Você está na página 1de 22

See

discussions, stats, and author profiles for this publication at: http://www.researchgate.net/publication/261568460

Rock uplift and exhumation of continental


margins by the collision, accretion and
subduction of buoyant and topographically
prominent oceanic crust
ARTICLE MAY 2014
DOI: 10.1002/2013TC003425

CITATION

DOWNLOADS

VIEWS

129

81

2 AUTHORS, INCLUDING:
Richard Spikings
University of Geneva
76 PUBLICATIONS 890 CITATIONS
SEE PROFILE

Available from: Richard Spikings


Retrieved on: 16 July 2015

PUBLICATIONS
Tectonics
RESEARCH ARTICLE
10.1002/2013TC003425
Key Points:
Plateau and ridge collision in
subduction zones drives rock uplift
The quantity of rock uplift is dominantly a function of buoyancy forces
Exhumation patterns are strongly
dependant on climate induced erosion

Correspondence to:
R. Spikings,
richard.spikings@unige.ch

Citation:
Spikings, R., and G. Simpson (2014),
Rock uplift and exhumation of
continental margins by the collision,
accretion, and subduction of buoyant
and topographically prominent oceanic
crust, Tectonics, 33, doi:10.1002/
2013TC003425.
Received 13 AUG 2013
Accepted 8 APR 2014
Accepted article online 10 APR 2014

Rock uplift and exhumation of continental margins


by the collision, accretion, and subduction
of buoyant and topographically
prominent oceanic crust
Richard Spikings1 and Guy Simpson1
1

Department of Earth Sciences, University of Geneva, Geneva, Switzerland

Abstract

Understanding the causes of rock and surface uplift is important because they control the
location of mountain building, depocenters, and drainage characteristics and can inuence climate. Here
we combine previous thermochronological data with eld observations to determine the amount of
exhumation, rock, and surface uplift that occurs in the upper plate of Central and South American subduction
zones during the collision, accretion, and subduction of oceanic plateaus and aseismic ridges. The collision of
buoyant and topographically prominent oceanic plateaus and ridges can drive at least 5 km of rock uplift within
2 Ma. Uplift appears to be an immediate response to collision and is generally independent of the slab dip.
The amount of rock uplift is controlled mainly by excess topography associated with the ridge (ultimately linked
to buoyancy) and erosion, while it is also inuenced by the strength of the subduction interface related to the
presence of volcanic asperities and overpressured sediments in the subduction channel. The quantity of
exhumation is strongly dependant on climate-induced erosion and the lifespan over which the topography is
uplifted and supported. Sediment draining into the trench may leave the elevated ridge axis sediment starved,
increasing the shear stresses at the ridge subduction interface, leading to positive feedback between ridge
subduction, rock uplift, and exhumation. Trench-parallel variations in exhumation have a direct impact on
exploration paradigms for porphyry-related metalliferous deposits, and it is likely that porphyry systems are
completely eroded by the impingement of plateaus and aseismic ridges within temperate and tropical climates.

1. Introduction
Subduction of oceanic crust can be considered as a steady state process when plate convergence rates,
plate coupling, the thickness, and buoyancy of the colliding oceanic lithosphere and slab suction forces
remain constant over periods of millions of years. Signicant changes in arc magma composition and
distribution, deformation, exhumation, and surface elevation of the continental crust can occur when one
or more of these variables changes [e.g., Gutscher et al., 2000; Spikings et al., 2001; Lamb and Davis, 2003;
Vannucchi et al., 2006; Von Huene and Ranero, 2009; Wallace et al., 2009; Mason et al., 2010; Morell et al.,
2011]. This contribution aims to quantify the rock uplift and exhumation experienced by continental crust
when it collides with thick (>20 km) oceanic plateaus and is subducted by topographically prominent and
buoyant, hot spot-derived aseismic ridges (Figure 1), using examples from Middle and South American
subduction zones. We also discuss the causal link between rock uplift, exhumation, plate coupling,
and climate.
The western margin of the South American Plate has been almost continually active since ~500 Ma along
an extensive strike length of ~6000 km and is currently overriding the heterogeneous Nazca Plate.
Systematic along-strike variations in maximum mean elevation and trench axial depth over trench-parallel
distances of >1000 km led Lamb and Davis [2003] to suggest that high elevations in the central Andes are
dominantly a function of sediment deposition in the trench and ultimately climate. However, dramatic
variations in rock uplift and exhumation of continental crust, leading to changes in erosion level, relief, and
elevation, are spatially correlatable with the collision and subduction of buoyant and elevated oceanic
crust over trench-parallel distances of ~200 km and less [e.g., Dominguez et al., 2000; Spikings et al., 2001;
Villagmez et al., 2011a]. Similar conclusions are reached by analogue sandbox models, which demonstrate
that the collision and subduction of aseismic ridges with continental crust can drive rock uplift in the upper
plate [e.g., Hampel et al., 2004]. Understanding the causes of rock and surface uplift along active plate

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

60E

80E

100E

120E

140E

10.1002/2013TC003425

160E

140W

160W

180E

120W

80W

100W

60W

20W

40W

0W

V
I

Fa

CC

60N

Y
W

45N

Az

Em
S

He
Cs

Og

30N

Ha

Mp

CV
C

15N

Co

MR

Ch

0N

O
Ro

Ni

Mc

Ca

Ma

Tu

RG

15S

Na
Iq

Wa

30S

JF

H
45S

Depth

K
-6

-5

-4

-3

Elevation
-2

-1

km

60S

Figure 1. Global elevation and bathymetry (excluding east Africa and the western Indian Ocean (~2060E) constructed using GeoMapApp and the Global MultiResolution Topography (GMRT) database [http://www.geomapapp.org; Ryan et al., 2009]. A selection of the most voluminous oceanic plateaus (including those
that have already accreted and form part of the upper plate) and aseismic ridges are highlighted in red, after Cofn and Eldholm [1994]. Oceanic plateaus: A, Austral
Plateau; Az, Azores Plateau; C, Caribbean Plateau; Cs, Canaries plateau; C V, Cape Verde plateau; H, Hikurangi Plateau; I, Iceland Plateau; K, Kerguelen Plateau; M,
Manihiki Plateau; Mc, Mascarene Plateau; MR, Magellan Rise; N, Naturaliste Plateau; O, Ontong Java Plateau; Og, Ogasawara Plateau; S, Shatsky Rise; T, Tuamotu
Plateau; V, Voring Plateau; W, Wrangelia Terrane; Y, Yakutat Terrane. Aseismic Ridges: B, Broken Ridge; Ca, Carnegie Ridge; Ch, Chagos-Laccadive Ridge; Co, Cocos
Ridge; Em, Emperor Seamount Chain; Fa, Iceland-Faeroe Ridge; Ha, Hawaii seamount chain; He, Hess Rise; Iq, Iquique Ridge; JF, Juan Fernandez Ridge; Ma, Marquesas
Islands; Mp, Mid-Pacic Mountains; Na, Nazca Ridge; Ni, Ninety East Ridge; RG, Rio Grande Rise; Ro, Robbie Ridge; Wa, Walvis Ridge.

margins is important because they control the location of mountain building, depocenters, and drainage
characteristics and can inuence climate [Willet et al., 2006]. Variations in exhumation expose different
crustal levels, facilitating an improved understanding of the composition and structure of the crust, and are
responsible for preserving or eroding primary ore deposits.
Multiphase 40Ar/39Ar, ssion track, and (U-Th)/He data acquired from granitoids and metamorphic rocks
in the Andean cordilleras of Colombia, Ecuador, the Western Cordillera and coastal ranges of Peru, and
along the highlands of southern Costa Rica have been used to determine the response of continental
crust to the collision, accretion, and subduction of the Caribbean Plateau and the Carnegie, Nazca, and
Cocos ridges. All of these indentors are composed of hot spot-related rocks that are positively buoyant
relative to normal, midocean ridge-derived crust, have crustal thicknesses (835 km) that exceed normal
oceanic crust, and therefore change the forces that act at the plate interface in subduction zones. The
paper is a review in the sense that no new data is presented. The thermochronological data have been
compiled from several independent studies, most of which were completed within the previous 15 years
[Spikings and Crowhurst, 2004; Spikings et al., 2001, 2005; Grfe et al., 2002; Wipf et al., 2008; Morell et al.,
2011; Villagmez and Spikings, 2012]. These studies focused on the consequences of ocean-continent
interaction in different locations, and a comparison of each location is useful for assessing the inuence
of slab dip and plate coupling, in addition to the role played by preexisting structure within the
upper plate.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

82

86

90

78

74

Nicaragua
Normal Fault
Thrust Fault

(hbl, ms, bt)

Costa Rica

NP

TM

Sierra Nevada
de Santa
Marta

Panam

OP

Zircon FT
Zircon (U-Th)/He
Apatite FT
Apatite (U-Th)/He

Pan am a FZ

40

80 0

Latitude

CAF

4
6

Ga la pa gos Rift

Colombia
2

4
7

15

10
12

GHS

4N

2N

11

EC
62mm/a

56 m m /y
r

AFB

Carnegie Ridge

12

Ecuador
2S

16.5
17.5

NAZCA

PLATE

19

.5

24

GG

4
1
4

29

GF

Peru

21

6N

UMV

WC

12
14

PF

CLIP
CC

Pa na m a FZ

60-70mm/yr

Malpelo
Ridge

10

NAZCA
PLAT E

Cocos
Ridge

Amount of
exhumation
since 15 Ma (km)

95 mm/yr

Age (Ma)
0

PLATE

BP

COCOS

40Ar/39Ar

CARIBBEAN PLAT E

4S

25

Elevation and bathymetry (meters)


-5000

-4000

-3000

-2000

-1000

1000

2000

3000

4000

5000

6000

0
150
300
Temperature of current
surface at 15 Ma (C)

Figure 2. Digital elevation model of Ecuador, Colombia, and southern Central America (GTOPO30) and bathymetry of the offshore regions, with illumination from the
east. Magnetic anomalies with approximate age in million years (Myr) are shown [Lonsdale and Klitgord, 1978]. The Campanian ocean-continent suture of the northern
Andes is shown as a thick black line. Plate vectors are from Trenkamp et al. [2002; northern Andean fore arc relative to cratonic South America]; Norabuena et al. [1999;
Nazca PlateSouth America], and DeMets et al. [1994; Cocos PlateCaribbean Plate]. The Sierra Nevada de Santa Marta is shown along the northern Colombian coast.
Plots show relationship between latitude and (i) age (labels show the number of samples dated by apatite (U-Th)/He and apatite ssion-track) acquired by various
thermochronological systems (data from Spikings and Crowhurst [2004], Spikings et al. [2000, 2001, 2005, 2010], and Villagmez and Spikings [2012]) and (ii)
palaeotemperatures and depths of the current surface at 15 Ma, assuming a geothermal gradient of 30C/km. AFB, Amazon Foreland Basin; BP, Burica Peninsular;
CAF, Cauca-Almaguer Fault; CC, Central Cordillera; CLIP, Caribbean Large Igneous Province (accreted allochthons); CPT, Choc-Panam Terrane; EC, Eastern Cordillera; GFZ,
Grijalva Fracture Zone; GG, Gulf of Guayaquil; GHS, Galapagos hotspot; NP, Nicoya Peninsular; OP, Osa Peninsular; PF, Peltetec Fault; TM, Talamanca Mountains; UMV,
Upper Magdalena Valley; WC, Western Cordillera.

2. Oceanic PlateauContinent Collision


Oceanic plateaus are ultramac-mac, large igneous provinces that reach thicknesses of ~35 km (e.g., Ontong
Java Plateau) [Mahoney and Spencer, 1991; Neal et al., 1997], are buoyant relative to normal oceanic lithosphere
of the same age (due to the presence of a thickened basaltic crust), and can reach widths of more than 1000 km
(Figure 1). A majority of U-Pb and 40Ar/39Ar ages of the Caribbean Plateau span 9288 Ma [Kerr et al., 1997; Sinton
et al., 1998; Vallejo et al., 2006, 2009; Villagmez et al., 2011b], and 40Ar/39Ar ages of the Wrangellia Plateau
(western Canada) span 230225 Ma [Greene et al., 2010]. These short time spans (5 Ma) are typical of other
globally distributed oceanic plateaus (Figure 1), indicating that they are rapidly emplaced. Emplacement usually
occurs above the early stages of plume-related, hot spot magmatism or within fast-spreading centers [Nakanishi
et al., 1999; Korenaga, 2005]. The thick, buoyant nature of plateaus relative to normal oceanic crust may restrict
the extent to which they can be subducted [e.g., Cloos, 1993; Mason et al., 2010], and consensus exists that
accretion of oceanic plateaus to continental plates is a signicant contributor to the growth of continental crust
[e.g., Richards et al., 1991; Rudnick, 1995; Puchtel et al., 1998] and can drive orogenesis (e.g., Laramide Orogeny)
[Livaccari et al., 1981]. The collision and subduction of oceanic plateaus beneath continental crust is currently
responsible for the highest coastal mountain ranges on earth (>5500 m), such as the Chugach-St. Elias range in

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

A
20
60

Late Cretaceous margin of Ecuador


(Eastern Cordillera)

Late Cretaceous margin of Colombia


(Central Cordillera)
20

AHePRZ

60

APAZ
100

140
Far northern
Ecuador (0.5N)

CLIP
collision

180
220
260

CLIP
collision

180

ZHePRZ

Far southern
Colombia (2N)

North
of 5N
South
of 5N

260

South
of 130S

300

300
80

140

220

North
of 130S

ZPAZ

Temperature (C)

Temperature (C)

100

60

40
Time (Ma)

20

80

60

40
Time (Ma)

20

Sierra Nevada de Santa


Marta (Colombia)
Talamanca Mtns
20

Temperature (C)

