Sei sulla pagina 1di 4

GEOPHYSICAL RESEARCH LETTERS, VOL. 30, NO. 5, 1243, doi:10.

1029/2002GL016877, 2003

Thickening the Altiplano crust by gravity-driven crustal channel flow


L. Husson and T. Sempere
Institut de Recherche pour le Développement, Lima, Perú
Received 4 January 2003; revised 4 February 2003; accepted 6 February 2003; published 12 March 2003.

[1] In the Central Andes, crustal thickness is not well Altiplano plateau is the most characteristic due to its 3600 m
correlated to upper crustal shortening. Only little shortening mean elevation and 60– 65 km crustal thickness [Isacks,
is documented in the upper crust of the 60 –65 km-thick 1988; Beck et al., 1996]. The Altiplano is bounded by the
Altiplano plateau, whose thickening and uplift were delayed Eastern Cordillera (EC), which originated mainly from
with respect to earlier and greater thickening in the adjacent tectonic shortening [e.g., Sheffels, 1990], and by the Western
Western and Eastern Cordilleras. Because crustal thickness Cordillera (WC), where cluster the volcanoes of the sub-
variations induce horizontal stress gradients and cause crustal duction arc. Both EC and WC crustal thicknesses are, at least
flow, a thickness-dependent channel flow is modeled here locally, over 70 km [Beck et al., 1996], but there is no
and applied to the Central Andes. In situ thickening is evidence that the WC, unlike the EC, originated from
assumed for both cordilleras, while the Altiplano crustal tectonic contraction only. West of the WC, which is largely
thickening is generated by lateral flow from these covered by Late Neogene volcanics, the Coastal Belt dis-
overthickened adjacent domains. A 8.1018 Pa s viscosity plays a west-tapering 65– 0 km-thick crust [ANCORP Work-
channel is predicted for crustal thicknesses exceeding 45– 50 ing Group, 1999] and a complex history. In contrast, the EC
km to match the estimated topographic evolution of the has resulted from the Oligocene-Miocene tectonic inversion
Central Andes. Thickening in the cordilleras was sufficient to of a Triassic-Jurassic rift system [Sempere et al., 2002]. East
generate a flow of 6.109 m3 per unit length toward the initially of the EC, between 18S and 23S, the Subandean Zone
30 – 35 km-thick Altiplano. I NDEX T ERMS : 3210 (SAZ) is a Neogene fold-and-thrust belt, the foredeep of
Mathematical Geophysics: Modeling; 5104 Physical Properties of which underlies the Chaco plain [e.g., Dunn et al., 1995].
Rocks: Fracture and flow; 8164 Tectonophysics: Evolution of the [4] Models of Andean uplift [Gregory-Wodzicki, 2000;
Earth: Stresses—crust and lithosphere; 8102 Tectonophysics: Kennan, 2000] consider that the onset of the WC uplift
Continental contractional orogenic belts; 8122 Tectonophysics: occurred 60 Ma ago, and that the EC uplift developed later
Dynamics, gravity and tectonics. Citation: Husson, L., and and slower (Figure 2a). This first stage of mountain growth
T. Sempere, Thickening the Altiplano crust by gravity-driven last until the Late Paleogene, when the WC and EC reached
crustal channel flow, Geophys. Res. Lett., 30(5), 1243, doi:10.1029/ elevations of 2000 m and 1000 m respectively. Uplift
2002GL016877, 2003. rates of both cordilleras have increased since 20 Ma, leading
to the present-day high elevations. Located between them,
the Altiplano uplift mainly occurred during the last 20– 10
1. Introduction Ma, and thus was delayed with respect to the cordilleras.
[2] The origin of the Altiplano plateau is poorly under-
stood. In the Bolivian Orocline, i.e. the central segment of
the Central Andes, upper crustal shortening estimates do not 3. Discrepancies Between Shortening Rates and
correlate well with total crustal thickness [Kley and Mon- Crustal Thicknesses
aldi, 1998]. The largest deviation is found for the 60– 65 [5] The present-day crustal thickness in the EC can be
km-thick [Beck et al., 1996] Altiplano crust: in this domain, explained by intense inversion and tectonic contraction of
the pre-Neogene total crustal thickness was 30– 35 km previously thinned crust [Sempere et al., 2002]. According
[Sempere et al., 2002] and only minor (<10 – 15%) upper to shortening estimates [e.g., Rochat et al., 1999], bulk
crustal shortening has occurred during the Neogene [Rochat strain in the EC is 0.4, in agreement with its 70 km crustal
et al., 1999]. This profound discrepancy means that homo- thickness. In the SAZ, minimum bulk strain is 0.45, but its
geneous crustal shortening cannot have been responsible for overall crustal thickening has been only moderate due to its
the Altiplano crustal growth. We address this issue by thin-skinned deformation.
testing the hypothesis that the Bolivian Orocline has devel- [6] The high crustal thickness of the WC has not been
oped heterogeneously from gravity-driven channel flow of satisfactorily explained yet: tectonic shortening, magma-
crustal material injected from overthickened areas of the tism, and possibly other in-situ crustal growth processes
Western and Eastern Cordilleras towards the Altiplano. (see, Lamb and Hoke [1997] for a review) have contributed
to crustal growth.
2. Relevant Characteristics of the Bolivian [7] Figure 3 compares the actual crustal thickness and the
Orocline crustal thickness predicted by assuming in situ crustal
thickening correlated to upper crustal bulk strain across
[3] The Bolivian Orocline (BO) is commonly divided into the BO. The SAZ presents a large excess of crustal volume,
a number of geomorphic zones (Figure 1), among which the whereas the Altiplano is characterized by a significant
deficit (bulk strain 0.12). Because this discrepancy cannot
Copyright 2003 by the American Geophysical Union. be explained by homogeneous crustal deformation, we
0094-8276/03/2002GL016877$05.00 explore the hypothesis that thickening of the Altiplano crust

