You are on page 1of 22

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/236646549

Oroclines: Thick and thin

Article  in  Geological Society of America Bulletin · May 2013


DOI: 10.1130/B30765.1

CITATIONS READS
87 1,216

3 authors:

Stephen T. Johnston Arlo B. Weil


University of Alberta Bryn Mawr College
181 PUBLICATIONS   3,566 CITATIONS    95 PUBLICATIONS   3,132 CITATIONS   

SEE PROFILE SEE PROFILE

Gabriel Gutierrez-Alonso
Universidad de Salamanca
328 PUBLICATIONS   6,481 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Stratigraphy and structure of the Mary River Group in the Tuktuliarvik area, northern Baffin Island, NU View project

Geology, Archaeology and UAVs View project

All content following this page was uploaded by Arlo B. Weil on 21 May 2014.

The user has requested enhancement of the downloaded file.


Oroclines: Thick and thin 18 8 8 2 013

CELEBRATING ADVANCES IN GEOSCIENCE

S.T. Johnston1,†, A.B. Weil2, and G. Gutiérrez-Alonso3


1 Invited Review
School of Earth & Ocean Sciences, University of Victoria, P.O. Box 3065 STN CSC, Victoria,
British Columbia V8P 4B2, Canada
2
Department of Geology, Bryn Mawr College, Bryn Mawr, Pennsylvania 19010, USA
3
Departamento de Geología, Universidad de Salamanca, Plaza de los Caídos s/n, 37008 Salamanca, Spain

ABSTRACT OROCLINES: INTRODUCTION genic curvature, and developed the methods


for classifying these systems. Our focus here
An orocline is a thrust belt or orogen that Understanding how map-view curvature of is on oroclines, how they might form, and their
is curved in map-view due to it having been geologic structures, ranging from individual implications for plate tectonics and global
bent or buckled about a vertical axis of rota- thrust sheets all the way to entire mountain belts geodynamics.
tion. Two distinct types of oroclines are recog- (Fig. 1), form and evolve is a first-order Earth Oroclines are most commonly inferred to
nized: progressive and secondary. Progressive system problem. Historically, studies of defor- have developed in response to an applied stress
oroclines are restricted to the scale of a thrust mation at the outcrop to map scale have relied oriented at a high angle to the long axis of an
sheet to thrust belt, are thin-skinned, and de- on profile views for understanding the kine- existing orogen (Ries and Shackelton, 1976).
velop during thrust sheet emplacement. Sec- matics and dynamics in compressional orogens Stress-perpendicular models include tensile
ondary oroclines are larger, occurring at the (Hobbs, 1914). The geologic map perspective bending in response to rollback of an adjacent
scale of an orogen, and are plate-scale features on the other hand has been chiefly used for slab (Russo and Silver, 1996; Linzer, 1996;
that affect crust and lithospheric mantle. Un- understanding the spatial distribution of geo- Lucente and Speranza, 2001; Johnston, 2004;
like progressive oroclines, which develop dur- logic structures. Thus, most orogenic systems are Rosenbaum and Lister, 2004; Schellart and
ing initial orogenesis and in response to the envisioned, and depicted, as longitudinal in na- Lister, 2004; Schellart et al., 2007; Stegman
same orogen-perpendicular stress responsible ture, and therefore their full four-dimensional et al., 2006; Cifelli et al., 2007, 2008), and com-
for thrust sheet emplacement, secondary oro- deformation fields are insufficiently character- pressive bending in response to wrapping about
clines are extra-orogenic, developing after ini- ized. This bias stems from the general simpli- a tectonic obstacle or indentor (e.g., Rankin,
tial orogenesis and in response to an orogen- fication of plane strain and the propensity for 1976; Thomas, 1977; Tapponnier et al., 1982;
parallel principal compressive stress that is the geometric construction of transverse sec- Eldredge et al., 1985; Cobbold and Davy, 1988;
oriented at a high angle to the stress responsi- tions in characterizing and describing orogenic Davy and Cobbold, 1991; Thomas and Whiting,
ble for orogen development. We present case systems. It was Carey (1955) who first recog- 1995; Stamatakos and Hirt, 1994; Keep, 2000).
studies of the Wyoming Salient, a progressive nized the importance of map-view deformation Such models are commonly employed to argue
orocline that characterizes the Sevier thrust of Earth’s crust and coined the term orocline for a “thin-skinned” interpretation of oroclines,
belt of the western United States, and the (Carey, 1955). However, the idea that significant place significant limits on the character and
coupled Cantabrian and Central Iberian oro- crustal and even lithospheric buckling and fold- tectonic setting of oroclines, and predict that
clines, which are linked secondary oroclines ing take place about vertical axes of rotation has oroclines form in response to the same tectonic
affecting the Variscan orogen of Iberia. The yet to be fully appreciated in studies of modern setting that gave rise to the orogen in the first
vertical-axis rotations involved in progressive and ancient mountain systems and in the growth place. An alternative explanation is that oro-
and secondary orocline formation are most of continental crust (e.g., Van der Voo, 2004; clines develop in response to an applied stress
readily quantified through paleomagnetic Gutiérrez-Alonso et al., 2004). oriented parallel to the long axis of an orogen.
analysis. Detailed three-dimensional palin- Map-view bends can be distinguished based Stress-parallel models require a plate-scale
spastic restoration that incorporates transla- on kinematics and the relative timing of curva- geodynamic interpretation in which orocline
tion rotation and strain can distinguish the ture (Weil and Sussman, 2004). Orogens and formation involves the entire lithosphere, with
role, if any, of primary curvature in progres- thrust belts characterized by primary curves are the additional requirement that the orocline in-
sive oroclines. The use of tectonic vectors, those in which the curvature is an inherited fea- volves two stages of development, because the
such as paleocurrent directions, provides a ture, present prior, during, and throughout the stress field responsible for orocline formation is
means of recognizing and characterizing the formation of the orogen. Oroclines, as first de- necessarily oriented at a high angle to the stress
initial geometry of secondary oroclines. Be- fined by Carey (1955), are map-view curves that field that gave rise to the initial orogen. Such a
cause secondary oroclines involve the entire developed in response to bending or buckling of switch in the orientation of the stress field begs
lithosphere, detailed studies of coeval meta- an existing orogenic belt about a vertical axis the question: Are there tectonic settings that can
morphism and magmatism provide a means of rotation. Previous reviews (e.g., Marshak, regularly give rise to such a stress-field rotation?
of constraining the fate of the mantle litho- 1988; Hindle and Burkhard, 1999; Marshak, We argue that there are two different types of
sphere during oroclinal buckling. 2004; Weil and Sussman, 2004) have, through oroclines: progressive and secondary (Fig. 2).
a focus on curved fold-and-thrust belts, defined Progressive oroclines (also termed progres-