Unconstrained

60
APAZ
100
140
180
80

60

40
Time (Ma)

20

Figure 3. Thermal history solutions for crystalline rocks of (a) the Eastern Cordillera of the Late Cretaceous margins of Ecuador
(Eastern Cordillera) [Spikings and Crowhurst, 2004; Spikings et al., 2000, 2001, 2010] and Colombia (Central Cordillera) [Villagmez
and Spikings, 2012], and (b) The Sierra Nevada de Santa Marta (northern Colombia) [Villagomez et al., 2011a] and the Talamanca
Mountains (Costa Rica; only the best-t solution is shown) [Grfe et al., 2002]. The solutions were obtained using Markov Chain
Monte Carlo simulations [Ketcham et al., 2007] and are considered to be good ts with values of the Kuipers statistics test of
>0.5. Kinetic parameters used in all studies are those of Flowers et al. [2009] and Reiners et al. [2004] for diffusion of He in zircon,
Farley [2000] for diffusion of He in apatite, and Ketcham et al. [1999] and Ketcham et al. [2007] for ssion track annealing
in apatite. The bold line shown for the Sierra Nevada de Santa Marta is the best t solution for the sample that yielded
the youngest apatite ssion track age. AHePRZ, Apatite helium partial retention zone; APAZ, Apatite partial annealing zone;
ZHePRZ, Zircon helium partial retention zone; ZPAZ, Zircon partial annealing zone; CLIP, Caribbean Large Igneous Province.

southeast Alaska (Yakutat Plateau) [Enkelmann et al., 2010] and the Sierra Nevada de Santa Marta in northern
Colombia (Caribbean Plateau; Figure 2) [Villagmez et al., 2011a]. The immediate response (i.e., within 12 Myr of
collision) of continental crust to the collision, accretion, and occasional obduction of oceanic plateaus onto
continental margins is investigated here.
2.1. Caribbean Plateau: Initial Collision With Continental Crust in the Early Cretaceous
The Caribbean Plateau formed as part of the Farallon Plate at equatorial latitudes [Acton et al., 2000; Luzieux
et al., 2006; Vallejo et al., 2009], was displaced eastward, and started colliding with northwestern South
America at 7573 Ma [Figure 2; Spikings et al., 2001, 2010; Vallejo et al., 2006; Villagmez and Spikings, 2012].
Accreted, detached allochthons of the Caribbean Plateau currently form the basement of the western
Andean cordilleras and fore arc of Ecuador and Colombia. Thermochronological data from the Late
Cretaceous margin of Ecuador [Spikings et al., 2001, 2010] and Colombia [Villagmez and Spikings, 2012]
constrain the amount of exhumation that was caused by the collision and accretion of the plateau.
Early CretaceousPaleozoic greenschist-amphibolite grade rocks exposed in the Eastern Cordillera of Ecuador
yield muscovite and biotite 40Ar/39Ar plateau dates, zircon, and apatite ssion track and apatite (U-Th)/He
dates that are all younger than 80 Ma (Figure 2) [Spikings et al., 2010], indicating that the current surface of the
paleomargin has been shallower than ~12 km since ~75 Ma. Markov Chain Monte Carlo modeling of the
thermochronological data shows that the rocks started to cool with rates of 36C/Myr at 7570 Ma (Figure 3
and Table 1), which was synchronous with (i) the deposition of terrigenous sediments in uvial systems at
SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

SPIKINGS AND SIMPSON

>50
?
?
?

<30
0.0320

0.0330

0.0140

60

>17
?

20

90
16
3

180
240
40

1513
150
150

11

165

150

30

36
13
44
17

180
200
220
260

Cooling Rate
(C/Myr)

7570
7560
7570
7560

Amount of
Cooling (C)

30

30

30
30

30

30
30
30

30

30
30
30
30

Geothermal
Gradient (C/km)

250, 30

100, 30
100, 30

200, 30

660, 30
660, 30
700, 30
700, 30

Strike Length,
c
Width (km)

>0.6, >1.7, 12750


?

0.7, 2.0, ?

3.0, 6.0, 18,000


0.5, 8.0, 24,000
0.1, 1.3

0.4, 5.5, 33,000

1.2, 6.0, 119,000


0.4, 6.7, 133,000
1.5, 7.3, 153,000
0.6, 8.7, 183,000

Exhumation
3
(km/Myr, km, km )

1.4
?

1.7

5.0
6.5
1.1

4.5

5.0
5.6
6.0
7.2

Isostatic
Rebound (km)

7.7, 0.10

8.5, 0.28

?
6.0, 0.20

0.3, 0.8

?
?
?

?
?
?
?

Surface Uplift
(km/Myr, km)

>0.6, >1.7
?

0.9, 2.8

3.0, 6.0
0.5, 8.0
0.1, 1.3

0.4, 5.5

1.2, 6.0
0.4, 6.7
1.5, 7.3
0.6, 8.7

Rock Uplift
(km/Myr, km)

All rate calculations are determined by dividing the length scale by time. No corrections have been made for heat advection. Volume calculations are simple and are exhumation multiplied by
strike length and width of the exhumed region.
b
Assumed value for the geothermal gradient determined from the measurements of Henry and Pollack [1988] (2040C/km; Peru and Bolivia); 30C is the maximum gradient reported by Springer
and Foerster [1998] in the Central Andes.
c
Width is average width of the Eastern Cordillera of Ecuador, Central Cordillera of Colombia and the Talamanca Mountains of Costa Rica.
d
Isostatic rebound is calculated using the procedure of Molnar and England [1990].
e
Estimate for surface uplift is provided when constraint stratigraphic horizons have been reported [e.g., Hsu, 1992].
f
Rock uplift is calculated as the sum of surface uplift and exhumation [Molnar and England, 1990].
g40
39
Ar/ Ar (muscovite, biotite, alkali feldspar), ssion track (zircon, apatite), and (U-Th)/He (zircon, apatite) data [Spikings et al., 2000, 2001, 2010; Spikings and Crowhurst, 2004; Villagmez and
Spikings, 2012].
h
Fission track (apatite) and (U-Th)/He (apatite) data from Wipf et al. [2008].
i
Approximate time period that the Nazca Ridge subducts any region of the upper plate [van Hunen et al., 2002].
j
Fission track (apatite) data from Grfe et al. [2002] and ssion track and (U-Th)/He data (apatite) from Morell et al. [2012].
k14
C dating from Sak and Fisher [2004].
l
Optically stimulated luminescence age from Gardner et al. [2013].
m14
C dating from Morell et al. [2011].

Caribbean Plateau Collision


Eastern Cordillera
of Ecuador (Arc)
Central Cordillera
of Colombia (Arc)
g
Carnegie Ridge Collision
Central Cordillera
of Colombia (12N; arc)
Eastern Cordillera of
Ecuador (0.5N1.0N; arc)
Eastern Cordillera of
Ecuador (4.0S; arc)
h
Nazca Ridge Collision
Northern Arequipa (fore arc)
Cocos Ridge Collision
j
Talamanca Mountains (arc)
Osa Peninsular
k
(outer fore arc)
Osa Peninsular
l
(outer fore arc)
Burica Peninsular
m
(outer fore arc)

Location

Time Period
(Ma)

Table 1. Summary of Rock Uplift, Exhumation, and Surface Uplift

Tectonics
10.1002/2013TC003425

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

~75 Ma (Tena Fm.) in the newly formed retroarc Amazon Foreland Basin (Figure 2), (ii) the termination of
east-facing arc magmatism beneath the Caribbean Plateau, as it approached from the west [Vallejo et al.,
2009], and (iii) up to 90 of clockwise rotation, and fragmentation of the indenting plateau, commencing at
~7573 Ma [Luzieux et al., 2006]. Elevated cooling rates continued until 60 Ma, by which time at least ~7 km
of crust had been exhumed [Spikings et al., 2010]. Fragmentation of buoyant oceanic plateaus during
collision with a subduction zone is consistent with the numerical models of Mason et al. [2010], which
solves the Stokes equation for creeping ow
Similarly, 40Ar/39Ar, ssion track, and (U-Th)/He analyses of greenschist-amphibolite grade. Early Cretaceous
and older igneous rocks in the Central Cordillera of Colombia (Figure 2) yield an identical result and reveal a
phase of rapid, exhumation-induced cooling of the paleomargin commencing at ~75 Ma (Figure 3 and
Table 1) [Villagmez and Spikings, 2012], which was synchronous with clastic deposition in the Upper
Magdalena basin (e.g., Montserrate Fm.).
2.2. Caribbean Plateau: Recent Interaction With the South American Plate
Although the Caribbean Plateau and South America have shared a plate boundary since ~75 Ma, the
isostatically uncompensated (+208 mGal Bouguer Anomaly) [Kellogg and Bonini, 1982; Cern, 2008] high
elevations of the Sierra Nevada de Santa Marta, northern Colombia (Figure 2), are a present-day consequence
of ocean plateau-continent collision. The Sierra Nevada de Santa Marta resides within a triangular faulted
block of continental crust and reaches a peak elevation of ~5.75 km within 40 km of the northern coastline of
South America. The youngest apatite ssion track age obtained from the lowest elevations is 16.3 1.8 Ma
(2), and the youngest period of cooling through the apatite partial annealing zone occurred at 15 1 Ma
(Figure 3) [Villagmez et al., 2011a], which is unexpected given the high elevations and relief, and an annual
rainfall of 12 m/yr [Arias and Morales, 1999], which should provide signicant erosive power. There has
been insufcient time for erosion to exhume the late Miocene and younger apatite (ssion track) partial
annealing zone to the surface, implying that rock uplift is greater than surface uplift by less than 1.5 km.
Clearly, recent rock uplift rates must have been extremely high.
Cern [2008] utilized gravity modeling to propose that the high mean elevation of the Sierra Nevada de Santa
Marta and isostatic disequilibrium may be maintained by rigid, underthrusted Caribbean lithosphere.
Furthermore, Miller et al. [2009] and Masy et al. [2011] interpret the shear wave velocity structure beneath the
Sierra Nevada de Santa Marta and Maracaibo Block as evidence of underthrusting of the Caribbean Plate,
which hosts the Caribbean Plateau, beneath the continental crust of the Sierra Nevada de Santa Marta. We
suggest that this probably occurred within the last 12 Ma [e.g., Villagmez et al., 2011a], which appears to
have been insufcient time for erosion rates to signicantly reduce the elevated topography. Thus, ocean
plateaucontinent collision along northern South America is currently driving rapid rock and surface uplift.

3. Consequences of Colliding and Subducting Oceanic Aseismic Ridges With


Continental Crust
Oceanic aseismic ridges are characterized by bathymetric prominence, high heat ow, and crustal
thicknesses of up to 35 km (e.g., Nazca Ridge; Figure 1) [Tassara et al., 2006], and form by the progressive
emplacement of magma at a common location into oceanic crust, either due to the presence of thermal
anomalies caused by impinging plumes or a consequence of compositional heterogeneities in the source
rocks [e.g., Sallars et al., 2005]. Low water depths above the Nazca Plate correlate closely with anomalously
thick oceanic crust, suggesting that at least 40% of the area of the Nazca Plate has been thermally uplifted
and has undergone magma emplacement during hot spot interaction [Tassara et al., 2006]. These generally
aseismic chains typically extend from source regions located at oceanic spreading centers, e.g., the Walvis
(South Atlantic Spreading Center), Cocos, and Carnegie Ridges (Cocos-Nazca Spreading Center), and are
buoyant relative to midocean ridge-derived oceanic crust (of similar age) due to the presence of supporting
depleted harzburgitic mantle [e.g., Canales et al., 2002; Van Hunen et al., 2004] and thickened crust [e.g.,
Walther, 2003]. As with oceanic plateaus, the buoyancy of oceanic ridges may interfere with the normal
subduction process and induce various effects on the adjacent continental margin.
Thermochronological data are presented from three regions of aseismic ridgecontinent interaction
dispersed along the western South American and Central American margin, which constrain the rock uplift