47 - 1
47 - 2 HUSSON AND SEMPERE: THICKENING THE ALTIPLANO

Figure 3. Deviation between the actual crustal thickness


(solid line, after Beck et al., 1996) and crustal thickness
predicted from estimates of in situ crustal growth in the
Central Andes (gray area); additional sources refer to other
processes for crustal growth (e.g. magmatism, underplating).

1978; Kirby, 1983]. Thus viscosity strongly decreases at


some critical depth in thick crusts. This in turns help to
understand why the upper crust is more likely described by
an elastic/brittle behavior and the lower crust by a ductile
behavior. Following Bird [1991], models have demonstrated
that a lateral flow in the middle to lower crust may explain
geological observations [e.g., Royden, 1996; McQuarrie
and Chase, 2000; Clark and Royden, 2000]. We model
Figure 1. Location map of the considered area of the the response of a low-viscosity material flowing into a
Central Andes and b) schematic lithospheric cross-section channel under a topographic load. Lateral flow variations
(after Beck et al., 1996; Myers et al., 1998, Yuan et al., induce vertical injection from the mid/lower crust and
2000). crustal thickness variations. The flow Q is assumed to occur
within a channel of constant thickness C and can be
described by a Poiseuille flow for a Newtonian fluid [e.g.,
originated by viscous flow from the overthickened WC and Turcotte and Schubert, 1982], by the equation
EC crusts.
C 3 @P
4. Viscous Mid/Lower Crustal Channel Flow Q¼ ; ð1Þ
12h @x
[8] Laboratory experiments suggest that the viscosity
exponentially decreases with temperature [e.g., Goetze, where h is the viscosity, and @P
@x the lateral pressure gradient.
We assume that the lateral pressure gradient only depends
on the lateral density variations related to crustal thickness.
The change in crustal thickness S through time is given by
the flow lateral variation and writes
 