E-mail: stj@uvic.ca the terminology necessary for describing oro- sive arcs by Sussman et al. [2004] and Weil

GSA Bulletin; May/June 2013; v. 125; no. 5/6; p. 643–663; doi: 10.1130/B30765.1; 14 figures.

For permission to copy, contact editing@geosociety.org


643
© 2013 Geological Society of America
Johnston et al.

A 111o
Cz Cenozoic basin fill r Pz WX CANADA
Cenozoic intrusion R ive
e Gr
TK Late Cret/Paleogene ak ain Cz
Sn P l os
V
K Cretaceous Up ent
Mz Triassic-Jurassic li re Montana
Pz Paleozoic Pz WX Pz Pz

Helena
Salient
Z Neoproterozoic Cr ld t
aw ra
fo

Zl Neoprot lower fo in Mz Mz
WX Archean/ rd
Paleoproterozoic
Wyoming
K K

hrust Belt
43o
K
Cz

Wyoming
Z

Prospec rby
Pz

Salient
Pocatello M Idaho
ea

Sevier T
Nevada
de TK
Mz

Da
t
Pz

Sali l Utah
Pa

Cz Pz
ri

ent
Utah Colorado
s

Cz

tra
Cen
Z
K Mz
Pz
Mz
Cz K
Mz
Hogsback
Bear
Lake

TK Arizona
Idaho 42o CA New
Utah Mexico
Pz Sevier Thrust Front
Pz N
Absaroka

Cz Laramide Fault/Uplis

Laramide Arch/Uplis
Willard

0 300
Pz Mexico

Pz
ford

Figure 1 (on this and following page).


Z Geological maps of (A) the Wyoming
Craw

Pz Salient progressive orocline (after


Green
River Yonkee and Weil, 2010a), and (B) the
Mz TK
Mz Basin coupled Ibero-Variscan secondary
oroclines (after Shaw et al., 2012).
Great (A) A generalized geologic map is
Salt Pz shown at left and a location map at
Lake Wyoming right. The location map shows the po-
WX
N K sition of the Wyoming Salient within
the Sevier thrust belt. Within the
Wyoming Salient, major thrust faults
K and folds curve from NW trends in
wood arch
Salt Lake Mz Mz Uinta-Coon the north to NE trends in the south.
City Pz
0 Abbreviations on following page.
50 km Zl
Pz

644 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

Figure 1 (continued). Geologi-


cal maps of (A) the Wyoming B MTSZ OUTCROPPING / COVERED
Variscan foreland

N
Salient progressive orocline fold-and-thrust belt

SZ
PT
(after Yonkee and Weil, 2010a), GTMZ Variscan outer hinterland
and (B) the coupled Ibero-Vari- fold-and-thrust belt
scan secondary oroclines (after Variscan hinterland core with
JPS CZ
Shaw et al., 2012). (B) A geologi- Z WALZ Cambro-Ordovician Magmasm
cal map of the Variscan orogen OC
BA OMZ
CIZ DRF Exoc Terranes with
of Iberia showing the major Ophiolites and High-P Rocks
DUF
tectonostratigraphic zones. Two Lower Allochthon-
SPZ JC Parautochthon
coupled oroclines are evident, Iberian

BCSZ
Autochthonous terrane with
including the convex-to-the- Massif strong Cadomian-Pan African

NP
west Cantabrian orocline in the PY imprint

F
north, and the convex-to-the- External thrust belt
east Central Iberian orocline CCR foredeep basin
to the south. BAOC—Beja- Bec deformaon front
Acebuches ophiolite complex; 300 km Variscan Strike-Slip
BCSZ—Badajoz-Cordoba shear Shear Zone
zone; CCR—Catalonian Coast
Ranges; CZ—Cantabrian zone; CIZ—Central Iberian zone; DRF—domain of recumbent folds; DUF—domain of upright folds; GTMZ—
Galician Tras-os-Montes zone; IC—Iberian Cordillera; JPSZ—Juzbado-Penalva shear zone; MTSZ—Malpica-Tui suture zone; NPF—North
Pyrennean Fault; OMZ—Ossa Morena zone; PTSZ—Porto-Tomar shear zone; PY—Pyrenees; SPZ—South Portuguese zone; WALZ—West
Asturian-Leonese zone.