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

and exhumation history of continental crust and oceanic plateaus, respectively, to the impact and subduction
of aseismic ridges.
3.1. The Carnegie Ridge
The Nazca Plate between the Cocos-Nazca Spreading Center and the Grijalva Fracture Zone is composed of
oceanic crust younger than 25 Myr old and hosts the aseismic Carnegie Ridge (Figure 2). The Grijalva Fracture
Zone separates Nazca crust (<25 Ma) from relict Farallon oceanic crust (>25 Ma) and also marks a sudden
spatial increase in water depth of ~1 km. The Carnegie Ridge is emergent at the currently active Galapagos
Hotspot and is orthogonally subducting beneath northwestern South America at a rate of ~56 mm/yr
[Norabuena et al., 1999; Trenkamp et al., 2002], where it has a total relief of >2000 m above the adjacent sea
oor. The buttressing continental crust is composed of accreted, emergent oceanic crust of the Caribbean
Plateau that is sutured against Paleozoic-Cretaceous metamorphic and igneous rocks of the Late
Cretaceous South American margin, which are exposed in the Eastern Cordillera of Ecuador and the Central
Cordillera of Colombia. The Late Cretaceous suture zone is highly faulted and hosts serpentinized mac
and ultramac rocks and altered high-pressure metamorphic sequences. The timing of collision of the
Carnegie Ridge with South America is highly debated, and published values range between 15 Ma [e.g., Pilger,
1984; Spikings et al., 2001] and 1 Ma [e.g., Lonsdale and Klitgord, 1978; see review in Michaud et al., 2009].
The spatial distribution of seismicity beneath central Ecuador shows that the Carnegie Ridge is hosted by a slab
that dips continuously at an angle of 2535 to a depth of 200 km [Guillier et al., 2001]. The measured slab
dip corroborates abundant Pleistocene-Present volcanic activity above the region subducted by the ridge (e.g.,
volcanos Antisana, Cayambe, Reventador, and Pichincha) [e.g., Bourdon et al., 2002].
Trench-parallel trends in multiphase 40Ar/39Ar, zircon, and apatite FT and apatite (U-Th)/He ages acquired
from the Eastern Cordillera of Ecuador and Central Cordillera of Colombia (Figure 2) clearly reveal a phase of
very high rates of rock uplift and exhumation during ~1513 Ma (6 km/Myr; Figure 3 and Table 1) [Spikings
et al., 2000, 2001; Villagmez and Spikings, 2012], which occurred at the same latitude as where the trench is
intersected by the subducting, eastward-migrating Carnegie Ridge. A signicant percentage of apatite FT and
(U-Th)/He ages from the northern Eastern Cordillera of Ecuador and the southern Central Cordillera of
Colombia are Miocene or younger, whereas almost all FT and (U-Th)/He ages acquired from south of S130
and north of N2 span between the Late Cretaceous and Eocene (Figure 2). This trend reects distinguishable
trench-parallel differences in the quantity of cooling of continental basement in the northern Andes
(Figures 2 and 3), which are interpreted as varied exhumation depths and rates as a direct response to the
collision and subduction of the Carnegie Ridge since ~15 Ma [Spikings et al., 2001, 2010]. Middle Miocene
cooling phases were accompanied by the contemporaneous deposition of clastic sediments in the Amazon
Foreland Basin (e.g., Arajuno Fm.) [Spikings et al., 2001], supporting interpretations that cooling was driven
by exhumation.
The current surface of some faulted blocks of the Eastern Cordillera of Ecuador at ~ N030 was at depths of
~8 km at 15 Ma, whereas rocks exposed between S030 and S130 resided at depths of ~1.3 km prior to
15 Ma (Table 1) [Spikings et al., 2010]. Further south, the current surface of the Eastern Cordillera was at
depths shallower than 1.3 km immediately prior to exhumation commencing at 15 Ma, and specic
locations may have been within a few hundred meters of the surface since ~30 Ma (Table 1). The reverse
trend can be seen along the strike of the Central Cordillera of Colombia, where exhumation depths since
15 Ma increase from north to south, toward the region of subduction of the Carnegie Ridge (Figure 2).
Faulted blocks along the western ank of the Central Cordillera in northern Colombia (~6N) exhumed from
depths as shallow as 1 km since ~15 Ma [Villagmez and Spikings, 2012], and Restrepo-Moreno et al. [2009]
report very low erosion rates (~0.04 mm/a) in the northern Central Cordillera since 25 Ma.
These along-strike trends in exhumation can be accounted for by dextral, transcurrent reactivation of the
Campanian suture (Peltetec Fault in Ecuador and the Cauca-Almaguer Fault in Colombia; Figure 2) [e.g., Hughes
and Pilatasig, 2002] as a result of the subduction of the Carnegie Ridge at 15 Ma. The trend of the suture bends
toward a northerly strike at ~ S3 and dextral displacement of the allochthonous oceanic region toward the
north-northeast along the Peltetec Fault drove transpression in northern Ecuador and southern Colombia,
leading to rock uplift and erosional exhumation. The same tectonic rearrangement is responsible for the onset
of extension, subsidence, and marine transgressions in southern Ecuador during 159 Ma [Hungerbhler et al.,
2002], due to the removal of the fore arc by lateral displacement. Clearly, the collision and subduction of the

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

Carnegie Ridge had a signicant impact on the vertical displacement and exhumation of the buttressing
continental plate.
The Choc-Panam terrane forms the basement to the region of northwestern Colombia (Figure 2), and
collided with the South American Plate at either 2325 Ma [Farris et al., 2011] or 1613 Ma [Duque-Caro, 1990],
or during the latest Miocene-Pliocene [e.g., Cediel et al., 2003; Newkirk and Martin, 2009; Montes et al., 2012].
Montes et al. [2012] report thermochronological data from the basement complex of Panam and conclude
that exhumation related cooling occurred during 4752 Ma and 129 Ma, the latter of which dates the time of
collision with South America. We do not consider elevated exhumation rates in northwestern South America
at 1513 Ma to be due to the collision of the Choc-Panam terrane for several reasons. First, the lack of
consensus on the timing of collision creates uncertainty when assessing its effect on South American crust
during 1513 Ma. Second, Restrepo-Moreno et al. [2009] and Villagmez and Spikings [2012] report very low
cooling rates and exhumation depths in the northern Central Cordillera of Colombia since 25 Ma [RestrepoMoreno et al., 2009; Villagmez and Spikings, 2012], directly inboard of the collision zone of the terrane,
corroborating the conclusions of Mann and Corrigan [1990] and Rockwell et al. [2010] that the Choc-Panam
terrane was a soft indentor. Finally, the most compelling argument that the Carnegie Ridge was responsible
for exhumation lies in the along-strike variations in cooling dates and exhumation depths since 15 Ma along
the Early Cretaceous palaeomargin of north-western South America (Figure 2). Exhumation depths are
approximately 1 order of magnitude higher directly above the subducted extent of the Carnegie Ridge,
compared to regions distal to the ridge, providing a compelling argument for a cause and effect relationship.
The timing of collision of the Carnegie Ridge with the North Andean margin has been a matter of
conjecture for over 20 years. Variations in coastal morphology, generated by very slow southward
migration of the Carnegie Ridge led Gutscher et al. [1999] to propose that the Carnegie Ridge has collided
with the trench since 8 Ma and caused the northeastward drift of the allochthonous oceanic terranes at
rates of 0.62 cm/yr [Freymueller et al., 1993; Egbue and Kellogg, 2010]. Late MiocenePliocene uplift,
recorded by shallowing upward trends in the Ecuadorian fore arc, led Daly [1989] to propose that the
collision started at ~8 Ma. Pedoja [2003] and Cantalamessa and Di Celma [2004] independently postulated
that ridge subduction began at the Pliocene-Pleistocene boundary (1.8 Ma), based on analyses of marine
terrace uplift. Lonsdale and Klitgord [1978] proposed that the Carnegie Ridge arrived at the trench about
1 Ma, based on interpretation of magnetic anomalies and bathymetry of the Cocos and Nazca plates. As
previously mentioned, Spikings et al. [2001, 2010] utilize the variation in cooling ages and exhumation
depths to suggest the ridge collided at 15 Ma.
Hungerbhler et al. [2002] report that the coastal block (west of the Peltetec Fault in Figure 1) started
migrating approximately northward at 15 Ma, resulting in the opening of marine basins (e.g., the Loja, Nabon,
and Cuenca basins), which have since been uplifted by the Eastern Cordillera [Hungerbhler et al., 2002].
Gutscher et al. [1999] postulate that the arrival of the Carnegie Ridge at the trench initiated displacement of
the coastal Block. Furthermore, the spatial coincidence of the variation in exhumation depths since 15 Ma
[Spikings et al., 2001, 2010] shows a remarkable spatial correlation with the subducted extent of the Carnegie
Ridge. Therefore, in this review we assume that the Carnegie Ridge initially collided with the trench at 15 Ma.
The Carnegie Ridge hosts signicant relief, and therefore, we propose that reactivation of the crust after
15 Ma was a consequence of subducting topographically prominent segments of the ridge, creating marine
terraces [Pedoja, 2003; Cantalamessa and Di Celma, 2004] and opening the intraarc rift of the Interandean
depression at ~5.5 Ma [Spikings and Crowhurst, 2004].
3.2. The Nazca Ridge
The ~1000 km long Nazca Ridge (Figure 4) forms part of the Nazca Plate, stands ~1.5 km above the surrounding
sea oor, and rst impinged on the western South American Plate at some point during 1115 Ma at
~1011S (central Peru) [Hampel, 2002; Rosenbaum et al., 2005]. The ridge is oriented oblique to the
trench and plate convergence direction, has migrated southward along the active plate boundary at a
rate of ~43 mm/yr, and is currently located at 1516S, where it is subducting beneath the northern
Arequipa Terrane with a ridge width of ~200 km. The coastal fore arc and Western Cordillera of the upper
plate consists of Cretaceous granitoids of the Coastal Batholith, which intrude isotopically juvenile crust
between 7 and 14S [Polliand et al., 2005], and Sunsas-aged (~1Ga) and Ordovician-Jurassic metamorphic
and magmatic rocks in the Arequipa Terrane.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics
Apatite FT

Time (Ma)

120

10.1002/2013TC003425

20

Nazca
Ridge

Region never
subducted by the
Nazca Ridge

North of the Nazca Ridge


HePRZ

C
100

80

APAZ
5

40
20

Above the Nazca Ridge


HePRZ

Nazca
Plate

APAZ

10

CB

100

Apatite (U-Th)/He

PERU

20

South of the Nazca Ridge


HePRZ

C
20

43mm/a

100

Ri
dg
e

40

APAZ

15

AT

zc
a

Region never
subducted by the
Nazca Ridge

Na

Time (Ma)

60

0
0

200

400

600

800

1000

140 120 100

1200

80

60

40

20

Time (Ma)

Southward distance along the coast


from the Ecuadorian border (km)

80

75

70

Figure 4. Apatite (U-Th)/He and ssion track data acquired from granitoids dispersed along the Peruvian coast [from Wipf et al., 2008]. Uncertainties on the apatite
(U-Th)/He dates are 10%. Migration rate of the Nazca Ridge is taken from Hampel [2002]. AT, Arequipa Terrane; CB, Coastal Batholith.

Wipf et al. [2008] report apatite (UTh)/He and FT ages of 149 Ma and 2880 Ma, respectively, from
Precambrian and Early Cretaceous crystalline rocks located along the coastline above the subducted Nazca
Ridge and from northern coastal Peru where the ridge has previously subducted (Figure 4). These ages are
younger than those obtained from river valleys located southeast of the Nazca Ridge, where the ridge has not
yet subducted, and indicate that elevated rates (0.7 km/Myr) of exhumation (assuming a geothermal
gradient of 30C/km; Table 1) were driven by ridge subduction, which removed 2 km of overburden in
response to changes in base level during surface uplift. These low exhumation depths, compared with the
upper plate above the Carnegie ridge (~8 km), corroborate the combined effects of the lower erosive power
of the climate along coastal Peru, and the short time period that the ridge subducts any given region along
the margin (~3.5 Ma) [Hsu, 1992; van Hunen et al., 2002] during its SE-directed displacement [e.g., Gutscher
et al., 2000; Hampel, 2002] relative to South America. Current annual rainfall along coastal Peru, including the
Arequipa Massif, is less than 5 cm/yr, and arid conditions have persisted throughout the Cenozoic [e.g.,
Strecker et al., 2006], limiting the erosive power of the climate relative to wetter climates in Ecuador and
Colombia. Pleistocene marine terraces crop out at a maximum elevation of ~800 m along coastal Arequipa
[Hsu, 1992; Machar and Ortlieb, 1992] above the present-day zone of subduction of the crest of the Nazca
Ridge, and Hsu [1992] recorded surface uplift rates of ~0.3 km/Myr over the previous 125 ka along coastal
Arequipa. Clearly, ridge subduction is driving rock uplift in that location, although the erosion rates are low,
permitting substantial surface uplift.
3.3. The Cocos Ridge
The ~1000 km long Cocos Ridge, located in the northwestern Panama Basin, forms part of the oceanic Cocos
Plate (Figure 2) and is subducting beneath southern Costa Rica at the Middle America Trench. Estimates of the
time of the onset of subduction of the ridge (a detailed discussion is presented in Morell et al. [2012]) range
from 85 Ma [e.g., Abratis and Wrner, 2001] to 5.53.0 Ma [Grfe et al., 2002], to as young as 10.5 Ma
[Lonsdale and Klitgord, 1978; Gardner et al., 1992]. More recently, Morell et al. [2012] combined ssion track
and (U-Th)/He data with plate reconstruction models to suggest that the onset of collision occurred within
the last 3 Ma. Similarly, Gardner et al. [2013] prefer an age of 31.5 Ma, based on modern plate reconstructions
and the distribution of the youngest crystallization ages of arc plutons. These dates are consistent with rapid
surface uplift along the outer fore arc during >2.5 Ma2.0 Ma [Vannucchi et al., 2013].
The Cocos Ridge originates close to the Cocos-Nazca spreading center [Barckhausen et al., 2001], at the
Galapagos Hotspot, has a width of 250500 km, a maximum crustal thickness that probably exceeds 21 km
and an overall density that is slightly lower than normal oceanic crust due to a thickened lower density upper