@S C 3 @ 1 @h
¼ grc ; ð2Þ
@t 12 @x h @x

where g is the gravitational acceleration and h the predicted


elevation (h varies through time by Airy-type isostasy with
@h rm rc @S
@t ¼ rm @t ), where rm and rc are the mantle and crust
densities, of 3300 kg m3 and 2800 kg m3 respectively).
Changes in crustal thickness are calculated assuming
incompressibility (r~ v where ~v is the velocity). In our
model, the negative lateral flow variations from the zones
Figure 2. a) Observed uplift for the E. and W. Cordilleras which tend to collapse are instantaneously compensated, i.e.
(solid lines) and Altiplano (gray dashed) [Kennan, 2000], the net balance of flowing material is always positive and the
and predicted Altiplano uplift for the thickness-dependent total volume increases through time due to shortening in the
viscosity attitudes shown in b). Only EC and WC uplifts are cordilleras. We assume that the warm flowing material,
inputs for the model, Altiplano is given as a reference. while injected in a thinner crust, cools and retrieves the
Heavy black curve is the best fit model, dashed lines and viscosity of the material embedding the channel (i.e.
dotted lines show the effect of varying Sc and hmin, undeformable). The channel thickness may range from 5
respectively. Predicted Altiplano elevations are given at its km to 25 km [Wernicke, 1990]; here C is fixed to 15 km (see
center. discussion). Viscosity exponentially relates to the thermal
HUSSON AND SEMPERE: THICKENING THE ALTIPLANO 47 - 3

Royden, 2000; McQuarrie and Chase, 2000; Medvedev


and Beaumont, 2001]. Viscosity estimates from full-2D
models involving more complex rheologies [e.g., Beaumont
et al., 2001] are also in the same range. The low viscosity
found for the channel is expected for temperatures higher
than 600– 700C, consistent with the observations of parti-
ally molten rocks within the crust of high plateaus.