Figure 2. A series of schematic block dia- A B


grams showing (A) the development of an
orogen or fold-and-thrust belt (D1) in re-
sponse to a stress (σ1, indicated with red
arrows) oriented at a high angle to the de-
veloping orogen. Gray indicates crustal
basement and mantle lithosphere; lighter D1σ1 D1 D2
colors and white indicate a supracrustal
stratigraphic package sitting depositionally
above the basement. (B) Progressive oroclines (D2) result from bending C
of thrust sheets and folds in response to the continued application of the
same stress field responsible for primary orogenesis and end down against
the basal décollement of the thrust belt. (C) Secondary oroclines develop in
response to the subsequent application of an orogen-parallel stress (indi-
cated by blue arrows), leading to the buckling of the entire orogen, includ-
ing its basement and mantle lithosphere, about a vertical axis of rotation.
Figure is modified from Weil et al. (2012b). D2σ1 D2

and Sussman [2004]) are: (1) restricted in scale pansive, and occur at the scale of an orogen; Distinguishing evidence between primary
and setting to thrust sheets, folds, and fold- (2) plate-scale structures that involve the crust curvature and oroclines is based on the kine-
and-thrust belts; (2) thin-skinned, with bending and likely the lithospheric mantle; (3) second- matic evolution of, and the structures formed
being restricted to the crust lying above a thrust ary, or extra-orogenic, meaning that buckling in response to shortening within, orogens and
or basal décollement; (3) progressive, meaning postdates and occurs after orogen formation, the relative timing of curvature acquisition with
that bending develops during and in response albeit commonly shortly after crustal thicken- respect to their protracted evolution. Thus, in-
to ongoing crustal shortening and thrust sheet ing; and (4) attributable to an orogen-parallel trinsic data are required for the analysis and
emplacement; and (4) attributable to the same compressive stress oriented at a high angle to the interpretation of finite rotations (McCaig and
strike-perpendicular, compressive stress respon- compressive stress responsible for the original McClelland, 1992). Traditionally, the best
sible for development of the bent thrusts, folds, orogen. Secondary oroclines therefore require way to quantify vertical-axis rotation has been
or fold-and-thrust belts. Hence, the stress re- a dramatic switch of the prevailing stress field, through paleomagnetic analysis (e.g., Irving and
sponsible for progressive oroclines lies in, and is and develop in response to a compressive stress Opdyke, 1965; Kotasek and Krs, 1965; Grubbs
parallel to, the axial surface that bisects the arc field oriented at a high angle to the axial surface and Van der Voo, 1976; Channell et al., 1978;
of curvature. Secondary oroclines are: (1) ex- that bisects the orocline’s arc of curvature. Van der Voo and Channell, 1980; Schwartz and

Geological Society of America Bulletin, May/June 2013 645


Johnston et al.

