Sie sind auf Seite 1von 22

Journal of South American Earth Sciences, VoL 7, Nos 3/4, pp.

241-262, 1994
Copyright 1994 Elsevier Science Lad & Earth Sciences & Resou~es Institute
Printed in Great Britain. All rights reserved
0895-9811194 $7.00 + 0.{30

Pergamon
0895-9811(94)00020-4

Consolidation of the American Cordilleras


E J. C O N E Y 1 a n d C . A . E V E N C H I C K 2
I Department o f Geosciences, University o f Arizona, Tucson, AZ, 85721, U S A
2Geological Survey o f Canada, 100 West Pender Street, Vancouver, BC, V6B 1R8, Canada

(Received October 1992; Revision Accepted August 1993)


Abstract--The continental margin orogenic systems of the western Americas are enormous features that formed along the Pacific
margins of the North and South American plates during late Mesozoic through Cenozoic time. There has been considerable debate
concerning their origin, and they are often compared with intra-oceanic fringing arc-trench systems more typical of the Australasian margins of the Pacific Ocean, in that both involve the subduction of oceanic lithosphere, often with similar convergent relative
motion vectors. The onset of orogenesis in the two Cordilleras, as shown in reversal of sedimentary polarity from sources generally on the continent to sources along the Pacific margin, seems to date from shortly after emplacement of the oldest oceanic crust
in that part of the Atlantic Ocean east of each continent - - i.e., about 170 Ma, or Middle Jurassic, in the case of the Central Atlantic, and about 135 to 100 Ma, or Early to mid-Cretaceous, in the case of the South Atlantic. These ages also seem to mark the onset
of westward motion of the two continents over the Pacific Ocean basin and subsequent crustal thickening and uplift, with development of thrust belts, foreland basins, and foredeeps. Prior to this prolonged westward drift, both margins had been convergent for
at least several hundred million years, but no massive mountain building had taken place. Instead, the margins were tectonically
"neutral," with typically submarine fringing arc-trench systems or shallow marine to continental margin arcs which stood "outboard" of shallow marine platformal shelves or basins whose main sedimentary polarity was from the continent. Although accretion of "suspect" terranes, high rates of convergence, and age of subducting lithosphere all may have influenced particularly local
tectonic response and/or phases of orogenic activity in the two chains, the absolute motion of the two continental margins over the
Pacific Ocean basin is considered to have been the dominant factor in Cordilleran tectonic evolution.
Resumen--Los sistemas orogrnicos de margen continental localizados en la regi6n occidental del Continente Americano constituyen enormes entidades geol6gicas formadas a 1o largo de la margen Pacffica de las placas tect6nicas de Norteamrrica y
Sudamrrica durante las eras Mesoz6ica y Cenoz6ica. Estos sistemas orogrnicos se formaron en tiempos relativamente recientes, a
pesar de que dichas margenes continentales han estado probablemente frente a la cuenca del Ocrano Pacffico durante la mayor
parle del Fanerozoico. En cambio, la margen Austroasi~ltica del Ocrano Pacffico presenta una interdigitaci6n de sistemas arcotrinchera intraoce,'inicos. El hecho de que ambas margenes del Pacffico presenten diferente evoluci6n, a pesar de estar relacionadas igualmente a procesos de subducci6n, y de presentar vectores de movimiento relativo de convergencia similares, contintia
siendo el centro de debate dentro del modelo de tect6nica de placas. Se concluye que el inicio del engrosamiento de la corteza y de
la orogrnesis tfpica de ambas cadenas (evidenciada por la inversi6n de la polaridad de la sedimentaci6n de ~ireas fuente generalmente en el continente a fuentes a lo largo de la mfirgen Pacffica) es contemporaneo con la formaci6n de la corteza oce~inica mils
antigua en esa parle del Ocrano Atlfintico, al este del continente. Por ejemplo el caso del Atlantico Central durante el Jurfisico
Medio, alrededor de 170 Ma, o en el caso del Atlantico Sur, en el Cret,'icico Temprano-Medio, entre los 134 y 100 Ma. Estas
fechas tambien parecen marcar el inicio del movimiento absoluto hacia el oeste de los dos continentes sobre la cuenca del Oc6ano
Pacffico y consecuentemente, el engrosamiento de la corteza, y el levantamiento y desarrollo de cinturones de cabalgaduras y
cuencas asociadas. A pesar de que la margen occidental de Amrrica era de tipo convergente varios millones de afios antes de este
periodo prolongado de deriva hacia el occidente, no origin6 la construcci6n de macizos montafiosos. En su lugar, la margen fur
tect6nicamente "neutral," con el tfpico desarrollo de sistemas arco-trinchera o arcos marinos someros o de margen continental,
que evolucionaron en el "frente" de las plataformas continentales o cuencas cuyos sedimentos provenfan del continente. Aunque
la acreci6n de terrenos "sospechosos," las tasas altas de convergencia y la edad de la lit6sfera subducida probablemente influenciaron en especial la tect6nica local o/y las fases de actividad orogrnica en las dos cadenas montafiosas, nosotros pensamos que el
movimiento "absoluto" de las dos mftrgenes continentales hacia la cuenca del Ocrano Pacffico ha sido el factor dominante de la
evoluci6n tectonica cordillerana.

INTRODUCTION

l i t h o s p h e r e is s u b d u c t e d b e n e a t h a n o t h e r o c e a n i c p l a t e ,
and belts of inter-continental

OROGENIC

collision. Although

our

that f o r m a l o n g c o n t i n e n t a l m a r -

u n d e r s t a n d i n g o f c o l l i s i o n a l s y s t e m s is f a i r l y s t r a i g h t f o r -

g i n s , w h e r e o c e a n i c l i t h o s p h e r e is s u b d u c t e d b e n e a t h c o n -

w a r d ( M o l n a r a n d L e o n - C a e n , 1988), t h e r e h a s b e e n c o n -

t i n e n t a l l i t h o s p h e r e , w e r e r e c o g n i z e d in t h e p l a t e t e c t o n i c

s i d e r a b l e d e b a t e a b o u t t h e o t h e r t w o ( C o n e y , 1973, 1987;

s y n t h e s i s as o n e e n d m e m b e r o f a triad o f c o n v e r g e n t p l a t e

U y e d a a n d K a n a m o r i , 1979; C h a s e , 1978; s e e a l s o D a l z i e l ,

tectonic

SYSTEMS

settings. The other two end members

are, of

1981), b e c a u s e w h e r e a s b o t h t h e s e e n d m e m b e r s

course, intra-oceanic arc-trench systems, where oceanic

involve

s u b d u c t i o n o f o c e a n i c l i t h o s p h e r e , t h e t e c t o n i c r e s p o n s e in

Address all correspondence to: Prof. Peter J. Coney: Tel [ 1] (602) 621-6017; Fax [1] (602) 621-2672.

Presented at the Fifth C i r e u m - P a c i f i e Terrane Conference, Santiago, Chile, 11-14 N o v e m b e r 1991.


241

242

P.J. CONEY and C.A. EVENCHICK


Fig. 1. The North American Cordillera. Stipple pattern is that part of the Cordillera
underlain by autochthonous North American Precambrian basement (after Coney et
al, 1980; Campa and Coney, 1983). The
remainder of the Cordillera, left white, is the
area underlain by "suspect" terranes.

4s.

Ii

o
i

Key to terranes discussed in the text:


An Angayuchum
Ax Alexander
BR BridgeRiver
C Caborca
Ca Cassiar
CG Chugach
Ch CacheCreek
Co Coahuilla
F Franciscan
G Guererro
GL Golconda
JK UpperJurassic to mid-Cretaceous
flysch
Ma Maya
N Nisling
P Peninsular
RM RobertsMountain
SM SlideMountain
St Stikinia
T Taku
W Wrangelia
YT YukonTanana

600
|

km

the upper plates of the systems can be markedly different,


as comparison of the Marianas arc-trench system of the
western Pacific and the Andean Cordillera of western
South America suggests. The principal long recognized
difference between the two is that intra-oceanic arc-trench
systems tend to be extensional, often with sea-floor
spreading in back-arc areas, whereas continental margin
systems tend to be compressional, with "arc-rear" foreland
fold and thrust belts. Furthermore, this difference seems to
exist despite similarities in relative convergent plate
motions across the boundaries. The issue is further con-

fused by the realization that most continental margin


mountain belts tend to include so-called "suspect" or
accreted terranes (Coney et al., 1980), leading to the idea
that collisions may occur even in continental margin orogenic systems (Nur and Ben-Avraham, 1982). Add to this
the fact that transcurrent, transpression, transtensional, and
even extensional episodes are caused by vagaries of evolving plate interactions along continental margins, and one
begins to understand how complex the histories of continental margin mountain systems can become.

Consolidation of the American Cordilleras


Continental margin orogenic systems such as the Cordilleras of the western Americas are enormous features.
Over 16,000 km long and up to 2000 km wide, with
crustal thicknesses of 60 km and more and average elevations locally over 4 km at various times in their evolution,
the American Cordilleras represent extraordinary tectonic
edifices. At issue here is the question of how and from
what are they built? One can think of them as comprising
an orogenic volume or mass. This mass has somehow been
constructed for the most part through folding, faulting,
metamorphism, and melting of pre-existing material from
some inherited continental crustal and lithospheric constitution originally found along the continental margin, and
various amounts of accreted oceanic and/or redistributed
continental material, and to which has been added juvenile
magmatic products direct from the sublithospheric mantle.
The term consolidation is a useful one for conceptualizing
this construction of orogenic mass. We review here certain
aspects of the tectonic history of the North American and
South American Cordilleras and place it in the context of
evolving motions of the two continents and their interactions with the Pacific Ocean plates - - our aim being to
establish the tectonic setting in which the two Cordilleras
were consolidated, with particular attention to when and
where consolidation began.
The North and South American Cordilleras are certainly the most extensive, best developed, and best preserved known examples of continental margin orogenic
systems of Phanerozoic time. Often thought of as typical
of the "peri-Pacific" realm, they are, of course, typical of
the eastern margin of the Circum-Pacific rim, where they
have formed along the leading edge of the two American
plates. The American Cordilleras have become the type
examples of continental margin mountain belts, and they
may in fact be the only Phanerozoic examples, at least the
only ones on such an enormous scale of space and time.
The longevity is worth emphasizing. The complete history
extends back into the Late Precambrian, for a duration of
about 750 My in the case of the North American Cordillera and probably at least 600 My in the case of the Andes.
This is apparently due to the seeming permanence of the
Pacific Ocean basin throughout Phanerozoic time (Coney,
1990, 1992), and probably back into the late Precambrian
to about 750 Ma when it may have first opened as a result
of the breakup of a Late Proterozoic supercontinent (Hoffman, 1991; Moores, 1991; Dalziel, 1991).
The principal features of the American Cordilleras
which we associate with the typical convergent continental
margin orogenic system, however, developed in Mesozoic-Cenozoic time, mainly between the Jurassic and the
present in the case of the North American Cordillera, and
between the Cretaceous and the present in the case of the
Andes. We focus on this history here, after an introductory
summary of the general character of the two chains and
the significant p r e - M e s o z o i c - C e n o z o i c events which
shaped the margins that evolved into today's American
Cordilleras.

243

GENERAL CHARACTER OF THE


AMERICAN CORDILLERAS
Although often considered in essence a continuous belt
of Pacific Rim orogenesis, the two Cordilleras are actually
considerably different in lithotectonic constitution and tectonic evolution, and there are some important differences
in timing as well. There are, however, numerous obvious
similarities between the two, and today's active Andes are
often viewed as a modern analog for the Cretaceous to
early Cenozoic history of the North American Cordillera
(Hamilton, 1969).

North American Cordillera


The North A m e r i c a n C o r d i l l e r a (Coney, 1989a)
stretches from the Bering Sea to the Caribbean over a distance of about 8000 km and is as much as 2000 km wide in
its central sector in the western United States (Fig. 1).
Elsewhere it is somewhat narrower, but on average the
width is 800 to 1000 km. Only at its northern and southern
extremities does it become narrower.
The lithotectonic constitution of the North American
Cordillera can be divided into two major crustal types
(Coney et al., 1980; Coney, 1989a). The first is that part of
the Cordillera underlain by the autochthonous Precambrian cratonic crust of the North American continent, the
second is that part made up of "suspect" or tectonostratigraphic terranes. Only about 30% of Cordilleran crust is
underlain by cratonic North America. This includes most
of the eastern higher ranges from the Alaska-Canada border south into northernmost Mexico, but also includes the
Colorado Plateau and the eastern parts of the Great Basin
in Nevada-Utah and the Basin and Range province in
southern Arizona and New Mexico and southward into
Sonora and Chihuahua in Mexico. Most of this "ancestral
North America" tract of the Cordillera is constructed from
upper Precambrian through lower Mesozoic rocks associated with the Cordilleran miogeoclinal terrace, which
was draped over the edge of North America's crystalline
Precambrian cratonic basement. The classic Rocky Mountain province of Wyoming-Colorado, and the Colorado
Plateau, however, are east of the miogeocline in thin cratonic shelf sequences of Paleozoic-Mesozoic age, so all of
Alaska, two-thirds of the Canadian Cordillera, about onehalf of the Cordillera in western United States, and about
80% of Mexico are underlain by the "suspect" terranes.
The lithotectonic character of that part of the Cordillera
underlain by "suspect" terranes can be further subdivided
into three major types (Coney, 1989a,b).
Terranes of continental margin affinity, including
apparently slightly displaced fragments of the Cordilleran miogeocline such as the Caborca and Cassiar terranes, and extensive quartz-feldspathic, now metamorphosed terranes, such as the Kootenay and the lower
part of the Yukon Tanana, whose original paleogeographic setting in the Cordillera, however, is less certain (Mortensen, 1992).

