You are on page 1of 29

Tectonophysics 321 (2000) 297325 www.elsevier.

com/ locate/tecto

Crustal dynamics and active fault mechanics during subduction erosion. Application of frictional wedge analysis on to the North Chilean Forearc
J. Adam *,1, C.-D. Reuther
Geologisch-Pala ontologisches Institut und Museum, Universita t Hamburg, Bundesstrasse 55, D-20146 Hamburg, Germany Received 23 August 1999; accepted for publication 13 February 2000

Abstract The forearc region of the non-accreting South American Plate margin in northern Chile is characterised by subduction erosion and regional uplift. Neotectonic deformation structures reect simultaneous extensional and contractional fault kinematics. In the outer forearc, where the brittle crust directly overlies the subducting Nazca Plate, the stress regime changes from extension in the upper part to compression at the base of the forearc wedge as seen in neotectonic surface structures and seismic data. In the inner forearc, surface structures indicate a compressional stress regime also aecting the western rim of the magmatic arc. This stress regime is limited to a brittle crustal wedge segment which overlies the ductile part of the inner forearc lithosphere. The shape of the two brittle forearc wedges at the leading edge of the South American Plate is controlled by exogenetic surface processes, internal deformation processes, contemporaneous basal tectonic erosion and underplating. Mechanical parameter sets controlling and reecting the recent tectonic processes and geometrical wedge segmentation within the forearc system are evaluated and applied to general frictional wedge models. The states of stress within the crustal wedges are controlled by spatial variations of the basal mechanical parameters in the down-dip direction of the forearc wedge base. The new models illustrate the fundamental kinematics and dynamic processes of Coulomb-type basal tectonic erosion and mass transfer modes along active non-accretive convergent margins. The frictional wedge models explain the dynamics of the simultaneous and contrary deformation processes aecting the forearc crust at the North Chilean convergent margin shaped by active basal tectonic erosion. 2000 Elsevier Science B.V. All rights reserved.
Keywords: basal tectonic erosion; forearc wedge dynamics; frictional wedge analysis; mass transfer; neotectonics; non-accretive margin

1. Introduction The North Chilean segment of the South American Plate boundary is a non-accretive margin characterised by subduction erosion ( Kulm
* Corresponding author. Tel.: +49-331-288-1317; fax: +49-331-288-1370. E-mail address: jogi@gfz-potsdam.de (J. Adam) 1 Present address: GeoForschungsZentrum Potsdam, Telegrafenberg, D-14473 Potsdam, Germany.

et al., 1977; Huene and Lallemand, 1990; Huene and Scholl, 1991). Subduction erosion of the continental crust at the leading edge of the South American Plate was rst considered by Rutland (1971) to explain the present position of the extinct Mesozoic magmatic arc along the Chilean coast. Previous magmatic arcs from the Jurassic and Cretaceous have been situated in the present forearc setting since Oligocene times. The complex conguration of the active margin and structure of the forearc lithosphere are consequences of

0040-1951/00/$ - see front matter 2000 Elsevier Science B.V. All rights reserved. PII: S0 0 4 0- 1 9 51 ( 0 0 ) 0 00 7 4 -3

298

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

about 200 km eastward migration of the magmatic arc and the related forearc/backarc system since Jurassic times (Scheuber and Reutter, 1992; Scheuber et al., 1994). Thus, neotectonic faults and crustal discontinuities are strongly inuenced by pre-existing structures. The present forearc is subdivided into an inner and an outer domain separated by the NS-trending Atacama fault system. Neotectonic surface and oshore structures are extensional in the outer forearc (Buddin et al., 1993; Reichert and CINCA Study Group, 1996). However, within the lower part of the outer forearc lithosphere, seismological investigations indicate compressional deformation

(Delouis et al., 1996). The inner forearc domain is compressional (Jolley et al., 1990; Buddin et al., 1993). Each geodynamic model is faced with the problem of the contrasting and contemporaneous fault kinematics. According to Armijo and Thiele (1990), the formation of extensional surface structures in the outer forearc region, e.g. on the Mejillones Peninsula is related to the Coastal Escarpment (Fig. 1). They interpret this asymmetric west-dipping normal fault system of crustal scale as a secondary eect of a dip variation at the subduction interface. This bend within the subducting plate, however, was not conrmed by later geophysical data (Comte et al., 1992).

Fig. 1. Main structural features and upper crustal stresses deduced from neotectonic faults in the North Chilean forearc region between 22S and 24S ( location of Rio Salado area=RS, Talabre Thrust=TT, Quebrada Diabolo=QD).

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

299

Wdowinski et al. (1989) and Wdowinski and OConnell (1991) attributed the extension in the forearc to the shear forces acting along the base of the lithospheric wedge and derive the deformation processes within the forearc from a ow model on a plate-tectonic scale. In this model, Wdowinski and OConnell (1991) assume that the apex of the asthenospheric wedge and the edge of the lithosphere are at the same horizontal position, which they identify as the trench. But their model does not separate the brittle from the ductile part of the forearc crust and does not cover the region between the trench and the tip of the asthenospheric wedge, which is the main focus of our investigations on a more regional scale. To investigate the mechanics of basal tectonic erosion and recent deformation processes between the non-accreting North Chilean margin and the magmatic arc, we extended the critical taper analysis on to framework forearc-wedge blocks (Reuther and Adam, 1996, 1997, 1998). We apply the critical taper analysis (e.g., Coulomb-wedge modelling), originally developed to analyse the mechanics of accretionary wedges and fold-andthrust belts, on to the brittle continental crust of an entire non-accretive forearc prism. The frictional wedge models demonstrate how spatial variations of the basal mechanical parameters in down-dip direction of the wedge base control the contrasting crustal stresses and neotectonic deformation processes. We develop a general model that explains the mechanics and mass transfer modes of high-friction basal subduction erosion along a non-accretive active margin.

upper plate probable causes of frontal erosion are (a) the subduction of an oceanic plate with an intense topography, e.g. fault scarps, horst-andgraben structures, seamounts and ridges, and (b) the fundamental reconguration of the margin due to a variation of accretionary processes through time. An inferred mechanism for basal erosion is the upward migration of water into fractures along the base of the upper plate (Marauchi and Ludwig, 1980; Lallemand et al., 1994). Becoming overpressured, the uids induce hydrofracturing that softens the base of the upper plate producing clasts and slivers forming a melange with subducting sediments. Along the North Chilean non-accretive margin, subduction erosion directly aects the framework rock at the toe of the upper plate. The strong trench retreat (200 km since the Jurassic; Scheuber et al., 1994) implies that sediment accretion and accretionary wedge formation played no or only a minor role in the conguration of this segment of the plate margin. Crustal wedge dynamics In general crustal wedges (accretionary wedges, orogenic wedges and fold-and-thrust wedges) develop in convergent settings by compressional deformation of rock material until they exceed the shear strength along a basal detachment. To exceed the basal shear strength and to initiate basal displacement, the wedge has to attain a critical taper (Chapple, 1978; Davis and Suppe, 1980; Davis et al., 1983; Dahlen, 1984; Dahlen et al., 1984; Dahlen and Suppe, 1988). In the critically tapered wedge the basal shear stresses are balanced by wedge internal stresses on the verge of shear failure. The crustal stresses are generated by gravitational forces caused by the topographic gradient, plate-tectonic forces resulting from frictional resistance at the subduction interface and/or from gravitational forces generated by orogenic features. The shape of a crustal wedge is controlled by various factors such as frontal or basal accretion of rock material, internal deformation, sedimentation, surface erosion and tectonic erosion. A change of one or more of these factors generates internal deformation of the wedge caused by internal stress release to regain or to maintain

2. General concepts Subduction erosion Subduction erosion has been recognised in various regions along the convergent margins of the Pacic (e.g., Scholl et al., 1977, 1980; Karig et al., 1983; Huene et al., 1985; Huene and Culotta, 1989; Lallemand and Le Pichon, 1987). Huene and Lallemand (1990) distinguish between frontal subduction erosion at the toe of the upper plate and basal subduction erosion along the base of the

300

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

stability. The mode of stress release and strain accommodation is governed by the depth-dependent rheology within the wedge. Upper crustal wedge systems like accretionary prisms and foldand-thrust belts are formed under brittle conditions and exhibit a plastic or Coulomb-type rheology. On a regional scale, dynamic processes of large, brittle crustal wedges can be modelled following the critical taper theory for non-cohesive wedges (Dahlen, 1984). Cohesion can be neglected, since it is insignicant compared with the increasing gravitational stresses with depth and the tectonic stresses that aect the wedge within a convergent setting.

3. Forearc wedge formation and plate-tectonic setting To explain the origin of the deformation processes within the brittle part of the forearc, we briey discuss the lithospheric structure and the forces acting on the North-Central Andean subduction system. According to the mechanical model of doubly vergent compressional orogens by Willett et al. (1993), the North Chilean Plate margin can be divided into three large lithospheric units ( Fig. 2): pro-wedge, orogenic plateau and retro-wedge. As proposed in the model of Wdowinski and OConnell (1991), the compression in the overriding plate arises from shear traction acting on the base of the lithosphere toward the asthenospheric wedge tip from both directions. Between the trench and the asthenospheric wedge tip, the lithosphere is sheared by the subducting Nazca Plate and between the asthenospheric wedge tip and the Brazilian shield, the base of the lithosphere is sheared by basal drag resulting from asthenospheric ow within the continental mantle. The thickened crust of the orogenic plateau of the North-Central Andes, formed by horizontal shortening of the thermally softened lithosphere, magmatic addition, lithospheric thinning, upper mantle hydration and tectonic underplating (Allmendinger et al., 1997), is in balance with the far-eld plate-tectonic compressional forces (Froidevaux and Isacks, 1984). Gravitational

forces resulting from plateau build-up act on the bounding pro-wedge and retro-wedge systems ( Willett et al., 1993). The pro-wedge is also aected by plate-tectonic stresses transmitted from the subducting Nazca Plate on to the overriding South American Plate due to high frictional resistance between the plates (Fig. 2). For northern Chile, between 18S and 24S, estimated plate coupling extends to depths of about 50 km ( Tichelaar and Ru, 1991). Lithospheric shear zones decouple the upper crust from the lower lithosphere aected by ductile deformation processes and are marked by an abrupt decrease of crustal shear strength. Probably, in the inner forearc segment, a shear zone segment ramps into the upper crust and forms an intracrustal detachment. (ICD; Fig. 2). Additionally, translithospheric shear zones like the Main Andean Thrust (MAT ), provide crustal thickening and lithospheric stacking of crustal and upper mantle segments. On the cratonic side, the MAT ramps into the upper crust and forms an intracrustal detachment beneath the Subandean belt. Thus, internal lithospheric stacking of lower crust segments leads to the initial formation of the external fold-and-thrust belt (Roeder and Chamberlain, 1995). The plateau-induced gravitational forces increase the contraction in the frictional-based bounding wedges and favour west-verging thrust formation in the pro-wedge and east-verging thrusts in the retro-wedge. This is conrmed by eld observations of west-verging neotectonic forethrusts in the inner forearc and by east-verging main thrusts in the Eastern Cordillera and the Subandean fold-and-thrust belt.