Van der Voo, 1984; Kent and Opdyke, 1985; passive-margin section, and an eastern thrust dence of bending. On a larger scale, there is no
Eldredge et al., 1985; Lowrie and Hirt, 1986; system with closer-spaced thrusts developed indication (for example, magmatism or regional
Miller and Kent, 1986a, 1986b; Eldredge and in the cratonic section. Thrust development metamorphism) that bending of the thrust belt
Van der Voo, 1988; Kent, 1988; Muttoni et al., and propagation direction were forelandward involved any portion of the crust or lithosphere
1998, 2000; Weil et al., 2000, 2001; Pueyo et al., (west to east), from Early Cretaceous to Paleo- beneath the basal décollement. The footwall
2003, 2004; Weil, 2006; Weil et al., 2010b). gene time, with synorogenic strata deposited in does, however, appear to have played a role in
Rotation analysis, if done at the appropriate foreland basins (Armstrong and Oriel, 1965; the development of the Wyoming Salient. Base-
scales (Yonkee and Weil, 2010a), can quantify Armstrong, 1968; Royse et al., 1975; Royse, ment uplifts characterize the north and south
the magnitudes and distributions of block rota- 1993; DeCelles, 1994, 2004). Major thrusts margins of the salient, and the greatest degrees
tions between individual sites and provide key display typical ramp-flat geometries and have of curvature are found adjacent to these uplifts
data to test kinematic models. Vertical-axis rota- associated large-scale fault-bend and fault- (Grubbs and Van der Voo, 1976; Schwartz and
tion, however, is just one component of deforma- propagation folds. Structural imbrication was Van der Voo, 1984; Craddock et al., 1988; Weil
tion, and thus any viable kinematic model must thin skinned, occurred above a basal décolle- et al., 2010b), and there is a significant transfer
also be consistent, and integrated with, other ment located within Lower Paleozoic strata, zone in the central portion of the salient related
geologic data (e.g., Gray and Stamatakos, 1997; and resulted in translation of imbricated strata to the Moxa Arch (Kraig et al., 1987, 1988).
Hindle and Burkhard, 1999; Kwon and Mitra, east toward the autochthonous North Ameri- These observations suggest that the basement
2004; Pueyo et al., 2004; Weil and Sussman, can craton. Major thrust and fold traces display uplifts acted as buttresses that served to impede
2004; Yonkee and Weil, 2010b; Pastor-Galán systematic map-view curvature (~90°), which the eastward advance of Sevier thrust sheets.
et al., 2012a). Few oroclines have been stud- defines the Wyoming Salient. The Gros Ventre Alternatively, thinning of the stratigraphic se-
ied in detail, and we necessarily focus on those and Uinta basement uplifts bound the salient quence toward the bounding uplifts may have
for which there are significant geological and to the north and south, respectively. Paleozoic hindered eastward advance of the adjacent thrust
paleomagnetic data. Two case studies are pre- strata thin across these basement uplifts, indi- sheets by limiting the growth of the critical
sented: (1) The Wyoming Salient of the Sevier cating that they are long-lived features of the Coulomb thrust wedge. Regardless, the char-
thrust belt, western United States, provides autochthonous foreland. acter of the footwall to the thrust belt appears
a type example of a progressive orocline; and intimately related to the development of the
(2) the coupled Cantabrian and Central Iberian The Wyoming Salient Wyoming Salient.
oroclines of the Variscan orogen of Iberia exem- Recent paleomagnetic, strain, and large-scale
plify the characteristics of secondary oroclines. Thrust sheets of the Wyoming Salient are con- structural patterns indicate that the Wyoming
Progressive oroclines are well understood and vex toward the foreland, the direction in which Salient began with a component of primary
studied; secondary oroclines are not. Hence, we the thrusts verge. The amount of curvature is curvature related to the geometry of the ini-
finish by exploring the tectonic implications of ~90°; the thrust sheets strike northwest in the tial sedimentary prism, and that eastern thrust
secondary oroclines, discuss the plate-tectonic north and become progressively more southwest sheets experienced 60%–80% progressive rota-
setting responsible for their development, and striking near their southern termination (Fig. 1). tion (Weil and Yonkee, 2009; Yonkee and Weil,
address the applicability of our model for sec- Total shortening across the Sevier belt is ~150– 2010a; Weil et al., 2010b). Progressive rotation
ondary orocline formation by assessing the An- 300 km (Royse, 1993; Yonkee, 1997; DeCelles appears related to along-strike changes in thrust
dean oroclines. and Coogan, 2006). The early emplaced Willard, slip directions and fold-thrust shortening, and to
Paris, and Meade thrust sheets within the more interaction of frontal thrusts with Laramide fore-
PROGRESSIVE OROCLINES: internal, structurally high, hinterland region land uplifts at the salient ends. Internal defor-
THE WYOMING SALIENT are more bent than the younger, more external, mation in the western thrust system was limited
structurally lower, foreland proximal Crawford, within upper levels above a strong quartzite
Regional Geology Absaroka, and Hogsback thrusts. In addition, the interval, whereas lower levels underwent com-
hinterland proximal thrust sheets show greater bined simple shear and vertical thinning that
The Sevier fold-and-thrust belt in northern internal strain (locally >30%), including the de- increased toward the base of the thrust sheet,
Utah to western Wyoming (Fig. 1) records a velopment of a strike-parallel cleavage and asso- likely reflecting a weak basal fault zone. Early
protracted geologic history, from accumula- ciated fractures, vein networks, and minor folds layer-parallel shortening and minor tangential
tion of a sedimentary prism, to development and faults (e.g., Crosby, 1969; Mitra and Yonkee, (strike-parallel) extension were widespread in
of an orogenic wedge and foreland basin sys- 1985; Craddock et al., 1988; Protzman and the eastern system. Layer-parallel shortening
tem, to final-stage thrusting that overlapped Mitra, 1990; Craddock, 1992; Mitra, 1994; Weil directions define a radial pattern that reflects
with development of Laramide foreland uplifts and Yonkee, 2009; Yonkee and Weil, 2010a). both initial dispersion about a regional E-W di-
(e.g., Rubey and Hubbert, 1959; Armstrong, Strain increases toward the salient ends and de- rection, and secondary rotation. Thin-skin fold-
1968; Burchfiel and Davis, 1972; Jordan, 1981; creases toward the foreland to very low values in thrust shortening in the Sevier belt developed
Boyer and Elliott, 1982; DeCelles et al., 2009; the frontal Hogsback thrust system (Yonkee and contemporaneously with, and was balanced
Yonkee and Weil, 2010a; Weil et al., 2010b). Weil, 2010a). The main strain consists of strike- by, lower-crustal shortening in a metamorphic
Sedimentary prism architecture was influenced normal shortening, and the principal shortening hinterland to the west. Overall, geologic obser-
by underlying basement structure, and included directions are subperpendicular to structural vations and paleomagnetic data indicate that
a west-facing passive-margin section that accu- trend (Fig. 3) (Yonkee and Weil, 2010a). There bending in the Wyoming Salient was “progres-
mulated above rifted basement and a more east- is no indication of bending of the footwall to the sive”; the bulk (up to 75%) of the observed bend
erly cratonic section that accumulated on old thrust belt. The basal décollement lies within is the result of rotation of originally more linear
continental crust. A western thrust system with Cambrian shales, and underlying, structurally thrust sheets, and bending appears to have been
extensive, thick thrust sheets developed in the deeper Cambrian and older strata show no evi- coeval with thrust sheet emplacement.

646 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

A Gros
Ventre B C D F
Arch

Darb
Me Prospe
y
ad
e
c t
Pa

Stable
r

Foreland
is

Sites (F)
Strain ellipse
parallel to
Ankareh bedding
Tr Pmag
LPS AMS fabric
med/high

low strain Weak


Lineaon K Pmag
Idaho very low
Moderate
Utah Lineaon
Absaroka

Hogsback
rd
Willa

ord
Crawf

Wyoming

Uinta Arch
0 50 km F

Figure 3. Maps of the Wyoming Salient showing deformation determinations for the Ankareh formation, including: (A) layer-parallel short-
ening (LPS) direction; (B) bedding-parallel strain ellipses estimated from reduction spots; (C) anisotropy of magnetic susceptibility (AMS)
Kmax directions; and (D) structurally restored site-mean paleomagnetic declinations for prethrusting Triassic (Tr) component (purple) and
Cretaceous (K) demagnetization component (green) with declinations plotted with respect to appropriate reference direction of 340° for
the Triassic and 350° for the Cretaceous components, respectively. Figure is modified after Yonkee and Weil (2010a) and Weil et al. (2010a).