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

Tectonics

10.1002/2013TC003425

crust [Meschede and Frisch, 1998; Walther, 2003], and elevated geothermal gradients compared to the host
lithosphere [Grevemeyer et al., 2004]. The buttressing rocks of Costa Rica are oored by thickened oceanic
crust of the Caribbean Plate, which have been modied by arc magmatism since the Oligocene [Drummond
et al., 1995], giving rise to the Middle America Volcanic Arc. The Talamanca Mountains of southern Costa Rica
lie directly inboard of where the ridge collides with the upper plate (Figure 2), and the highest elevations
(~3800 m) are aligned directly inboard of the crest of the ridge, suggesting a cause and effect relationship.
Similarly, the uvial channels which drain the Talamanca Mountains are signicantly steeper than within the
adjacent highlands of the Cordillera Central of Panama, where the Cocos Ridge is not being subducted,
indicating that rock uplift rates are higher above the Cocos Ridge [Morell et al., 2012]. Refraction [Stavenhagen
et al., 1998] and seismicity [Protti et al., 1994, 1995] data reveal no evidence that the subducted Cocos Ridge
resides on a at, horizontal slab. However, the seismicity data do show that the slab dip reduces from ~80 in
the northwest (beneath Nicaragua), where the Cocos Ridge does not subduct, to 60 beneath Central Costa
Rica and perhaps to as shallow as 10 beneath southern Costa Rica [Protti et al., 1994, 1995].
Grfe et al. [2002] utilized ssion track analyses of high-Cl apatites extracted from middle Miocene
granitoids of the Talamanca Mountains in southern Costa Rica to propose that the rocks of the Talamanca
Mountains have cooled from ~170C to surface temperatures (20C) since ~4 Ma, via erosional exhumation
that was a consequence of rock uplift, with the fastest cooling and exhumation rates occurring between 4
and 3 Ma. However, Morell et al. [2012] suggest that high cooling rates prior to 3 Ma are due to thermal
relaxation of the shallow magmatic intrusions. The thermal history models of Grfe et al. [2002] reveal 50C
of cooling since 3 Ma, which is equivalent to 1.52 km of exhumation (Table 1). These values are in
agreement with the interpretation of Morell et al. [2012], who report 12 km of erosional exhumation since
3 Ma. These same authors suggest that rock uplift was driven by (i) increased compressive stress that
formed during the collision of the ridge with the thick Caribbean Plate and (ii) increased positive buoyancy
forces on the upper plate.
Similarly, surface uplift and exhumation in the fore arc of Costa have been documented by numerous studies,
all of which conclude that elevated rates were driven by the collision and subduction of the Cocos Ridge or
isolated seamounts. These include (Table 1) (i) Vannucchi et al. [2006], who report the exhumation of Eocene
ophiolitic rocks and an Eocene tectonic mlange, (ii) Corrigan et al. [1990], who report 800 m of shallowing of
offshore, sedimentary rocks along the outer fore arc of the Burica Peninsular (Figure 2) during the Late
Pliocene, which they attribute to the collision and subduction of the Cocos Ridge, (iii) Morell et al. [2011], who
report surface uplift rates that range between 2.1 0.1and 7.7 0.5 mm/a since 40 ka14 ka, respectively,
from the Burica Peninsular, which overlies the leading edge of the Cocos Ridge where the sea oor deepens
by 2 km toward the southeast across the Panama Fracture Zone (Figure 2), (iv) Gardner et al. [2001], who
report surface uplift rates of 6 mm/a across the Nicoya Peninsular (Figure 2), in response to the collision and
subduction of seamounts, (v) Sak and Fisher [2004], who combine 14C dates of fossiliferous sedimentary rocks
and paleo sea level information to show that the Osa Peninsular (outer fore arc; approximately 20 km inboard
of the Middle America Trench) has undergone surface uplift at a high average rate of >6 mm/yr over the
previous 32 kyr, due to the subduction of bathymetrically prominent seamounts that form part of the Cocos
Ridge, and (vi) Gardner et al. [2013], who report surface uplift rates of 1.78.5 mm/a within the Osa Peninsular
over the previous 833 7 Ka, respectively, determined by a combination of 14C and optically stimulated
luminescence dating. These last observations are consistent with shallowing along the Tonga-Kermadec
subduction zone, and the Izu-Bonin trench, adjacent to subducting seamount chains [e.g., Scholz and Small,
2013]. Morell et al. [2011] conclude that surface uplift was caused by changes in crustal thickness due to the
subduction of the thick crust of the Cocos Ridge.

4. Discussion
4.1. Quantifying Rock Uplift
Pleistocene (2.5 Ma) marine terraces on the upper plate, located directly above the region of subduction
of the Nazca Ridge, are currently perched at an altitude of ~800 m [Hsu, 1992], and thermochronological
data suggest that up to 2 km of exhumation may have occurred in local valleys [Wipf et al., 2008],
indicating that a maximum estimate of rock uplift above the Nazca Ridge is ~2.8 km (Table 1). In contrast,
Pleistocene marine strata located along the coastline to the north and south of the Nazca Ridge are

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

10

Tectonics

10.1002/2013TC003425

either submarine or reside close to sea level. Utilizing apatite ssion track and (U-Th)/He data [Wipf et al.,
2008], we calculated a time-averaged (3 Ma) maximum rock uplift rate of ~0.9 km/Myr (Table 1) above the
region of the Nazca Ridge, since the Nazca Ridge subducted that region.
Sedimentary rocks do not constrain the quantity of surface uplift that occurred in the northern Eastern
Cordillera of Ecuador, or the southern Central Cordillera of Colombia between 15 and 13 Ma, due to the
collision of the Carnegie Ridge. However, the collision event drove ~6 km of exhumation during 1513 Ma,
constraining rock uplift to 6 km, yielding a high rock uplift rate of 3 km/Myr (Table 1) [Spikings et al., 2010].
A combination of geomorphological observations and apatite ssion track and (U-Th)/He data from the
Talamanca mountains suggests the Cocos Ridge has exhumed 1.52 km of crust through a geothermal
gradient of 30C/km, suggesting that total rock uplift has probably exceeded 1.52 km since 3 Ma (Table 1)
[Grfe et al., 2002]. Rock uplift rates were >0.6 km/Myr, which is lower than those obtained from the above
the Carnegie Ridge. However, it is likely that this rate is inaccurate and low because precise estimates for the
timing of collision of the ridge are lacking. The rate has been averaged over a period of 3 Ma, whereas
collision may have occurred after 3 Ma [Morell et al., 2012]. Furthermore, the quantity of surface uplift since
3 Ma in the Talamanca Mountains is certainly more than zero, while Quaternary surface uplift rates in the
outer fore arc range between ~6 and 19 8 mm/a [Gardner et al., 2001; Sak and Fisher, 2004; Morell et al., 2011]
(see section 4.3). These calculations assume that the Carnegie and Cocos ridges have contributed to rock
uplift of northern Ecuador and the Talamanca Mountains since 15 Ma and <3 Ma, respectively [Spikings et al.,
2001; Morell et al., 2012].
The collision of the Caribbean Plateau with northwestern South America drove a minimum of 67 km of rock
uplift during 7570 Ma along the entire paleomargin between ~5S and ~7N (present-day coordinates;
Figure 2), at a rate of >1.21.5 km/Myr [Spikings et al., 2001, 2010] (Table 1). Collision at ~75 Ma had a
profound impact on the South American Plate, because (i) it drove 69 km of rock uplift along ~1350 km of
continental margin during 7560 Ma, (ii) it was probably responsible for the present exposure of amphibolite
grade rocks in the cordilleras of the Northern Andes (e.g., the Totoras Amphibolite of the Western Cordillera
of Ecuador) [Vallejo et al., 2009], (iii) it gave rise to a hydrocarbon rich, continental retroforeland depocenter
(the Amazon Foreland Basin; Figure 2) via crustal loading, and (iv) it switched the prevailing drainage
direction across the northern South American Plate from east to west, to west to east, forming the primordial
phase of the Amazon catchment area.
4.2. Factors Driving Rock Uplift at Subduction Zones
The Caribbean oceanic plateau and the Carnegie, Nazca, and Cocos ridges all enhanced rock uplift along the
margins of the South American continent, and basement regions of Central America, which commenced
within 1 Ma of the start of collision. The amount of rock uplift is determined by (i) buoyancy of the subducted
oceanic crust, (ii) ridge topography, (iii) plate coupling and the strength of the subduction interface, (iv)
erosion rates, and (v) the preexisting rheology and structure of the buttressing continental crust, which may
cause strain partitioning along zones of relative weakness.
4.2.1. Buoyancy, Flat-Slab Subduction, and Accretion
Oceanic plateaus and ridges decrease the negative buoyancy of the subducting lithosphere, and thus, they
reduce the slab-pull force that ultimately drives subduction [Cloos, 1993]. Subduction of relatively buoyant
oceanic crust may be manifest as an along-strike reduction in the dip of a downgoing slab [Gerya et al., 2009],
as has been suggested in the central Andes where the Nazca ridge resides on a slab that is subducting with a
at conguration between 3S and 20S [e.g., Barazangi and Isacks, 1976; Pilger, 1981; Gutscher et al., 2000]. A
reduction in slab dip can increase plate coupling, leading to increased and landward migrating
compressional deformation of the continental crust and rock uplift [e.g., Ramos et al., 2002; LaFemina et al.,
2009]. However, the Carnegie Ridge is not hosted by a at slab, despite the fact that it is relatively buoyant
compared to midocean ridge-derived crust and occurs within younger lithosphere than that associated with
the Nazca ridge. Espurt et al. [2008] suggested that the Carnegie Ridge is causing its slab to atten and that
this process can take as long as 7 Ma. However, Spikings et al. [2001] present thermochronological and
sedimentological data that suggest the Carnegie Ridge collided at 15 Ma, and hence, there is no unequivocal
evidence that the slab is attening. Similarly, the Iquique Ridge, which subducts beneath the South American
Altiplano (Figure 1), is associated with a steep (Nazca) slab. Skinner and Clayton [2013] recently utilized
bathymetric data to conclude that there is very little correlation between the subduction of plateaus and

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

11

Tectonics

10.1002/2013TC003425

ridges, and historic zones of at subduction in South America. Tassara et al. [2006] utilized a threedimensional density model of the Andean margin to suggest that the positive buoyancy of subducted ridges
with respect to the surrounding slab is a necessary but solely insufcient condition for causing slab attening.
ODriscoll et al. [2012] conclude that a combination of factors, including enhanced buoyancy, continental
overriding, and enhanced suction above the slab are required to cause slab attening. Finally, Gerya et al.
[2009] utilize numerical modeling to show that attening of a slab by enhanced positive buoyancy can only
occur if the slab is already at a threshold critical state. Clearly, at-slab subduction does not always occur
when aseismic ridges are subducted, and it is not a prerequisite to drive high rates of rock uplift within the
continental lithosphere over periods of a million years or more (Table 1).
The dip of the slab that hosts the Cocos Ridge reduces from 60 to >10 from the northwestern to the
southeastern limits of subduction of the Cocos Ridge (northern and southern Costa Rica, respectively)
[Protti et al., 1994, 1995]. Lowering of the slap dip coincides with the presence of a fold and thrust belt
within the inner fore arc (Fila Costea Thrust Belt) and back arc (Limon back arc) within Costa Rica [Fisher
et al., 2004; Sitchler et al., 2007; LaFemina et al., 2009] suggesting that plate coupling increases as the slab
dip is reduced. Therefore, plate coupling may contribute more to the forces that drive rock uplift if the
introduction of an aseismic ridge to the trench causes the slab dip to reduce. However, Protti et al. [1994]
suggest that changes to the slap dip beneath Costa Rica are a result of trench-parallel changes in the age of
the subducted Cocos Plate and are not caused by the subduction of the Cocos Ridge. Elevated surface
uplifts in the outer fore arc (Osa and Burica Peninsulars) of Costa Rica are the result of subducting
bathymetric relief, and not by plate-boundary friction [Gardner et al., 2013].
In extreme cases, when the oceanic crust is especially thick (>17 km) and laterally extensive and the
lithosphere is relatively young (<10 Ma), relatively buoyant oceanic ridges or plateaus may inhibit subduction
[Cloos, 1993; Mason et al., 2010] and cause major plate reorganization. This may cause obduction of oceanic
crust onto the active margin and migration of the subduction zone, as may have occurred during accretion of
the Caribbean Plateau onto the northwestern edge of South America during the late Cretaceous [Spikings
et al., 2001]. The Caribbean plateau approached South America from the west [e.g., Pindell et al., 2005], and
the intervening remnant oceanic basin was consumed by west-dipping subduction beneath the plateau
[Vallejo et al., 2009]. This geometry favored accretion of the buoyant plateau leading to enhanced uplift of the
adjacent continental margin, during compression.
4.2.2. Ridge Topography
Even if buoyant oceanic plateaus and ridges are unable to block subduction or signicantly modify the dip
of the downgoing slab, their subduction will generate additional rock uplift as the excess topography
associated with a ridge is underthrust beneath the overriding continental margin [Cloos, 1993]. The
magnitude of this rock uplift is expected to be similar to the relief between the ridge or plateau top and the
adjacent sea oor, which is typically ~2 km. The fact that most ridges and plateaus have widths (>100 km)
that are at least as large as the likely exural wavelength of the overthrust margin [~100 km; Turcotte and
Schubert, 1982] makes it unlikely that this uplift can be resisted by the elastic rigidity of the overlying plate.
For the same reason, it is unlikely that deformation in the overriding plate will extend laterally far beyond
the limits of the subducted ridge/plateau. This is consistent with the lack of gravity anomalies where the
Cocos, Carnegie, and Nazca ridges pass beneath the continental margin. Thus, it seems likely that the
overthrust plate will be locally more deformed and uplifted as the topographic step associated with
buoyant oceanic crust is subducted beneath the leading edge of the margin, as observed in analogue
models [Lallemand et al., 1992; Hampel et al., 2004].
4.2.3. Plate Coupling
The Eastern Cordillera of Ecuador and the Central Cordillera of Colombia were ~200 km east of the trench
when the Carnegie Ridge collided at 15 Ma. Consequently, it is unlikely that the ridge topography alone alone
can account for rock uplift of the Andes of Ecuador and Colombia at 15 Ma, and it is likely that some
component of rock uplift can be attributed to enhanced plate coupling.
LeFevre and McNally [1985] and LaFemina et al. [2009] show that the impingement of aseismic ridges in the
Middle American Trench gave rise to increased plate coupling, seismicity, and crustal deformation.
Differential motions between the subducting and overriding plates generate shear stresses along the plate
interface that inuence the shape and uplift of the deforming upper plate [Dahlen, 1990]. The presence of
subduction zone earthquakes indicates that deformation at upper crustal levels is controlled by frictional

