the late cenozoic geodynamic evolution of the central segment of the andean subduction zone

19
305 ISSN 0016-8521, Geotectonics, 2009, Vol. 43, No. 4, pp. 305–323. © Pleiades Publishing, Inc., 2009. Original Russian Text © T.V. Romanyuk, 2009, published in Geotektonika, 2009, No. 4, pp. 63–83. INTRODUCTION The Andean subduction zone, as a convergent boundary between the oceanic Nazca Plate and the South American continent, is characterized by many almost extreme geological and geophysical parameters. The most important features are as follows. (1) The fold–thrust belt of the Andean mountain chains extends along the entire western margin of South America and is characterized by recent volcanic activity at its western margin (Fig. 1b). The Altiplano– Puna Plateau, at elevations of ~4 km masl, is situated in the central segment of this continental margin (10–30° S). The height difference of the topography in the central segment of the Andean subduction zone attains 13 km: the depth of the oceanic trench is ~7 km, whereas the height of the volcanoes in the Andes is ~6 km. (2) GPS experiments [58, 63, 69, 76] have shown that the oceanic Nazca Plate moves eastward with a velocity of 7–8 cm/yr, slightly deviating to the north. The South American Plate moves approximately to the northwest with a velocity of 1–2 cm/yr (http://sideshow. jpl.nasa.gov/mbh/series.html). Thus, the rate of west- ern drift is about 0.5–1.0 cm/yr. Both movements taken together ensure a rate of convergence in the Andean subduction zone of about 9 cm/yr (one of the world’s highest rates of subduction). (3) The Benioff seismofocal zone beneath the cen- tral Andes is traced down to a depth 670 km (Fig. 2a); earthquakes with magnitudes reaching 9.5 have been recorded here (Fig. 1a). (4) The Andean mountain system is characterized by extremely great crustal thickness (up to 70–75 km [104, 107]). In the Altiplano–Puna Plateau, the anom- aly of geoid is estimated at +60 m, while the Bouguer anomaly reaches –400 mGal [41, 64], characterizing one of the greatest lithospheric density anomalies on the Earth. The Late Cenozoic Geodynamic Evolution of the Central Segment of the Andean Subduction Zone T. V. Romanyuk Institute of Physics of the Earth, Russian Academy of Sciences, Bol’shaya Gruzinskaya ul. 10, Moscow, 123995 Russia e-mail: [email protected] Received March 26, 2007 Abstract—The presented model of the Late Cenozoic geodynamic evolution of the central Andes and the com- plex tectonic, geological, and geophysical model of the Earth’s crust and upper mantle along the Central Andean Transect, which crosses the Andean subduction zone along 21° S, are based on the integration of volu- minous and diverse data. The onset of the recent evolution of the central Andes is dated at the late Oligocene (27 Ma ago), when the local fluid-induced rheological attenuation of the continental lithosphere occurred far back of the subduction zone. Tectonic deformation started to develop in thick-skinned style above the attenu- ated domain in the upper mantle and then in the Earth’s crust, creating the bivergent system of the present-day Eastern Cordillera. The destruction of the continental lithosphere is correlated with ore mineralization in the Bolivian tin belt, which presumably started at 16° S and spread to the north and to the south. Approximately 19 Ma ago, the gently dipping Subandean Thrust Fault was formed beneath the Eastern Cordillera, along which the South American Platform began to thrust under the Andes with rapid thickening of the crust in the eastern Andean Orogen owing to its doubling. The style of deformation in the upper crust above the Subandean Thrust Fault changed from thick- to thin-skinned, and the deformation front migrated to the east inland, forming the Subandean system of folds and thrust faults verging largely eastward. The thickening of the crust was accom- panied by flows at the lower and/or middle crustal levels, delamination, and collapse of fragments of the lower crust and lithospheric mantle beneath the Eastern Cordillera and Altiplano–Puna Plateau. As the thickness of the middle and lower crustal layers reached a critical thickness about 10 Ma ago, the viscoplastic flow in the meridional direction became more intense. Extension of the upper brittle crust was realized mainly in gliding and rotation of blocks along a rhombic fault system. Some blocks sank, creating sedimentary basins. The rate of southward migration estimated from the age of these basins is 26 km/Ma. Tectonic deformation was accom- panied by diverse magmatic activity (ignimbrite complexes, basaltic flows, shoshonitic volcanism, etc.) within the tract from the Western Cordillera to the western edge of the Eastern Cordillera 27–5 Ma ago with a peak at 7 Ma; after this, it began to recede westward; by 5 Ma ago, the magmatic activity reached only the western part of the Altiplano–Puna Plateau, and it has been concentrated in the volcanic arc of the Western Cordillera during the last 2 Ma. DOI: 10.1134/S0016852109040050

Upload: t-v-romanyuk

Post on 03-Aug-2016

216 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

305

ISSN 0016-8521, Geotectonics, 2009, Vol. 43, No. 4, pp. 305–323. © Pleiades Publishing, Inc., 2009.Original Russian Text © T.V. Romanyuk, 2009, published in Geotektonika, 2009, No. 4, pp. 63–83.

INTRODUCTION

The Andean subduction zone, as a convergentboundary between the oceanic Nazca Plate and theSouth American continent, is characterized by manyalmost extreme geological and geophysical parameters.The most important features are as follows.

(1) The fold–thrust belt of the Andean mountainchains extends along the entire western margin ofSouth America and is characterized by recent volcanicactivity at its western margin (Fig. 1b). The Altiplano–Puna Plateau, at elevations of ~4 km masl, is situated in thecentral segment of this continental margin (10–

30°

S).The height difference of the topography in the centralsegment of the Andean subduction zone attains 13 km:the depth of the oceanic trench is ~7 km, whereas theheight of the volcanoes in the Andes is ~6 km.

(2) GPS experiments [58, 63, 69, 76] have shownthat the oceanic Nazca Plate moves eastward with a

velocity of 7–8 cm/yr, slightly deviating to the north.The South American Plate moves approximately to thenorthwest with a velocity of 1–2 cm/yr (http://sideshow.jpl.nasa.gov/mbh/series.html). Thus, the rate of west-ern drift is about 0.5–1.0 cm/yr. Both movements takentogether ensure a rate of convergence in the Andeansubduction zone of about 9 cm/yr (one of the world’shighest rates of subduction).

(3) The Benioff seismofocal zone beneath the cen-tral Andes is traced down to a depth 670 km (Fig. 2a);earthquakes with magnitudes reaching 9.5 have beenrecorded here (Fig. 1a).

(4) The Andean mountain system is characterizedby extremely great crustal thickness (up to 70–75 km[104, 107]). In the Altiplano–Puna Plateau, the anom-aly of geoid is estimated at +60 m, while the Bougueranomaly reaches –400 mGal [41, 64], characterizingone of the greatest lithospheric density anomalies onthe Earth.

The Late Cenozoic Geodynamic Evolution of the Central Segment of the Andean Subduction Zone

T. V. Romanyuk

Institute of Physics of the Earth, Russian Academy of Sciences, Bol’shaya Gruzinskaya ul. 10, Moscow, 123995 Russiae-mail: [email protected]

Received March 26, 2007

Abstract

—The presented model of the Late Cenozoic geodynamic evolution of the central Andes and the com-plex tectonic, geological, and geophysical model of the Earth’s crust and upper mantle along the CentralAndean Transect, which crosses the Andean subduction zone along 21

°

S, are based on the integration of volu-minous and diverse data. The onset of the recent evolution of the central Andes is dated at the late Oligocene(27 Ma ago), when the local fluid-induced rheological attenuation of the continental lithosphere occurred farback of the subduction zone. Tectonic deformation started to develop in thick-skinned style above the attenu-ated domain in the upper mantle and then in the Earth’s crust, creating the bivergent system of the present-dayEastern Cordillera. The destruction of the continental lithosphere is correlated with ore mineralization in theBolivian tin belt, which presumably started at 16

°

S and spread to the north and to the south. Approximately19 Ma ago, the gently dipping Subandean Thrust Fault was formed beneath the Eastern Cordillera, along whichthe South American Platform began to thrust under the Andes with rapid thickening of the crust in the easternAndean Orogen owing to its doubling. The style of deformation in the upper crust above the Subandean ThrustFault changed from thick- to thin-skinned, and the deformation front migrated to the east inland, forming theSubandean system of folds and thrust faults verging largely eastward. The thickening of the crust was accom-panied by flows at the lower and/or middle crustal levels, delamination, and collapse of fragments of the lowercrust and lithospheric mantle beneath the Eastern Cordillera and Altiplano–Puna Plateau. As the thickness ofthe middle and lower crustal layers reached a critical thickness about 10 Ma ago, the viscoplastic flow in themeridional direction became more intense. Extension of the upper brittle crust was realized mainly in glidingand rotation of blocks along a rhombic fault system. Some blocks sank, creating sedimentary basins. The rateof southward migration estimated from the age of these basins is 26 km/Ma. Tectonic deformation was accom-panied by diverse magmatic activity (ignimbrite complexes, basaltic flows, shoshonitic volcanism, etc.) withinthe tract from the Western Cordillera to the western edge of the Eastern Cordillera 27–5 Ma ago with a peak at7 Ma; after this, it began to recede westward; by 5 Ma ago, the magmatic activity reached only the western partof the Altiplano–Puna Plateau, and it has been concentrated in the volcanic arc of the Western Cordillera duringthe last 2 Ma.

DOI:

10.1134/S0016852109040050

Page 2: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

306

GEOTECTONICS

Vol. 43

No. 4

2009

ROMANYUK

The Central Andes are separated from the coastalregions by the Precordillera Fault Zone in the west andfrom the Chaco Plain by the Subandean Thrust-FaultZone in the east. In contrast to the South American Plat-form, which has remained stable since the Proterozoic[85, 99], the Andes were characterized by periods ofsedimentation, orogeny, rifting, and magmatic activity,at least from the Proterozoic [27, 93]. The onset ofrecent tectonomagmatic reactivation in the centralAndes, which is continuing now, is dated at the late Oli-

gocene [94]. High velocities of vertical and horizontalmovements, volcanism, and seismicity are noted atpresent.

The central Andes consist of five tectonic and oro-graphic provinces that extend parallel to the Pacificcoast and are subdivided into several subprovinces(Figs. 1, 2b): (i) the andesitic volcanic belt of the West-ern Cordillera, (ii) the high-elevated Altiplano–PunaPlateau, (iii) the fold–thrust complexes of the EasternCordillera, (iv) the IntraAndean, and (v) Subandean

75

°

W

70

°

65

°

10

°

20

°

30

°

20

mm/yr

Mw = 9.5; 1960

65

mm/yr

Mw = 8.0; 19951

km

2

km

3

km

77mm/yr (NUVEL-1A)GPS NZ-SA

South America

PACIFIC OCEAN

(a)

S

Main axis of Neogene magmatism

Boundary between hinterland and fold–nappe belt

70

°

W

80

°

W

10

AltiplanoPlateau

20

°

Central segment ofthe fold

-

nappe belt

Thin-skinneddeformation offoreland

Thick-skinneddeformation offoreland

PunaPlateau

(b) 40

°

S

Fig. 1.