244

P.J. CONEY and C.A. EVENCH1CK

Large-scale accretionary complexes such as the Franciscan, Pacific Rim, and Chugach terranes, which are
mostly found along or near the present Pacific margin,
and the more inboard Cache Creek and Bridge RiverHozameen terranes.
Magmatic-sedimentary terranes of oceanic affinity,
including such large bodies as the Stikine, Quesnellia,
and Guerrero terranes, which appear to be intra-oceanic
arc systems formed at various times during late Paleozoic and Mesozoic time; the Wrangellia terrane, probably an early Mesozoic oceanic plateau constructed on a
late Paleozoic intra-oceanic arc assemblage; and the
Slide Mountain-Angayuchum terranes, which appear to
be late Paleozoic ocean floor.
South American Cordillera
The South American, or Andean, Cordillera (Fig. 2)
stretches for nearly 8000 km from the Caribbean region
south to Tierra del Fuego (Mfgard, 1989; Ramos, 1988b;
Dalziel, 1986). Unlike the North American Cordillera, the
Andes are comparatively narrow, reaching their greatest
width, of about 800 km, in southern Peru-Bolivia-northern
Argentina and Chile. Elsewhere, in the north and south,
the width ranges from only 200 to 300 kin.
The lithotectonic crustal constitution of the Andean
Cordillera is still incompletely known, but it is seemingly
quite different from the North American Cordillera (Richards and Coney, 1991; Ramos et al., 1986; Restrepo and
Toussaint, 1988). The only part of the Andes founded on
exposed "autochthonous" ancient Precambrian basement
is the eastern Cordilleras of Colombia, Ecuador, and Peru.
Although it may also exist in Bolivia, this basement is not
exposed there. This older Precambrian basement has
yielded scattered isotopic ages which suggest a Grenville
(1.0 Ga) overprint on older protoliths, and the assumption
is that these rocks represent the distal western edge of the
Guianan-Brazilian shield. Southward through Argentina,
the basement for the eastern ranges is in part at least latest
Precambrian "Pan-African" in age. There are almost no
uppermost Precambrian to Lower and Middle Cambrian
rocks exposed in the northern and central Andean eastern
ranges; instead Upper Cambrian to Ordovician-Silurian
rocks, usually muds and sands now metamorphosed to
low-grade schists, are assumed to directly overlie the Precambrian substrate. There is no suggestion of a classic
long-lived carbonate-dominated miogeoclinal terrace
along the length of the Andes, like that which bordered
North America.
The rest of the Andes is made up of "suspect" terranes,
but quite different in character from western North America (Richards and Coney, 1991; Ramos et al., 1986). Most
of the so-called terranes presently identified seem to be
largely of continental margin affinity, or at least of continental aspect, whose protoliths and possible accretionary
ages are pre-Mesozoic. Many of these terranes are probably of South American origin, such as the Arequipa terrane (Forsythe et al., 1993), but several may be truly
"exotic," such as the Precordillera terrane, which may be a

piece of the Appalachian miogeocline transferred to South


America during the Taconic-Famatinian collision (Dalla
Salda et al., 1992b). The only large-scale Mesozoic-Cenozoic magmatic-sedimentary terranes of oceanic affinity in
the Andes are found in western Colombia and Ecuador
(Aspden and McCourt, 1986; Goossens and Rose, 1973)
and along the northern coast of Venezeula. The northern
Andes oceanic terranes seem to represent some combination of ocean floor and/or submarine arc sequences of
Mesozoic to Cenozoic age which accreted in ColombiaEcuador starting in Cretaceous time and continuing into
the Cenozoic. They seem to be related somehow to the
Caribbean arcs and ocean floor of similar age, and also to
the Guerrero terrane in southern Mexico (Ruiz et al.,
1991). Significant accretionary complexes of late Paleozoic to Mesozoic-Cenozoic age are found in central to
southem Chile (Herv6 et al., 1987).
PRE-CORDILLERAN HISTORY OF THE
NORTH AND SOUTH A M E R I C A N MARGINS
North American M a r g i n
The history of the Pacific margin of North America
extends over more than 750 My, apparently beginning
with a rifting event which either attenuated a pre-existing
margin or produced it by the breakaway of AntarcticaAustralia to open the Pacific Ocean basin (Dalziel, 1991;
Moores, 1991; Hoffman, 1991). This caused the development of a passive margin miogeoclinal terrace (Stewart,
1976; Gabrielse, 1972) which continued to subside uninterrupted until Late Devonian-Early Carboniferous time.
During this period of time there is no evidence that Alaska
or Mexico existed in anything like their present form, and
the miogeocline seems to have swung northeastward from
northwesternmost Canada into the Inuitian region of
northern Canada and southeastward to eastward from
southern California across northern Mexico into the Ouachita-Marathon margin of the present-day Gulf of Mexico.
During the Late Devonian-Early Carboniferous, the
development of the miogeocline was briefly interrupted by
the short-lived Antler " o r o g e n y " (Poole, 1975). In
Nevada-Idaho, an oceanic arc edifice either collided, or
collapsed a back-arc marginal basin, to obduct over the
distal edge of the miogeocline and shed a sharp but narrow
and short-lived detrital wedge eastward. The event seems
to be related somehow to the Inuitian orogeny of the Canadian Arctic, and a "detrital disturbance" in western Canada
with much chert pebble sands and conglomerates and
euxinic muds spread over large areas, apparently associated with extensional tectonics. An event of this age is also
recorded in magmatic-metamorphic-structural events in
several of the "suspect" terranes of western Canada, particularly the Yukon Tanana terrane (Mortensen, 1992). The
event, in any case, was soon over and the miogeocline was
reestablished to prograde over the minor wreckage of the
Antler edifice.
The miogeocline in the central and southwestern
United States and much of the craton interior to the east
and southeast was affected by the mostly mid- to Late Car-

Consolidation of the American Cordilleras

~.

"~ '~

%,

~.

'~

%:

A~

PC
CH

CE

lllO0 KM

MG
DW

Fig. 2. The South American Cordillera. Dash pattern is the Precambrian shield of South America. Light stipple is that part of
the Andes presumed to lie on authochthonous Precambrian basement of South America. Dark stipple shows the Mesozoic-Cenozoic oceanic terranes of the northern Andes (MCOT) and late
Paleozoic-early Mesozoic accretionary terranes (CE, Chiloe;
MG, Magallanes) and the Mesozoic-Cenozoic accretionary terrane (DW, Darwinia) of the southern Andes. CT0 Canta terrane;
the remainder of the terranes are pre-Mesozoic: AQ, Arequipa
terrane; CH, Chilenia terrane; P, Puna terrane; PC, Precordillera
terrane. Terranes after Richards and Coney (1991), Ramos et al.
(1986), and Dalla Salda et al. (1992).

245

boniferous completely a-magmatic ancestral Rockies


"event," which produced sharp linear uplifts and flanking
basins in a system of transpressive intraplate shear. The
deformation was apparently a distal effect of the OuachitaWichita orogenies resulting from the collision of the South
American part of Gondwanaland with the southern margin
of North America as Iapatus closed (Kluth and Coney,
1981). This collision, however, also had a significant
effect on the crustal constitution of Mexico in that the
Coahuila and Maya terranes, which make up most of eastern Mexico, were accreted at this time (Campa and Coney,
1983).
The late Paleozoic to early Mesozoic (Early Jurassic) is
one of the most enigmatic periods in the history of the
Cordilleran margin. Of particular concern is the Sonoma
"orogeny" of Nevada in Late Permian-Early Triassic time,
when another "oceanic" to distal continental margin
allochthon and associated outboard arc assemblages were
emplaced against and/or upon the miogeocline (Silberling,
1973). Southward in eastern Mexico a continental margin
arc developed in Permo-Triassic time. Many of the large
oceanic magmatic-sedimentary terranes, for example in
western Canada ( Monger e t al., 1992), record mostly submarine arc magmatism during late Paleozoic and early
Mesozoic time, most without any evidence of continental
influence (Samson and Patchett, 1991). Upper Paleozoic
to Mesozoic accretionary complexes as young as Jurassic
in age, such as the Cache Creek terrane, lie in the midst of
the collage. The paleogeography of all this is very obscure
and a topic of much discussion. In the southern Cordillera,
however, a Late Triassic (?) to Jurassic magmatic arc was
established upon North American continental crust and
extended the length of eastern central Mexico and into
northwestern South America (Coney, 1989a).
Consolidation of the North American Cordillera began
in the Middle to Late Jurassic and has continued to the
present. This is the Cordilleran orogeny, and almost the
entire edifice was constructed through large-scale thrusting and folding, extensive magmatism, and some metamorphism. Most of the contractional damage, in fact, was
completed by the early Cenozoic. Since the Eocene, the
original largely compressive edifice has been much modified by extensional collapse and transtensional rifting.
In summary, it is worth pointing out that the Pacific
margin of western North America has existed for about
750 My, but that the Cordillera now on its western edge
has been almost completely constructed since about 150
Ma. There is no evidence of any significant Cordilleranwide orogenic edifice on the North American margin prior
to Late Jurassic time. It seems impossible that the Pacific
margin of North America was not separated from spreading centers in the Pacific Ocean basin by convergent plate
tectonic settings for 600 My. This suggests that convergent
intra-oceanic arc trench systems must have lain offshore of
North America's western margin through most of Phanerozoic pre-Late Jurassic time.

246

P.J. CONEY and C.A. EVENCHICK

South American Margin


Unlike the North American Pacific margin, which was
mostly passive for almost 600 My before MesozoicCenozoic orogenesis, the South American margin seems
to have been tectonically active almost continuously at
one place or time or another through most of the Phanerozoic (Coney and Richards, 1991). As already mentioned,
even formation of the margin was not only later than North
America's but also more complex.
The Guianan-Brazilian shield seems to have broken
away from North America's eastern or northeastern margin by about 600 Ma as a late Middle Proterozoic supercontinent, which had been constructed from 1.0 Ga
Grenville age collisions, disintegrated beginning at about
750 Ma (Hoffman, 1991; Dalziel, 1992). Following Hoffman's perception of the process, with the exception of
North America itself, and Baltica and Siberia, which each
went their own ways, the remaining fragments, including
the Guianan-Brazilian shield, Antarctica-Australia-India,
and the several Archaean-Proterozoic nuclei that make up
Africa and southern South America, amalgamated through
Pan-African collisions to form Gondwanaland. This consolidation was finally completed during Cambrian time,
and Gondwanaland eventually became the southeastern,
southern, and southwestern margin of the Pacific Ocean
basin. The result of this is that the northern Andes formed
on the margin of the Guianan-Brazilian shield, apparently
following the trend of a Grenville overprint, but the South
American margin south of Bolivia was constructed from
Pan-African events (Ramos, 1988a).
Once formed, the South American margin seems to
have experienced almost continuous convergent tectonics
of one sort or another, including at least two and possibly
three collisions with what is now eastern North America,
through the remainder of Paleozoic time and into the early
Mesozoic. Scattered unconformities and radiometric ages
of plutonism, volcanism, and metamorphism, usually
overprinted by younger events, suggest widespread orogenic activity during the period extending from Middle
Cambrian in northern Argentina through the Ordovician in
Colombia, Bolivia, and northern Argentina. It has been
suggested that these events were caused by convergence
and collision with eastern North America to produce the
Taconic-Famatinian orogen (Dalla Salda et al., 1992a,b).
There is some evidence of deformation in Late SilurianEarly Devonian time in the eastern Cordillera of Colombia
(Restrepo-Pace, 1992), perhaps caused by another collision with eastern North America during the Acadian orogeny (Van Der Voo, 1988), or possibly by continued
convergence and interaction related to the earlier collision.
After its initial separation from Laurentia in the latest
Precambrian-Early Cambrian, the South American protoAndean margin may not have strayed far from what is
today eastern North America until late Paleozoic time
(Dalziel, 1992, per. comm., 1993). There is evidence of
significant Late Devonian-Early Carboniferous orogeny
along much of the Andes, particularly in Peru (M6gard,
1978). Then, in Carboniferous to early Mesozoic time a
magmatic arc formed along much of the western margin of