4. Forearc wedge conguration Based on eld observations and geophysical data, we correlate the contrasting deformation processes within the framework rock of the NorthChilean forearc region with the mechanical behaviour of two distinct frictional forearc wedges in dierent dynamic states of stress an outer forearc wedge and an inner forearc wedge. The forearc wedges are bounded at the top by the present-day topographic slope. The brittle outer

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

Fig. 2. Schematic mechanical model and lithospheric wedge conguration of the North-Central Andean subduction system from the Nasca Plate to the Brazilian shield (adopted after the mechanical model of doubly vergent compressional orogenes by Willett et al. (1993) and the lithospheric viscous ow model for the Andes from Wdowinski and OConnell (1991). The model illustrates the interaction of plate-tectonic stresses resulting from active subduction and gravitationally induced tectonic stresses by the thickened orogenic lithosphere. Stress transmission within the pro-wedge, orogenic plateau and retro-wedge system is indicated by white arrows, the dotted lines refer to maximum stress trajectories within the upper crust. The dashed segment of the subduction zone corresponds to the inactive subduction interface between the oceanic and continental crust. During subduction erosion the active subduction fault lies within the upper plate. Dashed bold lines=lithospheric shear zones; ICD=intracrustal detachment; MAT=Main Andean Thrust after Roeder and Chamberlain (1995); S=singularity; black arrow=active underplating; horizontal hatched area=ductile upper crust; vertical hatched area=rheological buer.

301

302

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

forearc wedge is bounded along its base by the subduction interface; the base of the brittle inner forearc wedge is marked by an intracrustal discontinuity acting as intracrustal detachment (Figs. 2 and 3). The outer forearc wedge ends at the Atacama Fault system and is backstopped by the inner forearc wedge. The brittle, inner forearc wedge ends at the thermally weakened magmatic arc, which forms a rheological buer acting like an intervertebral disk between the forearc wedge and the orogenic plateau. This rheological buer transmits gravitationally induced compressional stresses on to the brittle, inner forearc wedge, which are generated by the high-elevated Altiplano/Puna plateau (Fig. 2). In general, high subduction speeds of cold oceanic plates cause a signicant downward shift of the isotherms in the forearc segment of the overriding upper plate (Honda and Uyeda, 1983). Due to the high convergence rate of the Nazca Plate (about 9 cm/year; Minster and Jordan, 1978; DeMets et al., 1990) this downward shift of the isotherms is observable in the outer forearc wedge segment of northern Chile as conrmed by surface heat-ow density data and 2D thermal modelling (Springer, 1999). Therefore in the outer forearc wedge, thermally driven recrystallisation (power law creep rheology) is not to be expected until a depth of 3040 km which additionally is conrmed by the presence of interplate and intraplate crustal seismicity (Comte et al., 1992).

The inner forearc wedge shows compressional neotectonic deformations and extends from the Atacama fault zone to the western margin of the active magmatic arc. We analyse the deformation structures observed at the surface for the brittle subwedge overlying a midcrustal discontinuity at depths between 15 km and 25 km (Fig. 3) represented by a low-velocity zone in interpreted seismic sections between Tocopilla and Chuquicamata/ Calama (Schmitz, 1993; Wigger et al., 1994). As shown by thermal modelling (Springer, 1999) and crustal seismicity (Comte et al., 1992) for this upper crustal segment of the inner forearc lithosphere brittle behavior can be expected. The crustal wedge overlaying the intracrustal detachment is not aected by the upward deection of the isotherms in the thickened crust of the orogenic plateau due to the presence of an astenospheric wedge ( Wdowinski and Bock, 1994a,b; Springer 1999). Thus, for the outer forearc wedge and the upper crustal inner forearc wedge segment, the critical taper analysis following the Coulomb Navier failure law for brittle deformation (Paterson, 1978) or frictional sliding (Byerlee, 1978) can be applied. To investigate the fundamental processes of basal tectonic erosion at non-accreting active margins, in this study, the MejillonesAtacama section is chosen as an example for the application of the 2D critical taper approach. Geological (Scheuber and Reutter, 1992), seismological (Comte et al.,

Fig. 3. Forearc wedge conguration of the North-Chilean trench-arc system and states of active crustal stresses obtained from neotectonic and seismological data. White arrows=active compressional and extensional crustal stresses; vertical hatched=underthrusting of erosional debris in deformation or melange zone; oblique hatched=underplating of underthrusted material in basal duplex zone. LCD=lower crustal detachment, ICD=intracrustal detachment.

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

303

1992), and GPS data ( Klotz et al., 1999) indicate that in this segment of the Chile trench, oblique convergence (Minster and Jordan, 1978) is actually not signicantly aecting the active state of stress and deformation processes in the overriding forearc crust.

5. Wedge geometry and regional geological setting The outer forearc wedge extends over an average distance of 120 km from the trench axis in the west to the Atacama fault zone in the east (Figs. 1 and 3). The outer forearc wedge is limited at the base by the subduction zone with an average dip of b=10 towards the East according to seismic data of Wigger et al. (1994). Bathymetric data of Schweller et al. (1981) at latitudes 22.21S and 25.01S demonstrate inner trench slope inclinations of about a=7 characterising the toe of the outer forearc wedge. After a pronounced change within the oshore topography about 50 km landwards of the trench, the surface continues onshore into the western slope of the Coastal Cordillera with a more gentle inclination of a=3.5W characterising the internal segment of the outer forearc wedge. Neotectonic and active surface structures between the Chile trench and the Atacama fault zone reect extensional processes. In the frontal part of the outer forearc wedge, marine seismic investigations suggest extensive mass movements and block sliding along the steep, inner trench slope down into the trench (Reichert and CINCA Study Group, 1996). This slump material is trapped in horst-and-graben structures and is carried down into the subduction zone. The outer forearc wedge consists entirely of framework rock and includes palaeozoic igneous, metamorphic and folded sedimentary rocks of the Pre-Andean basement which is well exposed in the Coastal Cordillera and on the Mejillones Peninsula. The major part of the Coastal Cordillera consists of gabbros, granodiorites, mac to felsic dikes, tus and lavas associated with a JurassicEarly Cretaceous magmatic arc system (Scheuber and Reutter, 1992; Scheuber et al., 1994). Further forearc rocks are Lower Cretaceous and Cenozoic continental clastic rocks, marine

carbonates and calcareous sandstones (Ferraris and Di Biase, 1978). The entire succession of the onshore outer forearc is cut by neotectonic Wand E-dipping normal faults and large NS-trending tensile ssures (Armijo and Thiele, 1990; Buddin et al., 1993). This fault geometry indicates a subhorizontal s axis (minimum principal stress) 3 in an EW direction ( Figs. 1 and 3). Quaternary coastal uplift of the Mejillones Peninsula is indicated by the existence of high-level regressive terraces and shorelines (Armijo and Thiele, 1990). In the same region, at depths between 20 and 30 km beneath the coastal area, earthquake focal mechanisms conrm thrust faulting (Delouis et al., 1996) indicating a subhorizontal s axis (maximum prin1 cipal stress) slightly inclined to the West (Fig. 3). The inner forearc wedge extends over a distance of 210 km and is situated between the Atacama Fault and the western margin of the active magmatic arc (Figs. 1 and 3). Based on linear regression analyses of topographic data across the Precordillera and the Salar de Atacama basin, the average surface slope is a=1. The base of the brittle inner forearc wedge (ICD in Fig. 2, intracrustal discontinuity and ICD in Fig. 3) dips 7 to the east. As in the outer forearc wedge the crustal structures of the inner forearc wedge was also created by the Mid-Cretaceous and Late Cretaceous Paleogene arc systems. The wedge consists of lower Cretaceous volcanic rocks, mid-Cretaceous granodiorites, uplifted Precambrian to Lower Palaeozoic metamorphic rocks, Carboniferous to Permotriassic magmatic and sedimentary rocks, lower to mid-Jurassic carbonates and Cretaceous to recent marine/continental sediments with intercalated evaporitic layers (Reutter et al., 1988; Scheuber and Reutter, 1992; Scheuber et al., 1994) providing potential detachment horizons. The western rim of the active magmatic arc is dominated by Upper Miocene to Pleistocene ignimbrites and Neogene to recent andesites and dacites (De Silva, 1989). The inner forearc domain and the western margin of the active magmatic arc are characterised by compressional deformations. Buddin et al. (1993) describe thick-skinned thrust systems in the Pre-Cordillera with localised thin-skinned tectonics controlled by detachment horizons along evapo-

304

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

(a)

(b)
Fig. 4. (a) Fault escarpments on the southern Mejillones peninsula caused by post-Miocene and post-Pliocene extension of the outer forearc. (b) Post-late Miocene west-verging folds and thrust faults in the Rio Salado Valley (RS in Fig. 1). (c) East-verging Talabre backthrust at the western margin of the magmatic arc with Tumisa volcano in the background. This NS-trending thrust fault displaces the 8-m-thick Talabre ignimbrite (2.17 Ma) and the underlying Atana ignimbrite (4.09 Ma) of 35 m exposed thickness (TT in Fig. 1). (d) Low-angle west-verging and east-verging thrusts within the upper Miocene Sifon ignimbrite, Quebrada de Diabolo, western margin of Salar de Atacama (QD in Fig. 1).