Progressive Oroclines acted as buttresses that impeded eastward trans- Staal et al., 1998; Matte, 2001; Hatcher, 1989,
lation of Sevier thrust sheets, or influenced the 2002; Winchester et al., 2002; Franke et al.,
These observations suggest that progressive character of the sedimentary cover sequence and, 2005). Laurussia included the previously amal-
oroclines: (1) are limited in scale; (2) are a prod- in doing so, impacted the initial (producing a pri- gamated continents of Baltica and Laurentia,
uct of a strike-perpendicular stress; (3) develop mary curvature of 25% of the present-day geom- and the ribbon continent Avalonia that accreted
during and are coeval with thrust fault emplace- etry) and subsequent development of the thrust to, and formed the southern margin of, the
ment; (4) are thin-skinned and end down against belt (the additional 75% of observed curvature). northern continent (Rast et al., 1976; Trench and
the basal décollement underlying the thrust belt; Torsvik, 1992). The timing of orogeny is con-
and (5) are a response to the same stress respon- SECONDARY OROCLINES: strained by deformation and metamorphism of
sible for development of the thrust belt (Fig. 2). THE COUPLED IBERO-VARISCAN Devonian and Carboniferous strata, and by Car-
Hence, bends of thrust belts are not the result of OROCLINES boniferous foreland basin development (Paris
buckling of a preexisting structure. Instead the and Robardet, 1990; Martínez-Catalán et al.,
bends are “progressive” and develop during and Regional Geology 2007). Synkinematic remagnetization indicates
in response to the mechanisms responsible for that deformation was ongoing into the late Car-
thrust sheet emplacement. Therefore, the devel- The late Paleozoic Variscan orogen of West- boniferous and was complete prior to 310 Ma
opment of curvature is linked to the character of ern Europe is correlative with the Appalachian (Parés et al., 1994; Van der Voo et al., 1997; Weil
the crustal package being shortened, the nature orogen of eastern North America, developed in et al., 2000, 2001, 2010a, 2012a; Weil, 2006).
of the décollement, and the tectonic stresses being response to closure of the Rheic Ocean, and re- In Iberia, the Variscan orogen is divisible into
applied from the rear of the wedge. In the case cords the resulting collision of Gondwana with four zones (Fig. 1). These are, from foreland
of the Wyoming Salient, bending is linked to the Laurussia, forming the Pangea supercontinent to hinterland, the Cantabrian, West Asturian-
existence of long-lived basement highs that either (e.g., Martínez-Catalán et al., 1997, 2007; van Leonese, the Central Iberian, and the Galician

Geological Society of America Bulletin, May/June 2013 647


Johnston et al.

Tras-os-Montes zones (Martínez-Catalán et al., convex-to-the-east Central Iberian to the south ture results from buckling of a preexisting oro-
2007, and references therein). A little- to non- (Shaw et al., 2012) (Fig. 1). Both oroclines are gen that was originally linear or nearly so (Shaw
metamorphosed Lower Paleozoic passive-mar- “isoclinal” and exhibit 180° of curvature. The et al., 2012). Variscan thrust faults are continu-
gin sequence characterized by shallow-marine foreland fold-and-thrust belt occupies the core ous around the coupled oroclines, and vergence
to shoreface facies sedimentary sequences com- of the northern Cantabrian orocline; thrusts direction is a function of strike (Fig. 1). Simi-
prises the foreland Cantabrian zone (e.g., located in the orocline hinge verge toward the larly, orthogonal joint sets that developed during
Julivert, 1971; Julivert and Marcos, 1973; Pérez- east (in present-day coordinates), parallel to the thrusting, and that are oriented parallel and per-
Estaún et al., 1988). Foreland strata are imbri- orocline’s east-west–striking axial surface. The pendicular to the thrusts, have been shown to be
cated along a thin-skinned, cratonward-verging metamorphic West Asturian-Leonese and Cen- continuous around the Cantabrian orocline and
foreland fold-and-thrust belt that roots to the tral Iberian hinterland zones form the outer arc exhibit the full 180° of curvature (Fig. 4). Paleo-
west beneath the West Asturian-Leonese zone of the buckle. The opposite is true of the south- magnetic declination also varies as a function of
(Pérez-Estaún et al., 1991). The West Asturian- erly Central Iberian orocline, where the meta- strike around the Cantabrian orocline (Fig. 5)
Leonese and more internal Central Iberian zones morphic Central Iberian zone occupies the core (Weil et al., 2000, 2001). Limited paleomag-
consist of more distal facies sedimentary strata of the orocline, and structurally overlying netic data are similarly consistent with declina-
(Gutiérrez-Marco et al., 2002). Metamorphic oceanic allochthons of the Galician Tras-os- tion varying as a function of strike around the
grade and extent of granitoid plutonism increase Montes zone are preserved along the orocline’s Central Iberian orocline (Perroud et al., 1985;
from the Cantabrian to the Central Iberian zone, axial surface. The unmetamorphosed foreland, Ruffett, 1990; Perroud et al., 1991; Parés and
and the Central Iberian zone is characterized by though largely buried beneath younger strata, is Van der Voo, 1992). Palinspastic restoration of
ductilely deformed amphibolites- to granulite- distributed around the outer arc of the orocline, the Variscan orogen of Iberia to an originally
grade metamorphic rocks (Díez Balda et al., and thrust faults verge away from the orocline’s linear, north-south–striking geometry, yields a
1990; Rubio et al., 2013). Oceanic terranes of axial surface. The amount of penetrative strain 2300-km-long orogen that features, from east to
the Galician Tras-os-Montes zone lie structurally within the buckled orogen is, therefore, indepen- west, a west-facing Gondwana passive-margin
above and were thrust over the Central Iberian dent of the oroclines, with cleaved and foliated sequence that passes from proximal to increas-
zone (Martínez-Catalán et al., 1997). Termi- rocks of the highly strained orogenic hinter- ingly more distal facies to the west; an east-
nation of Variscan orogenesis was followed land occupying the core of the Central Iberian verging foreland fold-and-thrust belt; a more
by significant strike-slip faulting (Martínez- orocline and the external arc of the Cantabrian westerly metamorphic hinterland; and, at the
Catalán, 1990). Available age constraints limit orocline. Curvature is not readily related to any highest structural levels to the west, obducted
Variscan dextral and sinistral strike-slip faults to specific feature of the orogen being buckled, terranes of oceanic affinity (Fig. 6).
an age of 315–305 Ma (Dallmeyer et al., 1997; and preexisting stratigraphic and structural fea- Structural, stratigraphic, and paleomagnetic
Valle Aguado et al., 2005). Dextral and sinistral tures cannot be readily shown to have controlled data constrain orocline formation to a 10–20 m.y.
strike-slip faults are common in northern Iberia, the development or geometry of the oroclines. interval straddling the Carboniferous-Permian
whereas sinistral strike-slip faults are predomi- Synoroclinal structures are limited. There period boundary (Weil et al., 2010a; Pastor-
nant in southern Iberia, including the Badajoz- is no evidence of significant strike-parallel ex- Galán et al., 2011; Gutiérrez-Alonso et al., 2004,
Cordoba shear zones (Fig. 1). tension during orocline formation. Structures, 2011a, 2011b). The oroclines postdate Variscan
Two additional zones, the Ossa Morena and including individual thrust faults and strati- deformation and hence developed after 310 Ma.
more southerly South Portuguese zones, are graphic sequences have been mapped continu- Early Permian sandstones and conglomerates
recognized in the south of Iberia. The Ossa ously around the oroclines (Pérez-Estaún et al., unconformably overlie deformed strata, exhibit
Morena zone is geologically similar to, and has 1991). There is evidence of synoroclinal conical no evidence of involvement in the oroclines, and
been mapped as sharing a transitional contact folding (Pastor-Galán et al., 2012a) and joint are characterized by consistent paleomagnetic
with, the adjacent Central Iberian zone (Pereira set and cleavage formation (e.g., Pastor-Galán declination directions showing that orocline for-
and Silva, 2001; Solá et al., 2008), whereas the et al., 2011) in the core of the Cantabrian oro- mation preceded deposition (Weil et al., 2010a).
South Portuguese zone is interpreted as a por- cline. Stephanian conglomerates inferred to be Strata remagnetized during orocline formation
tion of Avalonia and is thought to represent part synoroclinal are characterized by joints that yield an inclination consistent with remagnetiza-
of the Laurussian continent (Lefort, 1988; Mar- fan about the orocline, whereas postoroclinal, tion, and hence orocline formation, between 310
tínez-Catalán et al., 1997; de la Rosa et al., 2002; Early Permian sandstones are characterized by and 290 Ma (Weil and Van der Voo, 2002). Oro-
Braid et al., 2012). The Ossa Morena and South orthogonal joints sets that strike consistently cline formation therefore postdates, but occurred
Portuguese zones lie south of the major sinis- north-south and east-west (Fig. 4) (Pastor-Galán shortly after (within 10–20 m.y.), the cessation
tral Cordoba-Badajoz shear zone; their original et al., 2011). It is, therefore, likely that the fan- of Variscan orogenesis.
paleogeographic location with respect to the rest ning Stephanian joints formed in response to the Magmatic, metamorphic, mineral deposit,
of the Variscan rocks of Iberia is a matter of con- stress field responsible for orocline formation. It and isotopic evidence points to orocline for-
jecture (e.g., Quesada and Dallmeyer, 1994), and is debatable whether the north-south and east- mation having involved the entire lithosphere
we exclude them from the subsequent discussion west orthogonal joint sets that characterize the (Fernández-Suárez et al., 2000; Gutiérrez-
of the Ibero-Variscan oroclines. Early Permian strata developed in response to Alonso et al., 2004, 2011a, 2011b; Ducea, 2011),
the same stress field. consistent with the results of analogue modeling
Coupled Ibero-Variscan Paleomagnetic and structural studies show of the orocline buckling process (Pastor-Galán
Secondary Oroclines that bending is secondary (Parés et al., 1994; et al., 2012b). Significant post-Variscan, syn-
Van der Voo et al., 1997; Kollmeier et al., 2000; oroclinal (310–285 Ma) magmatism spans the
The Variscan orogen of Iberia is buckled Weil et al., 2000, 2001; Aerden, 2004; Weil, entire orogen, includes gabbroic and granitic
into two, coupled oroclines, the convex-to-the- 2006; Weil et al., 2010a, 2012a; Pastor-Galán intrusions, and, in the north, becomes younger
west Cantabrian in the north, and the linked, et al., 2011), meaning that the observed curva- to the east toward the core of the Cantabrian

