation estimates suggest that the
the δ18O values of paleo–meteoric water from the Altiplano was bounded by mountain ranges up to 2000 m-
carbonates for different time intervals. Using their high by 9 Ma (Hartley, 2003). Evidence for significant
uncertainties the 11.5–10.3 Ma samples were deposited tectonic shortening (and implied uplift) prior to 10 Ma is
between elevations of −1700 m and +2400 m, the 7.6– overwhelming and supported by paleomagnetic data that
6.8 Ma samples between 700 and 3900 m, and the 6.8– provide ample evidence that large scale rotation and
5.4 Ma samples between 3000 and 4700 m. Given these oroclinal bending of the Andean forearc region occurred
error limits uplift between 10.3 and 6.8 Ma can only be prior to eruption of the plateau forming ignimbrites in
objectively constrained to be somewhere between 600 m southern Peru and northern Chile, i.e. between 20 and
and 6400 m. Garzione et al. (2006) supported their results 35 Ma (Roperch et al., 2006). Additional evidence for uplift
by claiming that the uplift rates they determined were of the Altiplano is provided by the presence of up to 2500 m
compatible with those based on fossil leaf morphology of conglomerates deposited between 25 and 10 Ma in the
[e.g. Gregory-Wodzicki, 2000]. However, in Bolivia and forearc and derived directly from the uplifting western edge
East Asia, the current leaf morphology method system- of the Altiplano [e.g. Victor et al., 2004; Wörner et al.,
atically underestimates altitudes when these are high (as it 2002]. In Southern Peru and Northern Chile, this sediment
reconstructs mean annual temperatures in error by as wedge has been covered by extensive plateau ignimbrites
much as 15 °C), because applying equations generated with volumes exceeding 1000 km3 (Wörner et al., 2000).
from forests in North America to unrelated forests is The age of these ignimbrites just west of the Garzione et al.
inaccurate (Kowalski, 2002). Low Miocene paleoaltitudes study area shows a narrow range around 19 Ma and
estimated by fossil leaf morphology in Andean Bolivia are suggests a short period of extensive crustal melting
therefore likely to be serious underestimations. (Wörner et al., 2002, 2000). This melting is difficult to
envisage in thin continental crust and indicates the
3. Evidence for surface uplift combined thermal effects of advective heat input into the
lower crust and crustal scale convection after crustal
Garzione et al. (2006) used the work of Kennan et al. thickening (Wörner et al., 2002, 2000; Babeyko et al.,
(1997) to suggest that widespread incision of 1.5–2.0 km 2002). The timing of these plateau ignimbrites then argues
A.J. Hartley et al. / Earth and Planetary Science Letters 259 (2007) 625–629 627

for substantial crustal thicknesses already in place and and/or west of the Altiplano, before ∼ 11 Ma and
resulting isostatic uplift prior to 20 Ma. independently of any lithospheric delamination.