The modern geodynamics of the Andean subduction zone. Panel (a): the integrated results of GPS observations at the westernmargin of the South American continent. Arrows indicate movements of stations relative to the rigid core of the South Americancontinent. To the north of

22°

S, the data from Cornell University, obtained over two years of observations(http://www.earth.nwu.edu/research/snapp.html) are presented. The arrow designated as GPS NZ-SA shows the motion of theNazca Plate toward South America, deduced from measurements at the stations located on the islands pertaining to the Nazca Plate.The arrow designated NUVEL-1A ~77 mm/yr shows the motion velocity of the Nasca Plate calculated from the NUVEL-1A model.To the south of

~22°

S, the data from [63], obtained over 1994–1996, are presented. The motion velocity of the Nazca Plate is esti-mated at 65 mm/yr. The hatched area covers the source zones of the two strongest earthquakes in the region; earthquake mechanismsare shown as a projection on the lower hemisphere. The schematic topography of the Cordilleras is shown in gray color. Panel (b):a generalized scheme of tectonic elements at the western margin of South America, after [60].

Page 3: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS

Vol. 43

No. 4

2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 307

72

°

W 71

°

70

°

69

°

68

°

67

°

66

°

65

°

25

°

24

°

23

°

22

°

21

°

64

°

Peru Bolivia

La PazThe Central Andean-

Argentina

Santiago

3.7 kmFig. 2b

(a)

Earthquakehypocenters, km

0–6060 –100100–200200–400400–700

(b)

PACIFICOCEAN

SouthAmerica

The Central

Oceanic trench Western

Cordillera

PunaPlateau

EasternCordillera and

Subandes

AltiplanoPlateau

70 57

80

52

43

4050

67

70

65

35 67 66 58

585967 51

5845425636

70

°

Á.‰. 60

°

1

2

3

4

5

6

7

Transect

Andean Transect

Fig. 2.

State-of-the-art seismic profiling and tectonic zoning of the central Andes. Panel (a): seismicity of the central Andean sub-duction zone. Earthquakes with magnitudes >5 and the line of the Central Andean Transect [1] are shown. The darkest domain inthe continuous-tone sketch of the of South American topography embraces elevations higher than 3.7 km. Panel (b): the main tec-tonic provinces (Western Cordillera, Altiplano and Puna plateaus, Eastern Cordillera, and Subandes) and the state-of-the-art seismicprofiling, after [12]. Numerals are depths of the Moho discontinuity. (

1

) DSS lines 1984–1989 on land; (

2

) DDS lines on land andpilot PISCO 1994 passive seismic experiments; (

3

) marine and on-land DSS lines and CINCA 1995 passive seismic experiments;(

4

) ANCORP-96 reflected-wave line and passive seismic experiments 1996; (

5

) PUNA 1997 passive seismic experiments;(

6

) BANJO-SEDA 1997 passive seismic experiments; (

7

) APVC 1999 passive seismic experiments.

30

°

20

°

10

°

S

20

°

S

Page 4: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

308

GEOTECTONICS

Vol. 43

No. 4

2009

ROMANYUK

domains, which are characterized by slightly deformedbasement and thin-skinned tectonics of sedimentarycover.

The central Andes, as one of the most extremeregions of the Earth, attract special attention of geolo-gists and geophysicits and are studied actively. In par-ticular, important international projects have beenimplemented here in the most recent decades: PISCO1994, LITHOSCOPE 1994, BANJO-SEDA 1995,CINCA 1995, ANCORP 1996, PUNA 1997, APVC1999, ISSA 2000, ECCO 2001, etc. These projects haveenhanced the knowledge of this region, especially itscentral sector (Fig. 2b) to a quite new level. For exam-ple, the seismic structure of the continental crust hasbeen examined by deep seismic sounding [79, 91, 105],seismic reflection methods [7–9, 109], converted waves[21, 108], seismic tomography with P-waves [12, 43,47, 92], and with both P- and S-waves [32, 75, 97].Detailed gravity measurements [41, 64] and magneto-telluric sounding [22, 34, 82] have been carried out.Integrating works on the thermal regime [48, 50, 95],geology and tectonics [6, 10, 16, 17, 19, 40, 44, 47, 49,51, 54, 55, 59–62, 65, 66, 68, 70, 72, 77, 90, 93, 94,101–103], petrology, and geochemistry [13, 31, 36, 56, 57]have been published. As a result, the central segment ofthe Andean subduction zone has become one of the beststudied regions of the Earth with respect to its deepstructure. The Central Andean Transect along

~21°

S,the best studied geotraverse, has been comprehensivelyexamined by a complex of methods.

The model of the Late Cenozoic geodynamic evolutionof the central Andes and the complex tectonic, geological,and geophysical model of the Central Andean Transectpresented in this paper are based on an extensive body ofcomprehensive information and demonstrate the struc-tural features of the crust and upper mantle in the centralsegment of the Andean subduction zone.

THE LATE CENOZOIC TECTONICS OF THE CENTRAL ANDES

As early as in the 1980s, it became clear that the lateOligocene is a special moment in the evolution of theAndean subduction zone [54, 94]. Approximately27 Ma ago, the Nazca and Cocos plates separated fromthe Farallon Plate and began their own movement withsharply increasing rate of convergence between theNazca and South American plates [78]. Further, duringthe Neogene and Quaternary, the geodynamic settingsalong the Pacific margin of South America underwentsubstantial rearrangement, including change of the sub-duction angle of the Nazca Plate in different segments;enhanced lateral shortening of the central Andes; acti-vation of magmatic activity (Fig. 3), etc.

At least three stages are recognized in the tectonicevolution of the central Andes (Fig. 4). At the first stage(27–19 Ma ago), thrusting and faulting of both westernand eastern vergence occurred in the Eastern Cordil-

lera, and the Altiplano–Puma Plateau began slowly ris-ing. During the second and the third stages, the forma-tion of westward-verging folds practically ceased, andthe eastern front of deformation started to migrateinland; 10 Ma ago it reached the present-day IntraAn-dean domain and 5 Ma ago, the Subandean domain.The third stage, which began about 10 Ma ago, is char-acterized by an abrupt increase in the intensity of ign-imbrite magmatism at the Altiplano–Puma Plateau(Fig. 3) and by uplift of the entire orogen continuinguntil now. No less than 2 km of uplift of the centralAndes pertain to the last 10 Ma [10, 35, 37, 44, 54].

Post-Oligocene tectonic deformations were accom-panied by magmatic activity, which at the initialmoment of activation covered the territory from thepresent-day Western Cordillera to the western bound-ary of the Eastern Cordillera. In particular, one of theworld’s largest Frailes ignimbrite complexes, dated at7 Ma, was located at the boundary between Altiplanoand the Eastern Cordillera. Beginning from 5 Ma ago,the magmatic activity receded to the west; it is currentlylocalized only in the Western Cordillera [13, 56, 57].

High-resolution seismic tomography [12, 19, 43, 92]and geochemical data [13, 56, 57] testify to possibledelamination of the continental lithosphere beneath theAltiplano–Puna Plateau in the central Andes. Togetherwith GPS evidence for slow migration of the Andesinland to the South American Platform, this implies theactive role of the continental lithosphere in the LateCenozoic activation of the central Andes.

THE CENTRAL ANDEAN TRANSECT: A COMPLEX MODEL

Initial Data

The composite tectonic, geological, and geophysi-cal model of the Central Andean Transect is shown inFig. 5. The structure of the continental crust is deducedfrom the DDS section [105]. The details of the structurein the coastal zone, in particular the thickness of low-velocity rocks on the shelf and the oceanic Moho dis-continuity, are given after [79]. The velocities in themantle wedge are shown according to [12, 43, 91]. TheBenioff zone down to a depth of 670 m (Fig. 2a)beneath the margin of the South American continent iscontoured from the hypocenters of strong earthquakes[23, 104] determined from the global seismic network.The surface of the subducted Nazca Plate near the Cen-tral Andean Transect within a depth interval down to50 km is mapped from the high-resolution determina-tions of hypocenters of weak earthquakes recorded bya local network of stations (Fig. 2b); within the depthinterval of 50–100 km it is mapped from processing ofreflected wave records [8, 9]; and below 100 km it ismapped from converted waves [21, 108]. The zones of ele-vated electric conductivity beneath the PrecordilleranFault Zone and the Western Cordillera are contoured fromthe results of magnetotelluric sounding [22, 34]. The

Page 5: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS

Vol. 43

No. 4

2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 309

20

°

21

°

22

°

23

°

24

°

25

°

26

°

70

°

W

69

°

68

°

67

°

66

°

65

°

64

°

S

Oxaya

Chair Kkollubasaltic

28–17 Ma

Bolivia

Rondal

Chile Argentina

Basalticfields

flows

Shoshonites

Queva

Aquilar

Aqua Dulce

La Coipa

Segerstrom

Altosde Pica

(a)

20

°

21

°

22

°

23

°

24

°

25

°

26

°

70

°

W 69

°

68° 67° 66° 65° 64°

S

17–12 Ma

(b)

20°

21°

22°

23°

24°

25°

26°

70° W 69° 68° 67° 66° 65° 64°

S

12–3 Ma

(c)

20°

21°

22°

23°

24°

25°

26°

70° W 69° 68° 67° 66° 65° 64°

S

3–0 Ma

(d)

CerroRico

Tasna

ChocayaMorokho

Bonete

Smalldomes >13

Queva

AguaEscondida

Galan

ValleAncho

19°

Morococola 7

Frailes 7

Pastos Grandes 3.6–8

Panizos 7

Guacha 4.8 Vilama Cerito 7–8.5Corarzull 6–7Rachaite>7

Pairique 11La Pacana 4–5.5

Queva 7–10V. AntofallaGroup 7–11

Coplapo

basaltic <7Galan4–6

flows

7–11 Laguna VerdeGroup 3–4

19°

The main present-day

volcanic zone

Shoshonites

AltiplanoPlateau

PunaPlateau

Alkalibasalts F

ol

d-

na

pp

eb

el

t

of

Ea

st e

r nC

or d

i ll e

r aShoshonites

Alkalibasalts

Vol

cani

c ar

c of

Wes

tern

Co

rdil

lera

Solo

ChaoTuzgle

Cerro Gale 2.2–2.7

Fig. 3. The largest dated magmatic fields (inscriptions) and centers (symbols), after [57]. Calderas filled with ignimbrites are des-ignated by circles and stratovolcanoes by squares. The size of the symbols is approximately proportional to the size of the magmaticsystems; age, Ma is shown near the names. The open circles in panel (c) indicate events younger than 7 Ma; the filled circles indicateevents older than 7 Ma. APVC is the Altiplano–Puna Volcanic Complex (Panizos–Palrique).

structure of the uppermost crust in the central Andes cor-responds to the integral geological section [73].