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

305

(c)

(d)
Fig. 4. (continued )

306

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

ritic layers. Structure and seismic stratigraphy of the Salar de Atacama basin were examined by Macellari et al. (1991). They explain the Atacama basin as a late Cretaceous halfgraben structure, which underwent an Eocene to Oligocene compressional inversion. Later compression is recognised in the Rio Salado Valley where west-verging faulted anticlines in upper Miocene strata are exposed ( Fig. 4a, loc. Fig. 1). Numerous extinct fault systems and crustal discontinuities within the forearc crust reect the complex deformation history of the forearc region since Jurassic times and have potential for reactivation under the active stress regime. The active compressional stage of the inner forearc is expressed in east-verging and west-verging thrusts aecting Pliocene and Quaternary deposits at the western margin of the Salar de Atacama (Jolley et al., 1990). Compressional deformation also aects the western margin of the active magmatic arc. Beneath the Tumisa volcano, in the Quebrada So ncor the 2.03 Ma (0.35) old Talabre ignimbrite ( K-Ar age obtained from multiple determinations; De Silva, 1989) is thrusted to the east (Fig. 4b, loc. Fig. 1); near Socaire within the Pliocene Tucu caro ignimbrite (3.2 Ma, K-Ar biotite age, Ram rez and Gardeweg, 1982) east-verging thrusts and pop-up structures are exposed. Impressive low-angle west- and eastverging thrusts are exposed within the upper Miocene Sifon ignimbrite at the western margin of the Salar de Atacama in the Quebrada de Diabolo area ( Fig. 4c, loc. Fig. 1). In the literature, the east-verging compressional faults are considered as forethrusts and the west-verging faults as backthrusts. No conducive explanations for the thrust faults in the inner forearc or for the recent contradictory deformation processes in the outer forearc have been proposed until now.

6. Application of critical taper analysis to the erosive North-Chilean forearc To explain the mechanics, the active state of stress and the dynamic evolution of the described forearc-wedge system, we apply the critical taper theory for non-cohesive Coulomb wedges by

Dahlen (1984). The critical taper models describe the dynamic relations between the wedge geometry and the acting stresses within the wedges (Fig. 5). The wedge shape is controlled by the balance of various geological processes, e.g. exogenetic surface processes, internal deformation, basal subduction erosion and underplating. Moreover, the critical taper will be inuenced by variation of internal mechanical conditions of the Coulomb wedge. In the critically tapered state, the wedge deformation is balanced between internal shear failure and basal frictional sliding (Coulomb rheology). Within a critical wedge, an equilibrium between three main elements exists (Fig. 5): (1) Basal traction of the compressive wedge. (2) Stresses acting at the rear end of the wedge. (3) Wedge geometry. The internal and basal tectonic stresses of the forearc wedges result from the internal shear strength of the wedge material, the frictional coupling of the convergent plates and the gravitationally induced crustal stresses of the thickened continental crust. Thus, the three model elements are determined by the regional geodynamic and plate-tectonic setting ( Fig. 2): (1) Basal traction of the compressive wedge. Regarding the brittle forearc crust we have to consider dierent basal conditions for the inner and outer forearc wedges. Basal traction of the outer forearc wedge results from the shear resistance along the subduction interface. Basal traction of the inner forearc wedge results from the shear resistance along the intracrustal detachment (intracrustal detachment; Fig. 5). (2) Stresses at the rear end of the forearc wedge system are caused by gravitationally induced stresses of the orogenic plateau. Within the brittle crust, these stresses are transmitted by the Western Cordillera/magmatic arc (rheological buer; Figs. 2 and 3). (3) Wedge geometry. The critical taper equations (see the Appendix) describe the theoretical critical taper for crustal wedges (a+b ) required crit for basal wedge transport without internal wedge deformation. The comparison of this required critical taper value with the actual wedge geometry ( Fig. 6) characterises the active dynamic state of

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

307

Fig. 5. Schematic cross-section of a critical non-cohesive crustal forearc wedge in an active subduction system (brittle part of the lithosphere), illustrating the internal state of stress, governed by three main elements: (1) basal traction of the compressive wedge; (2) stresses at the rear end of the wedge; and (3) wedge geometry. Additionally the required geometric and mechanical parameters for Critical Taper modelling are shown (modied after Dahlen, 1984).

the forearc wedge controlling internal stresses and the kinematics of active deformation structures (Adam, 1996). Additionally to the mathematical solution (Fig. 7ad, Table 1) of Dahlen (1984), we show the graphical solution of the critical wedge problem with Mohr stress circles presented by Lehner (1986). The Mohr circles (Fig. 8ad.) give a graphical presentation of the wedge mechanics and fault kinematics of the brittle forearc wedges. Due to the self-similar wedge geometry, the applied mechanical boundary conditions control the depth-independent orientation of the active deformation structures. Therefore the limiting states of stress (eective normal stress sn and shear stress t) can be obtained for any point (sn , t ) at z z depth z within the wedge segments. Using the pole construction method ( Terzaghi, 1943; Crans and Mandl, 1980) the Mohr circles provide the actual arrangement and the sense of movement for the sets of active faults within the forearc wedge

segments (Fig. 8ad). The pole has the useful property by which stresses determined by the second intersection with the Mohr circle of any line passing through the pole act on a plane oriented parallel to that line in physical space. Already Dahlen (1984) illustrates in a series of hypothetical wedge models, that variations in style of deformation at active convergent margins are the consequence of the degree of frictional coupling between the overriding and subducting plates. Small variations of basal friction can easily account for dierent deformation patterns. The succession of deformation processes from low to high friction in the hypothetical wedge consists of extension by normal faulting, sediment subduction without accretion, accretion and imbricate thrusting and subduction erosion. Long-term basal erosion only occurs in high frictional wedges on the verge of their existence limit (no contrast between basal and internal shear strength; strength ratio x=1, Eq. 7 in the Appendix).

308

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

Fig. 6. Schematic cross-section of the North-Chilean forearc-wedge system illustrating the relations between the present wedge geometry and basal mass transfer modes within the distinct forearc segments. Numbers in wedge segments refer to wedge-specic data sets in Table 1. ICD, intracrustal detachment separating the brittle inner forearc wedge from the lower crust; LCD, newly formed lower crustal detachment in the outer forearc wedge; dashed bold lines, inactive or future faults (vertical exaggeration of factor 2).

Even high frictional wedges with minor contrast between internal and basal shear strength (x<1) are able to adjust the critical taper by internal deformation for continuous basal transport without basal erosion. An increase in basal friction results in a decrease in contrast of basal and internal shear strength (x[ 1). The critical wedge geometry, as well as the dip of the direction of maximum principle stress s relative to the wedge base, steepens simulta1 neously. The actual orientation of the maximum principle stress s controls the position of the sets of 1 slip lines (potential and active shear planes) in the wedge. With the steepening s direction the set of 1 slip lines with trenchward mass transport (e.g. forethrusts) rotates to a more wedge base parallel direction, whereas the set of slip lines with arcward mass transport (e.g. backthrusts) steepens (Fig. 8a). If the contrast in shear strength disappears ( x= 1) within the stability diagram the former eld for

stable wedge geometries (a+b ) is reduced to a line (as shown in Fig. 7a). Therefore, for a particular set of mechanical parameters, for any possible value of b only one corresponding a exists to generate a stable wedge geometry. In this case the forethrust-related slip lines are rotating parallel to the wedge base, e.g. the subduction fault. A further increase in the basal shear strength ( x>1) shifts the wedge beyond the existence limit, basal wedge transport is stopped and a new detachment must be formed up in the wedge for subduction to continue. The forethrust-related set of slip lines paralleling the blocked wedge base will be preferred for any new detachment horizon. The precise location of the wedge internal detachment will be governed by any pre-existing structures and weakness zones (Dahlen, 1984). The wedge material in the footwall of the newly formed detachment is xed to the subducting oceanic crust and will be basally eroded and transported arcward.

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

309

7. Database and global parameters Our model parameters for northern Chile are presented in Table 1. The geometrical data, geological data and seismological data are given by the regional geology and structures. The geodynamic and mechanical parameters dier signicantly from the values in sedimentary accretionary complexes as the outer forearc wedge consists entirely of framework rock. Up to now all published values are considering submerged accretionary wedges (Lallemand et al., 1994). Invariant physical properties of our wedge models are summarised in the global parameter set. For the coecient of internal friction of the framework rock, we apply m=0.7 (angle of internal friction w=35 ) characterising a brittle continental crust in frictional equilibrium on pre-existing faults oriented favourably to the acting stress eld. The values used correspond to in situ determinations within the upper brittle crust at a depth of 8 km at the KTB borehole in Germany (Brudy et al., 1997). For the density of the framework rock we have chosen 2600 kg /m3 and for the density of the pore uid 1030 kg /m3. Within active crustal wedges, the ratio of the basal traction and internal strength controls wedge kinematics and dynamics, and is characterised by: x=t /|t| with (0x1). In the special case where b x>1, the basal traction t exceeds the internal b shear strength t and a continuous detachment at the wedge base cannot develop, the wedge cannot exist. Therefore, for an active crustal wedge it is an important assumption that the internal wedge is stronger than its base, which is expressed by the following inequality: m (1l ) m (1l). To b b customise critical taper calculations it is usual to x either the coecients of internal/basal friction (m, m ) or basal/internal pore uid pressure ratios b (l, l ). b Because accretionary wedges and thrust belts are made of heterogeneous sedimentary sequences with varying mechanical properties, normally the wedge strength will be modelled by the coecients of internal and basal friction (m>m ) and xed b basal and internal pore uid pressure ratios (l=l =constant) (Dahlen and Suppe, 1988; b Roeder, 1992; Adam, 1996). In contrast to these

accretive systems, we model the dierence between internal and basal shear strength in crustal wedges by variation of the internal and basal pore uid pressure ratios (ll ) and similar coecients of b internal and basal friction (m=m =0.7) reecting b a non-accretive forearc system made of prestructured framework rock. The ratio ( x ) and the related internal and basal pore uid pressure ratios (l, l ) are summarised b as the wedge-specic parameter set in Table 1. The applied values for the internal pore uid pressure ratio vary from l=0.42 (subaerial wedge with hydrostatic pore uid pressure), l=0.6 (submerged wedge) to l=0.8 (overpressured wedge). The basal pore uid pressure ratio l will be b adjusted to model dierent strength ratios of the particular forearc wedges. The resulting strength ratios describe dierent states of the wedge base from locked ( x=1.0) to weak ( x<0.4).