the southern Andes from northern Chile to south-central


Chile, then southeastward across northern Patagonia (Forsythe, 1982). An extensive tectonically related accretionary complex is preserved in central and southern Chile
(Forsythe, 1982; Herv6 et al., 1987). Scattered metamorphic radiometric dates of Permian age may record deformation in the northern Andes of Colombia and Venezuela,
presumably related to the collision forming Pangaea (Irving, 1975). In southern South America, the Gondwanide
fold belt formed during Triassic time, oblique to present
Andean trends, extending southeast across northern Patagonia, then into the Cape Fold Belt of South Africa and
thence into Antarctica, probably the result of accretion of
an arc-forearc and collapse of a marginal basin (Dalziel
and Grunow, 1992).
There is evidence for a marginal, magmatic arc of Triassic-Jurassic age along the Andean margin at least from
southern Peru southward into Patagonia (M6gard, 1989;
Ramos, 1988b; Herv6 et al., 1987). Rocks reflecting marginal submarine arc magmatism are widespread through
Chile into southern Peru. During the Cretaceous, however,
the arc moved northward through coastal Peru and the
eastern flank of the central Cordillera of Ecuador. In the
northern Andes of Colombia-Venezuela, continental red
beds and associated alkaline magmatism seem to record
rifting in this area from Triassic into Late Jurassic time.
An important point regarding these early Mesozoic magmatic arcs is that most seem to have been located outboard
of shallow marine seas. In other words, although subduction was taking place along the South American margin,
no massive Andean-wide consolidation, or mountain
building, was taking place. The major period of mountain
building, the Andean cycle of orogenesis, did not begin
until Cretaceous time (Dalziel, 1986).
MESOZOIC-CENOZOIC TECTONICS OF T H E
NORTH A M E R I C A N C O R D I L L E R A
Canadian Cordillera (Figs. 3 and 4)
Cordilleran Foreland Thrust Belt. One of the most
telling transitions in the depositional history of the North
American Cordilleran continental margin from Utah north
into northern British Columbia in Canada took place in the
Late Jurassic (Price and Mountjoy, 1970; Yorath, 1992).
Up until that time, with the exception of the two brief disturbances of the Antler and Sonoma "orogenies" discussed
earlier, the sedimentary polarity was from the craton
toward the presumed shelf-slope break of the continental
terrace for nearly 600 My. This means that the miogeoclinal edge of western North America was just below, or
never far above, sea level for most of this long period of
time, and there is no evidence of any significant oceanic
source areas beyond the shelf edge capable of shedding
debris up onto the continental terrace and platform for
most of this time. In the southern Canadian Cordillera, the
Late Jurassic transition is recorded in a reversal of sedimentary polarity within the Upper Jurassic Fernie Formation where the first debris from a western source appeared
in the upper part of the formation. This event marked the

Consolidation of the American Cordilleras

/
128

247

/;
...,~:.
,-z'J

NA
i
I
I
l
I
f
t

S
Bower
B~m

200 KM

BR

128

PR

Fig. 3. Generalized terrane map and map of morphotectonic belts (inset) of the Canadian Cordillera, after Gabrielse and Yorath (1991).
Key: AX, Alexander terrane; BR, Bridge River terrane; CA, Cassiar terrane; CC, Cache Creek terrane; CG, Chugatch terrane;
K, Kootenay terrane; KSFZ, King Salmon fault zone; N, Nisling terrane; NA, authochthonous cratonic North American craton;
Q, Quesnellia terrane; S, Stikinia terrane; SM, Slide Mountain terrane; W, Wrangellia terrane; YT, Yukon Tanana terrane.

end of the Cordilleran miogeocline and its passage into a


foreland basin, thus the beginning of the emergence of a
Cordilleran orogenic edifice somewhere to the west. In the
northern Canadian Cordillera, this transition occurred
somewhat later, in Early Cretaceous time.
The source for this debris seems to have been from a
region which is today the Omineca belt. This domain
eventually became the "hinterland" of the Cordilleran
foreland fold and thrust belt. The fold and thrust belt

evolved eastward, cannibalizing its own debris as it


migrated progressively across the old miogeoclinal prism
toward the cratonic interior. As the thrust fronts moved
relatively eastward, the entire folding and thrust faulting
miogeoclinal prism and its growing foreland basin debris
detached from the underlying Precambrian basement,
which was moving relatively westward beneath it (Price,
1981). Three great pulses of foreland basin sedimentary
debris are preserved in the foreland: the Kimmeridgian to

248

P.J. CONEY and C.A. EVENCHICK

c,~..,~.~

Cache Creek

Canadian Cordillera
0

. ~

Neogcne
Paleogene

I00

Cretaceous

om

.N

0
O

Jurassic
.)o0

Triaseic

Permian
--,J

30C

"~
N

Carboniferous
Devonian

40(
Ma

Turbidite and/or deeper


water sediments

.'~. .'~." .y. .y.

n u m m n
m m m m m
m m m m |

Limestone

Accrefionary prism

Sands and shales

Magmatic arcs partly submerged or near sea level

Foreland basins

Submarine basalt

Continental magmatic arcs

Plutons

iil

Metamorphic rocks

, ~,..~:,,. Meaamorphism

Orogenesis

Fig. 4. Generalized post-Devonian tectono-stratigraphic columns for the Canadian Cordillera (from Fig. 1; Correlation Charts; Gabrielse and Yorath, 1992;, and compilations in the Laboratory of Geotectonics, University of Arizona). The heavy black horizontal line is
the approximate age of oldest sea floor in the Central Atlantic Ocean. Key:a, Kootenay; b, Blairmore; c, Brazeau; d, Bowser Lake Gr.;
e, Sustut basin; f, Gravina assemblage; g, Hazelton assemblage; SM, Slide Mountain terrane.

Valanginian Fernie-Kootenay sequence, the Aptian to


Cenomanian Blairmore, and the Campanian through Paleocene Brazeau. Contractional deformation ended by the
early Eocene.
O m i n e c a Belt. The Omineca hinterland consists of
metamorphic and plutonic rocks exposed in a belt lying
just west of the generally unmetamorphosed foreland fold
and thrust belt (Monger et al., 1982; Gabrielse et al.,
1992). The belt consists of two wide culminations up to
400 km across, one in the southern Canadian Cordillera
and another in the north. Both of these culminations lie

west of prominent salients in the thrust belt to the east.


Between the two culminations, the belt narrows to less
than 200 km wide. The Omineca belt is generally structurally high and often, particularly in the southern culmination and locally in the narrow central portion, contains
exposures of Precambrian basement rocks of the North
American craton, which are known only at depth in the
thrust belt to the east below the basal detachment. The
rocks of the Omineca belt are now much modified by
Mesozoic-early Cenozoic deep-seated thrust faulting,
metamorphism to upper amphibolite grade, and Eocene
extensional faulting. The Eocene extensional detachment

Consolidation of the American Cordilleras


faulting is largely responsible for the present exposure of
the deeper seated rocks through tectonic denudation (Parrish et al., 1988), presumably related to earlier crustal
shortening and resulting thickening. In its narrower central
portion and in the western part of the northern salient, the
belt is much disrupted by major right strike-slip faulting,
such as along the Northern Rocky Mountain Trench and
Tintina fault system (Gabrielse, 1985). The Omineca belt
is also characterized by abundant mid-Cretaceous plutons
which, significantly, are chemically distinct, with isotopic
evidence of much crustal melting in their sources (Armstrong, 1988). Much of the Omineca belt seems to represent a whole-crust complex duplex stack (Brown et al.,
1986), locally antiformal, into which the foreland thrust
belt basal detachments to the east must have fed.
Except very locally, the oldest unmetamorphosed sediments found in the Omineca belt are early Cenozoic or
younger in age. Isotopic constraints on the timing of deformation suggest movement on some of the deep-seated
thrust surfaces as recent as Late Cretaceous to Paleocene.
Metamorphism associated with deformation reached a
peak in the Middle Jurassic (Archibald et al., 1983). All
evidence seems to indicate considerable erosional denudation, with the Omineca belt as the principal source of the
Upper Jurassic to Paleocene foreland basin deposits found
to the east and discussed earlier.
I n t e r m o n t a n e Terranes. West of the Omineca belt lies
the Intermontane belt, which is generally equivalent to
most of the composite Intermontane superterrane, or
"superterrane I" (Monger et al., 1982). The Intermontane
superterrane is made up of a number of suspect terranes,
almost all of whose lithotectonic successions carry a
marked "oceanic" or at best distal continental margin signature. Several, for example, are usually interpreted as
"intra-oceanic arc assemblages." This ground is classic in
North American Cordilleran tectonics. The several suspect
terranes it comprises are almost type examples of the
problem plaguing Cordilleran geologists since 1970, when
the seemingly indeterminant late Paleozoic-early Mesozoic paleogeographic relationships between each of the
terranes and with the Cordilleran continental margin, still
held secret, were first recognized by Monger and Ross
(1971). The debate still continues (for example, see Currie
and Parrish, 1993; Nelson and Mihalynuk, 1993). We have
neither the space nor the wish to address this major paleogeographic problem - - the late Paleozoic-early Mesozoic
deployment of these terranes. Instead, we simply pick
them up in the Jurassic, by which time most agree they are
all more or less accounted for in Cordilleran paleogeography (i.e., they were all amalgamated to one another and
the whole more or less "in contact" with North America),
and follow their consolidation into the Cordilleran orogenic edifice.

The most easterly of the terranes is the Kootenay,


which in its several subdivisions extends the length of the
Canadian Cordillera, and in its probable correlative, the
Yukon Tanana, continues into central Alaska (Mortensen,
1992). The possibly correlative Nisling terrane makes up
the metamorphic "core" of at least the central and northern
SAES 7 ; 3 - ~ B

249

part of the Coast belt and, as presently mapped, is continuous with the Yukon Tanana terrane. The Kootenay terrane
actually extends eastward into the Omineca belt; in fact,
locally it is thrust across the Omineca belt and occurs as
klippen on its east flank. The terrane consists generally of
a lower part made up of quartz-feldspathic meta-grits,
schists, and gneisses, and a more heterogeneous upper part
with much meta-volcanic content and orthogneisses.
These latter have yielded Late Devonian-Early Carboniferous ages, but other ages extend up into the Permian. No
assumed marine protoliths younger than Permian are
known from the terrane. The quartz-rich lower part is
assumed to be at least as old as early Paleozoic, but no
unequivocal Precambrian ages have been detected as yet,
except in detrital zircons. The widespread subhorizontal
tectonite fabrics typical of much of the terrane are intruded
by generally less deformed Upper Triassic to lower Middle Jurassic plutons. This suggests to Mortensen (1992)
that the fabric is Permo-Triassic in age. Locally, Upper
Triassic marine rocks sit as klippen on the terrane, often
associated with greenstones and ultramafic rocks that are
possibly part of the Slide Mountain terrane. Much of the
terrane has been locally severely affected by post-Middle
Triassic to possibly syn-Middle Jurassic thrust faulting,
Cretaceous intrusions, and major strike-slip faulting. The
Kootenay terrane is often assumed to be a distal, offshelf,
western facies equivalent to the Cordilleran miogeocline.
The fact, however, that the terrane records deformational
and magmatic events not recognized in the authochthonous miogeoclinal sequence to the east leaves it in "suspect" status. It is usually assumed that its tectonic
e m p l a c e m e n t onto North A m e r i c a took place in the
Middle Jurassic.
The Slide Mountain terrane in the central Canadian
Cordillera, where it is best known, consists of a tectonically interleaved sequence of radiolarian chert, basalt,
argillite, carbonate, diorite, gabbro, and ultramafic rocks
that range in age from at least Devonian to Late Triassic
(Harms et al., 1988). Possible correlative rocks of the
Angayuchum terrane in northern Alaska contain radiolarian cherts as young as earliest Jurassic (Coney and Jones,
1985). In northern British Columbia, the terrane is a klippe
(the Sylvester allochthon) sitting upon the duplexed carbonate platform sequence of the Cassiar terrane, a northward-displaced fragment of the Cordilleran miogeocline
that here is part of the Omineca belt. The Slide Mountain
terrane has a clear "oceanic" aspect and is usually interpreted as oceanic crust, or at least layer 2. Faunal and
paleomagnetic data suggest that the terrane was as far
south as the latitude of Mexico in Permian time (Richards
et al., 1991, and in press). The terrane includes marine
rocks as young as Late Triassic (possibly regionally as
young as Early Jurassic). In northern British Columbia,
Upper Triassic rocks may tie the terrane to a position close
to "Ancestral North America" (Gabrielse, 1991). The
Slide Mountain terrane was emplaced upon North America's margin prior to intrusion of Lower Cretaceous plutons
(Gabrielse, 1991). Possible correlatives were perhaps also
emplaced upon the Kootenay-Yukon Tanana terrane in
syn- to post-Late Triassic time, i.e., Middle Jurassic (?). In