The morphology and depth of the PrecoldilleranFault are shown in the style of [109] for 33° S. Regions

of strong seismic reflectors, visible as bright spots inrecord-section presentations [9], were revealed beneaththe Western Cordillera and at the base of the Uni–Ken-wani and San Vicente folds; the Quebrada Blanca spot

agoago

agoago

Kari Kari

Page 6: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

310

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

00.01

0

50

A ~40–19 Ma

SA Platform

ç2O?

0

50

km

km

~19 Ma

~19–10 MaåÄí

åÄF 10–5 Ma

0

50

0

50 km

WesternCordillera

Altiplano-Puna Plateau

EasternCordillera

IntraAndean Domain

Subandean Domain

SUT 5–0 Ma0

50 km

100

Flow perpendicular tofigure plane

200 400 600

0.1

1

10Paleoflora

11 Ma

Paleoflora17.5 Ma Uplifted

peneplanesurfaces 11 Ma Inversion

of uplift rateimplying

homogeneouscrustal

shortening

B

a

b

Uplift

mm/yr

ForelandAltiplanoEastern Cordillera Subandes

MohoR i g i d p l a t eV i s c o u s f l o w

Moho Continental lithosphere

50 km

H2O?Suggested

fluidimpact

attenuatingcontinentallithosphere

Domain ofmagmatic activity

Rate ofuplift

Magmauplift

Normal and thrust faults

‡ bLower

crust of SA Platforma) mafic and

b) eclogitized or undergone other

high-pressuretransformations

Fragments of delaminated

lower crust andlithospheric

mantle

Western Cordillera

Altiplano–Puna Plateau

EasternCordillera

IntraAndeanDomain

SubandeanDomain

km

I

I

II

III

III

rate,

Page 7: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 311

(QBBS) is the brightest. It is important that the Mohodiscontinuity is traced along the DSS line only frag-mentarily and is lost in some segments. Another appre-ciable feature is the considerable discrepancy in theposition of the Moho discontinuity determined fromrefracted and converted waves.

The slope and position of the Subandean ThrustFault in the upper crust were estimated from the resultsof a seismic survey [7]. Both the seismic reflectionmethod [4] and magnetotelluric sounding [82] testify tothe existence of a specific weakened zone at the exten-sion of the Subandean Thrust Fault, which plungesbeneath the region of thin-skinned tectonics in the Sub-andes. This zone is described in [70] as the bottom of afold complex.

The crustal structure in the zone transitional fromthe Subandes to the Chaco Plain is poorly studied byseismic survey and therefore is approximated by arbi-trary horizontal layers. The topography of the sedimen-tary basin bottom close to the Subandean Thrust Zonewas estimated from a density model [87].

The oceanic crust of the subducted Nazca Plate, 30–40 Ma in age, is approximated by basaltic layer 2C andgabbroic layer 3. The seismic velocities and densities inboth layers increase with depth. It is suggested thatlayer 2 pinches out by a depth of 20 km because thepores in the basalt are closed at a high pressure so thatthe density of the basalt approaches the density of thegabbro [24]. No data are available on the depth of theasthenosphere under the Nazca Plate. The thickness ofthis plate is accepted in correspondence to its age andtemperature [25].

Petrological Estimates for Some Deep Elementsof the Model

Four different P–T paths are recognized along theCentral Andean Transect: (1) the path of the oceaniclithosphere, about 40 Ma in age; (2) the path of layer 3(gabbroic layer) of the oceanic crust, pertaining to thesubducted Nazca Plate; (3) the path beneath the West-ern Cordillera and Altiplano Plateau; and (4) the path ofthe Precambrian South American continental platform(Fig. 6a). These paths are plotted on a diagram with thestability fields of mineral assemblages in the metabasicrocks (Fig. 6b); the lines of dry and wet solidus of peri-dotite, quartz–coesite transition, etc., are added. Thisplot makes it possible to estimate the state and densityof the matter at a depth. It should be noted that the fieldsof metamorphic facies of basic rocks are not applicable to

the upper and middle continental crust (paths 3 and 4),which is mainly sialic in composition.

Path 1: the oceanic lithosphere. Because data onthe seismic structure of the lithosphere of the NazcaPlate are not available, the calculated temperature dis-tribution [95] was used for its subdivision into layers.The stationary path (1) for the oceanic lithosphere,40 Ma in age, predicts a temperature of 700°ë at thebottom of the apparently elastic part of the plate at adepth of 50 km. The wet melting of peridotite at a depthof 70 km occurs at a temperature of 900°ë, while drymelting in the asthenosphere at a depth of 100 kmrequires 1300–1400°ë.

Path 2: the subducted oceanic crust (gabbroiclayer). According to [33], the Andean subduction zoneis cold. In the oceanic crust of the Nazca Plate, the gab-bro–eclogite transition is possible at a depth of 120 km.This estimate is consistent with the results of the studyof the Andean subduction zone using converted waves[108], which indicates that the oceanic crust is tracedhere to a depth of 120 km.

Path 3: the Western Cordillera and Altiplano.The temperature of the middle and lower crust beneaththe Altiplano Plateau remains uncertain; only the twoextreme variants are shown in Fig. 6b by dashed lines.The high-temperature path calculated in [95] corre-sponds to the stationary model. In this case, the temper-ature at the Moho discontinuity is estimated at~1200°ë. The low-temperature path corresponds to themiddle and upper crust, created over the last 27 Ma,mainly at the expense of the supply of cold material ofthe South American Platform. If this was the case, thetemperature at a depth of 70 km (Moho discontinuity)should be not higher than 700°ë and the P–T condi-tions at depth of 70 km (point T shown in Fig. 6b withaccount of 4-km height of mountains) should corre-spond to quartz–coesite transition. Complex seismicsurveying shows that the middle and lower crustbeneath the Altiplano is characterized by low andmedium velocities with the Poisson ratio varying fromlow to normal values. The crust is felsic or intermediatein composition [97] with a temperature of about 800°Cat the base. This estimate is consistent with a low(<10%) magmatic addition to the crust and the largelytectonic nature of its thickening. The temperature of thelower crust and the upper mantle fits the conditions ofgranulite-facies metamorphism (Fig. 6b) and wet par-tial melting of felsic rocks.

Fig. 4. The Late Cenozoic tectonic evolution of the Central Andes: (A) Tectonic horizontal shortening and vertical thickening of thecrust along the Central Andean Transect. The Roman numerals are stage numbers; see Fig. 5 for the abbreviations of faults. Ele-ments of the crust are shown by patterns and gray color; (B) suggested styles of crustal deformation explaining variable rates ofcrustal shortening and uplift in different domains of the central Andes, after [66]: (a) gray domains correspond to the rate of upliftingcalculated from the inverted rate of deformation on the assumption of homogeneous crustal thickening; over the last 10 Ma, theAltiplano and the Eastern Cordillera have experienced significant uplift with minimal horizontal shortening; (b) the South AmericanPlatform is thrust as a rigid block under the Subandes and the Eastern Cordillera; further, beneath the Altiplano, the lower crust andthe upper mantle underwent viscoplastic deformation. See Fig. 5 for other notations.

Page 8: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

312

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

5.6

5.9

5.6–

6.0

T

he o

bser

ved

grav

ity f

ield

(B

ougu

er a

nom

alie

s, m

Gal

; pr

ojec

tion

of th

e da

ta p

oint

s lo

cate

d w

ithin

the

trac

t 100

-km

wid

e al

ong

the

tran

sect

) M

odel

gra

vity

fie

ld

The

den

sity

of

the

cons

olid

ated

con

tinen

tal c

rust

, g/c

m3 a

vera

ged

in th

e ve

rtic

al d

irec

tion

(The

sam

e ex

cept

the

crus

t–m

antle

tran

sitio

n zo

ne)

(Onl

y fo

r th

e SA

Cra

ton)

3.0

2.9

2.8

–100

–200

Wes

tern

Cor

dille

raA

ltipl

ano

Plat

eau

Eas

tern

Cor

dille

ra

Intr

aAnd

ean

Dom

ain

Dom

ain

Cha

co P

lain

Sub-

ande

an

PCW

F

QB

BS

HV

ZL

VZ

UK

SVT

C M

AT

MA

FSU

TD

ista

nce,

km

200

300

500

900

Naz

caPl

ate 7,2

~ 7

cm/y

r

700°

C

900°

C

1250

°C

Oce

anic

asth

enos

pher

e

Ver

tical

sca

le h

oriz

onta

l sca

le =

2 :

1

4,5

5.0

8.1

3.5 6.

6?Pa

leoM

oho

6.8

8.1

7.5

H2O

7.0

7.5 8.0

Q =

250

Moh

o(c

onve

rted

wav

es)

7.2

7.2

6.6

–7.

2

6.0

8.7

5.0

–6 5.9(

?)

6.4

6.1

6.8

6.2

6.9

6.1

Moh

o (r

efra

ctio

n w

aves

The

cru

st o

f SA

Cra

ton

5.1

Bas

emen

t of

plat

form

Hig

h-ve

loci

ty d

omai

n w

ithV

p >

8.4

km

/s d

etec

ted

by

Depth, km

~ 1

cm/y

r

700°

C

900°

C

1250

°C

500°

C6.

7

8.2

6.2

5.9

8.1

100

200

seis

mic

tom

ogra

phy

Page 9: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 313

Path 4: the South American Platform. The calcu-lated temperature distribution [95] was used for subdi-vision of the continental mantle into layers. Stationarypath 4 for the continental lithosphere estimates the tem-perature at the bottom of the apparently elastic plate at700°ë (a depth of 70 km); the wet melting of peridotiteat a depth of 120 km requires 900°ë; at a depth of 200 km,the temperature reaches 1250°ë. The bottom of thecrust beneath the South American Platform is located ata depth of 40 km (temperature is estimated at ~500°ë)(Fig. 6b). For the onset of eclogitization or other high-pressure transformation of the lower crustal basicrocks, downgoing of the lower crust by 10 km or heat-ing by 100°ë is sufficient.

A Density Model

The gravity field in Bouguer reduction above thecentral Andes is characterized by a deep and flat mini-mum more than 1000 km in width (Fig. 5). Preliminarydensity models [41, 42, 45, 87] and additional calcula-tions for a model with constant densities in the crustallayers of the South American Platform and the oceaniccrust of the subducted Nazca Plate revealed a defi-ciency of almost 200 mGal and thus showed a high den-sity of the mantle beneath the Andes. Therefore, an ulti-mate density model has been constructed, with a down-ward increase in density of the layers in moving plates,taking the effect of metamorphism into account (Fig. 6b).The relationships of density versus depth in the layersselected in Fig. 6a are shown in Fig. 6c.