8. Critical taper models for the outer forearc wedge To model the contemporaneous but contrasting deformation processes in the outer forearc wedge a temporal variation of the basal conditions, as shown in the hypothetical wedge models from Dahlen (1984) is insucient. These models only allow similar internal stress conditions and resulting deformation processes within one distinct time interval. In our Coulomb-wedge model, contrasting states of internal stress within the wedge are controlled by spatial variations of the basal mechanical parameters in the down-dip direction of the forearc wedge base despite temporal variations. Continuous erosional mass transfer since Jurassic times is more easily achieved by a steadystate process with spatial variation of wedge dynamics than by multiple temporal variations, requiring a change of geodynamic or external factors. Because this region was not aected by signicant variations of geodynamic factors (e.g. rate and direction of convergence; Scheuber et al., 1994) over the last 10 m.y. and arid climatic conditions prevailed at least since the Late Quaternary, the present wedge geometry reects the dynamically stable state and the mechanical properties

310

Table 1 Summary of the geometric/geodynamic database, global/wedge-specic parameter sets and results of critical taper analysis concerning the three modelled forearc wedge segmentsa Outer forearc wedge Toe
50 km 7 10 (subduction interface) 17 70 km 3.5 10 (subduction interface) 13.5 210 km 1.0 7 (intracrustal discontinuity ICD) 8

Inner forearc wedge Internal outer forearc wedge

Geometry data set

Surface length normal to trench Topographic slope (a) Basal dip ( b ) Actual taper (a+b )

Geological data set Stress regime near surface deduced from neotectonic structures Recent geodynamic processes [Z: thrusts, seismic data Z[: locally at deformation front Basal tectonic erosion Z[: normal faults and ssures, Coastal Scarp, submarine slumps Crustal thickening, uplift of Coastal Range

[Z: fore- and backthrusts, folds Crustal thickening

Seismological data set Stress regime deduced from fault plane solution Seismological character of wedge base (SW-seismological window) Aseismic Aseismic subduction fault; entire wedge base above SW Z[: surface (Antofagasta quake 96)[Z: thrust faulting at 2030 km depth Strong seismic coupling of subduction fault; internal wedge base in SW

Seismic activity; epicentres trace wedge base above ICD

Global parameter set Coefficient of int./basal friction ( m=m ) b Density of frame work rock (d ) litho Density of pore uid 0.7 (w=w =35 ) b 2600 kg /m3 1030 kg /m3 0.7 (w=w =35 ) b 2600 kg/m3 1030 kg/m3

0.7 (w=w =35 ) b 2600 kg /m3 1030 kg /m3 Hypothetical

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

Wedge specic parameter set Internal pore uid pressure ratio (l) Basal pore fluid pressure ratio (l ) b Strength ratio base internal/ ( x ) 0.83 (calculated by taper geometry) 0.83 (overpressured) (l=l ) b 1.0 (no discrete wedge base) Basal erosion 17 (a =7.0 ) stable 27.5 0 (wedge internal s.p. d base) 55 On verge of existence limit, stable 0.77 0.8 0.68 (strong)

0.6 (submerged ) 0.7 0.54 (intermediate) Underplating 12 (a =2 ) crit 13.2 14.3 40.7 Critical

+ 7.9. : 0.hydrostatic) 0.6 0.49 (intermediate) Taper build up 10.4 (a =3.4 ) crit 11.8 15.7 39.3 Subcritical

0.7 0.35 (weak) Detached 8 ( a = 1 ) crit 8.3 19.2 35.6 Critical

Critical taper model results Critical taper (a+b ) crit Angle of s -direction/wedge base (y ) 1 b Forethrust-related s.p. ramp angle (d ) b Backthrust-related s.p. ramp angle (d ) b Dynamic state of actual wedge

Subduction 13.5 (a =3.5 ) crit 16.9 10.6 44.4 Critical

a SW, seismogenic window; ICD, intracrustal detachment; s.p., shear plane; [Z compressional stress regime; Z[ extensional stress regime.

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

311

and active dynamic processes of basal tectonic erosion. This concept is supported by the pronounced segmentation of the outer forearc wedge geometry. The wedge surface has a signicant slope break in the oshore area (50 km east of the trench axis) separating a toe segment from the internal outer forearc wedge (Figs. 3 and 6). This topographic feature cannot be explained by a geometrical variation along the subduction interface (wedge base) or by a change within the crustal structure of the upper plate. Thus, it implies dierent dynamic behaviours within the toe segment and the internal segment of the outer forearc wedge. 8.1. Critical taper model for the internal segment of the outer forearc wedge The model for the internal segment of the outer forearc wedge is used to limit the possible dynamic states (basal erosion to overcritical wedge extension) for a crustal forearc wedge characterised by a present taper of (a+b ) =13,5. The model topo parameters are summarised in Table 1. The wedgein Table 1) specic parameter set (parameter set describes a partly submerged crustal forearc wedge with a high frictional wedge base. Due to an internal pore uid pressure ratio l=0.6 characterising a partly submerged wedge, a high frictional base with an intermediate strength ratio ( x#0.5) can be modelled by a signicantly increased basal pore uid pressure ratio (l =0.7). This critical b taper model results in a stable to slightly overcritical wedge geometry for the internal segment of the outer forearc wedge with (a+b ) =12. This crit stable to slightly overcritical taper will be obtained by active underplating at the wedge base (see Fig. 6, set ). Contemporaneous wedge internal extensional deformation may occur to readjust the required minimum critical taper. This resulting wedge dynamic reects adequately the active geodynamic setting of the internal outer forearc wedge segment. Additionally, the critical taper results are graphically presented in a Mohr diagram (Fig. 8c), illustrating the general state of stress and active fault kinematics within the internal segment of the outer forearc wedge at depth z (s , t ). The stress z z

circle for the compressional state of stress shows a gently inclined subhorizontal maximum stress direction (s inclined by 13 to the wedge base) 1c and results in a detached wedge base with an active subduction interface on top of the oceanic crust. Contemporaneous steeper dipping thrust faults allow basal accretion and underplating of tectonically eroded forearc wedge material. Thrust faulting indicated by focal mechanisms of shallow earthquakes (depth30 km, Delouis et al., 1996) supports this dynamic model. This basal accretion of crustal material results in crustal thickening of the rear part of the outer forearc wedge and is a potential mechanism for the ongoing uplift of the coastal area. In contrast to this compressive basal accretion mechanism, neotectonic and active surface structures in the outer forearc show trench parallel extension. The normal faults are dynamically interpreted as a result of the extensional collapse of the slightly overcritical wedge due to continuous thickening by underplating. This is shown in our model by contemporaneous extensional and compressional limiting states of stress within the internal segment of the outer forearc. Near-surface extension with normal faulting would be favoured by overpressured crustal detachments. Finding these detachments is a task for further geophysical investigations. For a more detailed discussion and evaluation of the possible dynamic states of the partly submerged high frictional outer forearc wedge, the stability eld diagram for the applied wedge specic parameter set is plotted in Fig. 7c. The critical to slightly overcritical internal segment of the outer forearc wedge will plot inside the stability eld (open circle, Fig. 7c) near the lower boundary (critical minimum taper). This position near the lower minimum taper boundary reects the steady-state dynamic equilibrium between mass transfer and internal deformation processes in accretionary complexes and forearc wedges, dominantly shaped by frontal and basal accretive mass transfer modes. The extensional surface deformation is the expression of regional near-surface stress release within an overcritical but stable forearc wedge to maintain the required critical minimum taper. Pervasive

312

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

313

extensional deformation of the entire forearc crust can be excluded because it only occurs in maximum tapered critical wedges (upper boundary of stability eld). The theoretical maximum critical taper of the modelled forearc wedge segment with a basal dip b=10 is attained with a topographic slope a=22. This topographic slope is near the angle of repose of the forearc crust (a =25 ) max and may not be realised within the forearc setting even by a signicant increase of basal friction. Similar arguments are reasonable to exclude active tectonic erosion at the base of the internal outer forearc wedge as source for the thrust-related earthquakes near the wedge base. In terms of Coulomb wedge mechanics for the present forearc wedge geometry (a+b=13.5 ) nearly lithostatic internal and basal pore uid pressure ratios must be assumed to enable basal tectonic erosion of the continental forearc crust (see also next paragraph). 8.2. Critical taper model for the toe segment of the outer forearc wedge The taper increase from (a+b )=13.5 to 17 in the toe segment cannot be correlated with a changing crustal composition and structure of the upper plate or with a geometrical variation along the forearc wedge base. Thus, the segmentation must be the expression of the dierent dynamic states within the toe segment and the internal segment of the outer forearc wedge system. But, even for a submerged toe segment with strong basal traction (for example x=0.68 in Fig. 7b,

data set in Table 1) the required critical minimum taper (a+b ) =13.5 is signicantly lower crit than the present taper of (a+b ) =17. topo Therefore, the toe segment must represent the maximum taper on the verge of the existence limit due to continuous basal tectonic erosion. In this case (continuous subduction erosion driven only by Coulomb wedge mechanics and steady-state dynamic equilibrium of the outer forearc wedge geometry) the toe geometry reects the mechanical conditions of basal tectonic erosion along the North-Chilean convergent margin. With decreasing shear strength contrast, the stability eld will be reduced to a line. For a particular set of mechanical parameters and known basal dip b only a single topographic slope a exists to establish the maximum taper for a wedge suering basal erosion. Under these particular conditions for the known taper (a+b ) of the toe segment, it is possible to evaluate the mechanical parameters controlling active basal erosion by a structural approach. In the particular case of forearc wedges shaped by active Coulomb-type basal erosion, the required strength ratio x=1 is controlled by the relation of the pore uid pressure ratios (l=l =l ) b erosion which now is the only unknown parameter of the critical taper solution. In the stability eld diagram the parameter l xes the vertical position of erosion the stability line and can be adjusted by iteration. Applying this procedure to the toe segment (a= in Table 1), the stability 7, b=10, data set line ts the required toe geometry at

Fig. 7. Stability eld diagrams for the North-Chilean outer forearc-wedge system (surface slope a versus basal dip b ). (a) Toe segment of the outer forearc wedge at the verge of the existence limit without discrete wedge base and shaped by basal erosion. Increasing basal traction leads to a reduction of the stability eld resulting from the converging upper and lower limits. During basal erosion the stability eld is minimised to a line. In the state of basal erosion only one stable wedge conguration exists for any possible dip of the subduction zone. The open circle corresponds to the actual geometry of the toe segment of the outer forearc wedge. (b) Transitional segment of the outer forearc wedge during underthrusting of eroded toe material. In this case the outer forearc wedge reects the minimum critical taper. The open circle corresponds to the actual geometry of the internal segment of the outer forearc wedge. (c) Outer forearc wedge segment modelled as partly submerged wedge with high frictional wedge base. The open circle corresponds to the actual geometry of the internal segment of the outer forearc wedge and plots within the stability eld near the minimum taper boundary characterising a stable wedge geometry with a critical to overcritical taper due to active basal underplating accompanied by surcial extension. (d) Required pore uid pressure ratio depending on the variation of the angle of internal friction resulting in a stable wedge geometry for the present toe segment under basal erosion. The graph demonstrates that the variation of the angle of internal/basal friction over a wide range of suitable values in continental crust is a minor inuence on the required internal/basal pore uid pressure ratio for basal erosion.