6. Discussion
[13] Various processes have been suggested for plateau
formation, but they seldom explain all observations when
Figure 4. Conceptual model for the lateral injection of applied to the Andes. Other studies that favor that ‘‘blind’’
mid-lower crust below the Altiplano. brittle shortening would have occurred below the Altiplano
as deep as 26– 44 km [e.g., McQuarrie and DeCelles,
field; as the thermal regime of a thick crust increases due to 2001] are contradicted by laboratory experiments [e.g.,
radiogenic heat production, we now assume that the Goetze, 1978; Kirby, 1983] and by seismological data,
viscosity of the channel is a direct function of the crustal which both suggest that the lower crust is ductile, at least
thickness (neglecting any delay for thermal relaxation). On locally as shallow as 18 km [e.g., Yuan et al., 2000]. Mantle
this basis, the viscosity strongly decreases around a critical delamination [e.g., Platt and England, 1994] is insufficient
depth Sc, and a low-viscosity layer appears within thick to produce major uplift (<800 m, Husson [2001]) and does
crusts. We subsequently define the channel viscosity hch as not induce crustal thickening. Thickening by magmatic
a function of crustal thickness S by addition [Kay and Mpodozis, 2001] on the time scale of
  an orogenesis would involve considerable heat advection,
ðhmax  hmin Þ ðS  Sc Þ
hch ¼ hmin  1 þ tanh ; ð3Þ which would appear on the surface heat flow signal.
2 h
Eventually, homogeneous crustal deformation is incompat-
ible with the observed upper crust shortening.
[14] Although the average viscosity of the lithosphere is
5. Modeling the Altiplano Crustal Growth generally considered to be about 1021 to 1022 Pa s [e.g.,
[9] We now apply our model to the BO, in order to England, 1986; Wdowinski et al., 1989] the viscosity of the
characterize the critical thickness Sc for which a strong crust is highly variable with depth due to the elevated
viscosity decrease occurs, the length parameter over which crustal geotherm. Conversion of rheological envelopes
it occurs h, and to give estimates of the maximum and (after e.g., Ranalli and Murphy [1987]) into Newtonian
minimum viscosities hmin and hmax (see profiles Figure 2b). viscosity profiles indicates that viscosity in thick crust is
[10] As the crust thickens (in situ thickening, from t1 to t2, low (<1019 Pa s) at depth for typical geotherms. Our model
Figure 4), its base gradually melts as a function of thickness, supports the theoretical results for mid-lower crustal flow
and melted fractions are incorporated into the mid-lower within high plateaus. Significant channel weakening (<1019
crust, lowering its average viscosity to hch; this layer may Pa s) occurs for critical crustal thickness of 40 to 50 km,
correspond to the partially molten zone inferred within the while normal (e.g. 35 km thick) crust channel viscosity
Andean crust [e.g. Yuan et al., 2000, Figure 1]. The channel would be higher than 2.1020 Pa s. Note that a higher channel
is bounded laterally by the subducting Brazilian craton and viscosity for normal crust does not affect our results as most
Nazca plate. Density contrasts due to lateral thickness of the predicted Altiplano uplift occurred when both cor-
variations drive the channel-flow Q described by equation dilleras had thickened up to 40– 50 km, and 2.1020 Pa s is
(1) into the thinner Altiplano crust, inducing its uplift thus a lower bound. As emphasized by Clark and Royden
without any shortening. [2000], due to the uncertainty in channel thickness, the
[11] We calibrate the viscosity variations using the results for channel viscosity are nonunique, as they scale
present-day topography together with uplift estimates from with C3/h. Variations in C from 10 to 20 km must be
Kennan [2000] and Gregory-Wodzicki [2000]. Uplift rates balanced by variations in hmin and hmax from 2.4 1018 Pa s
are converted into in situ crustal growth rates from isostasy to 1.9 1019 Pa s and from 6.1019 Pa s to 9.5 1020 Pa s,
for both WC and EC (Figure 2a). The best fit between respectively, giving a range of channel viscosities of almost
estimates and the predicted evolution of the overall top- one degree of magnitude. Our estimates for the viscosity
ography occurs for a viscosity decrease from 2.1020 Pa s to profile are based on the assumption that temperature, and
8.1018 Pa s, for a critical crustal thickness Sc of 47 km, over thus viscosity, is a function of crustal thickness. But apart
a 10 km step h (Figure 2b). Uncertainty on WC and EC from radiogenic heating, various thermal processes are
elevation input values is 500 m (L. Kennan, personal heterogeneously distributed in orogens. Full-2D models
communication, 2002) inducing variations in the Altiplano could more likely integrate these processes. However, our
predicted elevation also of 500 m. Uncertainties in input poor knowledge of the distribution of heat sources limits
elevations correspond to variations in both Sc and h by integrative analysis, and simple models with restricted
5 km for similar Altiplano uplift rates. number of free parameters yield more informative results.
[12] This predicted viscosity also matches the results of [15] In the Central Andes, the heat flow is globally
previous depth-dependent or temperature-dependent mod- correlated to crustal thickness [Springer and Forster,
els, with a strong viscosity decrease (1019 Pa s in the 1998; Husson and Moretti, 2002]. To a first order approx-
channel) for crusts thicker than 40 – 50 km [Clark and imation, viscosity may be more conveniently described as
47 - 4 HUSSON AND SEMPERE: THICKENING THE ALTIPLANO