648 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

A 80
Permian Joint data Permian Joint data Post-Early Permian joint paern
60 m = –0.03 (α95= ±.08) m = 0.09 (α95= ±.08)
σr = 3.8 σr = 4.4

Joint set Orientaon – Joint ref


40 N = 9 N=9

20

–20

–40 Permian Joint data


m = 0.00 (α95= ±.08)
–60
σr = 4.6
Figure 4. Strike tests utilizing N=9
–80 S 1 (N-S)
Set
representative orientation data –100 –80 –60 –40 –20 0 20 40 60 80 0
100 SSet 2 (E-W)
for different joints sets around Structure Trend – STref
SSet 3 (~130°)
the northern Cantabrian oro-
cline of the coupled Ibero-Vari- B 80
Stephanian Joint data
Stephanian joint paern
scan oroclines: (A) data for 60 m = 0.72 (α95= ±.18)
σr = 12.7
three Permian and younger
Joint set Orientaon – Joint ref

40 N = 9
joints showing no correlation
between structural trend and 20
joint orientation; (B) data for 0
two Stephanian joint sets that
show some correlation between –20
structural trend and joint ori- –40 Stephanian Joint data
entation for these synbuckling m = 0.57 (α95= ±.12)
–60 σr = 7.0
joints; and (C) data for two N=8 P
Perpendicular set
pre-Stephanian joint sets that –80 P
Parallel set
–100 –80 –60 –40 –20 0 20 40 60 80 100
yield a one-to-one correlation Structure Trend – STref
between structural trend and
joint orientation, constraining
these joints to having formed C 100
Pre-Stephanian Joint data Pre-Stephanian joint paern
prior to oroclinal buckling. 80 m = 1.03 (α95= ±.06)
σr = 8.5
Figure is modified after Pastor- 60 N = 43
Joint set Orientaon – Joint ref

Galán et al. (2011) and Weil et al.