314

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

Fig. 8. Mohr diagrams, illustrating the general state of stress and active fault kinematics within the outer forearc wedge at depth z (sn , t ). (a) Compressional limiting state of stress within the maximum tapered stable toe segment of the outer forearc wedge on z z the verge of the existence limit due to maximum basal friction and basal tectonic erosion (a=23.6, a=7, b=10 ). (b) Compressional limiting state of stress within the minimum critical tapered transitional segment of the outer forearc wedge characterised by continuous subduction (underthrusting) of erosional debris from the base of the leading toe segment (a=9.1, a=3.5, b=10 ). (c) Extensional and compressional limiting states of stress within the critical to overcritical internal segment of the outer forearc wedge characterised by active basal underplating and near-surface extensional adjustment by normal faulting (a=3.0, a=3.5, b=10 ). (d) General state of stress and active fault kinematics within the subcritical inner forearc wedge at depth z.

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

315

l =0.83 (Fig. 7a). This uid pressure ratios erosion for active Coulomb-type basal erosion characterises an overpressured toe segment (see Fig. 6 ) and is in a similar range to characteristic l values, measured directly in submarine accretionary complexes (average value for l=0.88, Lallemand et al., 1994). Furthermore, we evaluate the sensitivity of this procedure for variations of the angle of internal and basal friction (in our model w=w =35 ). For b the present taper geometry we calculate l for erosion w values ranging from 20 w=w 50. The b results are graphically shown in Fig. 7d. The diagram shows that the determination of l is erosion almost independent from the angle of internal/basal friction. In the limits of geological applicable values (w#30 for accretionary complexes to w#40 continental crust without preexisting fractures; Byerlee, 1978), no signicant uctuation (l =0.830.02) can be observed. erosion The compressional state of stress and active fault kinematics within the toe segment are graphically presented in a Mohr diagram ( Fig. 8a). The present taper of the submerged toe segment (a+b=17 ) is illustrated as a shaded prism. The active basal subduction erosion results from similar mechanical properties in the toe segment and along its base (w=w [l=l ). The compressional limitb b ing state of stress within the toe segment is characterised by a relatively steep inclined subhorizontal maximum stress direction (s inclined 27.5 to 1c the wedge base) and the forethrust-related slip lines rotate parallel to the wedge base, e.g. the subduction interface on top of the oceanic crust. In this state of stress an additional minimum increase of the shear strength ratio ( x>1 if l >0.83) starts Coulomb-type basal tectonic erob sion. Because basal wedge transport on top of the Nasca Plate is stopped, a new detachment will be formed in the toe segment to act as a new active subduction interface. The forethrust-related set of slip lines will be preferred and the material in the footwall of the newly formed detachment will be basally eroded and transported arcward with the subducting oceanic plate in a high strain deformation zone or melange zone. The process is comparable with the subduction channel model for sediment subduction (Cloos and Shreve, 1988a,b).

Because in this special case the stability eld is reduced to a line, any minor variation of the boundary conditions shifts the wedge geometry into an unstable extensional or compressional limiting state of stress (see Fig. 7a). Therefore, a temporary unstable wedge geometry may be indicated by extensional or compressional deformation structures. Minor thrust faults, normal faults and slump structures in the frontal part of the toe segment, determined from reection seismic data (Reichert and CINCA Study Group, 1996), indicate this locally unstable taper. The extensional limiting state of stress and active normal faults (extensional adjustment) are also illustrated in Fig. 8a. 8.3. Critical taper model for the transient segment of the outer forearc wedge To allow a gradual variation of wedge dynamics from simultaneous basal erosion to underplating, a transitional critically tapered wedge segment is required which performs the mass transfer of erosional debris by basal transport and underthrusting in arcward direction. With increasing depth, the dynamic behaviour of the outer forearc wedge is controlled by the gradual decrease of basal coupling ( lowering of x, for example by strain hardening or dewatering processes; Moore and Byrne, 1987). Oceanic crust and erosive material together form the subducting footwall system of the transitional outer forearc wedge segment. Thus, beneath the transitional wedge segment, the active subduction interface of the upper plate/ lower plate system is located in the hanging wall of the subduction channel with erosional debris. Beneath the adjacent wedge segment in the coastal area, the basal mechanical variation in the down-dip direction triggers progressive footwall collapse. This process shifts the active subduction interface down into the underthrusted material or nally on top of the oceanic crust. For this reason, the erosional debris partly re-enters the internal outer forearc wedge by basal underplating. In northern Chile, the pronounced topographic break 50 km east of the trench axis is the surcial indication of this transition between the toe seg-

316

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

ment and the transitional outer forearc wedge segments. The maximum tapered, erosive toe segment builds up the steep lower and middle trench slope with a =7, whereas the upper trench slope toe with an average slope a =3.5 characterises the trans minimum tapered, stable transitional segment (see Fig. 6). The arcward limit of the transitional segment is uncertain, because the minor dynamic variation from stable-underthrusting to stableunderplating causes no signicant variation in the topographic slope. The initial conditions for the frictional wedge model of the transitional outer forearc wedge segment can be summarised as follows: (1) The model parameters of the transitional segment have to be obtained by a reasonable variation of the wedge-specic data set of the erosional toe segment (data set in Table 1), reecting the gradual transition of the mechanical conditions in the outer forearc wedge. (2) The present taper (a+b ) =13.5 charactrans terises a critical-shaped, stable wedge geometry of the transitional segment, which is required for basal wedge transport and active underthrusting of eroded forearc material. The arcward transition zone from active basal tectonic erosion to underthrusting is the consequence of the decreasing strength ratio ( x<1) within the external outer forearc wedge and is supercially expressed by a remarkable decrease of the middle trench slope (a =3.5 ). For estitrans mation of the bulk reduction of the strength ratio, the wedge-specic parameter x is gradually lowered until the resulting critical taper matches the observed wedge geometry of the transitional segment (a =a =3.5, b=10 ). Geologically, crit trans this process is comparable with the successive dewatering of the arcward thickening outer forearc wedge consisting of extensively prestructured framework rock. The start conditions of the procedure are given by the mechanical parameters of the erosive toe segment (l #0.8 overpreserosion in Table 1). sured, data set Adjusting the critical taper of the transitional wedge segment by decreasing the internal pore uid pressure ratio l demonstrates clearly that a minor reduction of the internal pore uid pressure ratio of about 4% (to l=0.77) is sucient to

support the minimum tapered, stable wedge geometry, to detach the transitional segment and to stop basal tectonic erosion. The entire wedgespecic parameter set for the transitional segment is summarised in Table 1 (parameter set ). It is remarkable that the small reduction of the internal pore uid pressure ratio results in a signicant drop of the strength ratio from x=1 to x=0.68, characterising a high-frictional forearc wedge with strong basal coupling. The state of stress and active fault kinematics within the transitional segment are shown in the Mohr diagram of Fig. 8b. The compressional limiting states of stress within the minimum tapered transitional segment is characterised by ongoing basal wedge transport. The compressional state of stress shows a gently inclined, subhorizontal maximum stress direction (s inclined 16.9 to the 1c wedge base) and results in a detached wedge base with an active subduction interface located in the hanging wall of the underthrusted erosional debris. Steeper dipping thrust faults are able to attain or readjust the critical wedge geometry by minor internal deformation. In the associated stability eld diagram the transitional segment plots at the lower minimum-taper boundary (open circle, Fig. 7b) reecting the typically stable geometry of Coulomb wedges in a pro-wedge setting. 8.4. Spatial variation of the stability eld in the outer forearc wedge of northern Chile and its geodynamic interpretation To show the dynamic variation of the outer forearc wedge from basal tectonic erosion to underplating, the stability eld diagrams of the three outer forearc wedge segments (toe , transitional , internal segment ) are summarised in Fig. 9. The progressive contraction of the stability eld in the trenchward direction is clearly shown in this overlay. The stability eld area decreases continuously from the internal outer forearc wedge segment (coastal area) to the transitional segment (upper trench slope) and nally in the erosional toe segment ( lower and middle trench slope), the upper and lower boundaries of the stability eld are merged into a stability-line. The intermediate position of the transitional

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

317

Fig. 9. Variation within the stability eld and derived geodynamic behavior of the outer forearc wedge (surface slope a versus basal dip b ) controlled by the changing mechanical conditions within the toe segment ( basal erosion) and the internal outer forearc wedge ( underthrusting and underplating and extensional adjustment). Index numbers of the three dynamic states refer to the wedge-specic parameter sets in Table 1.