In contrast to Eiler et al.'s (2006) suggestion that
4. Delamination model and implications for delamination was focused beneath the overthickened EC,
lithospheric processes Garzione et al. suggested that the eruption of mafic lavas
throughout the northern and central Altiplano between 7.5
Tectonic shortening is widely viewed as the primary and 5.5 Ma reflects removal of mantle lithosphere and
cause of crustal thickening in the Central Andes (Isacks, provides a minimum age for the event. However, none of
1988). Garzione et al. (2006) argued that the apparent the erupted products are in fact basaltic, and andesitic
uplift suggested by their results was too rapid to have magmatism also occurred at earlier times on the Bolivian
been generated by shortening-driven thickening, and Altiplano (Suarez-Soruco, 2000). In addition, these
therefore favored delamination as its cause. Such a “mafic” magmas are typically calc–alkaline, of low vol-
process can only occur however when the lower part of ume and do not carry the geochemical signature of
the lithosphere – that may include parts of the lower delamination magmatism (Kay and Kay, 1993). Mahl-
crust – has become gravitationally unstable due to burg-Kay et al. (1994) in fact found intra-plate signatures
lithospheric thickening (Kay and Kay, 1991). Delami- in mafic volcanic rocks from the Argentinian Puna, which
nation below the Corque Basin should thus have been a are indeed easily linked to delamination. In the Southern
consequence of thickening. However, if the Corque Puna case, however, delamination is concentrated locally
Basin was indeed at or below sea level at ∼ 11 Ma as and in fact has caused additional uplift since the average
suggested by Garzione et al. the Altiplano crust had elevation of this area is even higher than that of the
certainly not been thickened by then (it is now ∼ 55 km Bolivian Altiplano. None of these characteristics are
Beck et al., 1996 and, because an unthickened crust observed on the Bolivian Altiplano. A key point here is
implies an unthickened lithosphere, two obvious and that when delamination related magmatism occurs, it can
crucial questions arise: 1) why should the lower be identified. While delamination is thought to have
lithosphere start to delaminate while the local litho- occurred locally below the Southern Puna, there is no
sphere is still unthickened? And, 2) when (and how) was magmatic geochemical signature for delamination on the
the Altiplano crust thickened? To address these ques- Bolivian Altiplano. Delamination can therefore not be
tions, Eiler et al. (2006) proposed that overthickening invoked as the causal mechanism for uplift of the entire
produced by shortening in the Eastern Cordillera (EC) Altiplano.
triggered delamination there, producing uplift of the EC A significant calculation error should also be noted.
and eastern Altiplano and simultaneously enabling Garzione et al.'s Eq. (2), dh/dt = [(ρm −ρc) /ρc] dH/dt=
crustal flow from the EC into the Altiplano crust. Whilst [(ρm −ρc) /ρc] uH/W, allegedly links the rate of surface
the Altiplano crust is likely to have been thickened by uplift, dh/dt, to the rate of crustal thickening, dH/dt, as well
lower crustal flow from the overthickened Eastern and/or to the shortening rate, u, and the width of the deforming
Western Cordilleras (Wörner et al., 2002; Husson and region, W. This equation is however incorrect, as it is the
Sempere, 2003; Yang et al., 2003), evidence for removal density of the mantle, ρm, that should be the denominator
of mantle lithosphere beneath the EC has not been of the density ratio, instead of ρc, the density of the crust.
reported. If thickening of the Altiplano crust started at Let's define K as the ratio (uH/W) / (dh/dt), applying the
10.3 Ma following delamination in the EC and regionally values chosen for ρm and ρc by Garzione et al. (respectively
triggering crustal flow, this implies that overthickening 3300 and 2800 kg/m3) to their Eq. (2), K should have been
had reached a critical point in the cordillera(s) to the west found to be 5.6, and not 5.5 as published by these authors;
and/or east, and therefore that 1) these areas were already applying these values to the correct equation instead
at a high elevation, and 2) the lithosphere beneath the indicates K = 6.6. Fortunately, these errors have minor
Altiplano had behaved as an anomalous region “resist- consequences: applying the values chosen by Garzione et
ing” thickening. Although the Altiplano behaved al., dh/dt becomes 0.25 mm/yr (instead of 0.3), and 3 km of
somewhat anomalously during its Mesozoic and Ceno- uplift is achieved in 12 Myr (instead of 10) when the uplift
zoic evolution (Sempere et al., 2002), it is difficult to process is driven solely by tectonic shortening.
explain why lower crustal flow into the Altiplano crust
was hampered until delamination occurred beneath the 5. Conclusions
EC. Furthermore, this chain of reasoning paradoxically
implies that significant crustal thickening, and thus high We conclude that Garzione et al.'s (2006) model for
altitudes, had been acquired in the Central Andes, east large-scale rapid uplift of the Altiplano between 10.3 and
628 A.J. Hartley et al. / Earth and Planetary Science Letters 259 (2007) 625–629

6.8 Ma generated by crustal delamination is not supported Elger, K., Oncken, O., Glodny, J., 2005. Plateau-style accumulation of
by their data. In particular there are problems with the deformation: Southern Altiplano. Tectonics 24. doi:10.1029/
2004TC001675.