thickness-dependent (rather than truly temperature-depend- in Petroleum Basins of South America, edited by R. Tankard, S. Suarez,
and H. J. Welsink, AAPG Mem., 62, 523 – 543, 1995.
ent). Our model suggests that the present-day Altiplano England, P., Comment on ‘‘Brittle failure in the upper mantle during ex-
crust has been almost equally fed by both cordilleras tension of continental lithosphere’’ by Dale S. Sawyer, J. Geophys. Res,
(3.109 m3 per unit length flowing eastward and as much 91, 10,487 – 10,490, 1986.
Goetze, C., The mechanisms of creep in olivine, Philos. Trans. R. Soc.
westward), but an uneven distribution of thermal sources London, Ser. X, 288, 99 – 119, 1978.
could locally limit or enhance weakening and imply an Gregory-Wodzicki, K. M., Uplift history of the central and northern Andes:
asymmetric flow (only the cumulative flows are equal as A review, Geol. Soc. Am. Bull., 112, 1091 – 1105, 2000.
they vary through time). Husson, L., Dynamique et régime thermique des chaı̂nes de montagnes—
Application aux Andes centrales, Ph.D. thesis École Normale Supérieure
[16] About volume balance in the Central Andes, we de Lyon, France, 2001.
propose that: (i) Large amounts of excess crustal volume Husson, L., and I. Moretti, Thermal regime of fold and thrust belts-An
flow from the SAZ to the EC lower crust. For a 30 km-thick application to the Bolivian Sub Andean Zone, Tectonophysics, 345,
253 – 280, 2002.
basement shortened by 70 km to 150 km [e.g., Dunn et al., Isacks, B. L., Uplift of the central Andean plateau and bending of the
1995; Rochat et al., 1999], the volume accreted to the EC Bolivian orocline, J. Geophys. Res., 93, 3211 – 3231, 1988.
ranges between 2.1 109 and 4.5 109 m3 per unit length. (ii) Kay, S., and C. Mpodozis, Central Andean ore deposits linked to shallow
subduction systems and thickening crust, GSA Today, 11, 4 – 9, 2001.
The EC shows a fair correlation between upper crustal strain Kennan, L., Large-scale geomorphology of the Andes: Interrelationships of
and crustal thickness. However, the expected excess crustal tectonics, magmatism and climate, in Geomorphology and Global Tec-
material incorporated from the SAZ is difficult to evidence tonics, edited by M. A. Summerfield, pp. 167 – 199, John Wiley, New
as crustal mass also flows from the EC into the Altiplano. York, 2000.
Kirby, S. H., Rheology of the lithosphere, Rev. Geophys., 21, 1458 – 1487,
(iii) No significant in situ crustal growth can be invoked for 1983.
the Altiplano, and its crustal volume represents the injected Kley, J., and C. R. Monaldi, Tectonic shortening and crustal thickness in the
mass from the middle and lower crusts of the adjacent central Andes: How good is the correlation?, Geology, 26, 723 – 726, 1998.
Lamb, S. H., and L. Hoke, Origin of the high plateau in the central Andes,
cordilleras. This interpretation is supported by the timing Bolivia, South America, Tectonics, 16, 623 – 649, 1997.
(Altiplano uplift delayed with respect to the cordilleras), McQuarrie, N., and C. G. Chase, Raising the Colorado plateau, Geology,
which illustrates the time span necessary for viscous pro- 28, 91 – 94, 2000.
cesses to develop, and by its crustal growth without sig- McQuarrie, N., and P. DeCelles, Geometry and structural evolution of the
central Andean backthrust belt, Bolivia, Tectonics, 20, 669 – 692, 2001.
nificant upper crustal shortening. (iv) Ductile crust was Medvedev, S., and C. Beaumont, Models for mid and lower crustal channel
likely to flow into the Altiplano from the WC. In this area flows (abstract), Eos Trans. AGU, 82(47), Fall Meet. Suppl., Abstract
early shortening was possibly larger than currently eval- T12D-0937, 2001.
Meyers, S. C., S. Beck, G. Zandt, and T. Wallace, Lithospheric-scale structure
uated and that more recent strain was low; additional across the Bolivian Andes from tomographic images of velocity and at-