250

P.J. CONEY and C.A. EVENCHICK

southern British Columbia, the Slide Mountain terrane


was emplaced onto the Kootenay/North American margin
in early Middle Jurassic time (Klepacki, 1989).
The remainder of the Intermontane belt is made up of
three extensive terranes: Quesnellia, Cache Creek, and Stikinia (from east to west). All three extend nearly the entire
length of the Canadian Cordillera, and the most westerly
of the three, Stikinia, is the second largest suspect terrane
in western North America (the Guerrero terrane of western
Mexico is the largest). Quesnellia and Stikinia are both
considered upper Paleozoic to lower Mesozoic intraoceanic arc assemblages, whereas Cache Creek is usually
interpreted as an upper Paleozoic to lower Mesozoic
accretionary complex of largely oceanic materials.
The Quesnellia terrane contains marine sedimentary
rocks as young as Bajocian in the south and Pleinsbachian
in the north (i.e., middle Early Jurassic), as well as plutonic rocks as young as Middle Jurassic. Its contacts with
"Ancestral North America" in northern British Columbia
are either complex strike-slip faults that yield Early Cretaceous metamorphic ages, or locally an east-vergent thrust
fault that yields a Late Jurassic isotopic metamorphic age
(Gabrielse, 1991).
The Cache Creek terrane is for the most part in contact
with Quesnellia along major strike-slip faults which often
contain extensive mylonite (Gabrielse, 1991). Depositional relationships between it and Quesnellia of Late Tria s s i c a g e s u g g e s t that C a c h e C r e e k m a y be the
accretionary complex of the Quesnellia arc. The terrane
contains marine sedimentary rocks as young as Early
Jurassic, but it may have ceased being an accretionary
complex in the Late Triassic.
The Stikinia terrane, like Quesnellia, contains upper
Paleozoic to Lower Jurassic submarine volcanic sequences
and associated marine sediments. Plutonic rocks were
emplaced during Late Triassic to Middle Jurassic time
(Wheeler and McFeely, 1991). In northern British Columbia, the terrane is in contact with the Cache Creek terrane
along an important system of southwest-vergent thrust
faults, of which the King Salmon fault is the most important. These faults cut rocks as young as Middle Jurassic.
Elsewhere, contacts with more easterly terranes are usually strike-slip faults. Triassic rocks thought to be associated with the northwestern part of Stikinia are said to be
depositional on the Nisling terrane (Jackson et al., 1991 ).
Most important in Stikinia, however, for our purposes,
are extensive marine to fluvial sedimentary rocks of the
"overlapping" Middle Jurassic to mid-Cretaceous Bowser
"basin" (Eisbacher, 1981; Evenchick, 1991, 1992). The
Bowser basin sits astride the entire width of central Stikinia terrane in northern British Columbia. Its strata,
including the Bowser Lake Group and clastic top of the
underlying Hazelton Group, were deposited on the Lower
Jurassic Hazelton volcanic arc rocks typical of Stikinia.
The Bowser Lake Group contains chert clasts of Cache
Creek origin, and its general source areas were to the
north, northeast, east, and south. The sequence can thus be
viewed as a foreland basin deposit of the King Salmon
fault system, where Cache Creek terrane was thrust over
Stikinia starting in Middle Jurassic time (Ricketts et al.,

1992). The entire sequence was then intensely deformed


and uplifted by near-classic folds and thrusts in a belt with
east-vergent geometry, producing at least 160 km of shortening, in pre-mid-Cretaceous to early Cenozoic time
(Evenchick, 1991, 1992). The deformation produced a
narrow foreland basin, the Sustut basin, on its eastern
front. Detrital mica, presumably from the Omineca belt,
first appeared in mid-Cretaceous time (Evenchick, 1992).
Interestingly enough, metamorphic detritus arrived here
about the same time as it did in the foreland east of the
Omineca belt (Price and Mountjoy, 1970).
Coast Belt. The Coast belt extends for 1700 km along
the entire length of the Canadian Cordillera and is between
100 to 400 km wide (Monger et al., 1982). Its main characteristic is its composition, mainly metamorphic and plutonic rocks, and a generally higher average elevation
compared to the Insular belt to the west or the Intermontane belt to the east. The plutonic rocks range in age from
mostly Jurassic to Eocene. South of Prince Rupert, Middle
Jurassic to mid-Cretaceous plutons are found in the west
of the belt, while the eastern part is mostly mid-Cretaceous
to Eocene plutons. North of Prince Rupert, the Late Cretaceous to Eocene plutonic belt continues and becomes the
p r e d o m i n a n t suite ( A r m s t r o n g , 1988: W h e e l e r and
McFeely, 1991). Metamorphic screens and septa are found
throughout the Coast belt, and at least in the north some
have a "continental" isotopic and lithologic signature very
similar to the Yukon Tanana-Nisling-Kootenay terranes
discussed earlier. How this southward-extending "sliver"
of older protolith got into a position "outboard" of the oceanic Quesnellia-Cache Creek-Stikinia terranes has been
the principal Canadian conundrum of two decades.

The boundary between the Coast belt and the Intermontane belt lies along the western side of Stikinia for most of
its length. It is defined by strike-slip faults in the south and
by the eastern limit of widespread granitic rocks further
north (Gabrielse et al., 1992). East-vergent thrust faults in
high-grade rocks at the east side of the southern Coast belt
may be deeper expressions of the Skeena fold-thrust belt.
In this case, the fold belt roots westward into and beneath
the Coast belt, which then is an uplifted "hinterland" to the
Skeena fold belt (Evenchick, 1991, 1992). The age of
these structures is mainly mid-Cretaceous to early Tertiary. The western margin of the Coast belt is a west-vergent deep-seated thrust system also of Late Cretaceous to
early Cenozoic age, which places the Coast belt over the
various terranes of "Greater Wrangellia" or the Insular belt
(Crawford et al., 1987). A distinctive suite of high-pressure Late Cretaceous plutons and some of the highest
grade Phanerozoic metamorphic rocks known anywhere in
western North America are found in the upper plate of this
thrust system. Deep marine turbidite and submarine volcanic successions of the Upper Jurassic to mid-Cretaceous
Gravina assemblage are generally involved in these structures in the north (Coney and Jones, 1985). Southward,
this zone seems to broaden and may cross the Coast belt
into the mrlange and broken formations of the upper Paleozoic to Middle Jurassic Bridge River terrane and the
Upper Triassic to mid-Cretaceous Methow-Tyaughton ter-

Consolidation of the American Cordilleras


ranes. Viewed in this way, much of the western part of the
Coast belt south of Prince Rupert may actually be Greater
Wrangellia itself, while the eastern part is perhaps a complex fragment-filled "suture zone" between the outboard
Greater Wrangellia terranes and the Intermontane terranes.
Whatever the Coast belt represents, for our purposes it was
consolidated into the Cordillera by Late Cretaceous to
early Cenozoic uplift, both east- and west-vergent contractional telescoping, and accompanying magmatism.
Insular Belt. The Insular belt comprises two principal
terranes - - the Wrangellia and Alexander terranes - - the
whole of which is sometimes referred to as a composite
Greater Wrangellia, the Insular superterrane, or "superterrane II" (Monger et al., 1982). The Alexander terrane has
a complex history extending from the Late Proterozoic to
the Triassic, almost all of which suggests an oceanic setting (Samson and Patchett, 1991). The Wrangellia terrane
has upper Paleozoic submarine arc assemblages overlain
by a distinctive Triassic section, suggesting an oceanic
plateau. This in turn is overlain by Lower to Middle Jurassic arc volcanic rocks and, in the Queen Charlotte Islands,
an extensive marine basin (Thompson et al., 1990). Along
the complex northern boundary zone with the Coast belt
are found several assemblages such as the quite deep
marine U p p e r Jurassic to m i d - C r e t a c e o u s GravinaG a m b i e r assemblages, which have the character of a
flysch basin with submarine arcs, and the Permo-Triassic
"oceanic" Taku terrane. In the south, where the Insular
superterrane has crossed the Coast belt ,the Bridge River,
Cadwaller, and other smaller terranes, all of which have an
"oceanic" aspect, are tectonically emplaced against the
Methow-Tyaughton Upper Jurassic-Cretaceous flysch
basins that form the boundary with the Intermontane terranes. The west side of the Insular belt has a series of
accretionary complexes ranging in age from Late Jurassic
to Cenozoic, and seismic reflection studies across Vancouver Island suggest large amounts of similar material have
been tectonically underplated from the west beneath much
of the Insular belt (Clowes et al., 1987).
Most of the Insular belt was below or close to sea level
during much of Triassic and Early to Middle Jurassic time.
Deep-marine conditions prevailed east (i.e., inboard) of it
until the mid-Cretaceous, from at least just south of the
Canada-USA border northward into the central Alaska
Range. The belt seems to have been consolidated into the
Cordillera during the Late Cretaceous-early Cenozoic as
the Insular superterrane underthrust the Coast plutonic
belt. Much of its actual uplift may be due to massive tectonic underplating from the west.
S u m m a r y o f Western Canada. There is seemingly no
evidence of major mountain building in the Canadian Cordillera until sometime in the Middle to Late Jurassic. All
the tectonic belts preserve evidence of marine rocks at
least through the late Paleozoic, and the miogeocline on
North American crust and almost all the "suspect" terranes
had marine rocks into Triassic time and many into the
early Middle Jurassic. Some terranes, particularly in the
Insular belt, and certain assemblages found along its
boundary with the Coast belt preserve deep-marine facies

251

into the mid-Cretaceous. This is in spite of the fact that


there is a considerable opinion that most of the "suspect"
terranes, at least the Intermontane group, were more or
less amalgamated with one another and more or less "in
place," although farther south, with respect to North
America during Triassic to Middle Jurassic time. In other
words, the "collisions," if any, or the process of amalgamation and possibly even initial accretion do not seem to
have produced major tectonic-orogenic edifices that have
left any evidence. The one exception is the Yukon TananaKootenay-Nisling terrane, but its major deformational fabric may be Permo-Triassic in age.
The process of consolidation of the Canadian Cordillera seems to have begun in the region of the Omineca belt
and eastern Intermontane belt during Middle to Late Jurassic time. The initial event may have been obduction of thin
slices of the eastern terranes, particularly Kootenay, Slide
Mountain, and Quesnellia, onto the distal edge of the
North American margin during Middle Jurassic time. At
roughly the same time the Cache Creek terrane began to
overide Stikinia. Soon after, enough of this region had
thickened and risen far enough above sea level to shed
debris westward onto the Stikinia terrane in Middle Jurassic time and also eastward onto the foreland in Late Jurassic time. The orogenic belt grew from thrusting as the
cratonic interior drove relatively under the rising edifice
from the east, a process which continued almost uninterrupted through Paleocene time. Although Stikinia began to
underthrust Cache Creek in Middle Jurassic time, much of
Stikinia itself and the remainder of the terranes to the west
remained close to, or well below, sea level. Consolidation
then began across the western Intermontane and Coast belt
in mid-Cretaceous time as Stikinia drove relatively westward beneath the Coast belt and as the Insular terranes
drove relatively eastward beneath the Coast belt. Tectonic
underplating from west of the Insular terranes completed
the consolidation into Cenozoic time. All of this was
apparently taking place in a generally right oblique
transpressive regime, producing major right strike-slip
faulting and significant disruption and translation, particularly in Cretaceous to early Cenozoic time.
Western United States and Mexico
The tectonic setting of the western United States in earliest Mesozoic time is often portrayed as one of marginal
basins and arc systems "off edge" from the miogeoclinal
shelf. These marginal basins are said to have then collapsed in Middle to Late Jurassic time (Oldow et al.,
1989). Reversal of sedimentary polarity in the central and
northern parts of the western United States within the
former miogeoclinal domain has usually been recognized
in the widespread but thin Morrison Formation, which is
roughly Oxfordian-Kimmeridgian in age (Elison, 1991).
Recent studies, however, suggest that, locally, the first
westerly derived sediments may be actually as old as
Bajocian. Heller et al. (1986) suggested that the widespread but thin Morrison Formation was not a foreland
basin deposit but rather eroded from a thermally uplifted
"hinterland" to the west in western Utah-Nevada. In their

252

P.J. CONEY and C.A. EVENCHICK

view, thrusting on the foreland did not begin until midCretaceous time. Evidence for Late Jurassic to mid-Cretaceous metamorphism and deformation in the "hinterland"
is reported (Hodges et al., 1992), however, as well as the
development of the Middle Jurassic to Early Cretaceous
plutonism, which migrated eastward after the collapse of
the "marginal" basins and offshore arcs. Crustal thickening in the "hinterland" probably began here in Late Jurassic time, as it did to the north in Canada, but some of the
uplift could also be due to the thermal input from the magmatism and incipient subcrustal "erosion" of upper mantle
lithosphere. In any event, consolidation of the Cordillera
gradually spread eastward during Cretaceous time as
thrust fronts moved into the interior Cretaceous "seaway,"
or foreland basin, as the Sevier fold-thrust belt in western
Utah.
A significant aspect of this general eastward advance of
deformation and m a g m a t i s m is the classic Laramide
deformation of the cratonic shelf east of the Paleozoic
miogeocline in the Wyoming-Colorado Rocky Mountains
and the Colorado Plateau. This event is rather neatly
bracketed between the Late Cretaceous and the late
Eocene and produced the rather unique basement-cored,
thrust-bound, crustal-scale uplifts so typical of this region.
Magmatic patterns have suggested a flattening of Benioff
zone dip as the cause, probably related to very high convergent rates at this time (Coney, 1976; Coney and Reynolds, 1977; Engebretson et al., 1985). One aspect of the
Rocky Mountain region, including much of the western
Great Plains to the east, is the abnormally high elevations.
Remembering that this region was at or below sea level
until Late Cretaceous time, the problem is: why is it consolidated to such high elevations today? Laramide crustal
thickening can explain some of this but probably not all,
and certainly not 50-km-thick crust under the Great Plains
(Gregory and Chase, 1992). It is perhaps possible that the
lithosphere under this region was already thick before the
Laramide deformation, and very cold, since no orogenic
activity had occurred here since about 1.7 Ga. This cold
lithospheric root must have thinned, or delaminated during
Laramide orogeny, thus allowing the somewhat already
overthickened inherited crust to rise to its Laramide elevations, which have persisted until today except for collapse
of the Basin and Range in and around it.
As in western Canada, contractional deformation in the
western United States largely ceased by the late Eocene.
In the southwestern United States and Mexico, the
early to mid-Mesozoic history is quite different from that
discussed above; it is actually more similar to what we
describe below for northwestern South America, with
which this region was connected until Early Cretaceous
time. A well developed continental margin magmatic arc
of latest Triassic to Jurassic age came out of marginal
Sierran trends to the north and was spread across southern
Arizona into and through eastern Mexico and into northwestern South America. This arc was probably in part
responsible for the continental red-bed detrital disturbance
of the Upper Triassic to Lower Jurassic strata so typical of
northern Arizona and much of the Colorado Plateau. This
arc, however, seems to have been quite "neutral," tectoni-