isotope data due to the likelihood of fractionation resulting Garzione, C.N., Molnar, P., Libarkin, J.C., 2006. Macfadden, rapid late
from evaporation, the wide range of modern-day Alti- Miocene rise of the Bolivian Altiplano: evidence for removal of
plano δ18O values, and the lack of a change in δ18O mantle lithosphere. Earth Planet. Sci. Lett. 241, 543–556.
values between 10.3 and 6.8 Ma. The paleoelevation Gaupp, R., Kött, A., Wörner, G., 1999. Paleoclimatic implications of
Mio–Pliocene sedimentation in the high-altitude intra-arc Lauca
estimates determined by Garzione et al. between 10.3 and
Basin of northern Chile. Palaeogeogr. Palaeoclimatol. Palaeoecol.
6.8 Ma range between 600 m and 6400 m when un- 151, 79–100.
certainties are included, and structural, stratigraphic, Ghosh, P., Garzione, C.N., Eiler, J.M., 2006. Rapid uplift of the
sedimentological and geomorphological data from the Altiplano revealed in 13C–18O bonds in paleosol carbonates.
central Andes indicate instead that considerable shorten- Science 311, 511–515.
ing and likely surface uplift occurred prior to 10 Ma. The Gonfiantini, R., Roche, M.-A., Olivry, J.-C., Fontes, J.-C., Zuppi, G.
M., 2001. The altitude effect on the isotopic composition of tropical
delamination mechanism favoured byGarzione et al. rains. Chem. Geol. 181, 147–167.
(2006) is difficult to reconcile with the fact that dela- Gregory-Wodzicki, K.M., 2000. Uplift history of the Central and Northern
mination follows lithospheric thickening and that the local Andes; a review. Geol. Soc. Amer. Bull. 112 (7), 1091–1105.
crust cannot have been thickened prior to 10 Ma because – Hartley, A.J., 2003. Andean uplift and climate change. J. Geol. Soc.
in their model – the Altiplano was located at relatively (Lond.) 160, 7–10.
Husson, L., Sempere, T., 2003. Thickening the Altiplano crust by gravity-
low elevations; identification of the Eastern Cordillera as driven crustal channel flow. Geophys. Res. Lett. 30, 1243–1246.
an outlying locus of delamination, as proposed by Eiler Isacks, B.L., 1988. Uplift of the Central Andean Plateau and bending
et al. (2006), implicitly leads to the recognition that high of the Bolivian Orocline. J. Geophys. Res. 93 (4), 3211–3231.
altitudes existed in the Central Andes before 10 Ma. There Kay, R.W., Kay, S.M., 1991. Creation and destruction of lower
is however no geochemical signature for delamination continental crust. Geol. Rundsch. 80, 259–278.
Kay, R.W., Kay, S.M., 1993. Delamination and delamination
beneath neither the Bolivian Altiplano nor the EC, but it is magmatism. Tectonophysics 219 (1–3), 177–189.
clear that if delamination had taken place the magmatic L. Kennan, S.H. Lamb, L. Hoke, High-altitude palaeosurfaces in the
signature would be identified. In addition the thin crust Bolivian Andes; evidence for late Cenozoic surface uplift, in: M.
model of Garzione et al. is not supported by the geo- Widdowson (Ed.), Palaeosurfaces; Recognition, Reconstruction
chemical evidence which suggests a significant crustal and Palaeoenvironmental Interpretation, Geological Society Spe-
cial Publication, vol. 120, The Geological Society, London, 1997,
thickness was present prior to 20 Ma. pp. 307–323. Geological Society of London.
An alternative explanation of the oxygen isotope data Kowalski, E.A., 2002. Mean annual temperature estimation based on
of Garzione et al. is that the lack of a trend among the leaf morphology: a test from tropical South America. Palaeogeogr.
11.5–6.8 Ma dataset suggests that the samples have been Palaeoclimatol. Palaeoecol. 188, 141–165.