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

12

Tectonics

10.1002/2013TC003425

sliding. In this case, shear stresses are governed by the product of the effective normal stresses across the
subduction zone interface and the friction coefcient. The mean shear stress on a subduction zone interface
is unlikely to vary signicantly due to variations in friction coefcients, since most rocks have nearly uniform
values of ~0.6 [Byerlee, 1978]. However, the presence of sequences of overpressured sediments blanketing
the oceanic crust can play an important factor in reducing the shear stresses on the subduction zone
interface (due to reduction of the effective normal stress). Lamb and Davis [2003] argued that the presence/
absence of weak sediments linked to regional climatic differences exerted a rst order control on
topography in the Andes during the Cenozoic. Arid and wet climatic zones along western South America vary
on length scales of >1000 km (e.g., arid and drier zones extend between 5S and 30S), and (hyper)arid zones
contribute a relatively low volume of continent-derived sediment input to the trench, increasing plate
coupling, which can support higher mean maximum elevations. We believe that similar plate coupling effects
could be important during the subduction of prominent oceanic topography. Because oceanic ridges are
elevated, they are likely to be sediment starved relative to adjacent oceanic crust, leading to increased plate
coupling when subduction eventually occurs. In addition, oceanic ridges typically contain numerous volcanic
seamounts that are expected to act as high shear stress asperities (due to locally high normal stresses) during
subduction. Both possibilities are consistent with bathymetric data which clearly show that the Cocos,
Carnegie, and Nazca ridges are visually rougher than adjacent, deeper sea oor. Thus, we expect that the
sliding interface in the region of subducting oceanic ridges and plateaus could be locally relatively strong,
which would lead to additional rock uplift in the overlying plate relative to neighboring regions.
Espurt et al. [2008] used analogue modeling to suggest that slab attening increases the intraplate friction
force and migration of shortening and increased elevations in the upper plate. However, there is no
correlation between the rates and quantities of rock uplift and slab dip, and we discount slab dip as a
dominant control on rock uplift rates within continental arcs that are subducted by aseismic ridges.
The potential for sediment to modify the friction strength of the subduction interface raises the possibility of
complex interactions between erosion, climate, sedimentation, and deformation [Simpson, 2010]. For
example, enhanced rock uplift and erosion rates above subducting ridges will increase sediment supply
offshore. However, most sediment delivered to the trench is expected to migrate longitudinally off the axis of
the elevated oceanic ridge to deeper water (e.g., see the analogue models of Hampel et al. [2004]). By this
process, off-ridge portions of the subduction interface are expected to weaken relative to the interface in the
region of the ridge, which will drive further rock uplift and erosion where ridge subduction is occurring,
leading to positive feedback.
4.2.4. Rock Uplift Driven by Erosion
Erosion induces rock uplift due to isostatic adjustment. Depending on the densities of the crust and mantle,
approximately 800 m of rock uplift occurs for every 1000 m of erosion, assuming local isostasy (i.e., the surface is
only lowered by 200 m even though 1000 m of erosion has taken place) [e.g., Molnar and England, 1990].
Although this effect is counteracted by the exural rigidity of the lithosphere, isostasy will normally prevail if
erosion occurs over length scales exceeding ~100 km [Turcotte and Schubert, 1982]. However, even in the
absence of isostatic rebound (for example, if erosion occurs on short length scales), erosion may still strongly
inuence patterns of rock uplift if the crust deforms plastically. This is because erosion reduces the gravitational
work associated with rock uplift, which tends to localize further deformation and rock uplift [Simpson, 2004].
The Cocos, Carnegie, and Nazca ridges all have similar bathymetric relief (~2 km) relative to the directly
adjacent oceanic oor. Despite this, rock uplift rates were approximately 3 times higher in the upper plate
above the Carnegie Ridge (>3 km/Myr) compared to the Nazca Ridge (~1 km/Ma; Table 1). We hypothesize
that these differences in rock uplift rate could be due to the inuence of regional climatic variations on
erosion rates in the upper plate (see section 4.4). Erosion rates are low where the Nazca ridge is currently
being subducted due to the arid climate that probably persisted throughout much of the Cenozoic [Lamb
and Davis, 2003]. Thus, most uplift in this area is probably either related to underthrusting of excess ridge
topography or related to changes in plate coupling. In contrast, erosion rates were relatively elevated in the
upper plate above the Carnegie Ridge (Table 1), linked with the subtropical climate. These elevated erosion
rates would have assisted in uplifting rocks toward the surface, which would have been amplied above
subducting ridges. Rock uplift rates in the upper plate above the Cocos Ridge may have been as low as
0.6 km/Myr, which is inconsistent with elevated rates in a tropical, wet climate. However, we consider this
value to be inaccurate and too low, given the ambiguity regarding the timing of collision of the Cocos Ridge.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

13

Tectonics

10.1002/2013TC003425

Continental crust above the Carnegie Ridge in northern Ecuador has been exhumed by 8 km since the ridge
collided with South America at 15 Ma (Table 1). We estimate that erosion would have driven 6.5 km of rock
uplift. Surface uplift is unconstrained, so we estimate that total rock uplift is 8 km, and the 1.5 km decit can
be attributed to the topographic relief between the ridge top and the adjacent sea oor, which varies
between 1 and 2 km along the length of the ridge (Figure 2). Similar calculations for the Talamanca Mountains
and the northern Arequipa Terrane results in decits in rock uplift of >300 m and >1100 m, respectively,
which can also be accounted for by the elevation of the Cocos and Nazca ridges, respectively.
4.2.5. Preexisting Structure of the Upper Plate
Large differences existed in the precollisional structures of the Ecuadorian and Peruvian margins, which
would have inuenced how they responded to aseismic ridge collision and subduction. An extensive,
serpentinized Campanian terrane suture (Figure 2; Peltetec Fault) separates accreted, competent rocks of the
Caribbean Large Igneous Province from the Late Cretaceous, continental paleomargin of Ecuador (Eastern
Cordillera) and Colombia (Central Cordillera). Rock uplift during 1513 Ma may have been focused along the
Ecuadorian-Colombian border region via dextral transpressive displacement of the mechanically weak
suture, preferentially accommodating displacement related to collision and subduction of the Carnegie
Ridge [e.g., Spikings et al., 2010]. The trend of the suture bends toward a northerly strike at ~ S3 (Figure 2) and
dextral migration of the allochthonous crust toward the N-NE along the suture [Hughes and Pilatasig, 2002]
opened the Gulf of Guayaquil and removed crustal support for the rocks of the southern Eastern Cordillera of
Ecuador, resulting in subsidence in that region commencing at 15 Ma [Hungerbhler et al., 2002].
The Peruvian margin has not experienced terrane accretion since the opening of the Iapetus Ocean at ~550 Ma
[e.g., Cawood, 2005], and the Nazca Ridge has impacted and subducted approximately homogeneous crystalline
rocks with ages varying between Late Cretaceous and Neoproterozoic since 1511 Ma [Hampel, 2002;
Rosenbaum et al., 2005], and it is unlikely that convergence was accommodated by an individual fault system.
The preexisting structure of Ecuadorian crust may have focused vertical displacement to within a restricted
volume of crust, giving rise to higher rates of rock uplift than what was achieved within the adjacent, more
structurally homogeneous Peruvian crust. This process may also be partly responsible for signicantly higher
quantities of rock uplift and exhumation 200300 km inboard of the trench, within the cordilleras of Ecuador and
Colombia, relative to small increases in elevation (e.g., 360 m) [Pedoja et al., 2006] along the coastline.
4.2.6. The Relative Contribution of Individual Driving Forces
Underthrusting of excess ridge topography (12 km) combined with erosion can almost completely account
for the total rock uplift experienced above ridges that are subducted as part of the Nazca and Cocos plates.
This conclusion is consistent with that proposed by Gardner et al. [2013], who combine structural mapping
with 14C and optical luminescence dating to show that surface uplift of the outer fore arc above the Cocos
Ridge was dominated by bathymetric relief related to the ridge and not by plate-boundary friction.
Thermochronological data suggest that the onset of elevated exhumation rates is synchronous with the
onset of subduction of aseismic ridges, indicating that rock uplift is an immediate response to collision and
early subduction. Plate coupling appears to play a minor role compared to ridge topography and erosion,
although it may increase rock uplift rates further inland via strain partitioning into zones of structural
weakness during the initial stages of collision and subduction.
4.3. Comparison of Rock Uplift in Fore Arcs and Arcs
Unfortunately, we are not aware of any data that can be used to determine coeval exhumation rates over
periods of millions of years from an arc and its fore arc, and thus, we cannot provide a general set of rules
regarding the relative response of different regions of the arc to the collision of aseismic ridges. Gardner et al.
[2001], Sak and Fisher [2004], and Morell et al. [2011] report very high surface uplift rates of ~619 8 mm/a
within the fore arc of Costa Rica (Osa and Nicoya peninsulars), above the subducting Cocos Ridge. Rock uplift
rates should be even higher in the presence of erosion, which would exceed most rock uplift rates reported
from arcs inboard of aseismic ridges (Table 1). Similarly high surface uplift rates (up to 7.5 0.7 mm/a) are
reported from the fore arc of the Kamchatka-Kuril arc [Panz et al., 2013]. Quaternary elevated marine
terraces, which reach elevations of 800 m along coastal Peru [Machar and Ortlieb, 1992] and up to 360 m
along the coast of central Ecuador [Pedoja et al., 2006], also indicate rapid recent surface uplift in fore-arc
regions. However, these high surface uplift rates are averages for the last 30,000 to 1000 years, and rock uplift
rates reported from arcs are typically averaged over periods longer than 1 Ma. Given that stress accumulates

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

14

Tectonics

10.1002/2013TC003425

over long time scales compared to the high rate at which it is released (during earthquakes), it is not
surprising that rock uplift rates can be extremely high over relatively short time scales [e.g., Morell et al., 2011].
Despite the high, transient surface uplift rates, fore arcs above aseismic ridges do not typically achieve
the same elevations as adjacent arcs, implying that surface and rock uplift rates are actually lower in the
fore arc when averaged over a million years or more.
Higher rock uplift rates in arcs compared to fore arcs over periods of a million years or more may be a
consequence of several different factors. First, the elevated rates observed for fore arcs must exist only for
relatively short time periods (< 1 Ma) as ridges are underthrust beneath the outer margin of the upper
plate, while it is likely that these rates will reduce as the topographic step associated with the ridge passes
further downdip. Second, rock and surface uplift in the arc is likely to be inuenced not only by ridge
subduction but also by crustal thickening linked to magmatism [e.g., Atherton and Petford, 1993], crustal
shortening [e.g., Ramos et al., 2002], and the underplating of material derived from the trench or fore arc [e.g.,
Clift and Hartley, 2007].
4.4. Exhumation
4.4.1. Quantifying Exhumation and Erosion
No evidence exists within the northern Ecuadorian Eastern Cordillera, southern Colombian Central Cordillera,
or directly above the present-day subducted extent of the Nazca Ridge for signicant extension within the
emergent continental margin during the collision of the Caribbean Plateau and aseismic ridges. Therefore,
we consider all exhumation to be a consequence of erosion occurring as a result of changes in base level
during rock uplift, and thus, the quantity of exhumation partly depends on climate and the life span of the
forces that support the uplifted crust.
The collision of the Carnegie Ridge at 15 Ma drove 6 km of exhumation in the northern Eastern Cordillera of
Ecuador during 1513 Ma, at a rate of 3 km/Myr [Spikings et al., 2010], resulting in the erosion of >50000 km3
of continental crust between 0.5N and 2N. The Talamanca Mountains overlie the subducted extent of the
Cocos Ridge in an arc position and have experienced 5 km of exhumation during the previous 4 Myr [Grfe
et al., 2002], yielding an average exhumation rate close to 1 km/Myr and the erosion of ~38000 km3 of
continental crust. Exhumation rates were as high as ~3.7 km/Myr, during and immediately after the initial
collision (43 Ma; Table 1). Models for the thermal history [Wipf et al., 2008] show that the Nazca Ridge has
exhumed 2 km of crust in the upper plate since it started subducting its current location beneath northern
Arequipa (Figure 3) at ~3 Ma, yielding a signicantly slower exhumation rate of 0.7 km/Myr. Furthermore,
elevated exhumation rates above the present-day subducted extent of the Nazca Ridge are spatially
restricted to river valleys that separate elevated regions whose apatite (U-Th)/He data reveal a cooling history
that predates the impact of the Nazca Ridge (Figure 4). In contrast, the entire land surface above the Carnegie
Ridge was at depths that were hotter than ~120C prior to the impact of the Carnegie Ridge (Figure 2).
Spatially restricted and low exhumation depths above the Nazca Ridge, compared with the upper plate
above the Carnegie ridges, corroborate the lower erosive power of the hyperarid climate in coastal Peru,
compared to high annual rainfall in more northern latitudes. Current annual rainfall along the region of
coastal Peru, including the Arequipa Massif (Figure 4), is less than 5 cm/yr, and arid conditions have persisted
throughout the Cenozoic [e.g., Strecker et al., 2006], lowering the erosive power of the climate relative to
wetter zones in the cordilleras of Ecuador and Colombia (12 m/yr), resulting in low erosive exhumation rates.
Southeastward-directed displacement of the buoyant Nazca Ridge along the Peruvian margin at a rate of
~43 km/Myr [Hsu, 1992; van Hunen et al., 2002], combined with a ridge width of ~200 km, suggests that the
ridge subducts any given region along the margin for ~3.5 Myr. However, both the Carnegie and Cocos
ridges subduct the Ecuadorian and Middle American trenches at almost orthogonal angles, which
combined with plate reconstructions of the Nazca and Cocos plates [e.g., Meschede and Barckhausen, 2000]
suggests that they have subducted the same arc crust since they started subducting at 15 and ~3 Ma,
respectively. The N-NE migration of the Western Cordillera and fore arc of Ecuador (Figure 2) suggests that
the Carnegie Ridge has progressively subducted more southerly regions of the Colombian and Ecuadorian
fore arc [e.g., Gutscher et al., 2000], although the thermochronological data utilized in this study were
collected from the Eastern Cordillera, which is not migrating with the fore-arc block. Rock uplift above the
Carnegie Ridge has exceeded 8 km since it started subducting, which signicantly exceeds 2.8 km of rock
uplift above the Nazca Ridge. Therefore, the life span of the forces that drive rock uplift appears to directly