The abrupt changes of average density of the con-solidated crust from the Coastal Zone to the WesternCordillera and from the Subandean Zone to theIntraAndean Zone are shown in Fig. 5. The first featuredepends on the changed structure and temperatureregime of the crust, and the second one marks theboundary between the doubled and weighted crust ofthe South American Platform and the lightened uppercrust of the Andean mountain chain. The deep-seatedpositive density anomaly beneath the Altiplano isrelated to the crust–mantle transitional zone.

MANIFESTATION OF ANCIENT AND RECENT GEODYNAMIC PROCESSES IN THE PRESENT-

DAY LITHOSPHERE ALONG THE CENTRAL ANDEAN TRANSECT

The subducted Nasca oceanic plate and mantlewedge. The Nazca oceanic plate, being subductedbeneath the continental margin, undergoes dehydrationin the upper part of the plate and eclogitization of basaltand gabbro in the oceanic crust [29, 33, 39, 46, 80, 81,83, 96]. The fluid released from the slab gives rise toserpentinization of peridotite in the mantle wedge andits wet melting, leading to a decrease in seismic veloc-ities and reducing the Q-factor of the wedge [12, 92].

The upper crust of the central Andes. The uppercrust is a layer bounded from below by a high-velocitylayer (HVL) at a depth of ~20 km (Fig. 5). The uppercrustal layer is a lens that thickens to 20–25 km beneaththe mountains and pinches out near the shore and theSubandean Thrust Fault. This is a relict sedimentarybasin of the marginal continental type, formed in thePaleozoic at the present-day western periphery of theancient core of the continent of South America ornearby [17, 27]. The HVL that separates the upper crustof the Andes from the lower crust may be interpreted asa relic of the basified continental crust and/or oceaniccrust of the former marginal sea. The bottom of theHVL is treated as a paleoMoho [40].

The coastal zone. Presumably in the Late Jurassic,the volcanic island arc was thrust under the backarcsedimentary basin from the side of ocean [85], and thevolcanosedimentary crust between the coast andpresent-day volcanic arc of the Western Cordillera wasdoubled. The cold crust is characterized here byinclined seismic boundaries and high average seismicvelocity and density.

The Western Cordillera is a volcanic belt withspectacular post-Miocene volcanic activity. The fluidreleased from the subducted Nazca Plate disturbed thegeochemical equilibrium of the overlying rocks andprovoked wet melting of peridotite in the mantlewedge, producing suprasubduction magmatism.

The Altiplano Plateau is a high-standing (>3.7 km)slightly differentiated block. It is a basin filled with Ter-tiary and Quaternary sedimentary and volcanic rocks

Fig. 5. The integrated tectonic, geological, and geophysical model along the Central Andean Transect.See Fig. 2 for transect location. The relatively dark color corresponds to elevated seismic velocities and densities; the scales of tonesfor the crust and the mantle are different. Numerals are seismic velocities, km/s. The heavy solid lines in the upper crust are regionalwrench and thrust faults: PC, Precordilleran Fault Zone; WF, Western Fissure; UK, Uni–Kenwani Fault Zone; SV, San Vicente FaultZone; T, Tupiza Thrust Fault; C, Camargo Fault Zone; MAT, Main Andean Thrust Fault; MAF, Main Frontal Thrust Fault; SUT,Subandean Thrust Fault. The structure of the uppermost crust of the central Andes corresponds to the integrated geological section,after [73]. The black spot in the crust indicates the approximate position of bright reflectors (QBBS) in reflection record sections [9].HVZ is the high-velocity layer from DSS data at a depth of 20 km; its bottom is interpreted as a paleoMoho [40]. LVZ is the low-velocity zone of converted waves beneath the Altiplano Plateau interpreted as a domain of magmatic and fluid activity [108]. Thegray dots are zones of elevated electric conductivity contoured by magnetotelluric sounding [22]. The isotherms are shown after[95]; the isotherm of 900°C beneath the Altiplano is corrected in agreement with estimates in [97]. The domain marked by blackdots with the inscription Q = 250 corresponds to the reduced seismic quality, after [92]. The high-velocity (Vp > 8.4 km/s) domainin the continental lithosphere beneath the Altiplano (dark color) was detected by seismic tomography [43].

Page 10: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

314

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

~Depth, km

200

400

600

800

1000

1200

12345

1400

8

(a)

1

Bot

tom

of

appa

rent

ly e

last

icpo

rtio

n of

pla

te (

~700

°C)

12

2

33

4

4–7

Ecl

ogiti

zatio

n of

low

er c

rust

al

maf

ic r

ocks

Lith

osph

ere–

asth

enos

pher

e tr

ansi

tion

(700

–125

0°C

)

Con

solid

ated

cru

stof

the

Sout

h A

mer

ican

Cra

ton

Pressure, GPa

(b)

306090120

Gab

bro

inth

e su

bduc

ted

ocea

nic

crus

t

Sou

thA

mer

ican

Cra

ton

1

2

3

4

Wet

per

idot

ite

solid

us

Wes

tern

Cor

dille

raan

d A

ltipl

ano

Plat

eau

3

Dry

peri

dotit

e so

lidus

Tem

pera

ture

, °C

Qua

rtz–

coes

itetr

anst

ion

Con

tinen

tal M

oho

Oce

anic

Oce

anic

M

oho

Ec

LB

EB

Am

Gr

Gs

EA

Met

amor

phic

fac

ies:

Gs

– gr

eens

chis

tE

B –

epi

dote

-blu

esch

ist

LB

– la

wso

nite

-blu

esch

ist

EA

– e

pido

te-a

mph

ibol

iteA

m –

am

phib

olite

Ec

– ec

logi

teG

r –

gran

ulite

Den

sity

, g/c

m350

100

150

200

250

3.6

3.5

3.4

3.3

3.2

3.1

3.0

2.9

2.8

2.7

2.6

300

Low

er c

rust

alec

logi

te o

fth

e So

uth

Am

eric

anSu

bduc

ted

slab

(N

azca

Pla

te)

Subd

ucte

d lit

hosp

here

of

the

Sout

h A

mer

ican

Cra

ton

Ecl

ogiti

zatio

n of

ocea

nic

crus

t

Subd

ucte

d oc

eani

ccr

ust o

fth

e N

azca

Plat

e (g

abbr

o)

Low

er c

rust

of

the

Sout

hA

mer

ican

Cra

ton

Mid

dle

crus

t of

the

Sout

hA

mer

ican

Cra

ton

Upp

er c

rust

of

the

Sout

hA

mer

ican

Cra

ton

8

2

1

7

6

5

4

1

2

3

Dep

th, k

m

litho

sphe

re

Cra

ton

(c)

T

Page 11: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 315

and located between the still higher Western and East-ern Cordilleras [68].

The middle and lower crust beneath the WesternCordillera and Altiplano Plateau. The seismic surveydemonstrates very low velocity throughout the crustand the absence of HLV beneath the Western Cordilleraand Altiplano. The seismic survey did not reveal dis-tinct boundaries within the middle and lower crust,which is devoid of layered structure beneath the West-ern Cordillera and Altiplano. Because the electric resis-tance and seismic Q-factor sharply decreased in thisblock [12], it may be suggested that the present-day orrecent magmatic and/or fluid activity was predominanthere. This process obliterated the preceding structuralgrain. The crustal and upper mantle rocks are intrudedby intrusive bodies and metamorphosed. In particular, aspecial study using converted waves [26] established alow-velocity zone (Vs < 0.5 km/s) interpreted as a sill-like magmatic body 750–810 m thick beneath theAPVC volcanic complex (see Fig. 3 for the complexlocation) at a depth of 19 km.

As concerns the bright seismic reflectors in theupper crust (QBBS is the brightest spot, see Fig. 5), ithas been shown that these phenomena are recorded inthe domains of intense tectonic deformation andlocated in the transitional zone from elastic deforma-tion and brittle failure to the viscoplastic flow. Com-monly, these are root zones of newly formed activefaults accompanied by high seismicity [89]. In domainsof intense deformation, rocks accumulate elastic energy(a future earthquake source). When such domains aretransmitted by seismic waves, the medium irradiates apart of the stored energy due to nonlinear effects; this isexpressed in the recorded sections of reflected waves asarrivals with increased amplitudes, i.e., bright spots,which mark the strained state of the rocks and probablytheir saturation with fluids or magmatic melt. However,it cannot be ruled out that both processes are interre-lated.

The Moho discontinuity beneath the WesternCordillera and Altiplano Plateau. No sharp boundarybetween the crust and the mantle is detected beneath theWestern Cordillera and Altiplano. In the records ofrefracted seismic waves, Pn arrivals are almost not dis-cernible. Beneath the Altiplano, only two short reflec-tors are noted at a depth of about 70 km [105]. Theirposition deviates rather strongly from the position ofthe receiver function Moho at a depth of 60 km [107,108]. The isostatic Moho beneath the central Andes is

estimated at 66–72 km [42]. All estimates are close tothe maximum possible thickness of the crust controlledby the depth limit of quartz stability (65–75 km)deduced from quartz–coesite transition [28]. It is mostlikely that at this depth the crust–mantle boundary is agradient zone 10–15 km thick, transitional from largelycrustal to predominantly mantle rocks.

The Eastern Cordillera, IntraAndian, and Sub-andian zones are a fold-nappe belt mainly composedof Paleozoic and Mesozoic sedimentary rocks. The old-est rocks now exposed in the Eastern Cordillera are theOrdovician metasedimentary complexes sporadicallyoverlapped by Cretaceous and Tertiary sediments. Sil-urian to Triassic rocks crop out in the IntraAndeanZone, while Carboniferous to Pliocene rocks areexposed in the Subandean Zone.

Currently, this fold–nappe belt is amagmatic; how-ever, igneous rocks related to the initial stage of Andeantectogenesis are identified in the boundary zonebetween Altiplano and the Eastern Cordillera. Thismagmatic episode is not related directly to supra-subduction magmatism, and to a great extent is a prod-uct of melting of the upper 10–15 km of the crustcomposed of specific granitic series [74]. In particu-lar, S-granites derived from metasedimentary sourceshave been identified here.

The South American Platform moves as a rigidplate beneath the Subandes. Zones of plastic and vis-cous deformations appear at a depth, and the beds arethickening due to the folding. Mainly viscous and plas-tic deformations occur beneath the Altiplano [66].