cally speaking, and no major orogenic edifice was produced. A second "fringing" system of magmatic arcs, now
preserved on Baja California, for example (Sedlock,
1993), may have accreted in Late Jurassic time. Opening
of the Gulf of Mexico in Early to Middle Jurassic time was
east of the continental margin arc, and the motion may
have transformed northwestward across Mexico as the
M o j a v e - S o n o r a m e g a s h e a r (Anderson and Schmidt,
1983). All of this is widely covered by thin platformal carbonate banks, mostly of Early to mid-Cretaceous age. The
first major reversal of sedimentary polarity to westerly
derived orogenic fluvial-deltaic floods across most of
Mexico began in Late Cretaceous time, when an off-edge
Upper Jurassic to mid-Cretaceous largely submarine arc
system, the Guerrero terrane, either accreted or more
likely rose above sea level in Late C r e t a c e o u s time
(Campa and Coney, 1983). Deformation then spread eastward across Mexico clear to the Gulf of Mexico as the
Laramide-age foreland fold-thrust belt of the Sierra Madre
Oriental. As in most of western North America, contractional deformation largely ceased by the late Eocene.
Probable equivalents of the G u e r r e r o terrane swept
through the opening between separating North and South
America as the Greater Antilles oceanic arc system.
Except locally, most of the consolidation of the North
American Cordillera was complete by Eocene time. This
was then followed by continued strike-slip disruption and
extensional collapse of much of the original orogen, development of the Caribbean plate, etc., as plate interactions
changed due to partial overriding of East Pacific spreading
centers (an aspect of North American Cordilleran history
that we cannot pursue here).

M E S O Z O I C - C E N O Z O I C T E C T O N I C S OF THE
ANDEAN C O R D I L L E R A
In contrast to the North American Cordillera where, as
we have seen, the transition to mountain building took
place in the Late Jurassic, in the Andean Cordillera the
same type of transition took place, but in the mid- to Late
Cretaceous instead, some 50 My later (Dalziel, 1986). The
nature of the transition in the Andes, however, is similar to
that described above for western North America in that it
is characterized by a reversal in sedimentary polarity from,
for the most part, craton or easterly derived, in generally
platformal sediments prior to the transition, to mainly
westerly derived orogenic foreland deposits after. Particularly in the northern Andes, but probably only there, the
transition also seems to be characterized first by the
obduction of oceanic terranes onto the South American
margin. Once started, the consolidation of the Andean edifice progressed to the present mainly by eastward advance
of magmatic arcs and fold-thrust belt fronts into the
Andean foreland, development of foredeeps, and progressive crustal thickening and uplift (Isacks, 1988).

Consolidation of the American Cordilleras

253

A
C-M
% % % % % % % % % ~

Col.

S%
J #%J S%#%~%
S #%#%f % % ~
S % ~ S $ S .f S". J S S $
J~IIV% J
~J#J
. .~. .J

0
Ce

Am

GS

.~ . l .J . J . ~ .

%%%%
~JJJsS~J

!fU

Brazil

10S

Bol.
Ca

500 KM

AQ

Fig. 5. Generalized terrane map of the Andes of Peru, northern Bolivia, Ecuador, and southern Colombia (after Richards and Coney,
1991; M6gard, 1989). Key: A, Amaime terrane; Am, Amotape terrane; B, Baudo terrane; AQ, Arequipa terrane; C-M, Cauca-Macuchi
terrane; Ca, Canta terrane; P, Pinon terrane; Z, Zamora terrane. Stippled area is that part of the Andean Cordillera presumed to be underlain by authochthonous cratonic South America: GS, Guiana shield; Ce, Celeca magmatic arc.

Andes of Peru, Ecuador, and Southern Colombia (Figs.


5 and 6)
In the Andes of Colombia, the Triassic and Jurassic of
the South American margin were characterized by continental red beds and associated rhyolitic to andesitic volcanism (M6gard, 1987, 1989). In most reconstructions of
Pangaea, eastern Colombia is continuous with Mexico
during Triassic-Jurassic time, and some combination of
Pacific margin continental arc and an incipient rift setting
related to the opening of the Central Atlantic and Gulf of
Mexico is not at all unreasonable.
In Peru during the Jurassic, the South American continental margin was a shallow marine shelf with widespread
carbonate deposition (M6gard, 1989). In southern Peru, a
largely submarine volcanic arc stood to the west of this
carbonate platform, but no rocks of this age are known
from central western Peru. In southern Peru and southward
through Bolivia into Chile and Argentina, this arc was
constructed on continental crust.
In Colombia during the Late Jurassic, however, the
regime changed, and marine conditions entered from the

northwest, presumably as South America began to break


away from southern North America. In contrast to the central Andes, however, significant oceanic terranes make up
the Western and Coastal Cordilleras in Colombia-Ecuador.
These terranes apparently accreted onto South America
starting in the mid-Cretaceous and into the mid-Cenozoic.
The accretion of these oceanic terranes seems to have cutoff and confined the marginal seas on the west in the midto Late Cretaceous. The first conglomeratic "foredeep"
deposits are of early Eocene age (M6gard, 1989), but the
first reversal of sedimentary polarity may have been in
Late Cretaceous time (R. B. Allen, per. comm., 1993). In
Maastrichtian time in Ecuador, the marine basins became
continental foreland basins. In Colombia, the sedimentary
polarity shifted to western source areas by at least the
early Eocene, as continental red beds flooded the foreland
region at this time.
In latest Jurassic time and continuing until the Late
Cretaceous, a new regime was established in Peru, consisting of the marine sedimentary platformal Peruvian basin
on the cratonic margin with a northwest-trending low arch

254

P.J. CONEY and C.A. EVENCHICK

Northern Andes of Colombia


Cauca-Macuchi T.

l~et~t~n

Central Andes of Peru


Arequipa T.

Canta T.

Central Peru and


Eastern Cordillera

0
- ,.,~'~,
m
100

. . . .
"
i
-

200

Paleogene

_2"-_

=
---~-~

"Capas
Rojas"

~
Cretaceous

P u n t a Piedra

:
,

Jurassic
. . . . . . .

~_.~hocolate

Pucara

[ ]

"

Fig. 6. Generalized tectonostratigraphic columns for the Central Andes of Peru and northern Andes of Colombia (after Mrgard, 1978,
1989; and compilations in the Laboratory of Geotectonics, University of Arizona). Heavy black horizontal line on each set of columns
is approximate age of oldest sea floor in the South Atlantic ocean. See Fig. 4 for legend.
which separated the feature into two sub-basins. Thickest
in the west, the Late Jurassic through Late Cretaceous
western basin is composed of uppermost Jurassic basal
shales followed by pure-quartz Neocomian sandstones,
then mixed carbonate platform layers interbedded with
sandstones and shales, all capped by extensive carbonates
of Santonian age (Wilson, 1963). Eastward, the sections
are thinner and sandier in composition, and they start in
the Albian. The sedimentary polarity was mostly craton
derived. To the west of this almost miogeoclinal-like basin
was a "eugeosynclinal" submarine volcanic arc, the Canta
terrane (Richards and Coney, 1991), which has yielded
Early Cretaceous fossils (Mrgard, 1987). This magmatic
arc seems to pass offshore in northern Peru, but the Celica
arc, mainly of Late Cretaceous age, emerges from beneath
a Tertiary basin in southwestern Ecuador and seems to
have continued northward as a continental margin arc
founded on the already accreted Amotape terrane and
South American basement of the east flank of the Central
Cordillera. The transition from the eastern platformal
basin to the western submarine volcanic arc in Peru is
interpreted as a facies change, but much of it is obscured
by faulting and massive e m p l a c e m e n t of the Andean

batholith, which began at about 100 Ma and continued into


the Miocene, or it is covered by Cenozoic volcanic rocks.
The fact that the Arequipa terrane seems to lie outboard of
the southern extremity of the Canta terrane suggests the
arc was either on the edge of South America or not far offshore.
The transition from mainly carbonate shelf to mainly
fluvial continental red beds in Peru seems to have taken
place in the Late Cretaceous, and source areas are from the
west (Mrgard, 1989, his Fig. 11). This seems to mark the
emergence of the magmatic region west of the platform.
The consolidation of the arc region may have begun as
early as 100 Ma, based on folded arc material intruded by
dated plutons (Cobbing et al, 1981). Thrusting and deformation of granitoid masses is suggested by both geologic
and isotopic data from latest Cretaceous time, and the arc
assemblages may have been thrust against the platform
during this early stage of consolidation. Thereafter, a
major period of thrusting in the platform sequences spread
across the present high Andes, with quite tight, generally
east-vergent folding and thrusting that detached the Upper
Jurassic to m i d - C r e t a c e o u s stratigraphy from some
unknown basement (Coney, 1971a; Mrgard, 1978), per-

Consolidation of the American Cordilleras


haps during the Eocene and into the Oligocene. In general,
the earliest conglomeratic foredeep deposits are of this
age. Finally, thrusting moved eastward into the subAndean foreland during the Miocene to the present. Continental arc magmatism, both volcanic and plutonic, particularly since the latest Eocene, has also spread eastward.
Southward in Bolivia, similar conditions prevailed, but
the Late Jurassic to Late Cretaceous interior basins were
largely continental and were accompanied by extension
and rifting behind the Jurassic-Cretaceous marginal arcs
along the Pacific margin (Sempere et al., 1990; Sempere,
in press). Starting in the Late Cretaceous (Senonian), the
region became a foreland basin from a rising hinterland in
the west. This upland region was apparently deformed
mainly during the Late Oligocene to early Miocene, at
which time the foredeeps in the Eastern Cordillera formed.
The mafic and ultramafic terranes of the Pacific margin
of Colombia and Ecuador (Goossens and Rose, 1973) generally extend from the western flank of the Central Cordillera to the coast across the Western and/or Coastal
Cordilleras. The most inboard terrane, now narrow slivers
along the western flank of the Central Cordillera termed
the Amaime terrane (Aspden and McCourt, 1986; Richards and Coney, 1991), was apparently accreted by obduction during the mid-Cretaceous. In space, but not in time,
these accretions recall the relationships described in western Canada for the klippen of the Slide Mountain terrane.
They have MORB compositions, and even some komatiitic basalts are known. Jurassic and Early Cretaceous
K-Ar ages are reported. The Amotape terrane, a probable
"micro-continent" now found in northernmost coastal
Peru and southernmost Ecuador (M6gard, 1989), has a
pre-Permian metamorphic basement. It may have accreted
in mid-Cretaceous time, embedded in the Amaime terrane.
The oceanic terrane of the Western Cordillera in Colombia, termed the Cauca-Macuchi terrane, is described as
thick thrust sheets, some of which reach the Central Cordillera of Colombia. Cretaceous fossils are reported, and
emplacement of the terrane apparently had begun by Late
Cretaceous time. In western Ecuador, the complex oceanic
arc-ocean floor Pinon terrane of Cretaceous to Eocene age
apparently accreted to the margin during Late Cretaceous
and/or Eocene-Oligocene time. Most of the deformation
seen in the Andes of Ecuador is thought to coincide with
or post-date this accretion. Finally, the most outboard terrane, the Paleogene Baudo terrane, accreted on the Colombian m a r g i n in M i o c e n e time. T h r u s t i n g had now
advanced eastward into the sub-Andean foothills of both
Colombia and Ecuador, as in Peru.

Southern Andes
In the central southern Andes of Chile and Argentina a
continental margin arc existed amid a system of complex
shifting intra-arc marine basins during Jurassic time and
extending into the mid-Cretaceous (Ramos, 1988b; Herv6
et al., 1987). The marine intra-arc basins were strongly
affected by Pacific, and eventually South Atlantic, transgressions and regressions of global, eustatic nature. Presumably the volcanic centers themselves were above sea

255

level. Starting in late Neocomian time, the arc edifice


migrated eastward and apparently coalesced into a single
arc with a single "retro-arc" basin which, from this point
on, remained continental. Farther south in the southernmost Andes, the "back-arc" area was apparently extensional to the extent that oceanic crust was produced
(Dalziel, 1981). The point to be drawn from all this is that
much of the Andean margin during Jurassic and most of
Cretaceous time was very close to sea level, with back-arc
shallow seas and/or extensional basins, despite active subduction and the fact that the basement of the arc system
was continental crust.
The first sign of significant consolidation from folding
and thrusting within and east of the arc "massif' seems to
date to the mid- to Late Cretaceous, when the volcanic and
structural fronts had migrated into western Argentina, east
of which the foreland basin evolved. In southern Chile the
extensional back-arc basin collapsed, which as Dalziel
(1981) noted, correlated with the opening of the South
Atlantic Ocean. From the Late Cretaceous until the present
time, the t h r u s t i n g and folding has p r o g r e s s i v e l y
encroached across the foreland, apparently somewhat
spasmodically, with major pulses of contraction in the
Eocene, then latest Oligocene to the present. This last
phase, still very active, also saw segmentation of the chain
into flat slab and normal slab-dip portions, which seem to
strongly influence structural style in the evolving thrust
belt (Jordan and Gardeweg, 1989). The most distinctive of
these segments are the basement-cored, thrust-bound
uplifts of the Pampean ranges north of Mendoza, which
coincide with a flat-slab segment and are modern analogs
for the classic Laramide Rocky Mountain uplifts of the
central western United States (Jordan and Allmendinger,
1986).