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

15

Tectonics

10.1002/2013TC003425

relate to the duration of collision and subduction of an aseismic ridge, supporting the causal link between
rock uplift and aseismic ridges.
We hypothesize that the life span of forces that cause rock uplift at any xed location on the upper plate
above the Nazca Ridge have provided insufcient time to exhume rocks that were deeper than ~2 km.
Southeastward displacement of the Nazca Ridge generates a wave of rock uplift along the Peruvian margin,
although its life span at any given location is too short to exhume rocks deeper than ~2 km by erosion, within
the hyperarid climate.
4.4.2. Exhumation Patterns and the Preservation of ore Deposits Along Active Margins
Rosenbaum et al. [2005] show that the temporal and spatial distribution of porphyry-related Miocene deposits in
Peru and northern Chile corresponds with the arrival of aseismic ridges and plateaus at the subduction zone. This
may occur due to mobilization of fertile magmas from the MASH zone, leading to the exsolution of metal-rich
uids at shallow levels. Alternatively, metalliferous uids may be derived from melting of young, at slabs, forming
adakitic magmas, which are frequently associated with porphyry deposits. [e.g., Chiaradia et al., 2009]. The
preservation of numerous bonanza-grade porphyry-related deposits along the Western Cordillera of Peru (e.g.,
Yanacocha) can be attributed to low amounts of erosional exhumation.
The erosion of 5.5 km of crust from the Late Cretaceous paleomargin of Colombia and Ecuador between 2N
and 0, since 15 Ma renders the region barren to porphyry-related deposits that formed due to the
emplacement of plutons in the upper crust prior to 15 Ma, and no porphyry deposits older than 15 Ma have
been documented in that region. Conversely, we predict that any porphyry-related ore deposits that formed
along the pre-15 Ma arc of the Northern Andes, south of 0 and north of 2N, have not been removed by
erosion. This hypothesis is consistent with the presence of bonanza-grade porphyry-related gold deposits in
southern Ecuador that are associated with Jurassic plutons (e.g., Fruta del Norte; Figure 2). Furthermore, we
suggest that middle Miocene strata in the proximal Amazon Foreland Basin regions of the northern
Ecuadorian and southern Colombian Andes (the sub-Andean zones) are prospective for placer gold, due to
extensive unroong of the metalliferous Jurassic arc via the erosion of >50000 km3 of crust (Table 1).
Collision and accretion of the laterally extensive Caribbean Plateau drove exhumation rates of >1 km/Myr
along the entire Eastern Cordillera of Ecuador and Central Cordillera of Colombia during 7570 Ma. The
approximately symmetrical distribution of coarse-grained molasse to the east (Tena Fm.) and west (Yunguilla
Fm.) [Vallejo et al., 2009] of the Eastern Cordillera of Ecuador suggests that exhumation was accompanied by
surface uplift. The removal of >6 km of continental crust [Spikings et al., 2010] brought medium grade
metamorphic rocks to the surface (e.g., amphibolites of the Agoyan Formation of the Eastern Cordillera of
Ecuador), providing a metamorphic (kyanite and staurolite) detrital component for the foreland molasse of
the Oriente (Ecuador) and Magdalena (Colombia) basins. Exhumation over distances of ~1375 km along strike
due to collision of the laterally extensive Caribbean Plateau eroded ~300,000 km3 of crust during 7560 Ma,
assuming that a 30 km wide corridor (approximate average width of the Eastern Cordillera of Ecuador and
Central Cordillera of Colombia) was exhumed along the margin, and probably eroded some preexisting
Jurassic, porphyry-related metalliferous deposits. Therefore, we predict that Campanian-Maastrichtian
sedimentary rocks in the contemporary foreland basins may also host placer deposits.

5. Conclusions
A combination of thermochronological data and eld observations from along the western margin of the
South American Plate and Central America suggests that
1. The collision and subduction of topographically prominent (e.g., typically ~12 km elevation above the
adjacent sea oor) and thick (crustal thicknesses up to 35 km) aseismic oceanic plateaus and volcanic
ridges with continental plates signicantly increases rock uplift rates in the upper plate. The collision of
large oceanic plateaus can drive at least 6 km of rock uplift within 5 Ma of the collision event, giving
rise to some of the highest coastal mountain ranges on Earth. The subduction of aseismic ridges can
also drive 6 km of rock uplift within 2 Ma, leading to elevated topography over trench-parallel
distances of ~200 km.
2. The enhancement of rock uplift above subducting ridges and plateaus is probably dominated by the
combined effects of anomalous ridge topography and surface erosion of the upper plate. Rock uplift rates
can be augmented by an increase in the strength of the subduction zone interface related to the presence

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

16

Tectonics

3.

4.

5.

6.

Acknowledgments
The authors are grateful for the assistance provided by W. Winkler, J-P. Burg,
B. Beate, D. Seward, C. Vallejo, and G.
Ruiz during eld work in Ecuador, and A.
Kammer, D. Villagomez, A. Mora, A.
Beltran, and L. Quiroz during eld work
in Colombia. The manuscript was
improved by the thorough and helpful
reviews of J. Kellogg and P. Vannucchi.
M-A. Gutscher is thanked for comments
related to an early version of the
manuscript. Funds for the eld work in
Ecuador and Colombia were awarded to
R.S. by the Swiss NSF (200021100079
and 200021103438).

SPIKINGS AND SIMPSON

10.1002/2013TC003425

of volcanic seamount asperities and the lack of a highly overpressured sedimentary cover on the elevated
ocean oor.
Rock uplift occurs as an immediate response to collision, and the quantity of rock uplift does not appear
to be a function of slab dip, corroborating our observation that plate coupling does not provide the
dominant driving force for rock uplift.
The amount and spatial extent of exhumation that occurs is strongly dependant on climate-induced
erosion and the life span over which topography is uplifted and maintained, which is itself controlled by
the angle between the topographic feature being subducted and the coast line and by the plate
convergence direction. Sediments eroded from above a subducting ridge are likely to return to the trench
leaving the ridge axis sediment starved due to its elevated bathymetry. This will tend to increase shear
stresses on the subduction interface where ridge subduction is occurring relative to adjacent regions,
leading to a positive feedback between ridge subduction, rock uplift, erosion, and sediment supply.
Nevertheless, excess ridge topography and erosion can account for most of the total rock uplift experienced above ridges that are subducted.
The collision of plateau and aseismic ridges can induce signicant trench-parallel differences in rock
uplift, exhumation, and elevation, on length scales of perhaps 1000 km and 200 km, respectively.
Variations in exhumation have a direct impact on exploration paradigms for metalliferous deposits
because it is unlikely that porphyry-related ore deposits are preserved above zones that have been
subducted by aseismic ridges, if (i) the porphyritic intrusions predate the timing of collision and (ii) the
region has experienced temperate-tropical precipitation since collision.
Lamb and Davis [2003] conclude that mean maximum elevations in the Andes are dominated by trenchparallel variations in shear stress at the plate interface, which they correlate with climatic zones and
trench sedimentation over length scales of >1000 km. However, subduction of prominent ridge
topography, and the collision of oceanic plateaus, drives signicant trench-parallel variations in rockuplift and exhumation over much shorter length scales.

References
Abratis, M., and G. Wrner (2001), Ridge collision, slab-window formation, and the ux of Pacic asthenosphere into the Caribbean realm,
Geology, 29, 127130, doi:10.1130/0091-7613(2001)029<0127:RCSWFA>2.0.CO;2.
Acton, G. D., B. Galbrun, and J. W. King (2000), Paleolatitude of the Caribbean Plate since the Late Cretaceous, in Proceedings of the Ocean
Drilling Program, Scientic Results, vol. 165, edited by R. M. Leckie et al., pp. 149173, Texas A & M University, Ocean Drilling Program,
College Station, Tex.
Arias, A., and C. J. Morales (1999), General Geological Map of the Department of Cesar, explanatory notes, 89 pp., Ministry of Mining and
Energy, Ingeominas, Bogot.
Atherton, M. P., and N. Petford (1993), Generation of sodium-rich magmas from newly underplated crust, Nature, 362, 144146, doi:10.1038/
362144a0.
Barazangi, M., and B. L. Isacks (1976), Spatial distribution of earthquakes and subduction of the Nazca plate beneath South America, Geology,
4, 686692, doi:10.1130/0091-7613(1976)4<686:SDOEAS>2.0.CO;2.
Barckhausen, U., C. R. Ranero, R. von Huene, S. C. Cande, and H. A. Roeser (2001), Revised tectonic boundaries in the Cocos plate off Costa
Rica: Implications for the segmentation of the convergent margin and for plate tectonic models, J. Geophys. Res., 106, 19,20719,220,
doi:10.1029/2001JB000238.
Bourdon, E., J.-P. Eissen, M. Monzier, C. Robin, H. Martin, J. Cotton, and M. L. Hall (2002), Adakite-like lavas from Antisana Volcano (Ecuador):
Evidence for slab melt metasomatism beneath Andean northern volcanic zone, J. Petrol., 43, 199217, doi:10.1093/petrology/43.2.199.
Byerlee, J. (1978), Friction of rocks, Pure Appl. Geophys., 116, 615626.
Canales, J. P., G. Ito, R. S. Detrick, and J. Sinton (2002), Crustal thickness along the western Galapagos Spreading Center and compensation of
the Galapagos Swell, Earth Planet. Sci. Lett., 203, 311327, doi:10.1016/S0012-821X(02)00843-9.
Cantalamessa, G., and C. Di Celma (2004), Origin and chronology of Pleistocene marine terraces of Isla de la Plata and of at, gently dipping
surfaces of the southern coast of Cabo San Lorenzo (Manab, Ecuador), J. South Am. Earth Sci., 16, 633648, doi:10.1016/j.jsames.2003.12.007.
Cawood, P. A. (2005), Terra Australia Orogen: Rodinia breakup and development of the Pacic and Iapetus margins of Gondwana during the
Neoproterozoic and Paleozoic, Earth Sci. Rev., 69, 249279, doi:10.1016/j.earscirev.2004.09.001.
Cediel, F., R. P. Shaw, and C. Caceres (2003), Tectonic assembly of the Northern Andean Block, in The Circum-Gulf of Mexico and the Caribbean:
Hydrocarbon Habitats, Basin Formation, and Plate Tectonics, AAPG Mem., vol. 79, edited by C. Bartolini, R. T. Bufer, and J. Blickwede, pp. 815848,
Mem. Am. Assoc. Pet. Geol., Tulsa, Okla.
Cern, J. (2008), Crustal structure of the Colombian Caribbean Basin and margins, PhD dissertation, University of South Carolina, Columbia, 165 p.
Chiaradia, M., O. Mntener, B. Beate, and D. Fontignie (2009), Adakite-like volcanism of Ecuador: Lower crust magmatic evolution and
recycling, Contrib. Mineral. Petrol., 158, 563588, doi:10.1007/s00410-009-0397-2.
Clift, P. D., and A. J. Hartley (2007), Slow rates of subduction erosion and coastal underplating along the Andean margin of Chile and Peru,
Geology, 35, 503506, doi:10.1130/G23584A.1.
Cloos, M. (1993), Lithospheric buoyancy and collisional orogenesis: Subduction of oceanic plateaus, continental margins, island arcs,
spreading ridges, and seamounts, Geol. Soc. Am. Bull., 105, 715737, doi:10.1130/0016-7606(1993)105<0715:LBACOS>2.3.CO;2.
Cofn, M. F., and O. Eldholm (1994), Large igneous provinces: Crustal structure, dimensions, and external consequences, Rev. Geophys., 32,
136, doi:10.1029/93RG02508.