in Fig. 9) is comparable with segment (circle the long-term, static-stable state of non-erosive outer forearc wedges at non-accretive margins. These non-erosive outer forearc wedges seem to be controlled by uniform basal mechanical conditions and the lack of notable exogenetic mass transfer and signicant internal deformation. Otherwise, if a mechanical variation along the base of a non-erosive outer forearc wedge develops, the former static-stable state will be transformed into a dynamic-stable state. Now the interaction of tectonic deformation and mass transfer establish a dynamically controlled equilibrium of wedgestabilising and destabilising processes. This dynamically controlled state is indicated by a pronounced geometric segmentation and by a complex, active stress regime, as shown by observations of the erosive North-Chilean outer forearc. In the North-Chilean outer forearc wedge system this dynamic equilibrium provides stable taper variations (a+b ) about 5 (from 12 in the

coastal area up to 17 near the trench axis) shifting around an intermediate, basic taper of (a+b )= 13.5 (indicated by the transitional segment, in Fig. 9). The increased basal coupling in a trenchward direction leads to a reduction of the stability eld. As a consequence, the lower minimum-taper boundary ascends and the wedge taper increases by internal shortening and mass transfer to adjust to the required stable minimum taper (indicated by the grey arrow in Fig. 9). Finally, if the stability eld is reduced to a line, the maximum taper is adjusted and basal tectonic erosion begins (toe segment in Fig. 9). Towards the arc, the stability eld is enlarged as a result of decreasing basal coupling. This lowers the minimum taper boundary. Thus, the wedge taper will be narrowed through extensional deformation and surcial mass transfer until the new minimum tapered state is reached (indicated by the white arrow in Fig. 9). In our model, the minimum taper for the internal outer forearc wedge ( in Fig. 9) is not yet adjusted because the present, slightly overcritical taper ( in Fig. 9) is controlled by continuous basal underplating. Therefore, an important result is that high frictional forearc wedges are capable of building up a long-term, dynamically stable wedge geometry over a wide range from intermediate to strong basal coupling. Non-accretive outer forearc wedges, characterised by a very strong base, seem to be extremely sensitive to minor variations of the wedge-specic mechanical parameters. Small mechanical modications can initiate basal tectonic erosion and, consequently, shift an active margin from a static-stable geodynamic state with minor deformation into a contrasting dynamically stable geodynamic state with strong deformation and signicant orogenic mass transfer.

9. Critical taper model for the inner forearc wedge In contrast to the mainly submerged outer forearc wedge system, which is situated entirely upon the subducting oceanic Nasca Plate, the subaerial inner forearc wedge is mechanical embedded in the overriding South American Plate, probably mechanical decoupled from the ductile crust

318

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

through a slightly arcward dipping intracrustal discontinuity (Fig. 3). The present taper of the inner forearc wedge (a+b=8 ) is signicantly smaller than in the outer forearc wedge system. The inner forearc wedge is modelled as a crustal forearc wedge with a high frictional wedge base, characterised by an intermediate strength ratio ( x#0.5). The wedge-specic parameter set is summarised in Table 1 (set ). Due to subaerial wedge conditions, a hydrostatic internal pore uid pressure ratio l=0.42 is assumed for the model. In this case, the intermediate strength ratio ( x#0.5) of the high frictional base is modelled by a slightly increased basal pore uid pressure ratio (l =0.6). This critical taper b model results in an unstable subcritical wedge geometry for the inner forearc wedge with (a+b ) =10.4. The present average topographic crit slope a=1.0 is signicantly smaller than the critical topographic slope a =3.4. In this subcritical crit state the inner forearc wedge is basally attached and has to build up a critically tapered, stable wedge geometry by active compression and internal shortening by thrusting and folding (see Fig. 6, set ). The state of stress at depth z (sn ,t ) and the z z active fault kinematics within the subcritically tapered inner forearc wedge are shown in the Mohr diagram of Fig. 8d. The modelled compressional limiting state of stress within this narrow wedge is characterised by a subhorizontal maximum stress axis (s inclined 11.8 to the wedge 1c base). The subcritical inner forearc wedge is under active compression and internal deformation processes favour active forethrusts and backthrusts (Fig. 8d) expressed by neotectonic west- and eastverging thrusts. A small-scale eld structure reecting this state of stress for conjugate thrust faults is shown in Fig. 4d. Active thrust faulting in the inner forearc wedge is characterised by outof-sequence thrusts, favoured by numerous evaporitic layers within the rock succession and previous structures which will be reactivated when tting the actual fault geometry and mechanical conditions. The result of the applied parameter set , which indicate a subcritical wedge increasing its taper by internal thrusting and folding, explains adequately the neotectonic eld observations.

The divergence of more than 2 between the present and required topographic slope indicates a subcritical state far from the minimum taper geometry. To shift the subaerial inner forearc wedge into a critical state a dramatical decrease of the strength ratio from a high frictional, intermediate wedge base with x#0.5 to a low frictional, weak wedge base with x#0.35 is required (wedgespecic parameter set for a hypothetical detached inner forearc wedge in Table 1). Therefore, the results of the frictional wedge model for the inner forearc wedge should be considered under more general aspects of lithospheric modelling. Some aspects and open questions for future modelling of inner forearc dynamics are: (1) Is the upper crustal, brittle deformation mechanically decoupled from ductile deformation in the lower lithosphere of the inner forearc, so that frictional wedge model results may interpreted quantitatively? (2) Is the wedge base thermally weakened, so that a low frictional model is more suitable to describe the dynamics of the inner forearc? (3) Is it more appropriate to use numerical models with composite, thermally controlled rheologies (brittle-ductile) for analysing the dynamics of the inner forearc wedge?

10. Active fault mechanics and mass transfer in the North-Chilean forearc system The regional correlation between the active state of stress and the active deformation structures between the Chilean trench and the Western Cordillera are summarised in Fig. 10. Similar mechanical properties within the toe segment and along its base prohibit a discrete wedge base and control Coulomb-type basal tectonic erosion. This results in the formation of a basal melange zone below a wedge-internal detachment in the outer forearc crust (LCD in Fig. 10). Within a subduction channel (Fig. 10, ), the erosional debris is transported arcward with the subducting Nazca Plate beneath the transitional segment of the outer forearc wedge (Fig. 10, ). Basal accretion of parts of the erosional debris

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

319

Fig. 10. Schematic dynamic cross-section summarising the kinematics of the main active tectonic structures and the active states of stress at the North-Chilean forearc (numbers refer to explanation in Section 10).

is controlled by the decreasing basal coupling in the down-dip direction of the wedge base. Underplating of the downcarried erosional debris forms crustal stacks at the internal base of the outer forearc-wedge (Fig. 10, ). In the onshore outer forearc of northern Chile, this mechanism controls the regional uplift. Trench parallel near surface extension by normal faults are interpreted as extensional collapse of the slightly overcritical wedge (Fig. 10, ). The upper crustal rocks are detached along large listric normal faults in the hanging wall of overpressured horizons and slide into the trench ( Fig. 10, ). Synchronous post-

Pliocene/Pleistocene west-verging forethrusts and east-verging backthrusts (Fig. 10, ) in the inner forearc and the western rim of the recent magmatic arc reect compressional internal deformation corresponding to a subcritical crustal wedge.

11. General implications of Coulomb-wedge modelling for mass transfer modes at erosive margins In terms of Coulomb wedge mechanics frontal and basal tectonic erosion occur only under partic-

320

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

ular conditions. Generally, frontal erosion processes are caused by the short-term modication of the wedge shape due to subducting topographic asperities of the oceanic crust (horst-and-graben structures, seamounts or aseismic ridges, Huene and Lallemand, 1990). Our models describe the processes and boundary conditions of continuous, long-term basal tectonic erosion at non-accretive margins completely in terms of Coulomb wedge mechanics. Regarding particular mechanical conditions (strong basal coupling, x=1) Coulomb wedges will suer signicant basal erosion over geological time scales, independently from external eects as frontal subduction of asperities. The mechanical conditions for Coulomb-type basal erosion, particularly along the wedge base, cannot be maintained over an extended region in the down-dip direction of the outer forearc wedge. For this reason, basal tectonic erosion as dominant mass transfer mode should occur in the neartrench segment of high-frictional outer forearc wedges, characterised by a typical maximum-taper wedge geometry as observed in northern Chile. Consequently, outer forearc segments at nonaccretive margins, which are shaped through longterm basal tectonic erosion, have to adjust a typical mass transfer pattern in arcward direction. This mass transfer pattern is governed by three dierent modes consisting of: (1) basal erosion (toe segment); (2) subduction/underthrusting (transitional segment); and (3) underplating of erosional material (internal segment). Each of the three contrasting mass transfer modes generates characteristic deformation structures and wedge tapers. As shown by the Coulomb-wedge analysis of the North-Chilean outer forearc system, the transition between the dierent dynamics states and their mass transfer modes is controlled by a small modication of the shear strength contrast between the crustal wedge and its base, which is probably depth-dependent. This dynamical succession of mass transfer modes (erosion, underthrusting and underplating), as observed along the erosive North-Chilean margin, is similar to typical high frictional wedge systems (accretion, underthrusting and underplat-

ing), intensely investigated at accretive margins (Gutscher et al., 1998a,b) and in analogue sandbox experiments ( Kukowski et al., 1994).

12. Structural approach for evaluation of the mechanical conditions for Coulomb-type basal erosion at non-accretive margins In the state of basal tectonic erosion of the forearc wedges, only a particular wedge geometry establishes a dynamic-stable state during tectonic destruction and erosive mass transfer reecting its mechanical conditions. Because in the erosional state the (a+b ) stability eld of the wedge is reduced to a stability line, only one corresponding topographic slope a exists for any basal dip b. Thus, the present geometry of erosional wedge segments is the key to evaluate the signicant mechanical parameters controlling basal tectonic erosion. The application of this procedure on to the erosive toe segment of the North-Chilean outer forearc wedge shows, for a wide range of crustal properties, that Coulomb-type basal tectonic erosion indicates a remarkable overpressure of the basal and internal pore uid pressure ratios (l #0.8). It can be demonstrated that the erosion oversteepend maximum taper geometry up to (a+b )=17 is controlled by basal tectonic erosion, because these taper values cannot be established by tectonic deformation and mass transfer for geologically reasonable mechanical parameter sets of the continental forearc crust.