Metamorphism in the subducted continental crustgradually increases the density and seismic velocity. Inthe lower crust, high-pressure metamorphism (Fig. 6)increases the density of rocks up to the mantle values.Because of this, the refractive Moho discontinuity islost beneath the Eastern Cordillera and the structuralbottom of the subducted crust of the South AmericanPlatform does not coincide with the currently recordedrefractive seismic M surface, being located at a greaterdepth between the Altiplano and the Eastern Cordillera(Fig. 5).

The continental upper mantle. The tectonic short-ening of the crust in the central Andes implies a corre-sponding rearrangement in the continental lithosphere.As follows from the geochemical data, the lithosphereof the South American Platform is traceable westwardnot farther than to 65–66° W [3]. The seismic data and

Fig. 6. Composition and density of the lithosphere in the Central Andes: (a) schematic section along the Central Andean Transectand location of P–T paths (numerals in circles correspond to the numerals in panel (b); numerals in squares correspond to the numer-als in panel (c)); (b) P–T stability fields of mineral assemblages pertaining to most important metamorphic facies (basaltic protolith),after [80]; dry and wet peridotite solidus lines and line of quartz–coesite equilibrium, after [28]; P–T paths (numerals in circles):(1) oceanic lithosphere, (2) gabbroic layer of the subducted oceanic crust, (3) lithosphere beneath the Western Cordillera and theAltiplano Plateau, (4) continental lithosphere of the South American Platform; the horizontal dashes on the paths mark the Mohodiscontinuity; (c) density versus depth relationships obtained from inversion of gravity field in some segments of the Central AndeanTransect; each symbol on the plot corresponds to one block of the model; the patterns of the symbols match the patterns of layersin panel (a); the proved and inferred relationships are shown by solid and dashed lines, respectively.

Page 12: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

316

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

Shoreline

(a) 22°S

24

SubductedNazca Plate

Crustalmagma chambers

Crust–mantletransition zone

Coastal Cordillera

Cordillera

Western

Lower crustaleclogite bodies

Sinking of lithosphericfragments into the mantle Calderas

Puna Plateau

EasternSubandes

Cordillera

70° W 68 66

(b)

AtacamaUKF

1 2 Lipez

3 45

67

8

9

101112

13

AF

14/15

Arizaro

Santa Maria

IF

16

17 18

20

TTF19 Pipanaco

La Riojia

1

2

68° W 66° 64°

22°S

23°

24°

26°

27°

28°

29°

(c)18°

50 40 30 20 10 S

20°22°24°26°28°30°

Latitude

Age, Ma

26km/Ma

Corque (6–14 km)Lipez (>4 km)

Atacama (3.7 km)Antofalla (1.3 km)Hombre Muerto (5 km)

Santa Maria (2–5 kmRipanaco (0.5 km)

La Riojia (0–0.4 km)

Fig. 7. Neogene deformation of the Puna Plateau andthe southern part of the central Andes, after [86].(a) Meridional extension of the upper crust in the PunaPlateau increased 10 Ma ago by reaching the criticalthickness of the crust. Arrows indicate general direc-tions of shortening and elongation in the brittle uppercrust. Extension is realized in strike-slip offsets androtation of blocks in the rhombic fault system.Calderas filled with ignimbrites are controlled by thefaults formed in the transtensional regime. Delami-nated fragments of the lower crust and lithosphericmantle of the South American Platform sank into themantle. (b) The present-day rhombic fault system ofthe Puna Plateau. The near-meridional left-lateralfaults parallel to the general Andean trend are black;the NE-trending right-lateral faults are gray. The faultsystems make up rhombic deformation domains,including sedimentary basins (dotted). The mecha-nisms of earthquakes are shown in the projection onthe lower hemisphere; the dots correspond to the direc-tion of shortening. The main faults (letters in figure):UKF, Uni–Kenwani Fault; AF, Acazoque Fault; TTF,Tucuman Transform Fault; IF, Iconza Transform Fault.Volcanic calderas (numerals in figures; age, Ma isshown in parentheses): 1, Pastos Grandes (8.1–5.4);2, Panizos (7.9–6.7); 3, Coruto (6.2); 4, Vilama Cerito(8.9–8.5); 5, Pairique (11.2); 6, Guacha (4.1); 7, Purico(1.3); 8, La Pacana (5.8–2.4); 9. Coranzuli (6.6);10, Ramadas (8.8–8.5); 11, Aguas Calientes (10.0–10.5); 12, Negra Muerta (7.4); 13, Galan (2.2);14, Wheelright (8.8–6.1); 15, Laguna Escondida (8.8–6.1);16, Cerro Bayo (8.5); 17, Mulas Muertas (6.1); 18, SanFrancisco; 19, Farallon Negro (12.6–8.6); 20, Incapillo(2.9–1.1); (1) direction of horizontal shortening;(2) direction of maximum extension close to calderas;(c) latitudinal variations of age, Ma and depth, km(numerals in parentheses) of separate rhombic basinsin the southern portion of the central Andes. The dashedline corresponds to the average rate of propagation of thesouthward deformation front over 26 km/Ma; the basindepth decreases to the south consistently with reducedcrustal shortening across the Andes.

Page 13: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 317

Santo DomingoSan Rafael

QuenamariSaritaCondoriquna

Palca XIFabulosa

MilluniChacaltaya

Rosario

Migration oore mineralization

with time

Plateau Altiplano

Large and superlargeore deposits(numerals are age, Ma)

APVCLarge ignimbritefields and their age, Ma

Chojila 19.5Kellhuani 21.3

VilocoCaracoles 24.9

ChicoteKami

Japo

Santa Fe–Morococola 20

Huanuni 20Llallague-

Catavi Colquechaca 22.6Cerro Rico

de Potosi 13.8Frailes

7 Caracota

Tatasi Tasna

Chorolque 16

Pulacaya 14 IscaiascaChilcobija

Chocaya 16

APVC11-7Loma Blanka 7

Pirquitas

Pun

a

El–Laco 5.3–1.6 Eastern C

ordillera

Pastos-Grandes 6.3

SerroGalle2–3

Plat

eau

Bolsa Negra

Oruno 16

Morococola

WesternCordillera

Central Andean Transect

7

Age of oremineralization

middle Mioceneearly Miocenelate OligoceneLate Triassic-Early Jurassic

16°

18°

20°

26° S

72° W 70° 68° 66°

Fig. 8. Centers of ore mineralization in the Bolivian tin belt migrating in time. Geological age is shown after [74]; the isotopic age,Ma of large deposits is after [2]; the location of large ignimbrite complexes corresponds to Fig. 3. The average rate of meridionalmigration to the south is 44 km/Ma.

in tin belt

Page 14: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

318

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

geochemistry of the igneous rocks testify to the delam-ination of fragments of the continental lithospherebeneath the Altiplano–Puna Plateau. The high-velocity(Vp > 8.4 km/s) domain in the mantle beneath Altiplanois interpreted as accumulation of high-density eclogiticinclusions or other high-pressure metamorphic rocks inthe downgoing continental lithosphere of the SouthAmerican Platform.

MECHAMISM AND DRIVING FORCES OF LATE CENOZOIC TECTONIC ACTIVATION

IN THE CENTRAL ANDES

According to [60], the tectonic horizontal shorten-ing of the crust in the backarc domain of the EasternCordillera and Subandes occurs along the entire west-ern coast of the South American continent and reachesa maximum of 230 km (250–290 km by other esti-mates) in the central Andes. The intense recent and cur-rent magmatic activity in the Altiplano–Puna Plateauand the Western Cordillera hampers estimation ofdeformation values. The estimates of crustal shorteningin the coastal zones at latitude 23° S were obtained onlyrecently, and the minimum total shortening in theBolivian Andes was estimated to be no less than 500 kmover 70 Ma [71, 72]. The kinematically balanced mod-els [15, 16, 30, 71, 72, 77, 90] show that these valuesare quite sufficient to explain the existing thickness ofthe crust.

Various geodynamic scenarios have been suggestedto explain the thickened crust in the central Andes. Thesubdivision of the Late Cenozoic evolution of the cen-tral Andes into stages was likely proposed for the firsttime in [54], when it was established that at the firststage the thickening occurred due to simple shearingdistributed more or less uniformly throughout the crust[5]. The regional thrust zone (Subandean Thrust Fault)was formed during the second stage. The further thick-ening was realized mainly by doubling caused bythrusting of the South American Platform under theupper crust of the Andes along the Subandean Fault.The subsequent models detailed and specified the prin-cipal two-stage model with involvement of the conti-nental lithosphere. The model of ablative two-sidedsubduction [84, 98] and models of crustal shortening inthe subduction zone, taking the climatic features of thecentral Andes [14, 15, 67] in account, should be noted.Although these models explain the crust thickening inqualitative terms, they cannot explain many docu-mented features of this process, in particular the lateOligocene jump of the locus of deformation far back ofthe subduction zone and the existence of the relativelyundeformed block of the Altiplano–Puna Plateaubetween the volcanic arc and fold–nappe belt; thechange of deformation from thick-skinned type at theinitial stages to the thin-skinned type at the late stages,etc. The idea of thermal attenuation of the continentallithosphere far back of the subduction zone as a primor-dial cause of the Late Cenozoic tectonic activation in

the central Andes was probably stated first in [54] andthen considered in detail in [103]. In terms of thismodel, the initial compression arises in the mantleabove the domain of the thermally attenuated lithos-phere. As the crust was thickening, the deformationlocus shifted from the mantle into the crust. When thecrustal thickness reaches a maximum possible value(the �65-km crust becomes gravitationally unstable),the zone of deformation starts to migrate toward a rela-tively thin crust and the foldbelt widens (inland migra-tion of deformation front).

However, the idea of thermal attenuation of the con-tinental lithosphere is poorly consistent with the lowmagmatic activity in the Eastern Cordillera in the lateOligocene. If the initial attenuation of the continentallithosphere was thermal, deformation should be moreintense in the Altiplano, where the Late Cenozoic mag-matism is intense and the present-day heat flow is high;the plateau, however, has remained little deformed untilnow. At the same time, the location of domains of thick-skinned deformation is in agreement with the geometryof the subducted slab. Tectonic reconstruction showsthat the Nazca Plate beneath the central Andes flattenedin the Oligocene and remained flat into the Miocene(Fig. 3a) just in the domains of tectonic activation. Inthe post-Miocene time, the flattened subduction zonemigrated southward and after that the zone of thick-skinned deformation shifted in the meridional direc-tion. At present, this zone embraced a wide region backof the subduction zone at 30–35° S (Fig. 1b).