2014. American Geophysical Union. All Rights Reserved.

17

Tectonics

10.1002/2013TC003425

Corrigan, J., P. Mann, and J. C. Ingle (1990), Forearc response to subduction of the Cocos Ridge, Panama-Costa Rica, Geol. Soc. Am. Bull., 102,
628652, doi:10.1130/0016-7606(1990)102<0628:FRTSOT>2.3.CO;2.
Dahlen, F. A. (1990), Critical taper model of fold-and-thrust belts and accretionary wedges, Ann. Rev. Earth Planet. Sci., 18, 5599, doi:10.1146/
annurev.ea.18.050190.000415.
Daly, M. C. (1989), Correlations between Nazca/Farallon Plate kinematics and forearc basin evolution in Ecuador, Tectonics, 8, 769790,
doi:10.1029/TC008i004p00769.
DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein (1994), Effect of recent revisions to the geomagnetic reversal time scale on estimates of
current plate motions, Geophys. Res. Lett., 21, 21912194, doi:10.1029/94GL02118.
Dominguez, S., J. Malavieille, and S. E. Lallemand (2000), Deformation of accretionary wedges in response to seamount subduction: Insights
from sandbox experiments, Tectonics, 19, 182196, doi:10.1029/1999TC900055.
Drummond, M. S., M. Bordelon, J. Z. De Boer, and M. J. Defant (1995), Igneous petrogenesis and tectonic setting of plutonic and volcanic
rocks of the Cordillera de Talamanca, Costa Rica-Panama, Central American Arc, Am. J. Sci., 295, 875919, doi:10.1130/SPE295-p35.
Duque-Caro, H. (1990), Neogene stratigraphy, paleoceanography and palaeobiology in northwest South America and the evolution of the
Panama Seaway, Palaeogeogr. Palaeoclimatol. Palaeoecol., 77, 203234, doi:10.1016/0031-0182(90)90178-A.
Egbue, O., and J. Kellogg (2010), Pleistocene to Present North Andean escape, Tectonophysics, 489, 248257, doi:10.1016/j.tecto.2010.04.021.
Enkelmann, E., P. K. Zeitler, J. I. Garver, T. L. Pavlis, and B. P. Hooks (2010), The thermochronological record of tectonic and surface process
interaction at the Yakutat-North American collision zone in southeast Alaska, Am. J. Sci., 310, 231260, doi:10.2475/04.2010.01.
Espurt, N., F. Funicello, J. Martinod, B. Guillaume, V. Regard, C. Faccenna, and S. Brusset (2008), Flat subduction dynamics and deformation of
the South American plate: Insights from analog modeling, Tectonics, 27, TC3011, doi:10.1029/2007TC002175.
Farley, K. A. (2000), Helium diffusion from apatite: General behavior as illustrated by Durango uorapatite, J. Geophys. Res., 105, 29032914,
doi:10.1029/1999JB900348.
Farris, D. W., C. Jaramillo, G. Bayona, S. A. Restrepo-Moreno, C. Montes, A. Cardona, A. Mora, R. J. Spakman, M. D. Glascock, and V. Valencia
(2011), Fracturing of the Panamanian Isthmus during initial collision with South America, Geology, 39, 10071010, doi:10.1130/G32237.1.
Fisher, D. M., T. W. Gardner, P. B. Sak, J. D. Sanchez, K. Murphy, and P. Vannuchi (2004), Active thrusting in the inner forearc of an erosive
convergent margin, Pacic coast, Costa Rica, Tectonics, 23, TC2007, doi:10.1029/2002TC001464.
Flowers, R. M., R. A. Ketcham, D. L. Shuster, and K. A. Farley (2009), Apatite (UTh)/He thermochronometry using a radiation damage
accumulation and annealing model, Geochem. Cosmochim. Acta, 73, 23472365, doi:10.1016/j.gca.2009.01.015.
Freymueller, J. T., J. N. Kellogg, and V. Vega (1993), Plate motions in the north Andean region, J. Geophys. Res., 98, 21,85321,863, doi:10.1029/
93JB00520.
Gardner, T. W., D. Verdonck, N. M. Pinter, R. Slingerland, K. P. Furlong, T. F. Bullard, and S. G. Wells (1992), Quaternary uplift astride the aseismic
Cocos Ridge, Pacic coast, Costa Rica, Geol. Soc. Am. Bull., 104, 219232, doi:10.1130/0016-7606(1992)104<0219:QUATAC>2.3.CO;2.
Gardner, T., et al. (2001), Holocene forearc block rotation in response to seamount subduction, southeastern Pennsular de Nicoya, Costa
Rica, Geology, 29, 151154, doi:10.1130/0091-7613(2001)029<0151:HFBRIR>2.0.CO;2.
Gardner, T. W., D. M. Fisher, K. D. Morell, and M. L. Cupper (2013), Upper-plate deformation in response to at slab subduction inboard of the
aseismic Cocos Ridge, Osa Peninsular, Costa Rica, Lithosphere, 5, 247264, doi:10.1130/L251.1.
Gerya, T. V., D. Fossati, C. Cantiene, and D. Seward (2009), Dynamic effects of aseismic ridge subduction: Numerical modelling, Eur. J. Mineral.,
21, 649661, doi:10.1127/0935-1221/2009/0021-1931.
Grfe, K., W. Frisch, I. M. Villa, and M. Meschede (2002), Geodynamic evolution of southern Costa Rica related to low-angle subduction of the
Cocos Ridge: Constraints from thermochronology, Tectonophysics, 348, 187204, doi:10.1016/S0040-1951(02)00113-0.
Greene, A. R., J. S. Scoates, D. Weis, E. C. Katvala, S. Israel, and G. T. Nixon (2010), The architecture of oceanic plateaus revealed by the volcanic
stratigraphy of the accreted Wrangelia oceanic plateau, Geosphere, 6, 4773, doi:10.1130/GES00212.1.
Grevemeyer, I., et al. (2004), Fluid ow through active mud dome Mound Culebra offshore Nicoya Peninsula, Costa Rica: Evidence from heat
ow surveying, Mar. Geol., 207, 145157, doi:10.1016/j.margeo.2004.04.002.
Guillier, B., J.-L. Chatelain, E. Jaillard, H. Yepes, G. Poupinet, and J.-F. Fels (2001), Seismological evidence on the geometry of the orogenic
system in central-northern Ecuador (South America), Geophys. Res. Lett., 28, 37493752, doi:10.1029/2001GL013257.
Gutscher, M. A., J. Malavieille, S. Lallemand, and J. Y. Collot (1999), Tectonic segmentation of the North Andean margin: Impact of the
Carnegie Ridge collision, Earth Planet. Sci. Lett., 168, 255270, doi:10.1016/S0012-821X(99)00060-6.
Gutscher, M.-A., W. Spakman, H. Bijwaard, and E. R. Engdahl (2000), Geodynamics of at subduction: Seismicity and tomographic constraints
from the Andean margin, Tectonics, 19, 814833, doi:10.1029/1999TC001152.
Hampel, A. (2002), The migration history of the Nazca Ridge along the Peruvian active margin: A re-evaluation, Earth Planet. Sci. Lett., 203,
665679, doi:10.1016/S0012-821X(02)00859-2.
Hampel, A., J. Adam, and N. Kukowski (2004), Response of the tectonically erosive south Peruvian forearc to subduction of the Nazca Ridge:
Analysis of three-dimensional analogue experiments, Tectonics, 23, TC5003, doi:10.1029/2003TC001585.
Henry, S. G., and H. N. Pollack (1988), Terrestrial heat ow above the Andean Subduction Zone in Bolivia and Peru, J. Geophys. Res., 93,
15,15315,162.
Hsu, J. T. (1992), Quaternary uplift of the Peruvian coast related to the subduction of the Nazca Ridge: 13.5 to 15.6 degrees south latitude,
Quat. Int., 15/16, 8797, doi:10.1016/1040-6182(92)90038-4.
Hughes, R. A., and L. F. Pilatasig (2002), Cretaceous and Tertiary terrene accretion in the Cordillera Occidental of the Andes of Ecuador,
Tectonophysics, 345, 2948, doi:10.1016/S0040-1951(01)00205-0.
Hungerbhler, D., M. Steinmann, W. Winkler, D. Seward, A. Egez, D. Peterson, U. Helg, and C. Hammer (2002), Neogene stratigraphy and
Andean geodynamics of southern Ecuador, Earth Sci. Rev., 57, 75124, doi:10.1016/S0012-8252(01)00071-X.
Kellogg, J., and W. Bonini (1982), Subduction of the Caribbean plate and basement uplifts in the overriding South American plate, Tectonics,
1, 251276, doi:10.1029/TC001i003p00251.
Kerr, A. C., J. Tarney, G. F. Marriner, A. Nivia, and A. D. Saunders (1997), The Caribbean-Colombian Cretaceous igneous province: The internal
anatomy of an oceanic plateau, in Large Igneous Provinces: Continental, Oceanic and Planetary Flood Volcanism, Geophysical Monograph,
vol. 100, edited by J. J. Mahoney and M. F. Cofn, pp. 123144, AGU, Washington, D. C., doi:10.1029/GM100p0123.
Ketcham, R. A., R. A. Donelick, and W. D. Carlson (1999), Variability of apatite ssion-track annealing kinetics: III. Extrapolation to geological
time scales, Am. Mineral., 84, 12351255.
Ketcham, R. A., A. Carter, R. A. Donelick, J. Barbarand, and A. J. Hurford (2007), Improved modeling of ssion-track annealing in apatite, Am.
Mineral., 92, 799810, doi:10.2138/am.2007.2281.
Korenaga, J. (2005), Firm mantle plumes and the nature of the core-mantle boundary region, Earth Planet. Sci. Lett., 232, 2937,
doi:10.1016/j.epsl.2005.01.016.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

18

Tectonics

10.1002/2013TC003425

LaFemina, P., T. H. Dixon, R. Govers, E. Norabuena, H. Turner, A. Saballos, G. Mattoli, M. Protti, and W. Strauch (2009), Fore-arc motion and
Cocos Ridge collision in Central America, Geochem. Geophys. Geosyst., 10, Q05S14, doi:10.1029/2008GC002181.
Lallemand, S. E., J. Malavieille, and S. Calassou (1992), Effects of oceanic ridge subduction on accretionary wedges: Experimental modeling
and marine observations, Tectonics, 11(6), 13011313, doi:10.1029/92TC00637.
Lamb, S., and P. Davis (2003), Cenozoic climate change as a possible cause for the rise of the Andes, Nature, 425, 792797, doi:10.1038/
nature02049.
LeFevre, L. V., and K. C. McNally (1985), Stress distribution and subduction of aseismic ridges in the Middle America Subduction Zone,
J. Geophys. Res., 90, 44954510, doi:10.1029/JB090iB06p04495.
Livaccari, R. F., K. Burke, and A. M. C. Sengor (1981), Was the Laramide Orogeny related to subduction of an oceanic plateau?, Nature, 289,
276278, doi:10.1038/289276a0.
Lonsdale, P., and K. D. Klitgord (1978), Structure and tectonic history of the eastern Panama Basin, Geol. Soc. Am. Bull., 89, 981999,
doi:10.1130/0016-7606(1978)89<981:SATHOT>2.0.CO;2.
Luzieux, L. D. A., F. Heller, R. Spikings, C. F. Vallejo, and W. Winkler (2006), Origin and Cretaceous history of the coastal Ecuadorian forearc
between 1N and 3S: Paleomagnetic, radiometric and fossil evidence, Earth Planet. Sci. Lett., 249, 400414, doi:10.1016/j.epsl.2006.07.008.
Machar, J., and L. Ortlieb (1992), Plio-Quaternary vertical motions and the subduction of the Nazca Ridge, central coast of Peru,
Tectonophysics, 205, 97108, doi:10.1016/0040-1951(92)90420-B.
Mahoney, J. J., and K. J. Spencer (1991), Isotopic evidence for the origin of the Manihiki and Ontong Java oceanic plateaus, Earth Planet. Sci.
Lett., 104, 196210, doi:10.1016/0012-821X(91)90204-U.
Mann, P., and J. Corrigan (1990), Model for late Neogene deformation in Panama, Geology, 18, 558562, doi:10.1130/0091-7613(1990)
018<0558:MFLNDI>2.3.CO;2.
Mason, W. G., L. Moresi, P. G. Betts, and M. S. Miller (2010), Three-dimensional numerical models of the inuence of a buoyant oceanic plateau
on subduction zones, Tectonophysics, 483, 7179, doi:10.1016/j.tecto.2009.08.021.
Masy, J., F. Niu, A. Levander, and M. Schmitz (2011), Mantle ow beneath northwestern Venezuela: Seismic evidence for a deep origin of the
Mrida Andes, Earth Planet. Sci. Lett., 305, 396404, doi:10.1016/j.epsl.2011.03.024.
Meschede, M., and U. Barckhausen (2000), Plate tectonic evolution of the Cocos-Nazca spreading center, in Proceedings of the Ocean Drilling
Program, Scientic Results, vol. 170, edited by E. A. Silver, G. Kimura, and T. H. Shipley, pp. 110, International Ocean Drilling Program,
College Station.
Meschede, M., and W. Frisch (1998), A plate-tectonic model for Mesozoic and Early Cenozoic history of the Caribbean Plate, Tectonophysics,
296, 269291, doi:10.1016/S0040-1951(98)00157-7.
Michaud, F., C. Witt, and J.-Y. Royer (2009), Inuence of subduction on the Carnegie volcanic ridge on Ecuadorian geology: Reality and ction,
Geol. Soc. Am. Mem., 204, 217228.
Miller, M. S., A. Levander, F. Niu, and A. Li (2009), Upper mantle structure beneath the Caribbean-South American plate boundary from
surface wave tomography, J. Geophys. Res., 114, B01311, doi:10.1029/2007JB005507.
Molnar, P., and P. England (1990), Late Cenozoic uplift of mountain ranges and global climate change: Chicken or egg?, Nature, 346, 2934,
doi:10.1130/0091-7613(1990)018<1173:SUUORA>2.3.CO;2.
Montes, C., et al. (2012), Evidence for middle Eocene and younger land emergence in central Panama: Implications for Isthmus closure, Geol.
Soc. Am. Bull., 124, 780799, doi:10.1130/B30528.1.
Morell, K. D., D. M. Fisher, T. W. Gardner, P. LaFemina, D. Davidson, and A. Teletzke (2011), Quaternary outer fore-arc deformation and uplift
inboard of the Panama Triple Junction, Burica Peninsular, J. Geophys. Res., 116, B05402, doi:10.1029/2010JB007979.
Morell, K. D., E. Kirby, D. M. Fisher, and M.-V. Soest (2012), Geomorphic and exhumational response of the Central American Volcanic Arc to
Cocos Ridge subduction, J. Geophys. Res., 117, B04409, doi:10.1029/2011JB008969.
Nakanishi, M., W. W. Sager, and A. Klaus (1999), Magnetic lineations within Shatsky Rise, northwest Pacic Ocean: Implications for hot spottriple junction interaction and oceanic plateau formation, J. Geophys. Res., 104, 75397556, doi:10.1029/1999JB900002.
Neal, C. R., J. J. Mahoney, L. W. Kroenke, R. A. Duncan, and M. G. Petterson (1997), The Ontong Java Plateau, in Large Igneous Provinces:
Continental, Oceanic and Planetary Flood Volcanism, Geophysical Monograph, vol. 100, edited by J. J. Mahoney and M. F. Cofn, pp. 183216,
AGU, Washington, D. C., doi:10.1029/GM100p0183.
Newkirk, D. R., and E. E. Martin (2009), Circulation through the Central American Seaway during Miocene carbonate crash, Geology, 37, 8790,
doi:10.1130/G25193A.1.
Norabuena, E. O., T. H. Dixon, S. Stein, and C. G. A. Harrison (1999), Decelerating Nazca-South America and Nazca-Pacic plate motions,
Geophys. Res. Lett., 26, 34053408, doi:10.1029/1999GL005394.
ODriscoll, L. J., M. A. Richards, and E. D. Humphreys (2012), NazcaSouth America interactions and the late Eocenelate Oligocene at-slab
episode in the central Andes, Tectonics, 31, TC2013, doi:10.1029/2011TC003036.
Pedoja, K. (2003), Les terrasses marines de la marge Nord Andine (Equateur et Nord Prou): Relations avec le contexte geodynamique, PhD
thesis, Univ. Paris VI, Paris, 413 p.
Pedoja, K., L. Ortlieb, J. F. Dumont, M. Lamothe, B. Ghaleb, M. Auclair, and B. Labrousse (2006), Quaternary coastal uplift along the Talara Arc
(Ecuador, Northern Peru) from new marine terrace data, Mar. Geol., 228, 7391, doi:10.1016/j.margeo.2006.01.004.
Panz, D., C. Gaedicke, R. Freitag, M. Krbetschek, N. Tsukanov, and B. Baranov (2013), Neotectonics and recent uplift at Kamchatka and
Aleutian arc junction, Kamchatka Cape area, NE Russia, Int. J. Earth Sci., 103, 903916, doi:10.1007/s00531-012-0830-z.
Pilger, R. H. (1981), Plate reconstructions, aseismic ridges, and low-angle subduction beneath the Andes, Geol. Soc. Am. Bull., 92, 448456,
doi:10.1130/0016-7606(1981)92<448:PRARAL>2.0.CO;2.
Pilger, R. H. (1984), Cenozoic kinematics, subduction and magmatism: South American Andes, J. Geol. Soc., 141, 793802, doi:10.1144/
gsjgs.141.5.0793.
Pindell, J., L. Kennan, W. V. Maresch, K.-P. Stanek, G. Draper, and R. Higgs (2005), Plate-kinematics and crustal dynamics of circum-Caribbean
arc-continent interactions: Tectonic controls on basin development in Proto-Caribbean margins, Geol. Soc. Am. Spec. Pap., 394, 752.
Polliand, M., U. Schaltegger, M. Frank, and L. Fontbote (2005), Formation of intra-arc volcanosedimentary basins in the western ank of the
central Peruvian Andes during Late Cretaceous oblique subduction: eld evidence and constraints from U-Pb ages and Hf isotopes, Int. J.
Earth Sci., 94, 231242, doi:10.1007/s00531-005-0464-5.
Protti, M., F. Gundel, and K. McNally (1994), The geometry of the Wadati-Benioff zone under southern Central America and its tectonic signicance: Results from a high-resolution local seismographic network, Phys. Earth Planet. Interiors, 84, 271287, doi:10.1016/0031-9201
(94)90046-9.
Protti, M., et al. (1995), The March 25, 1990 (Mw = 7.0, Ml = 6.8), earthquake at the entrance of the Nicoya Guly, Costa Rica: Its prior activity,
foreshocks, aftershocks, and triggered seismicity, J. Geophys. Res., 100, 20,34520,358, doi:10.1029/94JB03099.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