Summary of the Andean Cordillera


Along the entire length of the Andean Cordillera, over
a distance of almost 7000 km, construction of the Andean
orogenic edifice began in mid- to Late Cretaceous time
and extended into the early Cenozoic (Dalziel, 1986). This
is recorded in accretion of oceanic terranes, collapse of
marginal basins, termination of widespread marine conditions, reversals of sedimentary polarity from easterly craton-shield derived to westerly arc massif, or hinterland,
derived materials, initiation of folding and thrusting and
concomitant crustal thickening and resulting uplift, and
the initial development of foreland fold-thrust belts and
associated foreland basins. In the case of the Andes, particularly from northern Peru south, it is also clear that a
convergent continental margin plate tectonic regime
existed for over I00 My during Jurassic and much of Cretaceous time before the consolidation of the Andean Cordillera began. Evidence for this is found in the the well
developed continental margin magmatic arcs standing on
continental crust, but never far above sea level, usually
with shallow marine platforms or deeper extensional backarc basins to the east and/or intra-arc marine basins. In
other words, in spite of the convergent continental margin
plate tectonic setting, no major orogenic edifice was pro-

256

P.J. CONEY and C.A. EVENCHICK

duced until Late Cretaceous time. Finally, in the northern


Andes, and there alone, consolidation seems to have
begun after the accretion of oceanic terranes.
DISCUSSION AND CONCLUSIONS
The evidence shows that during the early Mesozoic, the
Pacific margins of North and South America were either
convergent continental plate margins (South America) or
seemingly were convergent continental margins in part
and perhaps had "off-edge" intra-oceanic arcs at some still
undefined distance in other sectors (North America). In
spite of this, no continuous Cordillera-wide mountain
chain developed in either case until the Late Jurassic in
western North America and until the Late Cretaceous in
the Andes. Thus, these margins were in some sense "neutral" (Dewey, 1980), meaning that although in active convergent settings, or close to them, no orogenic edifice, at
least compared to the mighty edifices that followed, was
produced during a period of at least 120 My in the case of
the Andean Cordillera, and something on the order of at
least 60 My in the case of the North American Cordillera.
What caused the transition, in both cases, that initiated the
consolidation and resulting mountain building and what
sustained the process once started?
This difference in tectonic response between the early
Mesozoic and the late Mesozoic-Cenozoic in the American Cordilleras has been described in various terms not
only in the western Americas but elsewhere. It has been
described, for example, as a Marianas-type convergent set-

Western N. America

ting as against a Chilean-type setting (Uyeda and Kanamori, 1979). The assumption is that the former is a "low
compressive stress" setting while the latter is a "high compressive stress" setting, an assumption actually supported
by seismology. In geological terms, this is also based in
part on the fact that the "back-arc" regions in the lowstress settings can be extensional, or at least neutral,
whereas the back-arc regions of the high-stress settings are
usually compressional, with foreland thrust belts common,
even though the relative convergent plate motions across
the boundaries in question may be very similar.
Explanations offered for the difference in tectonic style
in the two settings have ranged over the following in the
past 20 years:
age of subducting lithosphere (Molnar and
Atwater, 1978);
rate of convergence (Pardo-Casas and Molnar,
1987);
buoyant "asperities" on the down-going plate;
collisions and/or accretions of "exotic" terranes
(Nur and Ben-Avraham, 1982); and
"absolute" motion of the upper plate (Coney,
1971, 1973, 1978; Chase, 1978).
The debate has often revolved around the possible importance of variations in Benioff zone dip, which of course
can be thought of as influenced by all the above. It is not
our purpose here to exhaustively review this long-standing
debate in regional tectonics. We wish, however, to make a
few o b s e r v a t i o n s based on what we believe can be

Western

Subduction and
arc activity
T

100~

Subduction and
arc activity

"fore-<leap"

S. America

"fore-deep"

C
Reversal of sad.
'back-arc Sea"

Reversal of sed. pol.


"back-arc sea"

200
Breakaway from Gondwanaland
Fig. 7. Generalized tectono-sedimentary response on the forelands of the North and South American Cordilleras equilibrated in time to
the age of oldest sea floor (breakaway from Gondwanaland) in the Atlantic Ocean east of each of the continents.

I I A

80 W

80 W

Mid-Cretaceous

....... 11............

Relative motion c o n v e r g e n c e vector


in e m / y r

(locally includes transform)

Convergent m a r g i n s

S p r e a d i n g systems

Continental c r u s t

Basin a n d Range Province

F o r e l a n d basin deposits

Continental m a r g i n orogenesis

"Neutral" a n d / o r
"off-edge" m a g m a t i c arcs

80W

Late Cretaceous

Fig. 8. Advance of the American plates over the Pacific Ocean basin between the Middle Jurassic and the present in the "hot spot" frame. All reconstructions are centered on a fixed 80W. Spreading
centers in the Pacific generalized after Cole (1990) and Engebretson et al. (1985); relative motion vectors after Cole (1990); positions of the American continents manipulated after Scotese and Denham
(1988), with modifications after May and Butler (1986) and Cole (1990).

Ear

/////

Jurassic

e-~

t-~

258

P.J. CONEY and C.A. EVENCHICK

A
6 cm/yr

Continental margin stationary


in "hot spot" frame

"Neutnl" Off-Edge
MagmaticArc
/
_ I~

Sed. Polarity

Arc-Rmu"Basin

Plutons

ContinentalCrust/
6 cm/yr

B 10-2oMy

, ~

Reversalof SedimenhtryPolarity

9 cm/yr
%%%%%%%%%%%1~
%%%%%%%%%%%%%%%%%%%~
%%%%%%%%%%%%%%%%%%%%%%%%%~

Continental margin advances over adjacent ocean


in "hot spot" frame at 3 cm/yr
Intra-plate shortening at 0.5 cm/yr = 200 Km

C 3o-4oMy

8.5 cm/yr

.~

Fore.deep
/

ForelandFold-ThrustBelt
6 cm/yr

Ngl~e:Position of trench in C has moved


1,200 km to left in "hot spot" frame
during 40 My Since A

Fig. 9. Idealized sketches of the transition from "neutral" to compressive states of stress in the development of continental
margin orogens.

observed in the reconstruction of tectonic evolution in the


American Cordilleras.
We see no evidence that collision of "exotic terranes"
explains the transition from "low stress" to "high stress"
convergent tectonic settings in either of the American Cordilleras. The most compelling case against this model is
the Andean Cordillera, where from at least southern Peru
south, and probably from northern Peru south as well, no
large colliding terranes of Mesozoic-Cenozoic age are recognized (Dalziel, 1986; Dalziel and Forsythe, 1985). In
the northern Andes, the case has been made that accretion
of the mafic-ultramafic terranes found there does coincide
with initiation and subsequent pulses of contraction in
Colombia-Ecuador, but the similarity in timing throughout
most of the Andean chain of the mid- to Late Cretaceous
tectonic transition argues against this as a "driving force"
for the entire Andean Cordillera. The case has also been
made in western Canada in that accretion of "superterrane
I" coincides with initiation of deformation and metamorphism there, but as discussed earlier, the evidence seems
to be that the Intermontane terranes were already areal-

gamated to one another and some were tied to North


America well before the Late Jurassic. Thus, the collage
was already more or less loosely in place before consolidation began. In any event, a Himalayan model of indentation of collisional terranes to cause 100 My of telescoping
in western North America is impossible since, given the
reasonably well constrained relative motions of the FaralIon and/or Kula plates with respect to North America, the
terranes must have b e c o m e detached from whatever
"Pacific plates" they were part of almost instantaneously
(otherwise Wrangellia, for example, would have indented
far into central Canada by the end of Laramide deformation in the Eocene, whereas in fact it is splattered along the
Pacific margin from Idaho to southern Alaska; Coney,
1987).
Increases in rates of convergence are a possible cause
of inscreases in intensity of deformation in the continental
margins. This can be demonstrated along the Andean margin, particularly since the late Eocene (Pardo-Casas and
Moinar, 1987), and the Laramide orogeny in western
North America seems to correlate with high rates of con-

Consolidation of the American Cordilleras


vergence (Coney, 1976, 1978; Engebretson et al., 1985).
Although it is difficult to control convergence vectors back
as far as the two transitions, no significant change is noted
in the North American case (Engebretson et al., 1985). In
the case of South America, the passage of the PhoenixFarallon ridge southward down the Andes from the midCretaceous to the Late Cretaceous can actually yield a
decrease in relative convergence (see Fig. 8) as the triple
junction passes southward (Cole, 1990). Also, contacts
along the convergent margins of seamount chains and
other "asperities" could explain local changes in Benioff
zone dip, and thus have influenced local tectonic responses
and timing in the two Cordilleras, but these cannot explain
the widespread synchronous character of the transitions
over such large distances. In other words, changes in ages
of subducting lithosphere, asperities, and local, usually
fast evolving, relative plate interactions along continental
margins are probably the cause of "tectonic noise" certainly of local importance, but not the fundamental cause
of massive Cordilleran-wide consolidation that we seek.
We are, however, impressed by the following observation. In the case of both Cordilleras, the initiation of the
transition to "high stress" regime (i.e., the start of crustal
thickening from thrusting and folding, reversals of sedimentary polarity, etc.) seems to have come about 20-30
My after the first evidence for sea-floor spreading in the
portions of the Atlantic Ocean that lie east of the two continents (Figs. 4, 6, 7). In the case of North America, the
consensus is that the oldest oceanic crust between Africa
and North America in the Central Atlantic is of early
Middle Jurassic age (i.e., roughly Bathonian, which is
probably about 170 Ma; Vogt and Tusholke, 1989). The
age of the Fernie and Morrison Formation sedimentary
polarity reversal is about 150 Ma. In the case of South
America, the age of the oldest sea floor in the South Atlantic Ocean between Africa and South America is a little
younger than the Jurassic-Cretaceous boundary in the
south to about somewhere within the Cretaceous "magnetic quiet zone" farther north (i.e., between about 130 Ma
in the south and 110 Ma(?) in the north; Cande et al.,
1988), while the usual age of tectonic transition in the
Andes is between about 100 and 80 Ma. Seen in both Cordilleras, this correlation is interesting.
In most attempts to reconstruct the motions of the continents during Mesozoic-Cenozoic time, using (1) the
magnetic anomaly records from the ocean floors, (2) the
various APW paths of the continents, and (3) the somewhat imprecisely controlled so-called "hot spot" frames of
reference, the two times of initiation of sea-floor spreading
in the Central and South Atlantic Oceans also coincide in a
general way with initiation of significant prolonged absolute motion of the respective continents generally westward over the Pacific Ocean basin (Scotese and Denham,
1988; Cole, 1990). The details of timing and motions in
the various models vary some, mostly because it is very
difficult to control "absolute" motions of the various plates
before 80 Ma. In any event, in the kinematic scheme
shown here (see Fig. 8 ), prior to breakaway, while the two
continents were still part of Pangaea or Gondwanaland,
North and South America seem to have been relatively

259

stationary or moving northward with respect to the Pacific


Ocean basin in the "hot spot" frame. In some reconstructions (as in May and Butler, 1986), a clockwise rotation of
Pangaea pushed the southern Andes over the Pacific during Triassic-Early Jurassic time, but caused it to retreat
again in Late Jurassic-Early Cretaceous time (Cole, 1990).
In any event, as soon as "breakaway" started the two continents began a prolonged advance over the Pacific Ocean
floor in an "absolute" sense, pushing trenches ahead of
them, which has continued to the present time. North
America has probably moved 5000 km over the Pacific
since the Middle Jurassic, and South America probably
4000 km since the Early Cretaceous. This "absolute"
motion in itself, probably in conjunction with changes in
relative convergence vectors, was conducive to the initiation and continuation of "high stress" regimes in the upper
plates of the convergent plate margin settings and the
resulting contractional consolidation of the two Cordilleras during late Mesozoic-Cenozoic time.
The delay in the beginning of consolidation seen in
both Cordilleras (Fig. 7) is probably due, in part, to the
fact that it simply takes time to weaken and thicken the
crust on the leading edge of the continent, given the usual
rates of intraplate telescoping observed in continental margin contractional orogens (Fig. 9). These rates are usually
on the order of 0.2-0.5 cm/yr, and assuming original
crustal thicknesses of 25-30 km in the loose collages of
western North America during the Jurassic, a simple calculation shows that it would take 20 to 30 My for that
crust to thicken sufficiently to cause the uplift that would
produce the observed detrital foreland deposits. The subsequent encroachment of the spreading uplift and migration of foreland thrust belts eastward into the interior, and
the resulting delay in the formation of foredeeps, such as
noted by Heller et al. (1986), are probably explained by
Molnar's least-work principle (Molnar and Leon-Caen,
1988): When crustal thicknesses reach a certain point in
the softened magmatic arc-core massifs, or evolving "hinterlands," given the forces involved, it is easier for the orogen to grow wider rather than continue to thicken and
grow higher. Similarly, the role of the loose collages of
"suspect" terranes in western Canada and in the northern
Andes was probably quite "passive." They were simply
scraped up and consolidated as, like a snowplow, the
American continents pushed them ahead during the early
stages of the "advance" over the Pacific. To use a cruder
simile, they were like "bugs on the windshield" and played
no major dynamic role in Cordilleran consolidation. Similarly, the westward motion of the Americas must have carried them over "graveyards" of subducted oceanic crust
and thus thermally disturbed asthenosphere (Engebretson
et al., 1992). This may have caused thermal or mechanical
"erosion," or perhaps somehow triggered delamination, of
subcrustal upper mantle lithosphere, thus weakening and
heating the leading edge of the American plates and also
contributing to uplift, as in the case of the Laramide Rocky
Mountains discussed earlier.
In conclusion, an important factor in the initiation and
evolution of consolidation of the American Cordilleras
was the advance of the American plates over the Pacific

260

E J. CONEY and C . A . EVENCHICK

O c e a n b a s i n . A c c r e t i o n s , v a g a r i e s in r e l a t i v e m o t i o n vectors a l o n g t h e m a r g i n s , c h a n g e s in B e n i o f f z o n e dip, a n d
accidents of inherited crustal constitution and previous
h i s t o r y all p r o b a b l y e x p l a i n t h e v a r i a t i o n s in t i m i n g a n d
c h a r a c t e r o f t h e t e c t o n i c r e s p o n s e s e e n in t h e c o m p l e x
evolution of the two Cordilleras since consolidation
began.