Mahlburg-Kay, S., Coira, B., Viramonte, J., 1994. Young mafic back-
subject to fractionation due to evaporation resulting in an
arc volcanic rocks as indicators of continental lithospheric
underestimation of paleoelevation. This means that delamination beneath the Argentine Puna plateau, central Andes.
either instantaneous uplift occurred at 6.8 Ma, or that J. Geophys. Res. 99, 24323–24339.
the large-scale uplift between 10.3 and 6.8 advocated by McQuarrie, N., Horton, B.K., Zandt, G., Beck, S., DeCelles, P.G.,
Garzione et al. (2006) is unwarranted This latter view is 2005. Lithospheric evolution of the Andean fold-thrust belt,
supported by abundant geological data, which point to Bolivia and the origin of the central Andean plateau. Tectono-
physics 399, 15–37.
significant shortening, uplift and erosion in this region Roperch, P.T., Sempere, T., Macedo, O., Arriagada, C., Fornari, M.,
before 10 Ma. Tapia, C., García, M., Laj, C., 2006. Counterclockwise rotation of
Late Eocene–Oligocene forearc deposits in southern Peru and its
References significance for oroclinal bending in the Central Andes. Tectonics 25.
doi:10.1029/2005TC001882.
Babeyko, A.Y., Sobolev, S.V., Trumbull, R.B., Oncken, O., Lavier, L. Sébrier, M., Lavenu, A., Fornari, M., Soulas, J.-P., 1998. Tectonics and
L., 2002. Numerical models of crustal scale convection and partial uplift in Central Andes (Peru, Bolivia and northern Chile) from
melting beneath the Altiplano–Puna plateau. Earth Planet. Sci. Eocene to present. Géodynamique 3 (1–2), 85–106.
Lett. 199, 373–388. Sempere, T., Carlier, G., Soler, P., Fornari, M., Carlotto, V., Jacay, J.,
Beck, S.L., Zandt, G., Myers, S.C., Wallace, T.C., Silver, P.G., Drake, L., Arispe, O., Néraudeau, D., Cárdenas, J., Rosas, S., Jiménez, N.,
1996. Crustal thickness variations in the Central Andes. Geology 25, 2002. Late Permian–Middle Jurassic lithospheric thinning in Peru
407–410. and Bolivia, and its bearing on Andean-age tectonics. Tectono-
Clapperton, C.M., 1993. Quaternary Geology and Geomorphology of physics 345, 153–181.
South America. Elsevier, Amsterdam, p. 779. Suarez-Soruco, R., 2000. Geological Compendium of Bolivia. Rev.
Eiler, J., Garzione, C.N., Ghosh, P., 2006. Response to comment on Tec. Yacim. Pet. Fisc. Boliv. 18 (part A), 1–76.
“Rapid uplift of the Altiplano revealed through 13C–18O bonds in J.-C. Thouret, G. Wörner, B. Singer, A. Finizola, Valley evolution,
paleosol carbonates”. Science 314, 760b. uplift, volcanism and related hazards on the western slope of the
A.J. Hartley et al. / Earth and Planetary Science Letters 259 (2007) 625–629 629

central Andes, EGS-AGU-EU Joint Assembly, Nice, abstr. (2003) 22°S): implications for magmatism and tectonic evolution of the
EAE03-A-10498 (TS7). central Andes. Rev. Geol. Chile 27, 205–240.
Victor, P., Oncken, O., Glodny, J., 2004. Uplift of the western Altiplano Wörner, G., Uhlig, D., Kohler, I., Seyfried, H., 2002. Evolution of the West
plateau: evidence from the Precordillera between 20° and 21°S Andean Escarpment at 18°S (N. Chile) during the last 25 Ma: uplift,
(northern Chile). Tectonics 23. doi:10.1029/2003TC001519. erosion and collapse through time. Tectonophysics 345, 183–198.
Wörner, G., Hammerschmidt, K., Henjes-Kunst, F., Lezaun, J., Wilke, Yang, Y.Q., Liu, M., Stein, S., 2003. A 3-D geodynamic model of
H., 2000. Geochronology (40Ar–39Ar–, K–Ar, and He-exposure- lateral crustal flow during Andean mountain building. Geophys.
ages) of Cenozoic magmatic rocks from Northern Chile (18°– Res. Lett. 30, 2093–2096.

You might also like