19

Tectonics

10.1002/2013TC003425

Puchtel, I. S., A. W. Hofmann, K. Mezger, K. P. Jochum, A. A. Shchipansky, and A. V. Samsonov (1998), Oceanic plateau model for continental
crustal growth in the Archaean: A case study from the Kostomuksha greenstone belt, NW Baltic Shield, Earth Planet. Sci. Lett., 155, 5774,
doi:10.1016/S0012-821X(97)00202-1.
Ramos, V. A., E. O. Cristallini, and D. J. Perez (2002), The Pampean at-slab of the Central Andes, J. South Am. Earth Sci., 15, 5978, doi:10.1016/
S0895-9811(02)00006-8.
Reiners, P. W., T. L. Spell, S. Nicolescu, and K. A. Zanetti (2004), Zircon (U-Th)/He thermochronometry: He diffusion and comparisons with
40
39
Ar/ Ar dating, Geochem. Cosmochim., 68, 18571887, doi:10.1016/j.gca.2003.10.021.
Restrepo-Moreno, S. A., D. A. Foster, D. F. Stockli, and L. N. Parra-Snchez (2009), Long-term erosion and exhumation of the Altiplano
Antioqueo, Northern Andes (Colombia) from apatite (U-Th)/He thermochronology, Earth Planet. Sci. Lett., 278, 112, doi:10.1016/j.
epsl.2008.09.037.
Richards, M. A., D. L. Jones, R. A. Duncan, and J. DePaolo (1991), A mantle plume initiation model for the Wrangelia ood basalt and other
oceanic plateaus, Science, 254, 263267, doi:10.1126/science.254.5029.263.
Rockwell, T. K., R. A. Bennett, E. Gath, and P. Franceschi (2010), Unhinging and indentor: A new tectonic model for the internal deformation of
Panama, Tectonics, 29, TC4027, doi:10.1029/2009TC002571.
Rosenbaum, G., D. Giles, M. Saxon, P. G. Betts, R. F. Weinberg, and C. Duboz (2005), Subduction of the Nazca Ridge and the Inca Plateau:
Insights into the formation of ore deposits in Peru, Earth Planet Sci. Lett., 239, 1832, doi:10.1016/j.epsl.2005.08.003.
Rudnick, R. L. (1995), Making continental crust, Nature, 378, 571577.
Ryan, W. B. F., et al. (2009), Global multi-resolution topography synthesis, Geochem. Geophys. Geosyst., 10, Q03014, doi:10.1029/
2008GC002332.
Sak, P. B., and D. M. Fisher (2004), Effects of subducting seaoor roughness on upper plate vertical tectonism: Osa Peninsular, Costa Rica,
Tectonics, 23, TC1017, doi:10.1029/2002TC001474.
Sallars, V., P. Charvis, R. Flueh, and J. Bialas (2005), Seismic structure of the Carnegie Ridge and the nature of the Galapagos hotspot,
Geophys. J. Int., 161, 763788, doi:10.1111/j.1365-246X.2005.02592.x.
Scholz, C. H., and C. Small (2013), The effect of seamount subduction on seismic coupling, Geology, 25, 487490.
Simpson, G. D. H. (2004), Role of river incision in enhancing deformation, Geology, 32, 341344.
Simpson, G. D. H. (2010), Formation of accretionary prisms inuenced by sediment subduction and supplied by sediments from adjacent
continents, Geology, 38, 131134, doi:10.1130/G30461.1.
Sinton, C. W., R. A. Duncan, M. Storey, J. Lewis, and J. J. Estrada (1998), An oceanic ood basalts province within the Caribbean plate, Earth
Planet. Sci. Lett., 155, 221235, doi:10.1016/S0012-821X(97)00214-8.
Sitchler, J. C., D. M. Fisher, T. W. Gardner, and M. Protti (2007), Constraints on inner forearc deformation from balanced cross sections, Fila
Costea thrust belt, Costa Rica, Tectonics, 26, TC6012, doi:10.1029/2006TC001949.
Skinner, S. M., and R. W. Clayton (2013), The lack of correlation between at slabs and bathymetric impactors in South America, Earth Planet.
Sci. Lett., 371, 15, doi:10.1016/j.epsl.2013.04.013.
Spikings, R. A., and P. V. Crowhurst (2004), (U-Th)/He thermochronometric constraints on the late Miocene-Pliocene tectonic development of the northern Cordillera Real and the Interandean Depression, Ecuador, J. South Am. Earth Sci., 17, 239251, doi:10.1016/j.
jsames.2004.07.001.
Spikings, R. A., D. Seward, W. Winkler, and G. M. Ruiz (2000), Low-temperature thermochronology of the northern Cordillera Real, Ecuador:
Tectonic insights from zircon and apatite ssion track analysis, Tectonics, 19(4), 649668, doi:10.1029/2000tc900010.
Spikings, R. A., W. Winkler, D. Seward, and R. Handler (2001), Along-strike variations in the thermal and tectonic response of the continental
Ecuadorian Andes to the collision with heterogeneous oceanic crust, Earth Planet. Sci. Lett., 186, 5773, doi:10.1016/S0012-821X(01)00225-4.
Spikings, R. A., W. Winkler, R. A. Hughes, and R. Handler (2005), Thermochronology of allochthonous terranes in Ecuador: Unravelling the
accretionary and post-accretionary history of the Northern Andes, Tectonophysics, 399, 195220, doi:10.1016/j.tecto.2004.12.023.
Spikings, R. A., P. V. Crowhurst, W. Winkler, and D. Villagomez (2010), Syn- and post accretionary cooling history of the Ecuadorian
Andes constrained by their in-situ and detrital thermochronometric record, J. South Am. Earth Sci., 30, 121133, doi:10.1016/j.
jsames.2010.04.002.
Springer, M., and A. Foerster (1998), Heat-ow denisty across the Central Andean subduction zone, Tectonophysics, 291, 123139.
Stavenhagen, A. U., E. R. Flueh, C. Ranero, K. D. McIntosh, T. Shipley, G. Leandro, A. Schulze, and J. J. Danobeitia (1998), Seismic wide-angle
investigations in Costa RicaA crustal velocity model from the Pacic to the Caribbean coast, Zentralbl. Geol. Palaeontol. Teil 1, 36, 393408.
Strecker, M., A. Mulch, C. E. Uba, A. K. Schmitt, and P. Chamberlain (2006), Late Miocene onset of the South American Monsoon, Am. Geophys.
Union Fall Meeting, San Francisco, Calif.
Tassara, A., H.-J. Gtze, S. Schmidt, and R. Hackney (2006), Three-dimensional density model of the Nazca plate and the Andean continental
margin. J. Geophys. Res., 111, B09404, doi:10.1029/2005JB003976.
Trenkamp, R., J. N. Kellogg, J. T. Freymueller, and H. P. Mora (2002), Wide plate margin deformation, southern Central America and northwestern South America, CASA GPS observations, J. South Am. Earth Sci., 15, 151171, doi:10.1016/S0895-9811(02)00018-4.
Turcotte, D. L., and G. Schubert (1982), Geodynamics, pp. 456, Cambridge Univ. Press, Cambridge.
Vallejo, C., R. A. Spikings, L. Luzieux, W. Winkler, D. Chew, and L. Page (2006), The early interaction between the Caribbean Plateau and the NW
South American Plate, Terra Nova, 18, 264269, doi:10.1111/j.1365-3121.2006.00688.x.
Vallejo, C., W. Winkler, R. A. Spikings, L. Luzieux, F. Heller, and F. Bussy (2009), Mode and timing of terrane accretion in the forearc of the Andes
in Ecuador, Geol. Soc. Am. Mem., 204, 197216.
Van Hunen, J., A. P. van der Berg, and N. J. Vlaar (2002), On the role of subducting oceanic plateaus in the development of shallow at
subduction, Tectonophysics, 352, 317333, doi:10.1016/S0040-1951(02)00263-9.
Van Hunen, J., A. P. van der Berg, and N. J. Vlaar (2004), The impact of the South-American plate motion and the Nazca Ridge subduction on
the at subduction below South Peru, Geophys. Res. Lett., 29, 1690, doi:10.1029/2001GL014004.
Vannucchi, P., D. M. Fisher, S. Bier, and T. W. Gardner (2006), From seamount accretion to tectonic erosion: The formation of Osa Mlange and
the effects of Cocos Ridge subduction in southern Costa Rica, Tectonics, 25, TC2004, doi:10.1029/2005TC001855.
Vannucchi, P., P. B. Sak, J. P. Morgan, K. Ohkushi, and K. Ujiie (2013), Rapid pulses of uplift, subsidence, and subduction erosion offshore
Central America: Implications for building the rock record of convergent margins, Geology, 41, 995998, doi:10.1130/G34355.1.
Villagmez, D., and R. Spikings (2012), Thermochronology and tectonics of the Central and Western Cordilleras of Colombia: Early
CretaceousTertiary evolution of the Northern Andes, Lithos, 160, 228249, doi:10.1016/j.lithos.2012.12.008.
Villagmez, D., R. Spikings, A. Mora, G. Guzman, G. Ojeda, E. Cortes, and R. van der Lelij (2011a), Vertical tectonics at a continental crustoceanic plateau plate boundary zone: Fission track thermochronology of the Sierra Nevada de Santa Marta, Colombia, Tectonics, 30,
TC4004, doi:10.1029/2010TC002835.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

20

Tectonics

10.1002/2013TC003425

Villagmez, D., R. Spikings, T. Magna, A. Kammer, W. Winkler, and A. Beltrn (2011b), Geochronology, geochemistry and tectonic evolution of
the Western and Central cordilleras of Colombia, Lithos, 125, 875896, doi:10.1016/j.lithos.2011.05.003.
Von Huene, R., and R. Ranero (2009), Neogene collision and deformation of convergent margins along the backbone of the Americas in
Shallow Subduction, Plateau Uplift and Ridge and Terrane Collision, in Backbone of the Americas, Geol. Soc. Am. Mem., vol. 204,
edited by S. M. Kay et al., pp. 6783, Geol. Soc. Am. Inc., Boulder, Colo.
Wallace, L. M., S. Ellis, and P. Mann (2009), Collisional model for rapid fore-arc block rotations, arc curvature, and episodic back-arc rifting in
subduction settings, Geochem. Geophys. Geosyst., 10, Q05001, doi:10.1029/2008GC002220.
Walther, C. H. E. (2003), The crustal structure of the Cocos ridge off Costa Rica, J. Geophys. Res., 108(B3), 2136, doi:10.1029/2001JB000888.
Willet, S. D., F. Schlunegger, and V. Picotti (2006), Messinian climate change and erosional destruction of the central European Alps, Geology,
34, 613616, doi:10.1130/G22280.1.
Wipf, M., G. Zeilinger, D. Seward, and F. Schlunegger (2008), Focused subaerial erosion during ridge subduction: Impact on the geomorphology in south-central Peru, Terra Nova, 20, 110, doi:10.1111/j.1365-3121.2007.00780.x.

SPIKINGS AND SIMPSON

2014. American Geophysical Union. All Rights Reserved.

21

Você também pode gostar