processes like magmatic addition or underplating might tenuation for P and S waves, J. Geophys. Res., 103, 21,233 – 21,252, 1998.
have contributed to its crustal growth. Platt, J. P., and P. C. England, Convective removal of lithosphere beneath
mountain belts: Thermal and mechanical consequences, Am. J. Sci., 294,
[17] The recent vertical uplift of the coastal area of 307 – 336, 1994.
southern Peru and northern Chile can also be explained Ranalli, G., and D. C. Murphy, Rheological stratification of the lithosphere,
by a similar process of lateral crustal flow from the over- Tectonophysics, 132, 281 – 295, 1987.
thickened WC toward the coast. The issue of the Altiplano Rochat, P., G. Hérail, P. Baby, and G. Mascle, Bilan crustal et contrôle de la
dynamique erosive et sédimentaire sur les mécanismes de formation de
crustal thickening should also be considered in 3D, as l’Altiplano, C. R. Acad. Sci., Ser. X, 328, 189 – 195, 1999.
along-strike ductile crustal flow may have participated in Royden, L., Coupling and decoupling of crust and mantle in convergent
crustal thickening of the northern and southern Central orogens: Implications for strain partitioning in the crust, J. Geophys. Res.,
101, 17,679 – 17,705, 1996.
Andes. Sempere, T., et al., Late Permian-Middle Jurassic lithospheric thinning in
Peru and Bolivia and its bearing on Andean-age tectonics, Tectonophy-
[18] Acknowledgments. This study is a contribution of IRD’s UR sics, 345, 153 – 181, 2002.
104 (Continental Lithospheric Deformation in Convergence Zones and Sheffels, B. M., Lower bound on the amount of crustal shortening in the
Material Transfers). We acknowledge S. Wdowinski and S. Medvedev for central Bolivian Andes, Geology, 18, 812 – 815, 1990.
their reviews. Springer, M., and A. Forster, Heat flow density across the central Andean
subduction zone, Tectonophysics, 291, 123 – 139, 1998.
Turcotte, D. L., and G. Schubert, Geodynamics: Applications of Continuum
References Physics to Geological Problems, 450 pp., John Wiley, New York, 1982.
ANCORP Working Group, Seismic reflection image revealing offset of Wdowinski, S., R. O’Connell, and P. England, A continuum model of
Andean subduction-zone earthquake locations into oceanic mantle, Nat- continental deformation above subduction zones: Application to the An-
ure, 397, 341 – 344, 1999. des and the Aegean, J. Geophys. Res., 94, 10,331 – 10,346, 1989.
Beaumont, C., R. A. Jamieson, M. H. Nguyen, and B. Lee, Himalayan Wernicke, B. P., The fluid crustal layer and its implications for continental
tectonics explained by injection of a low-viscosity crustal channel dynamics, in Exposed Cross-Sections of the Continental Crust, edited by
coupled to focused surface denudation, Nature, 414, 738 – 742, 2001. M. H. Salisbury and D. M. Fountain, pp. 509 – 544, Kluwer Acad.,
Beck, S. L., G. Zandt, S. C. Maers, T. C. Wallace, P. G. Silver, and Norwell, Mass., 1990.
L. Drake, Crustal scale variations in the central Andes, Geology, 24, Yuan, X., et al., Subduction and collision processes in the central Andes
407 – 410, 1996. constrained by converted seismic phases, Nature, 408, 958 – 961, 2000.
Bird, P., Lateral extrusion of lower crust from under high topography, in the
isostatic limit, J. Geophys. Res., 96, 10,275 – 10,286, 1991. 
Clark, M. K., and L. H. Royden, Topographic ooze: Building the eastern L. Husson, Laboratoire de Géologie, Ecole Normale Supérieure, 24 rue
margin of Tibet by lower crustal flow, Geology, 28, 703 – 706, 2000. Lhomond, 75231 Paris cedex 05, France. (husson@geologie.ens.fr)
Dunn, J., K. G. Hartshorn, and P. W. Hartshorn, Structural styles and T. Sempere, I. R. D.-Perú, casilla 18-1209, Lima 18, Peru. (sempere@
hydrocarbon potential of the sub Andean thrust belt of southern Bolivia, terra.com.pe)

Potrebbero piacerti anche