Acknowledgments--We are grateful to T. Harms for encouragement to


pursue these ideas, and to H. Gabrielse, J. W. H. Monger, and J. O.
Wheeler of the Canadian Geological Survey, Vancouver, for discussions
and insight over a number of years which have helped our perceptions of
the Canadian Cordillera. David Richards provided much information on
the Andes and reviewed an early draft of this paper. Peter DeCelles
offered encouragement and preprints to clarify points on the western
United States. Discussions with Robert Butler and David Bazard clarified
paleomagnetic matters. Discussion with W. R. Dickinson was very helpful. We are also deeply grateful to Francisco Herv6 (Universidad de
Chile) and Victor Ramos (Universidad de Buenos Aires) for discussions
at the Vth Circum-Pacific Terrane Conference in Santiago and a very
informative and productive field excursion across the southern Andes of
Chile and Argentina. The manuscript received very constructive and
helpful formal reviews from R. B. Allen, lan Dalziel, Hubert Gabrielse,
and J. W. H. Monger. BHP Minerals International, San Francisco, and the
Canadian Geological Survey, Vancouver, provided partial support for this
work.

REFERENCES
Anderson, T. H., and Schmidt, V. A., 1983. The evolution of Middle
America and the Gulf of Mexico-Caribbean Sea during Mesozoic
time. Bulletin of the Geological Society of America 94, 941-966.
Archibald, D. A., Glover, J. K., Price, R. A., Farrar, E., and Carmichael,
D. M., 1983. Geochronology and tectonic implications of magmatism
and metamorphism, southern Kootenay arc and neighboring regions,
southeastern British Columbia, Part I: Jurassic to mid-Cretaceous.
Canadian Journal of Earth Sciences 20, 1891-1913.
Armstrong, R. L., 1988. Mesozoic and early Cenozoic magmatic evolution of the Canadian Cordillera. In: Processes in Continental Lithospheric Deformation (edited by S. P. Clark, Jr., B. C. Burchfiel, and J.
Suppe). Geological Society of America, Special Paper 218, 55-91.
Aspden, J. A., and McCourt, W. J., 1986. Mesozoic oceanic terrane in the
central Andes of Colombia. Geology 14, 415-418.
Brown, R. L., Journeay, J. M., Lane, L. S., Murphy, D. C., and Rees, C.
J., 1986. Obduction, backfolding and piggyback thrusting in the metamorphic hinterland of the southeastern Canadian Cordillera. Journal
of Structural Geology 8, 255-268.
Campa, M. F., and Coney, P. J., 1983. Tectonostratigraphic terranes and
mineral resource distribution in Mexico. Canadian Journal of Earth
Sciences 20, 1040-1051.
Cande, S. C., LaBrecque, J. L., and Haxby, W. E, 1988. Plate kinematics
of the South Atlantic: Chron 34 to present. Journal of Geophysical
Research 93, 479-492.
Chase, C. G., 1978. Extension behind island arcs and motions relative to
hot spots. Journal of Geophysical Research 83, 5385-5387.
Clowes, R. M., Brandon, M. T., Green, A. G., Yoreth, C. J., Brown, A. S.,
Kanasewich, E. R., and Spenser, C., 1987. Lithoprobe-Southern Vancouver Island: Cenozoic subduction complex imaged by deep seismic
reflections. Canadian Journal of Earth Sciences 24, 31-51.
Cobbing, E. J., Pitcher, W. S., Wilson, J. J., Baldock, J. W., Taylor, W. P.,
McCount, W., and Shelling, N. J., 1981. The Geology of the Western
Cordillera of Northern Peru. Institute of Geological Sciences, Over~as Memoir 5, 143 p.
Cole, G. L., 1990. Models of Plate Kinematics Along the Western Margin
of the Americas: Cretaceous to Present. Unpublished PhD dissertation, University of Arizona, Tucson, AZ, USA, 460 p.
Coney, P. J., 1971a. Structural evolution of the Cordillera Huayhuash,
Andes of Peru. Bulletin of the Geological Society of America 82,
1863-1884.

Coney, E J., 1971b. Cordilleran tectonic transitions and motion of the


North American plate. Nature 233, 462-465.
Coney, P. J. 1973. Plate tectonics of marginal foreland thrust-fold belts.
Geology 1, 13 I- 134.
Coney, P. J., 1976. Plate tectonics and the Laramide orogeny. New Mexico Geological Society, Special Paper 6, 5-10.
Coney, P. J., 1978. Mesozoic-Cenozoic Cordilleran plate tectonics. In:
Cenozoic Tectonics and Regional Geophysics of the Western Cordillera (edited by R. B. Smith and G. P. Eaton). Geological Society of
America, Memoir 152, 33-50.
Coney, P. J., 1987. Circum-Pacific tectonogenesis in the North American
Cordillera. In: Circum-Pacific Orogenic Belts and Evolution of the
Pacific Ocean Basin (edited by J. W. H. Monger and J. Francheteau).
American Geophysical Union, Geodynamics Series 18, 59-70.
Coney, P. J., 1989a. The North American sector of the Circum-Pacific. In:
The Evolution of the Pacific Ocean Margins (edited by Z. BenAvraham). Oxford Monographs on Geology and Geophysics 8, 4352.
Coney, P. J., 1989b. Structural aspects of suspect terranes and accretionary tectonics in western North America. Journal of Structural
Geology 11, 107-125.
Coney, P. J., 1990. Terranes, tectonics and the Pacific Rim. In: Proceedings, Pacific Rim Congress 90, pp. 19-30. Australian Institute of Mining and Metallurgy, Parkville, VIC, Australia, 746 p.
Coney, P. J., 1992. The Lachlan belt of eastern Australia and CircumPacific tectonic evolution. Tectonophysics 214, 1-25.
Coney, P. J., and Jones, D. L., 1985. Accretion tectonics and crustal structure in Alaska. Tectonophysics 119, 265-283.
Coney, P. J., and Reynolds, S. J., 1977. Cordilleran Benioff zones.
Nature 270, 403-406.
Coney, P. J., and Richards, D. R., 1991. The Andean Cordillera and
Circum-Pacific tectonic evolution. Comunicaciones (Santiago) 42,
54-58.
Coney, P. J., Jones, D. L., and Monger, J. W. H., 1980. Cordilleran suspect terranes. Nature 188, 329-333.
Crawford, M. L., Hollister, L. S., and Woodsworth, G. J., 1987. Crustal
deformation and regional metamorphism across a terrane boundary,
Coast Plutonic Complex, British Columbia. Tectonics 6, 343-361.
Currie, L., and Parrish, R. R., 1993. Jurassic accretion of Nisling terrane
along the western margin of Stikinia, Coast Mountains, northwestern
British Columbia. Geology 21,235-238.
Dalla Salda, L., Cingolani, C., and Varela, R., 1992a. Early Paleozoic
orogenic belt of the Andes in southwestern South America: Result of
Laurentia-Gondwana collision? Geology 20, 617-620.
Dalla Salda, L. H., Dalziel, I. W. D., Cingolani, C. A., and Varela, R.,
1992b. Did the Taconic Appalachians continue into southern South
America? Geology 20, 1059-1062.
Dalziel, I. W. D., 1981. Back arc extension in the southern Andes: A
review and critical reappraisal. Philosophical Transactions of the
Royal Society of London 300, 319-335.
Dalziel, I. W. D., 1986. Collision and Cordilleran orogenesis: An Andean
perspective. In: Collision Tectonics (edited by M. P. Coward and A.
C. Ries). Geological Society of London, Special Publication 19, 389404.
Dalziel, I. W. D., 1991. Pacific margins of Laurentia and East AntarcticaAustralia as a conjugate rift pair: Evidence and implications for an
Eocambrian supercontinent. Geology 19, 598-601.
Dalziel, I. W. D., 1992. On the organization of American plates in the
Neoproterozoic and the breakout of Laurentia. GSA Today 2, 237241.
Dalziel, !. W. D., and Forsythe, R. D., 1985. Andean evolution and the
terrane concept. In: Tectonostratigraphic Terranes of the CircumPacific Region (edited by D. G. Howell). Circum-Pacific Council for
Energy and Mineral Resources, Earth Science Series 1,565-581.
Dewey, J. E, 1980. Epidsodicity, sequence and style at convergent plate
boundaries. In: The Continental Crust and Its Mineral Deposits,
(edited by D. W. Strangway). Geological Association of Canada,
Special Paper 20, 553-574.
Eisbacher, G. H., 1981. Late Mesozoic-Paleogene Bowser molasse and
Cordilleran tectonics, western Canada. In: Sedimentation and Tectonics in Alluvial Basins (edited by A. D. Miall). Geological Association of Canada, Special Paper 23, 125-157.

Consolidation of the American Cordilleras


Elison, M. W., 1991. lntracontinental contraction in western North America: Continuity and episodicity. Bulletin of the Geological Society of
America 103, 1226-1238.
Engebretson, D. C., Cox, A., and Gordon, R. G., 1985. Relative Motions
Between Oceanic and Continental Plates in the Pacific Basin. Geological Society of America, Special Paper 206, 59 p.
Engebretson, D. C., Kelley, K. P., Cashman, H. J., and Richards, M. A.,
1992. 180 million years of subduction. GSA Today 2 (5), 93-100.
Evenchick, C. A., 1991. Geometry, evolution, and tectonic framework of
the Skeena Fold Belt, north-central British Columbia. Tectonics 10,
527-546.
Evenchick, C. A., 1992. The Skeena Fold Belt: A link between the Coast
Plutonic Complex, the Omineca Belt and the Rocky Mountain Fold
and Thrust Belt. In: Trust Tectonics (edited by K. R. McClay), pp.
365-376. Chapman and Hall, London, England, UK, 447 p.
Forsythe, R., 1982. The late Paleozoic to early Mesozoic evolution of
southern South America: A plate tectonic interpretation. Journal of
the Geological Society of London 139, 671-682.
Forsythe, R. D., Davidson, J., Mpodozis, C., and Jesininkey, C., 1993.
Lower Paleozoic relative motion of the Arequipa block and Gondwana: Paleomagnetic evidence from Sierra de Almeida of northern
Chile. Tectonics 12, 219-236.
Gabrielse, H., 1972. Younger Precambrian of the Canadian Cordillera.
American Journal of Science 272, 521-536.
Gabrielse, H., 1985. Major dextral transcurrent displacements along the
northern Rocky Mountain trench and related lineaments in northcentral British Columbia. Bulletin of the Geological Society of
America 96, 1-14.
Gabrielse, H., 1991. Late Paleozoic and Mesozoic terrane interactions in
north-central British Columbia. Canadian Journal of Earth Sciences
28, 947-957.
Gabrielse, H., and Yorath, C. J., 1991. DNAG #4: The Cordilleran orogen in Canada. Geoscience Canada 16, 67-83.
Gabrielse, H., and Yorath, C. J. (editors), 1992. Geology of the Cordilleran Orogen in Canada. Geological Society of America, Decade of
North American Geology G2, 844 p.
Gabrielse, H., Monger, J. W. H., Wheeler, J. O., and Yorath, C. J., 1992.
Part A: Morphogeological belts, tectonic assemblages, and terranes.
In: Geology of the Cordilleran Orogen in Canada (edited by H. Gahrielse and C. J. Yorath). Geological Survey of Canada, Geology of
Canada 4, 15-28. [Also: Geological Society of America, The Geology of North America G2.]
Goossens, P. J., and Rose, W. I., 1973. Chemical composition and age
determination of tholeiitic rocks in the basic igneous complex, Ecuador. Bulletin of the Geological Society of.America 84, 1043-1052.
Gregory, K. M., and Chase, C. G., 1992. Tectonic significance of paleobotanically estimated climate and altitude of the late Eocene erosion
surface, Colorado. Geology 20, 581-585.
Hamilton, W., 1969. The volcanic central Andes - - A modern model for
the Cretaceous batholiths and tectonics of western North America. In:
Proceedings of the Andesite Conference (edited by A. R. McBirney).
Oregon Department of Geology and Mineral Resources, Bulletin 65,
175-184.
Harms, T. A., Nelson, J. L., and Bradford, J., 1988. Geological transect
across the Sylvestor allochthon north of Blue River, northern British
Columbia. In: Geological Fieldwork 1987. British Columbia Ministry of Energy, Mines and Petroleum Research, Publication 1988-1,
245-248.
Heller, P. L., Bowdler, S. S., Chambers, H. P., Coogan, J. C., Hagen, E.
S., Shuster, M. W., Winslow, N. S., and Lawton, T. E, 1986. Time of
initial thrusting in the Sevier orogenic belt, Idaho-Wyoming and
Utah. Geology 14, 388-391.
Herv6, F., Godoy, E., Parada, M. A., Ramos, V., Rapela, C., Mpodozis,
C., and Davidson, J., 1987. A general view on the Chilean-Argentine
Andes, with emphasis on their early history. In: Circum-Pacific Orogenic Belts and Evolution of the Pacific Ocean Basin (edited by J. W.
H. Monger and J. Francheteau). American Geophysical Union, Geodynamics Series 18, 97-114.
Hedges, K. V., Snoke, A. W., and Hurlow, H. A., 1992. Thermal evolution of a portion of the Sevier hinterland: The northern Ruby Mountains-East Humboldt Range and Wood Hills, northeastern Nevada.
Tectonics 11, 154-164.