The poor correlation of the localization of tectonicdeformation with the manifestation of magmatism andthe good correlation with the geometry of the sub-ducted slab show that abrupt local mechanical attenua-tion of the continental lithosphere originally was not ofthe thermal nature. The fluid effect upon the continentallithosphere seems to be the most probable. Either thelow-angle subducted Nazca Plate itself [46] or the man-tle regions disturbed by this plate might be sources offluid. For example, the lower layer of the upper mantle(400–670 km), where reservoirs of water-saturatedrocks are expected [20, 100], could be such regions.The impact of fluid on orogenic processes is suggestedin many provinces [11, 52].

The tectonic flows in the middle and lower crust [38,53, 106], along with delamination of the lower crustand the continental lithosphere [37, 56], were otherimportant processes in the evolution of the centralAndes. Indeed, all seismic studies show that a high-velocity layer (Vp = 7.0 km/s), presumably mafic incomposition, is located at the bottom of the crustbeneath the Eastern Cordillera. However, the lowvelocities of seismic waves beneath the Altiplano–Punaadmit only an insignificant contribution of mafic rocksto the crust, consistent with the idea of eclogitization ofthe lower crust in the course of underthrusting of theSouth American Platform. The eclogitic layer, detachedfrom the overlying rocks, sank into the mantle together

Page 15: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 319

with fragments of the underlying continental lithos-phere and therefore did not participate in the formationof the middle and lower crust beneath the Altiplano andPuna (Figs. 4, 7a). The collapse of eclogite into themantle most likely was caused by the gravitationalinstability of a heavy layer consisting of eclogitizedlower crust and the continental lithospheric mantle andlying on the relatively light lithospheric layer. This sce-nario is difficult to prove by direct arguments, but con-trasting variations of the Moho depth from 65 to 80 km(Fig. 2) and Pn seismic velocity immediately under theMoho discontinuity [19, 107] beneath the Altiplano–Puna Plateau and the western margin of the EasternCordillera may serve as indirect evidence.

The variation of the intensity and type of magma-tism in the volcanic arcs is commonly referred to thechange in the regime of the subduction zone. In partic-ular, the eruption of an enormous volume of ignim-brites 7–11 Ma ago is explained by variable geometryof the Benioff zone or plate kinematics, and the rotationof the upper crustal blocks in the Andean Orogenaround a vertical axis is ascribed to bending of the oro-gen (Andean Orocline) [54]. However, as has beenshown in [86], the collapse calderas filled with ignim-brites are genetically related to the faults that broke thePuna Plateau and the southern part of the Altiplano intorhombic segments (domains) partly occupied by sedi-mentary basins (Fig. 7b). The rhombic basins started toform in the Eocene and propagated to the south at a rateof ~26 km/Ma (Fig. 7c). Thus, at the initial stages, thehorizontal shortening across the strike of the orogenwas compensated by thickening of the crust. By reach-ing the maximum possible thickness at the onset of thethird stage, the meridional tectonic flows in the middleand lower crust beneath the Altiplano–Puna Plateaubecame more intense and brought about longitudinalextension of the orogen. Fault zones that developed inthe transtensional regime served as conduits for theascent of viscous silicic melts from the deep crust. Thedetailed 3D kinematic model of the central Andean seg-ment of the subduction zone, which takes into accountboth lateral and longitudinal movement of materail[106], predicts rotation of upper crustal blocks by 6–25°, in qualitative agreement with paleomagnetic andother data [18, 88].

The deposits of the Bolivian tin belt are related toplutonic and subvolcanic rocks, including S-granites,localized at the western margin of the Eastern Cordil-lera and in the Altiplano–Puna Plateau (Fig. 8). TheCenozoic deposits become younger to the south, con-sistent with meridional migration of the front of tec-tonic reworking of the middle and lower crust and theunderlying lithospheric mantle.

CONCLUSIONS

Summing up the available data, the following prin-cipal events should be pointed out in the Late Cenozoicgeodynamic evolution of the central Andes.

(1) In the late Eocene, the local rheological attenua-tion of the continental lithosphere, presumably causedby fluid impact, occurred in the present-day centralAndes far back of the subduction zone. The lateral com-pression of the subduction zone led to deformation ofthe lithosphere above the weakened domain. The locusof deformation thus migrated upward from the mantleinto the crust.

(2) Induced by tectonic pulse from a depth, tectonicdeformation in the Oligocene developed in thick-skinned style, creating the symmetric bivergent systemof the Eastern Cordillera. The deformation of the crustresulted in its thickening. At the initial stage, this pro-cess was not accompanied by fast uplift of crustalblocks and formation of high-mountain topography.Most likely, high-density lower crustal rocks held thecrust in the submerged state. Beneath the Altiplano, thecrust was thickened and heated due to magma ascent,which ensured the slow isostatic emergence of thisblock.

(3) Approximately 19 Ma ago, the gently dippingSubandean Thrust-Fault Zone arose beneath the East-ern Cordillera, and the lithosphere of the South Ameri-can Craton started to thrust under the Andes along thiszone. Over a short time, the thickness of the crustalmost doubled, giving rise to the isostatic uplift of theEastern Cordillera. The upper crust above the Suban-dean Thrust Fault experienced thin-skinned deforma-tion with the formation of east-verging folds andnappes. The deformation front rapidly migrated east-ward inland to the continent

(4) Approximately 10 Ma ago, the crust attained thegreatest possible thickness. Viscoplastic tectonic flowsin the middle and lower crust started to develop alongthe orogen axis, and a rhombic fault system was formedin the upper crust. The relative elongation in the uppercrust was realized in the rotation of blocks around thevertical axis and displacements along faults. Faulting inthe transtensional regime created conduits for magma,which erupted in the form of giant ignimbrite fields.

(5) Fragments of the lower crust and subcrustallithospheric mantle beneath the Eastern Cordillera andAltiplano–Puna Plateau were detached from the overly-ing rocks and sank into the mantle. This process hascontinued until now.

(6) The character of the Late Cenozoic magmatismof the Altiplano–Puna Plateau is determined largely bythermomechanical transformation of the lithosphere ofthe Andean Orogen and is only indirectly related to thekinematics of the subducted Nazca Plate.

ACKNOWLEDGMENTS

This study was partially supported by the RussianFoundation for Basic Research, project nos. 04-05-65092 and 07-05-00109.

Page 16: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

320

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

REFERENCES

1. T. V. Romanyuk and Yu. L. Rebetsky, “Density Inhomo-geneities, Tectonics, and Stresses in the 21° S AndeanSubduction Zone,” Fiz. Zemli 37 (2), 23–57 (2001) [Izv.Physics Solid Earth 37 (2), 120–140 (2001)].

2. D. V. Rundquist, A. V. Tkachev, Yu. Gatinsky, et al.,Large and Superlarge Ore Deposits, Vol. 1: Global Ten-dencies of Localization (IGEM RAS, Moscow, 2006)[in Russian].

3. S. Aitcheson, R. Harmon, S. Moorbath, et al., “Pb Iso-topes Define Basement Domains of the Altiplano, Cen-tral Andes,” Geology 23, 555–558 (1995).

4. R. W. Allmendinger, “Tectonic Development, South-eastern Border of the Puna Plateau, NorthwesternArgentine Andes,” Geol. Soc. Amer. Bull. 97 (9), 1070–1082 (1986).

5. R. W. Allmendinger and T. Gubbels, “Pure and SimpleShear Plateau Uplift, Altiplano-Puna, Argentina andBolivia,” Tectonophysics 259, 1–14 (1996).

6. R. W. Allmendinger, T. E. Jordan, M. S. Kay, andB. L. Isacks, “The Evolution of the Altiplano–Puna Pla-teau of the Central Andes,” Ann. Rev. Earth. Planet. Sci.25, 139–174 (1997).

7. R. W. Allmendinger and T. R. Zapata, “Imaging theAndean Structure of the Eastern Cordillera on Repro-cessed YPF Seismic Reflection Data,” in Proceedings ofXIII Congreso Geologico Argentino y III Congreso deExploracion de Hidrocarburos (Actas II, 1996),pp. 125–134.

8. “ANCORP Research Group. Seismic Reflection Imageof the Andean Subduction Zone Reveals Offset of Inter-mediate-Depth Seismicity into Oceanic Mantle,”Nature 349, 341–344 (1999).

9. “ANCORP Working Group. Seismic Imaging of a Con-vergent Continental Margin and Plateau in the CentralAndes (Andean Continental Research Project 1996(ANCORP-96)),” J. Geophys. Res., 108 (B7), 23–28(2003) doi: 10.1029/2002JB001771.

10. M. H. Anders, W. K. M. Gregory, and M. Spiegelman,“A Critical Evaluation of Late Tertiary AcceleratedUplift Rates for the Eastern Cordillera, Central Andesof Bolivia,” J. Geol. 110, 89–100 (2002).

11. E. V. Artyushkov, N.-A. Morner, and D. L. Tarling, “TheCause of Loss of Lithospheric Rigidity in Areas Farfrom Plate Tectonic Activity,” Geophys. J. Int. 143,752–776 (2000).

12. G. Asch, B. Schurr, S. Luth, et al., “Structure and Rhe-ology of the Upper Plate from Seismological Investiga-tions,” in Deformation Processes in the Andes. Interac-tion between Endogenic and Exogenic Processes dur-ing Subduction Orogenesis. Report for the ResearchPeriod 1999–2001 (Potsdam, 2001), pp. 177–204.

13. W. Avila-Salinas, “Petrological and Tectonic Evolutionof the Cenozoic Volcanism in Bolivian Western Andes,”in Andean Magmatism and Its Tectonic Setting, Ed. byR. S. Harmon and C. W. Rapela (Geol. Soc. Amer. Spec.Paper, 1991), pp. 245–258.

14. J. P. Avouac and E. V. Burov, “Erosion as a DrivingMechanism of Intracontinental Mountain Growth,”J. Geophys. Res. 101 (B8), 17 747–17 769 (1996).

15. A. Y. Babeyko and S. V. Sobolev, “Quantifying Differ-ent Modes of the Late Cenozoic Shortening in the Cen-tral Andes,” Geology 33 (8), 621–624 (2005).

16. P. Baby, Ph. Rochat, G. Mascle, and G. Herail, “Neo-gene Shortening Contribution to Crustal Thikening inthe Back Arc of the Central Andes,” Geology 25 (10),883–886 (1997).

17. H. Bahlburg and F. Herve, “Geodynamic Evolution andTectonostratigraphic Terranes of Northwestern Argen-tina and Northern Chile,” Geol Soc. Amer. Bull. 109 (7),869–884 (1997).

18. M. E. Beck, Jr., “On the Mechanism of Crustal BlockRotations in the Central Andes,” Tectonophysics 299,75–92 (1998).

19. S. L. Beck and G. Zandt, “The Nature of Orogenic Crustin the Central Andes,” J. Geophys. Res. 107 (B10), 22–30 (2002), doi: 10.1029/2000JB000124.

20. D. Bercovici and S.-I. Karato, “Whole-Mantle Convec-tion and the Transition-Zone Water Filter,” Nature 425,39–44 (2003).