13. Global correlation with other present-day erosive margins In the summarised stability eld diagram for active accretionary wedges and non-accretive margins ( Fig. 11) we have correlated our results, deduced from the North-Chilean outer forearc wedge, with the wedge geometries of various present-day active margins. Stability elds are plotted for typical sedimentary accretionary wedges with (Lallemand et al., 1994), low frictional base crustal forearc wedges with high frictional wedge (a,b) and nally, the stability line for base

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

321

Fig. 11. Stability eld diagram for active subduction-related accretionary wedges and crustal forearc wedges with mean tapers of 30 transects across active continental margins (modied after Lallemand et al., 1994). Stability elds are plotted for typical sedimentary accretionary wedges with low frictional base (mean friction angles and pore uid pressure ratios from measurements and structural (a+b) crustal forearc wedges with high-frictional wedge base, and stability line considerations from Lallemand et al., 1994), for forearc wedges on the verge of the existence limit shaped by Coulomb-type basal tectonic erosion (mean friction angles and pore derived and for calculated by the present wedge geometry of the North-Chilean outer forearc system, uid pressure ratios for hatched corridor indicates vertical shift of the erosional stability line by minor variation of the poreuid pressure ratio). Data set for mean tapers of active margins from Lallemand et al., 1994 (table 2, p. 12 039). (i) accretive; B1 South Barbados, B2 Barbados, Mq Martinique, Gu Guadeloupe, Bb Barbuda, Hi Hikurangi, Na Nankai, Or Oregon, CA Central Aleutian; (ii) intermediate: Mn Manila, Su Sumba, SK South Kermadec, Ka Kashima, J1 Japan 37, J2 Japan 3940, J3 Japan 4010, J4 Japan 4040, SK southern Kurile, Pe Peru, NH New Hebrides; (iii) non-accretive: NB New Britain, T1 Tonga 19, T2 Tonga 20, T3 Tonga 23, Os Osbourn, NK North Kermadec, NC2 northern Chile. NC1-Toe toe segment, NC1-OFW internal segment of outer forearc wedge (this paper).

forearc wedges shaped by Coulomb-type basal tectonic erosion. The mean friction angles and pore uid pressure ratios for the stability elds are calcuare derived and for the stability line

lated by the present wedge geometry of the NorthChilean outer forearc system (NC1-Toe and NC1-OFW in Fig. 11). In addition, the mean tapers of 28 transects across active continental

322

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325

margins are plotted (compiled in Lallemand et al., 1994). The geometric data for the typical accretionary and intermediate accretionary wedges are clustered near the minimum tapered boundary of the associated stability elds. Therefore, these wedge geometries are reecting the dynamically stable state of accretionary and crustal wedges, which develop in a pro-wedge setting at active margins. But the most important result is given by the stability line for erosive wedges, which correlates with the geometric properties of numerous nonaccretive margins all over the world. The stability line, deduced by the wedge geometry of the erosive toe segment in northern Chile, shows a strong correlation with the wedge geometries of various present-day erosive margins. The data from the active margins of Tonga (19S ) and New Britain plot on to the erosive stability line, the data from the active margins of North Kermadec and Osbourn in the corridor (hatched area in Fig. 11) close to the erosive stability line. This is a strong indication for Coulomb-type basal erosion as the principal deformation processes at these non-accretive margins and for the validity of our general concept of Coulomb-type basal erosion and the deduced structural approach for evaluating the mechanical conditions for Coulomb-type basal erosion at non-accretive margins.

gins. Therefore, Coulomb-type basal erosion of outer forearc wedges seems to be a fundamental process at non-accretive margins. Future detailed analyses and global correlation of these erosive margins should improve our model. Furthermore, it is essential to study the deformation processes and tectonic mass transfer during Coulomb-type basal erosion by numerical and analogue-sandbox modelling.

Acknowledgements Field work for this study was carried out within the project TP A1, Sonderforschungsbereich 267 Deformationsprozesse in den Anden, Freie Universita t Berlin, Technische Universita t Berlin and GeoForschungsZentrum Potsdam, nancially supported by the Deutsche Forschungsgemeinschaft, Bonn. We thank S. Lallemand and S. Wdowinski for critical and constructive comments on an earlier version of this paper.

Appendix: Numerical procedures: critical taper calculations for non-cohesive Coulomb wedges In the present study, the computations for the critical taper of non-cohesive wedges follow the classical approach and numerical procedures of Dahlen (1984). The mathematical results are shown in Table 1 (critical taper model results) and are graphically illustrated by Mohr circle presentations. In Appendix 1 all procedures for the calculations are summarised. The computation and presentation of the procedures are generated by the software program MC (trademark of Mathsoft, Inc.), licensed to C.-D. Reuther). The software workow requires a prior denition of units and quantities. Therefore, the equations are shown with an example data set and the order of presentation is determined by software requirements and diers from the order given in the referenced literature.

14. Conclusions Long-term, Coulomb-type basal tectonic erosion occurs under clearly determined mechanical conditions and generates a dynamic-stable wedge geometry and wedge segmentation with characteristic deformation and mass transfer patterns. Consequently, it is possible to get new insights into the dynamics of present day erosive outer forearc systems from structural considerations. As shown for northern Chile, the analysis of the diagnostic features (e.g. the wedge geometry and segmentation) by frictional wedge modelling gives extensive information about the internal and basal mechanical properties of the outer forearc wedge which is additionally conrmed by a global geometric comparison with other non-accretive mar-

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325 Appendix 1 Units: MPa)106 Pa; kPa)103 Pa; Grad)p/180 rad

323

Example data set: internal outer forearc wedge (data set 3, submerged, intermediate base) Global parameter set: d )1030 kg m3; d )2600 kg m3; w lith w)35 Grad, m)tan (w); m=0.7; w =35 Grad; m )tan(w ); m =0.7 b b b b Wedge specific parameter set: l=0.6 (submerged ); l =0.7 b (intermediate base) Critical topographic slope: a) 2 Grad Numerical procedures for critical taper calculations for non-cohesive Coulomb wedges (Dahlen, 1994): 1l b [1] m )m b b 1l [2] a)arctan

A B GC C C

Eective coecient of basal traction (Dahlen, 1994, eq. 13)

w )arctan ( m ) w =27.7 Grad b b b m )tan (w ) m =0.53 b b b a=3 Grad Y =1.1 Grad 0 Y =13.2 Grad b b=10 Grad Modied slope angle (Dahlen, 1994; eq. 10) Angle between maximum stress direction and wedge surface (Dahlen, 1994; eq. 9) Angle between maximum stress direction and wedge base (Dahlen, 1994; eq. 19) Required dip of wedge base to obtain the critical taper for given topographic slope Critical taper x=0.54 d =14.3 Grad b d =40.7 Grad b Ratio of basal to internal shear strength Ramp angle of forethrust-related wedge internal shear planes Ramp angle of backthrust-related wedge internal shear planes

1(d /d ) w lith tan (a) 1l sin (a) 1 a [3] Y )1 arcsin 0 2 2 sin (w)

[4] Y )1 arcsin b 2

sin (w ) b 1 w 2 b sin (w) [5] b)Y Y a b 0 [6 ] a+b=12 Grad [7] x)(1+m2 )1/2 sin (2Y ) b [8] d )1 arctan (1/m)Y b b 2

D D

[9] d )1 arctan (1/m)+Y b 2 b

References
Adam, J., 1996. Kinematik und Dynamik des neogenen Faltenund Deckengu rtels in Sizilien (Quantizierung neotektonischer Deformationsprozesse in der zentralmediterranen Afro-Europa ischen Konvergenzzone). Berliner Geowiss. Abh. A185, 171. Allmendinger, R.W., Jordan, T.E., Kay, S.M., 1997. The evolution of the Altiplano-Puna plateau of the Central Andes. Annu. Rev. Earth Planet. Sci. 25, 139174. Armijo, R., Thiele, R., 1990. Active faulting in northern Chile: ramp stacking and lateral decoupling along a subduction plate boundary? Earth Planet. Sci. Lett. 98, 4061. Brudy, M., Zoback, M.D., Fuchs, K., Rummel, F., Baumgartner, J., 1997. Estimation of the complete stress tensor to 8 km depth in the KTB scientic drill holes: Implications for crustal strength. J. Geophys. Res. B8, 1845318475. Buddin, T.S., Stimpson, I.G., Williams, G.D., 1993. North Chilean forearc tectonics and Cenozoic plate kinematics. Tectonophysics 220, 193203. Byerlee, J., 1978. Friction of rocks. Pure Appl. Geophys. 116, 615626.

Chapple, W.M., 1978. Mechanics of thin-skinned fold and thrust belts. Geol. Soc. Am. Bull. 89, 11891198. Cloos, M., Shreve, R.L., 1988a. Subduction-channel model of prism accretion, Melange formation, sediment subduction and subduction erosion at convergent plate margins: Part I. Background and description. Pure Appl. Geophys. 128, (34), 455500. Cloos, M., Shreve, R.L., 1988b. Subduction-channel model of prism accretion, Melange formation, sediment subduction and subduction erosion at convergent plate margins: Part II. Implications and discussion. Pure Appl. Geophys. 128, (34), 501545. Comte, D., Pardo, M., Dorbath, L., Dorbath, C., Haessler, H., Rivera, L., Cisternas, A., Ponce, L., 1992. Crustal seismicity and subduction morphology around Antofagasta, Chile: preliminary results from a microearthquake survey. Tectonophysics 205, 1322. Crans, W., Mandl, G., 1980. On the theory of growth faulting, part II (a). J. Petrol. Geol. 3, 209236. Dahlen, F.A., 1984. Noncohesive critical Coulomb wedges: an exact solution. J. Geophys. Res. 89, 1012510133. Dahlen, F.A., Suppe, J., 1988. Mechanics, growth, and erosion