261

Hoffman, P. E, 1991. Did the breakout of Laurentia turn Gondwanaland


inside out.'?Science 252, 1409-1412.
Irving, E. M., 1975. Structural Evolution of the Northernmost Andes,
Colombia. U.S. Geological Survey, Professional Paper 846, 47 p.
lsacks, B. L., 1988. Uplift of the central Andean plateau and bending
of the Bolivia orocline. Journal of Geophysical Research 93,
321 !-3231.
Jackson, J. L., Gehrels, G. E., Patchett, P., J., and Mihalynuk, M. G.,
1991. Stratigraphic and isotopic link between the northern Stikine
terrane and an ancient continental margin assemblage, Canadian Cordillera. Geology 19, 1177-1180.
Jordan, T. E., and Allmendinger, R. W., 1986. The Sierras Pampeanas of
Argentina: A modern analog of Laramide deformation. American
Journal of Science 286, 737-764.
Klepacki, D. W., 1989. Linkages between the western North America
miogeocline and adjacent marginal terranes in the central Kootenay
Arc. Geological Society of America, Abstracts with Programs 21,
102.
Kluth, C. F., and Coney, P. J., 1981. Plate tectonics of the Ancestral
Rocky Mountains. Geology 9, 10-15.
May, S. R., and Butler, R. E, 1986. North American Jurassic apparent
polar wander, implications for plate motion, paleogeography, and
Cordilleran tectonics. Journal of Geophysical Research 91,519-544.
M6gard, E, 1978. Etude Gdologique des Andes du Peru Central. Memoires de I'ORSTROM 86, 310 p.
M6gard, E, 1987. Cordilleran Andes and marginal Andes: A review of
Andean geology north of the Africa elbow (18os). In: Circum-Pacific
Orogenic Belts and Evolution of the Pacific Ocean Basin (edited by J.
W. H. Monger and J. Francheteau) American Geophysical Union,
Geodynamics Series 18, 71-96.
M6gard, E, 1989. The evolution of the Pacifc Ocean margin in South
America north of the Arica Elbow (18S). In: The Evolution of the
Pacific Ocean Margins (edited by Z. Ben-Avraham). Oxford Monographs on Geology and Geophysics 8, 108-230.
Molnar, P., and Atwater, T., 1978. Interarc spreading and Cordilleran tectonics as alternates related to the age of subducted oceanic lithosphere. Earth and Planetary Science Letters 41, 330-340.
Molnar, P., and Lyon-Caen, H., 1988. Some simple physical aspects of
the support, structure, and evolution of mountain belts. In: Processes
in Continental Deformation (edited by S. P. Clark, Jr., B. C. Burchfiel, and J. Suppe). Geological Society of America, Special Paper
218, 179-208.
Monger, J. W. H., and Ross, C. A., 1971. Distribution of fusulinaceans in
the western Canadian Cordillera. Canadian Journal of Earth
Sciences 8, 259-278.
Monger, J. W. H., Price, R. A., and Tempelman-Kluit, D. J., 1982. Tectonic accretion and the origin of the two major metamorphic and plutonic welts in the Canadian Cordillera. Geology 10, 70-75.
Monger, J. W. H., Wheeler, J. O., Tipper, H. W., Gabrielse, H., Harms, T.,
Struik, U C., Campbell, R. B., Dodds, C. J., Gehrels, G. E., and
O'Brien, J., 1992. Part B: Cordilleran terranes. In: Geology of the
Cordilleran Orogen in Canada (edited by H. Gabrielse and C. J.
Yorath). Geological Society of America, Geology of North America
G2, 281-328.
Moores, E. M., 1991. Southwest U. S.-East Antarctica (SWEAT) connection: A hypothesis. Geology 19, 425-428.
Mortensen, J. K., 1992. Pre-mid-Mesozoic tectonic evolution of the
Yukon-Tanana terrane, Yukon and Alaska. Tectonics 11,836-853.
Nelson, J. and Mihalynuk, M., 1993. Cache Creek ocean: Closure or
enclosure. Geology 21, 173-176.
Nur, A., and Ben-Avraham, Z, 1982. Displaced terranes and mountain
building, In: Mountain Building Processes (edited by K. J. Hsu), pp.
73-84. Academic Press, New York, NY, USA, 263 p.
Oldow, J. S., Balley, A. W., Ave Lallemant, H., and Leeman, W. P., 1989.
Phanerozoic evolution of the North American Cordillera, United
States and Canada. In: The Geology of North America: An Overview
(edited by A. W. Balley and A. R. Palmer). Geological Society of
America, Geology of North America A, 139-232.
Pardo-Casas, and Molnar, P., 1987. Relative motion of the Nazca (Faralion) and South American plates since Late Cretaceous time.
Tectonics 6, 233-248.

262

P.J. CONEY and C . A . EVENCHICK

Parrish, R., Carr, S. D., and Parkinson, D. L., 1988. Extensional tectonics
of the southern Omenica belt, British Columbia and Washington.
Tectonics 7, 18 I-212.
Poole, E G., 1975. Flysch deposits of the Antler foreland basin, western
U.S. Society of Economic Paleontologists and Mineralogists, Special
Publication 22, 58-82.
Price, R. A., 1981. The Cordilleran foreland fold and thrust belt in the
southern Canadian Rocky Mountains. In: Thrust and Nappe Tectonics
(edited by K. R. McClay and N. J. Price). Geological Society of London, Special Publication 9, 427-448.
Price, R. A., and Mountjoy, E., 1970. Geologic structure of the Canadian
Rocky Mountains between Bow and Athabasca Rivers; A progress
report. In: Structure of the Southern Canadian Cordillera (edited by
J. O. Wheeler). Geological Association of Canada, Special
Publication 6, 7-25.
Ramos, V. A., 1988a. Late Proterozoic-early Paleozoic of South America
- - A collisional history: Episodes 11, 168-174.
Ramos, V. A., 1988b. The tectonics of the Central Andes; 30 to 33 latitude. In: Processes in Continental Deformation (edited by S. P. Clark,
Jr., B. C. Burchfiel, and J. Suppe). Geological Society of America,
Special Paper 218, 31-54.
Ramos, V. A., Jordan. T. E., Allmindinger, R. W., Mpodozis, C., Kay, S.
M., Cotes, J. M., and Paima, M., 1986. Paleozoic terranes oftbe central Argentine-Chilean Andes. Tectonics 5, 855-880.
Restrepo. J. J., and Toussaint, J. E, 1988. Terranes and continental accretion in the Colombia Andes. Episodes 11, 189-193.
Restrepo-Pace, P. A., 1992. Petrotectonic characterization of the Central
Andean terrane, Colombia, Journal of South American Earth
Sciences 5 (1), 97-116.
Richards, D. R., and Coney, P. J., 1991. Andean suspect terranes.
Comunicaciones (Santiago) 42, 194-197.
Richards, D. R., Buffer, R. F., and Harms, T. A., 1991. Paleomagnetism of
the late Paleozoic Slide Mountain terrane, northern and central British Columbia [abstract]. EOS, Transactions of the American Geophysical Union 72, 128.
Richards, D. R., Butler, R. E, and Harms, T. A., in press. Paleomagnetism
of the late Paleozoic Slide Mountain terrane, northern and central
British Columbia. Canadian Journal of Earth Sciences, in press.
Ricketts, B. D., Evenchick, C. A., Anderson, R. G., and Murphy, D. C.,
1992. Bowser Basin, northern British Columbia: Constraints on the
timing of initial subsidence and Stikinia-North America terrane interactions. Geology 20, 1119-I 122.
Ruiz, J., Centeno-Garcia, E., Coney, P. J., Patchett, P. J., and OrtegaGutierrez, E, 1991. Geology and geochemistry of the Guerrero terrune, western Mexico, and its possible correlation with rocks underlying the Greater Antilles and western Andes of Venezuela,
Colombia, and Ecuador. Geological Society of America, Abstracts
with Programs 23, A218.
Samson, S. D., and Patchett, P. J., 1991. The Canadian Cordillera as a
modern analogue of Proterozoic crustal growth. Australian Journal of
Earth Science 38, 595-611.
Scotese, C. R., and Denham, C. R., 1988. Terra Mobilis [software program]. Earth in Motion Technologies, Austin, TX, USA.

Sedlock, R. L., 1993. Mesozoic geology and tectonics of blueschist and


associated oceanic terranes in the Cedros-Vizcaino-San Benito and
Magdalena-Santa Margarita regions, Baja California, Mexico. In:
Mesozoic Paleogeography of the Western United States-H (edited by
G. Dunne and K. McDougall). SEPM (Society for Sedimentary Geology), Pacific Section, Book 71, 113-126.
Sempere, T., in press. Kimmeridgian? to Paleocene tectonic evolution of
Bolivia .In: Cretaceous Tectonics o f the Andes (edited by J. A.
Salfity). Earth Evolution Sciences Monograph Series, Vieweg Pub,
Wiesbaden, Germany.
Sempere, T., Herail, G., Oiler, J., and Bonhomme, M. G,, 1990. Late Oligocene-early Miocene major tectonic crisis and related basins in
Bolivia. Geology 18, 946-949.
Silberling, N. J., 1973. Geologic events during Permo-Triassic time along
the Pacific margin of the U. S. In: Permian and Triassic Systems and
Their Mutual Boundary (edited by S. Logan and E Hills), p. 345-362.
Alberta Society of Petroleum Geology, Calgary, AL, Canada.
Stewart, J. H., 1976. Late Precambrian evolution of North America - Plate tectonic implications. Geology 4, I l - 15.
Thompson, R. I., Haggart, J. W., and Lewis, E D., 1990. Late Triassic
through early Tertiary evolution of the Queen Charlotte Basin, British
Columbia, with a perspective on hydrocarbon potential. In: Evolution
and Hydrocarbon Potential of the Queen Charlotte Basin, British
Columbia (edited by G. J. Woodsworth). Geological Survey of Canada, Paper 9t1-10, 3-29.
Uyeda, S., and Kanamori, H., 1979. Back-arc opening and the mode of
subduction. Journal of Geophysical Research 84, 1049-1062.
Van der Voo, R., 1988. Paleozoic paleogeography of North America,
Gondwana, and intervening displaced terranes: Comparisons of
paleomagnetism with paleoclimatology and biogeographical patterns.
Bulletin of the Geological Society of America 100, 311-324.
Vogt, P. R., and Tucholke, B. E., 1989. North Atlantic Ocean basin:
Aspects of geologic structure and evolution. In: The Geology of North
America: An Overview (edited by A. W. Bailey and A. R. Palmer).
Geological Society of America, Geology of North America A, 53-80.
Wheeler, J. O., and McFeely, E (compilers), 199 l. Tectonic Assemblage
Map of the Canadian Cordillera and Adjacent Parts of the United
States of America. Geological Survey of Canada, Map 1712A, Scale
1:2,000,000.
Wilson, J., 1963. Cretaceous stratigraphy of the central Andes of Peru.
Bulletin of the American Association of Petroleum Geologists 47,
1-34.

Yorath, C. J., 1992. Upper Jurassic to Paleogene assemblages. In: Geology of the Cordilleran Orogen in Canada (edited by H. Gabrielse
and C, J. Yorath). Geological Society of America, Geology of North
America G2, 329-371.

Das könnte Ihnen auch gefallen