21. G. Bock, B. Schurr, and G. Asch, “High-ResolutionImage of the Oceanic Moho in the Subducting NazcaPlate from P–S Converted Waves,” Geophys. Rev. Lett.27 (23), 3929–3932 (2000).

22. H. Brasse, P. Lezaeta, K. Schwalenberg, et al., “TheBolivian Altiplano Conductivity Anomaly,” J. Geophys.Res. 107 (5) (2002), doi: 10.1029/2001JB000391.

23. T. Cahill and B. Isacks, “Seismicity and Shape of theSubducted Nazca Plate,” J. Geopys. Res. 97 (B12), 17503–17 529 (1992).

24. R. L. Carlson and C. N. Herrick, “Densities and Poros-ities in the Oceanic Crust and Their Variations withDepth and Age,” J. Geophys. Res. 95, 9153–9170(1990).

25. R. L. Carlson and H. P. Johnson, “On Modeling theThermal Evolution of the Oceanic Upper Mantle: AnAssessment of the Cooling Plate Model,” J. Geophys.Res. 99 (B2), 3201–3214 (1994).

26. J. Chmielowski, G. Zandt, and C. Haberland, “The Cen-tral Andean Altiplano-Puna Magma Body,” Geophys.Rev. Lett. 26 (6), 783–786 (1999).

27. T. T. Cladouhos, R. W. Allmendinger, B. Coira, andE. Farrar, “Late Cenozoic Deformation in the CentralAndes: Fault Kinematics from the Northern Puna,Northwestern Argentina and Southwestern Bolivia,”J. South Amer. Earth Sci. 7 (2), 209–228 (1994).

28. M. Cloos, “Lithospheric Buoyancy and CollisionalOrogenesis: Subduction of Oceanic Plateaus, Continen-tal Margins, Island Arcs, Spreading Ridges, and Sea-mounts,” Geol. Soc. Amer. Bull. 105, 715–737 (1993).

29. J. H. Davies and D. J. Stevenson, “Physical Model ofSource Region of Subduction Zone Volcanics,” J. Geo-phys. Res. 97 (B2), 2037–2070 (1992).

30. P. G. de Celles and B. K. Horton, “Early to Middle Ter-tiary Foreland Basin Development and the History ofAndean Crustal Shortering in Bolivia,” Geol Soc. Amer.Bull. 115 (1), 58–77 (2003).

31. S. L. de Silva, “Altiplano-Puna Complex of the CentralAndes,” Geology 17, 1102–1106 (1989).

32. C. Dorbath and F. Masson, “Composition of the Crustand Upper-Mantle in the Central Andes (19–30° S)

Page 17: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 321

Inferred from P-Wave Velocity and Poisson’s Ratio,”Tectonophysics 327, 213–223 (2000).

33. T. A. Dumitru, “Effect of Subduction Parameters onGeothermal Gradients in Fore-Arcs, with an Applica-tion to Franciscan Subduction in California,” J. Geo-phys. Res. 96 (B1), 621–641 (1991).

34. F. Echternacht, S. Tauber, M. Eisel, et al., “Electromag-netic Study of the Active Continental Margin in North-ern Chile,” Phys. Earth Planet. Int. 102, 69–87 (1997).

35. H. Ege, E. R. Sobel, E. Scheuber, and V. Jacobshagen,“Exhumation History of the Southern Altiplano Plateau(Southern Bolivia) Constrained by Apatite FissionTrack Thermochronology,” Tectonics 26 (TC1004),(2007), doi: 10.1029/2005TC001869.

36. P. Francis and C. Hawkesworth, “Late Cenozoic Ratesof Magmatic Activity in the Central Andes and TheirRelationships to Continental Crust Formation andThickening,” J. Geol. Soc. London 151, 845–854(1994).

37. C. N. Garzione, P. Molnar, J. C. Libarkin, andB. J. MacFadden “Rapid Late Miocene Rise of theBolivian Altiplano: Evidence for Removal of MantleLithosphere,” Earth Planet. Sci. Lett. 241, 543–556(2006).

38. M. Gerbault and J. Martinod, “Possible Orogeny-Paral-lel Lower Crustal Flow and Thickening in the CentralAndes,” Tectonophysics 399, 59–72 (2005).

39. T. Gerya and D. A. Yuen, “Rayleigh–Taylor Instabilitiesfrom Hydration and Melting Propel Cold Plumes” atSubduction Zones, Earth Planet. Sci. Lett. 212, 47–62(2003).

40. P. Giese, E. Scheuber, F. Schilling, et al., “CrustalThickening Processes in the Central Andes and the Dif-ferent Natures of the Moho Discontinuity,” J. SouthAmer. Earth Sci. 12, 201–220 (1999).

41. H.-J. Goetze and A. Kirchner, “Gravity Field at theSouth American Active Margin (20 to 29° S),” J. SouthAmer. Earth Sci. 10 (2), 179–188 (1997).

42. H.-J. Goetze, B. Lahmeyer, S. Schmidt, and S. Strunk,“The Lithospheric Structure of the Central Andes (20–26° S) as Inferred from Interpretation of Regional Grav-ity,” in Tectonics of the Southern Central Andes, Ed. byK.-J. Reutter, E. Schenber, and P. J. Wigger (Springer,Berlin, 1994), pp. 7–21.

43. F. M. Graeber and G. Asch, “Three-Dimensional Mod-els of P-Wave Velocity and P- to S-Velocity Ratio in theCentral Andes by Simultaneous Inversion of LocalEarthquake Data,” J. Geophys. Res. 104 (B9), 20 237–20 256 (1999).

44. K. M. Gregory-Wodzicki, “Uplift History of the Centraland Northern Andes; a Review,” Geology 112, 1091–1105 (2000).

45. J. A. Grow and C. O. Bowin, “Evidence for High-Den-sity Crust and Mantle beneath the Chile Trench Due tothe Descending Lithosphere,” J. Geophys. Res. 80 (11),1449–1458 (1975).

46. M.-A. Gutcher, R. Maury, J.-P. Eissen, and E. Bourdon,“Can Slab Melting Be Caused by Flat Subduction?,”Geology 28 (6), 535–538 (2000).

47. M.-A. Gutcher, W. Spakman, H. Bijwaard, andE. R. Engdahl, “Geodynamics of Flat Subduction: Seis-

micity and Tomographic Constraints from the AndeanMargin,” Tectonics 19 (5), 814–833 (2000).

48. V. M. Hamza and M. Munoz, “Heat Flow Map of SouthAmerica,” Geothermics 25 (6), 599–646 (1996).

49. A. J. Hartley, G. May, G. Chong, et al., “Developmentof a Continental Forearc: A Cenozoic Example from theCentral Andes, Northern Chile,” Geology 28, 331– 334(2000).

50. S. G. Henry and H. N. Pollack, “Terrestrial Heat Flowabove the Andean Subduction Zone in Bolivia andPeru,” J. Geophys. Res. 93 (B12), 15 153–15 162(1988).

51. B. K. Horton, B. A. Hampton, and G. L. Waanders,“Paleogene Synorogenic Sedimentation in the AltiplanoPlateau and Implications for Initial Mountain Buildingin the Central Andes,” Geol. Soc. Amer. Bull. 113,1387–1400 (2001).

52. E. Humphreys, E. Hessler, K. Dueker, et al., “HowLaramide-Age Hydration of North American Litho-sphere by the Farallon Slab Controlled SubsequentActivity in the Western United States,” in The Lithos-phere of Western North America and Its GeophysicalCharacterization, Ed. by S. L. Klemperer and W. G. Ernst(Bellwether, 2003), George A. Thompson Volume Int.Book Series, Vol.7, pp. 524–544.

53. L. Husson and T. Sempere, “Thickening the AltiplanoCrust by Gravity-Driven Crustal Channel Flow,” Geo-phys. Rev. Lett. 30 (5), 1243–1246 (2003).

54. B. L. Isacks, “Uplift of the Central Andean Plateau andBending of the Bolivian Orocline,” J. Geophys. Res. 93,3211–3231 (1988).

55. T. E. Jordan and M. Gardeweg, “Tectonic Evolution ofthe Late Cenozoic Central Andes (20–30° S),” in TheEvolution of the Pacific Ocean Margins, Ed. by Z. Ben-Avraham (Oxford Univ. Press, New York, 1989),pp. 193–206.

56. S. Kay, B. Corira, and J. Viramonte, “Young Mafic BackArc Volcanic as Indicator of Continental LithosphericDelaminating beneath the Argentine Puna Plateau, Cen-tral Andes,” J. Geophys. Res. 99, 24 323–24 339 (1994).

57. S. M. Kay, C. Mpodozis, and B. Corira, “Magmatism,Tectonism, and Mineral Deposits of the Central Andes(22–33° S Latitude),” in Geology and Ore Deposits ofthe Central Andes Ed. by B. J. Skinner (Soc. Econ.Geol. Spec. Publ., 1999), pp. 27–59.

58. E. Kendrick, M. Bevis, R. J. Smalley, and B. Brooks,“An Integrated Crustal Velocity Field for the CentralAndes,” Geochem. Geophys. Geosystem. 2 (11),(2001), doi: 10.1029/2001GC000191,2001.

59. J. Kley, “Transition from Basement-Involved to Thin-Skinned Thrusting in the Cordillera Oriental of South-ern Bolivia,” Tectonics 15 (4), 763–775 (1996).

60. J. Kley and C. R. Monaldi, “Tectonic Shortening andCrustal Thickness in the Central Andes: How Good Isthe Correlation?” Geology 26 (8), 723–726 (1998).

61. J. Kley, C. R. Monaldi, and J. A. Salfity, “Along-StrikeSegmentation of the Andean Foreland: Causes and Con-sequences,” Tectonophysics 301, 75–94 (1999).

62. J. Kley, J. Muller, S. Tawackoli, et al., “Pre-Andean andAndean-Age Deformation in the Eastern Cordillera ofSouthern Bolivia,” J. Am. Earth Sci. 10, 1–19 (1997).

Page 18: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

322

GEOTECTONICS Vol. 43 No. 4 2009

ROMANYUK

63. J. Klotz, G. W. Michel, G. Khazaradze, and B. Heinze,“GPS-Based Deformation Measurements and Model-ing,” in Deformation Processes in the Andes. Report forthe Research Period 1999–2001 (Posdam, 2001),pp. 363–392.

64. M. Koesters, H.-J. Goetze, S. Schmidt, et al., “GravityField of a Continent–Ocean Transition Mapped fromLand, Air, and Sea,” EOS Trans. AGU 78 (2), 13–16(1997).