40
(2012b).
20

–20

–40
Pre-Stephanian Joint data
P
Perpendicular set
–60 m = 1.16 (α95= ±.10)
Parallel set
P
σr = 13.4
–80 N = 28
–100
–100 –80 –60 –40 –20 0 20 40 60 80 100
Structure Trend – STref

orocline (Fig. 7) (Fernández-Suárez et al., 2000; of asthenospheric upwelling and melting in re- eval with orocline formation, consistent with the
Gutiérrez-Alonso et al., 2011a). Isotopic studies sponse to lithospheric thinning around the outer removal of the lithospheric mantle coeval with
demonstrate that pre-Variscan mantle-derived arc of the Cantabrian orocline and thickening orocline formation (Colmenero and Prado, 1993;
magmas were derived from melting of an ancient of the mantle and its subsequent delamination Gómez-Fernández et al., 2000; Martín-Izard
(Nd model ages of 1.0 Ga or greater) lithospheric beneath its core. Elevated heat flow consistent et al., 2000; Weil and Van der Voo, 2002; Arenas
mantle, whereas post-Variscan melts yield Nd with lithospheric mantle removal is indicated and Martínez-Catalán, 2003; Gasparrini et al.,
model ages of 0.3 Ga (Gutiérrez-Alonso et al., by widespread metamorphism of syn-Variscan 2003, 2006a, 2006b; Schneider et al., 2008).
2011b). Orocline formation was, therefore, co- sedimentary strata and coal-bearing sequences. Reconciling paleomagnetic and geological
eval with removal of the ancient lithospheric Significant epithermal gold mineralization, mas- data within orogens characterized by secondary
mantle underpinning the Gondwana margin and sive dolomitization reactions, abnormally high oroclines has proven difficult. The Gondwanan
its replacement by asthenosphere. The eastward coal ranks, and widespread remagnetization stratigraphy involved in the Ibero-Variscan oro-
younging of intrusions can be explained in terms point to a crustal-scale hydrothermal event co- clines can be correlated along the length of the

Geological Society of America Bulletin, May/June 2013 649


Johnston et al.

A C
100
CAA paleomagnec data
80 m = 0.98 ( α95= ±.06)
σr = 18

Paleomagnec Declinaon – Dec ref


60 N = 116

??? 40

20
1 2 3 4 5 6 7 8 9 10

B 0

–20

–40

–60

–80

–100
–100 –80 –60 –40 –20 0 20 40 60 80 100
Structure Trend – STref

Figure 5. (A–B) Schematic geological maps of the Cantabrian portion of the Variscan orogen: (A) palinspastically restored
to its pre-orocline geometry; and (B) in its present-day geometry after oroclinal buckling. Fill patterns represent different
tectonic units and include: (1) West Asturian-Leonese; (2) Pisuerga-Carrión; (3) Narcea antiform; (4) Somiedo-Correcillas/
Valsurbio; (5) Esla-Lois-Cigüera; (6) Sobia-Bodón/Central Coal Basin; (7) Ponga; (8) Picos de Europa; (9) unconform-
able Upper Carboniferous rocks; and (10) Mesozoic and Tertiary cover. Red arrows in A indicate orogen-normal principal
compressive stress responsible for Variscan orogeny, while blue arrows in B indicate orogen-parallel stress responsible for
Cantabrian orocline formation. (C) Paleomagnetic data from around the Cantabrian orocline show a one-to-one correlation
in a strike test, indicating that the remanence was acquired prior to oroclinal buckling of the orogen. Figure is modified after
Weil et al. (2001, 2012b). CAA—Cantabria-Asturias arc.

Variscan orogen of Western Europe, and collec- and lithospheric mantle; and (4) are not attrib- By definition, secondary oroclines are extra-
tively has been referred to as Armorica. While utable to the same stress state responsible for orogenic, because they postdate the orogenic
faunal and stratigraphic studies appear to pre- initial orogenesis. Hence, orogen-scale oroclines event responsible for the construction of the
clude separation of Armorica from Gondwana are secondary structures and reflect buckling orogen that was subsequently buckled. How-
(Robardet, 2002, 2003), regional paleomagnetic of a preexisting orogen about vertical axes of ever, timing constraints suggest that the time
studies have been used to suggest that Armorica rotation as originally defined by Carey (1955). span separating initial orogenesis and orocline
lay 40° north of Gondwana in the Devonian and In the case of the Ibero-Variscan oroclines, the formation is small, and the rate at which oro-
imply the Ordovician to Devonian growth of a preexisting orogen included a foreland fold- clines form is tectonically rapid. The Ibero-Vari-
major ocean separating Armorica and Gond- and-thrust belt, a metamorphic hinterland, and scan oroclines formed after 310 Ma, the age of
wana (Van der Voo, 1982; Tait et al., 1997) obducted oceanic terranes, and the pre-oroclinal the youngest Variscan structures, but before
(Fig. 8). Resolution of this apparent inclination orogen was a linear feature >2300 km long and ca. 290 Ma, the age of the youngest postoro-
anomaly and the return of Armorica to paleo- 300 km wide. The coupled Alaskan oroclines clinal strata. Orocline formation was, therefore,
latitudes coincident with Gondwana are con- of the Cordilleran orogen of western North completed within 20 m.y. of the termination of
strained to have occurred coeval with Variscan America formed by buckling of a preexisting orogenesis. Suprasubduction-zone ophiolites
orogeny and the subsequent formation of the orogen that was geometrically similar to the are similarly emplaced within 20 m.y. of litho-
Ibero-Armorican oroclines. Ibero-Variscan orogen (Johnston and Gutierrez- sphere formation. Is there a link?
Alonso, 2010), and that was over 3000 km long Secondary oroclines commonly form shortly
Secondary Oroclines and greater than 300 km wide (Johnston, 2001, after the completion of initial orogenesis. The
2008). The Carpathian–Balkan sector of the Al- Alaskan oroclines of the North American Cor-
The observations reviewed here suggest that pine orogen included a foreland fold-and-thrust dillera formed between 85 and 55 Ma, shortly
orogen-scale oroclines: (1) are a product of an belt and obducted ophiolite, and it formed a lin- after mid-Cretaceous (125–90 Ma) fold-and-
orogen-parallel stress; (2) are extra-orogenic, de- ear belt 2300 km long and 200 km wide prior thrust belt formation (Johnston, 2001, 2008).
veloping after and postdating initial orogenesis; to buckling into the coupled Carpathian–Balkan The Miocene Carpathian–Balkan coupled oro-
(3) are plate-scale structures, involving the crust oroclines (Shaw and Johnston, 2012). clines formed in the immediate aftermath of