324

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325 Neogene thrust tectonics, Atacama Desert, northern Chile. J. Geol. Soc., London 147, 769784. Karig, D.E., Kagami, H., Shipboard Scientists 1983. Varied response to subduction in Nankai Trough and Japan Trench forearcs. Nature 304 (5922), 148151. Klotz, J., Angermann, D., Michel, G.W., Porth, R., Reigber, C., Reinking, J., Viramonte, J., Perdomo, R., Rios, V.H., Barrientos, S., Barriga, R., Cifuentes, O., 1999. GPS-derived deformation of the Central Andes including the 1995 Antofagasta M =8.0 earthquake. Pure Appl. Geophys. 154, w 709730. Kulm, L.D., Schweller, W.J., Masias, A., 1977. A preliminary analysis of the subduction processes along the Andean continental margin 6 to 45. In: Talwani, M., Pitman, W.C. ( Eds.), Island Arcs Deep Sea Trenches and Back-arc Basins. Maurice Ewing Ser. vol. 1. AGU, Washington D.C., pp. 285302. Kukowski, N., Huene, R.v., Malavielle, J., Lallemand, S.E., 1994. Sediment accretion against a butress beneath the Peruvian continental margin as simulated with sandbox modeling. Geol. Rundsch. 83, 822831. Lallemand, S., Le Pichon, X., 1987. Coulomb wedge model applied to the subduction of seamounts in the Japan Trench. Geology 15, 10651069. Lallemand, S.E., Schnu rle, Ph., Malavieille, J., 1994. Coulomb theory applied to accretionary and nonacretionary wedges: possible causes for tectonic erosion and/or frontal accretion. J. Geophys. Res. 99, B6, 1203312055. Lehner, F.K., 1986. Comments on Noncohesive Critical Coulomb Wedges: An exact solution by F.A. Dahlen. J. Geophys. Res. 91, B1, 793796. Macellari, C.E., Su, M.J., Townsend, F., 1991. Structure and seismic stratigraphy of the Atacama Basin, Northern Chile. 6 Congreso Geologico Chileno Resumenes ampliados. Servicio Nacional de Geologia y Mineria, Chile Santiago. pp. 133137. Marauchi, S., Ludwig, W.J., 1980. Crustal structure of the Japan Trench: The eect of subduction of ocean crust, legs 5657. Init. Rep. Deep Sea Drill. Proj. 1 5657, part 1, 463469. Minster, J.B., Jordan, T.H., 1978. Present-day plate motions. J. Geophys. Res. 83, B11, 53315354. Moore, J.C., Byrne, T., 1987. Thickening of fault zones: a mechanism of melange formation in accreting sediments. Geology 15, 10401043. Paterson, M.S., 1978. Experimental Rock Deformation: The Brittle Field. Springer-Verlag, New York. Ramirez, C.F., Gardeweg, M., 1982. Hoja Toconao, Regio n de Antofagasta. Carta Geolo gica de Chile, 1:250 000, Servicio Nacional de Geolog a y Mineria, Santiago, Chile 54, 121. Reichert, C., CINCA Study Group 1996. Initial results of combined geoscientic investigations o- and onshore the active north-Chilean continental margin. Troisie ` me Symp. Int. Ge odynamique. Andine, Saint-Malo, pp. 107108. Reuther, C.D., Adam, J., 1996. Forearc dynamics and neotectonic arc deformation, Central Andes, Northern Chile. Ext. Abstr. Third ISAG, St. Malo ( France)., 219222.

of mountain belts. In: Clark, S.P., Burcheld, B.C., Suppe, J. ( Eds.), Processes in Continental Lithospheric Deformation, Geol. Soc. Am. Spec. Pap. 218, 161178. Dahlen, F.A., Suppe, J., Davis, D., 1984. Mechanics of foldand-thrust belts and accretionary wedges: Cohesive Coulomb Theory. J. Geophys. Res. 89, B12, 1008710101. Davis, D., Dahlen, F.A., Suppe, J., 1983. Mechanics of foldand-thrust belts and accretionary wedges.. J. Geophys. Res. 88, 11531172. Davis, D.M., Suppe, J., 1980. Critical taper in mechanics of fold-and-thrust belts. Geol. Soc. Am. Abstr. Prog. 12, 410. Delouis, B., Cisternas, A., Dorbath, L., Rivera, L., Kausel, E., 1996. The Andean subduction zone between 22 and 25S (northern Chile): precise geometry and state of stress. Tectonophysics 259, 81100. DeMets, C., Gordon, R.G., Argus, D.F., Stein, S., 1990. Current plate motions. Geophys. J. Int. 101, 425478. DeSilva, S.L., 1989. Geochronology and stratigraphy of the ignimbrites from the 2130S to 2330S portion of the Central Andes of Northern Chile. J. Volcanol. Geotherm. Res. 37, 93131. Ferraris, F.B., DiBiase, F., 1978. Carta Geologica de Chile, 1:250 000, Hoja Antofagasta, Inst. Invest. Geol., Santiago, 48 pp. Froidevaux, C., Isacks, B.L., 1984. The mechanical state in the Altiplano-Puna segment of the Andes. Earth Planet. Sci. Lett. 71, 305314. Gutscher, M.A., Kukowski, N., Malavieille, J., Lallemand, S., 1998a. Episodic imbricate thrusting and underthrusting, analog experiments and mechanical analysis applied to the Alaskan accretionary wedge. J. Geophys. Res. 103, B5, 1016110176. Gutscher, M.A., Kukowski, N., Malavieille, J., Lallemand, S.E., 1998b. Material transfer in accretionary wedges from analysis of a systematic series of analog experiments. J. Struct. Geol. 20, 407416. Honda, S., Uyeda, S., 1983. Thermal process in subduction zone a review and preliminary approach on the origin of arc volcanism. In: Shimozuru, D., Yokoyama, I. (Eds.), Arc Volcanism: Physics and Tectonics. Terra Scientic Publishing Company ( TERRAPUB), Tokyo, pp. 117140. Huene, R.v., Culotta, R.C., 1989. Tectonic erosion at the front of the Japan Trench convergent margin. Tectonophysics 160, 7590. Huene, R.v., Lallemand, S., 1990. Tectonic erosion along the Japan and Peru convergent margins. Geol. Soc. Am. Bull. 102, 704720. Huene, R.v., Scholl, D.W., 1991. Observations at convergent margins concerning sediment subduction, subduction erosion, and the growth of continental crust. Rev. Geophys. 29, 279316. Huene, R.v., Kulm, L.D., Miller, J., 1985. Structure of the frontal part of the Andean convergent margin. J. Geophys. Res. 9, B7, 54295442. Jolley, E.J., Turner, P., Williams, G.D., Hartley, A.J., Flint, S., 1990. Sedimentological response of an alluvial system to

J. Adam, C.-D. Reuther / Tectonophysics 321 (2000) 297325 Reuther, C.D., Adam, J., 1997. Dynamics and active fault mechanics during subduction erosion in the outer forearc region of Northern Chile. Ext. Abstr. VIII Congreso Geologico Chileno, Antofagasta (Chile), 5. Reuther, C.D., Adam, J., 1998. Krustendynamik und neotektonische Deformationen am aktiven su damerikanischen Plattenrand in Nord-Chile. Mitt. Geol. Pala ont. Inst. Univ. Hamburg 81, 143162. Reutter, K.J., Giese, P., Go tze, H.J., Scheuber, E., Schwab, K., Schwarz, G., Wigger, P., 1988. Structures and crustal development of the Central Andes between 21 and 25S, In: Bahlburg, H., Breitkreuz, C., Giese, P. ( Eds.), The Southern Central Andes. Lect. Notes Earth Sci. 17. Springer, Berlin Heidelberg, pp. 231261. Roeder, D., 1992. Thrusting and wedge growth, Southern Alps of Lombardia (Italy). Tectonophysics 207, 199243. Roeder, D., Chamberlain, R.L., 1995. Structural geology of sub-Andean fold and thrust belt in northwestern Bolivia. In: Tankard, A.J., Suaacuterez, R., Welsink, H.J. ( Eds.), Petroleum Basins of South America. AAPG Memoir 62, 459479. Rutland, R.W.R., 1971. Andean orogeny and sea oor spreading. Nature 233, 252255. Scheuber, E., Reutter, K.J., 1992. Magmatic arc tectonics in the Central Andes between 21 and 25S. Tectonophysics 205, 127140. Scheuber, E., Bogdanic, T., Jensen, A., Reutter, K., 1994. In: Reutter, K.J., Scheuber, E., Wigger, P.J. ( Eds.), Tectonic Development of the North Chilean Andes in Relation to Plate Convergence and Magmatism since the Jurassic. Tectonics of the Southern Central Andes. Springer, HeidelbergBerlinNew York, pp. 2348. Schmitz, M., 1993. Kollisionsstrukturen in den Zentralen Anden: Ergebnisse refraktionsseismischer Messungen und Modellierung krustaler Deformationen. Berl. Geowiss. Abh. 20(B), 1127. Scholl, D.W., Marlow, M.S., Cooper, A.K., 1977. Sediment subduction and oscraping at Pacic margins. In: Talwani, M., Pitman III, W.C. ( Eds.), Island Arcs, Deep Sea

325

Trenches and Back-arc Basins. Maurice Ewing Ser. 1. AGU, Washington, pp. 199210. Scholl, D.W., Huene, R.v., Vallier, T.L., Howell, D.G., 1980. Sedimentary masses and concepts about tectonic processes at underthrust ocean margins. Geology 8, 564568. Schweller, W.J., Kulm, L.D., Prince, R.A., 1981. Tectonics, structure, and sedimentary framework of the PeruChile Trench. Mem. Geol. Soc. Am. 154, 323349. Springer, M., 1999. Interpretation of heat-ow density in the Central Andes. Tectonophysics 306, 377395. Terzaghi, K., 1943. Theoretical Soil Mechanics. John Wiley, New York. Tichelaar, B.W., Ru, L.J., 1991. Seismic coupling along the Chilean subduction zones. J. Geophys. Res. 96, B7, 1199712022. Wigger, P.J., Schmitz, M., Araneda, M., Asch, G., Baldzuhn, S., Giese, P., Heinsohn, W.D., Martinez, E., Ricaldi, E., Rower, P., Viramonte, J., 1994. Variation in the crustal structure of the southern central Andes deduced from seismic refraction investigations. In: Reutter, K.J., Scheuber, E., Wigger, P.J. ( Eds.), Tectonics of the Southern Central Andes. Springer, HeidelbergBerlinNew York, pp. 2348. Willett, S., Beaumont, C., Fullsack, P., 1993. Mechanical model for the tectonics of doubly vergent compressional orogens. Geology 21, 371374. Wdowinski, S., Bock, Y., 1994a. The evolution of deformation and topography of high elevated plateaus. 1. Model, numerical analysis and general results. J. Geophys. Res. 99, B4, 71037119. Wdowinski, S., Bock, Y., 1994b. The evolution of deformation and topography of high elevated plateaus. 2. Application to the Central Andes. J. Geophys. Res. 99, B4, 71217130. Wdowinski, S., OConnell, R.J., 1991. Deformation of the Central Andes (1527S) derived from a ow model of subduction zones. J. Geophys. Res. 96, B7, 1224512255. Wdowinski, S., OConnell, R.J., England, P., 1989. A continuum model of continental deformation above subduction zones: application to the Andes and the Aegean. J. Geophys. Res. 94, B8, 1033110346.

You might also like