65. M. Kono, Y. Fukao, and A. Yamamoto, “MountainBuilding in the Central Andes,” J. Geophys. Res. 94,3891–3905 (1989).

66. S. Lamb, “Active Deformation in the Bolivian Andes,South America,” J. Geophys. Res. 105 (B11), 25 627–25 653 (2000).

67. S. Lamb and P. Davis, “Cenozoic Climate Change as aPossible Cause for the Rise of the Andes,” Nature 425,792–797 (2003).

68. S. Lamb and L. Hoke, “Origin of the High Plateau in theCentral Andes, Bolivia, South America,” Tectonics 16(4), 623–649 (1997).

69. L. Leffer, A. Mao, T. Dixon, et al., “Constraints on thePresent-Day Shortering Rate across the Central EasternAndes from GPS Measurements,” Geophys. Rev. Lett.24, 1031–1034 (1997).

70. R. A. Marret, R. W. Allmendinger, R. N. Alonso, andR. E. Drake, “Late Cenozoic Tectonic Evolution of thePuna Plateau and Adjacent Foreland, NorthwesternArgentine Andes,” J. South Amer. Earth Sci. 7 (2), 179–207 (1994).

71. N. McQuarrie, “Initial Plate Geometry, ShorteningVariations, and Evolution of the Bolivian Orocline,”Geology 30, 867–870 (2002).

72. N. McQuarrie and G. H. Davis, “Crossing the SeveralScales of Strain-Accomplishing Mechanisms in theHinterland of the Central Andean Fold-Thrust Belt,Bolivia,” J. Struct. Geol. 24, 1587–1602 (2002).

73. D. Mertmann, E. Scheuber, H. Ege, et al., “Tectono-Sedimentary Evolution of the Southern Altiplano: BasinEvolution, Thermochronology and Structural Geology,”in Deformation Processes in the Andes. Interactionbetween Endogenic and Exogenic Processes duringSubduction Orogenesis. Report for the Research Period1999–2001 (Potsdam, 2001), pp. 25–50.

74. M. S. J. Mlynarczyk and A. E. Williams-Jones, “TheRole of Collisional Tectonics in the Metallogeny of theCentral Andean Tin Belt,” Earth Planet. Sci. Lett. 240,656–667 (2005).

75. S. C. Myers, S. Beck, G. Zandt, and T. Wallace, “Litho-spheric-Scale Structure across the Bolivian Andes fromTomographic Images of Velocity and Attenuation for P-and S-Waves,” J. Geophys. Res. 103 (B9), 21 233–21252 (1998).

76. E. Norabuena, L. Leffler-Griffin, A. Mao, et al., “SpaceGeodetic Observations of Nazca–South America Con-vergence along the Central Andes,” Science 279, 358–362 (1998).

77. N. Okaya, S. Tawackoli, and P. Giese, “Area-BalancedModel of the Late Cenozoic Tectonic Evolution of theCentral Andean Arc and Back Arc (20–22° S),” Geol-ogy 25 (4), 367–370 (1997).

78. F. Padro-Casas and P. Molnar, “Relative Motion ofNazca (Farallon) and South American Plates since LateCretaceous Time,” Tectonics 6, 233–248 (1987).

79. R. Patzwahl, J. Mechie, A. Schulze, and P. Giese, “Two-Dimensional Velocity Models of the Nazca Plate Sub-duction Zone between 20 and 25° S from Wide-AngleSeismic Measurements during the CINCA95 Project,”J. Geophys. Res. 104 (B4), 7293–7318 (1999).

80. S. M. Peacock, “Blueschist-Facies Metamorphism,Shear Heating, and P–T–t Paths in Subduction ShearZones,” J. Geophys. Res. 97 (B12), 17693–17707(1992).

81. S. M. Peacock, “The Importance of Blueschist–EclogiteDehydration Reactions in Subducting Oceanic Crust,”Geol. Soc. Amer. Bull. 105, 684–694 (1993).

82. M. C. Pomposiello, J. Booker, L. Shenghui, et al., “IsThick-Skinned Tectonics Due to Weak Faults?,” in Pro-ceedings of the 23rd General Assembly IUGG 19 July–30 July 1999 (Birmingham, 1999), Vol. 2, p. B139.

83. S. C. Ponko and S. M. Peacock, “Thermal Modeling ofthe Southern Alaska Subduction Zone: Insight into thePetrology of Subducting Slab and Overlying MantleWedge,” J. Geophys. Res. 100 (B11), 22 117–22 128(1995).

84. D. C. Pope and S. D. Willett, “Thermal-MechanicalModel for Crustal Thickening in the Central AndesDriven by Ablative Subduction,” Geology 26 (6), 511–514 (1998).

85. C. W. Rapela, R. J. Pankhurst, C. Casquet, et al., “EarlyEvolution of the Proto-Andean Margin of South Amer-ica,” Geology 26 (8), 707–710 (1998).

86. U. Riller, O. Oncken, and M. Strecker, “Late CenozoicTectonomagmatic and Kinematic History of the Puna,”in Deformation Processes in the Andes. Interactionbetween Endogenic and Exogenic Processes duringSubducting Orogenesis. Report for the Research Period1999–2001 (Potsdam, 2001), pp. 81–94.

87. T. V. Romanyuk, H.-J. Goetze, and P. F. Halvorson,“A Density Model of Andean Subduction Zone,” Lead-ing Edge, (February), 264–268 (1999).

88. P. Roperch, M. Fornari, G. Herail, and G. V. Parraguez,“Tectonic Rotations within the Bolivian Altiplano:Implications for the Geodynamic Evolution of the Cen-tral Andes during the Late Tertiary,” J. Geophys. Res.105, 795–820 (2000).

89. T. Ryberg and G. Fuis, “The San Gabriel MountainsBright Reflective Zone: Possible Evidence of YoungMid-Crustal Thrust Faulting in Southern California,”Tectonophysics 286, 31–46 (1998).

90. M. Schmitz, “A Balanced Model of the Southern Cen-tral Andes,” Tectonics 13 (2), 484–492 (1994).

91. M. Schmitz, K. Lessel, P. Giese, et al., “The CrustalStructure beneath the Central Andean Forearc and Mag-matic Arc as Derived from Seismic Studies: the PISCO94 Experiment in Northern Chile (21–23° S),” J. SouthAmer. Earth Sci. 12, 237–260 (1999).

92. B. Schurr, G. Asch, A. Rietbrock, et al., “Complex Pat-terns of Fluid and Melt Transport in the Central AndeanSubduction Zone Revealed by Attenuation Tomogra-phy,” Earth Planet. Sci. Lett. 215, 105–119 (2003).

93. T. Sempere, R. F. Butler, D. R. Richards, et al., “Stratig-raphy and Chronology of Upper Cretaceous–Lower

Page 19: The Late Cenozoic geodynamic evolution of the central segment of the Andean subduction zone

GEOTECTONICS Vol. 43 No. 4 2009

THE LATE CENOZOIC GEODYNAMIC EVOLUTION OF THE CENTRAL 323

Paleogene Strata in Bolivia and Northwest Argentina,”Geology 109 (6), 709–727 (1997).

94. T. Sempere, G. Herail, J. Oller, and M. Bonhomme,“Late Oligocene–Early Miocene Major Tectonic Crisisand Related Basins in Bolivia,” Geology 18, 946–949(1990).

95. M. Springer, “Heat-Flow Density across the CentralAndean Subduction Zone,” Tectonophysics 306, 377–395 (1999).

96. R. J. Stern, “Subduction Zones,” Rev. Geophys. 40 (4),3.1–3.13.1012 (2002), doi: 10.1029/2001RG000108.

97. J. L. Swenson, S. L. Beck, and G. Zandt, “Crustal Struc-ture of the Altiplano from Broadband Regional Wave-form Modeling: Implications for the Composition ofThick Continental Crust,” J. Geophys. Res. 105, 607–621 (2000).

98. W. Tao and R. J. O’Connell, “Albative Subduction:A Two-Sided Alternative to the Conventional Subduc-tion Model,” J. Geophys. Res. 97, 8877–8904 (1992).

99. R. M. Tosdal, “The Amazon–Laurentian Connection asViewed from the Middle Proterozoic Rocks in the Cen-tral Andes, Western Bolivia and Northern Chile,” Tec-tonics 15 (4), 827–842 (1996).

100. M. Van der Meijde, F. Marone, and S. van der Lee,“Seismic Evidence for Water Atop the MediterraneanTransition Zone,” EOS Trans. AGU, 85(47), Fall Meet-ing Supplement, Abstract T3 F-03, 2004.

101. R. von Huene, I. A. Pecher, and M.-A. Gutscher,“Development of the Accretionary Prism along Peruand Material Flux after Subduction of Nazca Ridge,”Tectonics 15 (1), 19–33 (1996).

102. R. von Huene and D. W. Scholl, “Observations at Con-vergent Margins Concerning Sediment Subduction,

Subduction Erosion, and the Growth of ContinentalCrust,” Rev. Geophys. 29 (3), 279–316 (1991).

103. S. Wdowinski and Y. Bock, “The Evolution of Defor-mation and Topography of High-Elevated Plateaus.1. Model, Numerical Analysis, and General Results.2. Application to the Central Andes,” J. Geophys. Res.99 (B4), 7121–7130 (1994).

104. D. Whitman, “Moho Geometry beneath the EasternMargin of the Andes, Northwest Argentina, and ItsImplications to Effective Elastic Thickness of theAndean Foreland,” J. Geophys. Res. 99 (B8), 15277–15289 (1994).

105. P. Wigger, M. Schmitz, M. Araneda, et al., “Variation inthe Crustal Structure of the Southern Central AndesDeduced from Seismic Refraction Investigations,” in Tec-tonics of the Southern Central Andes, Ed. by K.-J. Reutter,E. Schenber, and P. J. Wigger (Springer, Berlin, 1994),pp. 23–48.

106. Y. Yang, M. Liu, and S. Stein, “A 3-D GeodynamicModel of Lateral Crustal Flow during Andean Moun-tain Building,” Geophys. Rev. Lett. 30 (21), 2093(2003), doi: 10.1029/2003GL018308.

107. X. Yuan, S. V. Sobolev, and R. Kind, “Moho Topogra-phy in the Central Andes and Its Geodynamic Implica-tion,” Earth Planet. Sci. Lett. 199 (3–4), 389–402(2002).

108. X. Yuan, S. V. Sobolev, R. Kind, et al., “Subduction andCollision Processes in the Central Andes Constrainedby Converted Seismic Phases,” Nature 408 (21), 958–961 (2000).

109. T. R. Zapata and R. W. Allemendinger, “Growth StrataRecords of Instantaneous and Progressive Limb Rota-tion in the Precordillera Thrust Belt and Bermejo Basin,Argentina,” Tectonics 15 (5), 1065–1083 (1996).

Reviewers: N. V. Koronovsky and M. G. Lomize