650 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

A 125 B Can
t abr
ian
WEVB N Oro
clin
e
100 paleo-current data
m = 1.20 (α95= ±.06)
75 σr = 25
Paleo-Current Direcon – PCDref

N = 43

nd
r la
50

nte
ges

l hi
sem ous
bla

a
c as on

ern
ani chth
25

Int
oce Allo
0 BC Cen

d
tra

n
SZ

ela
lI ber

For
ian
–25 Oro
clin
N
e

ges
sem ous
bla
c as o n
–50

ani chth
0 150 300 km

oce Allo
–75
Cross bed Variscan strike-slip
16 foresets shear zones

d
d

r l an
–100

rlan
12 Other Alpine deformaon front

d
nte

n
nte

ela
8 unidireconal

l hi
l hi

For
4 data Axial traces of oroclines

a
–125

ern
r na
Data of

Ex t
Inte
–100 –75 –50 –25 0 25 50 75 100 indeterminable Trace of oroclinal beam
unidireconality
Structure Trend – STref
Figure 6. (A) A strike test utilizing paleocurrent vectors for Lower Ordovician passive-margin strata shows a one-to-one correlation
between change in flow direction and structural trend around the coupled Cantabrian and Central Iberian oroclines. WEVB—West
European Variscan Belt. (B) Maps showing the paleocurrent data plotted at left on the coupled Ibero-Variscan oroclines in their current
geometry, and, at right, after palinspastic restoration of the orogen to a linear geometry, yielding a uniform offshore current direction
to the west (current coordinates). Figure is modified after Shaw et al. (2012) and Weil et al. (2012b). PTSZ—Porto-Tomar shear zone;
BCSZ—Badajoz-Cordoba shear zone.

Oligocene to Miocene fold-and-thrust belt for- examples include the coupled North Alaskan In the North American Cordillera, palinspastic
mation (Burtman, 1986; Shaw and Johnston, and Kulukbuk Hills oroclines (Johnston, 2001), restoration of the coupled Alaskan oroclines
2012). The Upper Miocene to Holocene which affect crust characterized by the west- restores Cordilleran crust 3000 km to the south
Calabrian orocline began forming immediately ern North American Cordilleran inclination (Johnston, 2001, 2008) and fully explains the
after significant Miocene crustal shortening and anomaly (Beck and Noson, 1972; Irving et al., Cordilleran inclination anomaly (Johnston,
fold-and-thrust belt formation along the Apen- 1996); and the coupled Carpathian and Balkan 2008). Similarly, palinspastic restoration of the
nine chain (Cifelli et al., 2008; Johnston and oroclines (Burtman, 1986; Shaw and Johnston, coupled Carpathian-Balkan secondary oroclines
Mazzoli, 2009). 2012), which are spatially and temporally corre- resolves the Eastern Mediterranean inclination
Formation of the Ibero-Armorican oro- lated with the Eastern Mediterranean inclination anomaly (Shaw and Johnston, 2012). Conflict-
clines within a 20 m.y. window requires tec- anomaly. Geologically based interpretations, ing interpretations of paleomagnetic and geo-
tonically rapid rates of translation. Buckling including stratigraphic correlations and faunal logical data may therefore be reconciled by
of the 2300-km-long segment of the Variscan provinciality, are commonly used to argue recognition of the significant translations re-
orogen that makes up the Ibero-Variscan oro- that crust affected by inclination anomalies is quired by, and the palinspastic implications of,
clines requires >1100 km of relative translation autochthonous, and that the paleomagnetic in- secondary oroclines.
of the southwestern end of the southern limb of clinations do not provide a robust measure of The removal of the mantle lithosphere under-
the Central Iberian orocline toward the north- paleolatitude. However, secondary oroclines re- pinning the Variscan orogen in Iberia during
eastern end of the north limb of the Cantabrian quire significant mobility of the buckled orogen, orocline formation implies that the formation of
orocline. A 10 m.y. window for orocline for- and if the belt has a long-axis trend that deviates orogen-scale oroclines is a plate-scale process.
mation implies rates of relative translation of from east-west in its paleogeographic reference The space problems inherent in buckling a
11 cm/yr, near the top end of the plate-tectonic frame, then significant latitudinal translation lithospheric-scale beam about a vertical axis
speed limit. A 20 m.y. window allows for a more will occur. As shown already, the south end of of rotation are significant (Pastor-Galán et al.,
manageable rate of 5.5 cm/yr, but this requires the Ibero-Variscan orogen moved >1100 km 2012b). Should the lithospheric mantle under-
that orocline formation began immediately upon relative to the north end during secondary oro- pinning the Variscan orogen have remained
the cessation of Variscan orogenesis. cline formation. Both ends of the Ibero-Variscan in place during formation of the two isoclinal
Orogens characterized by secondary oro- orogen cannot, therefore, have been fixed and Ibero-Variscan oroclines, an enormous volume
clines are commonly also characterized by incli- autochthonous with respect to cratonic Gond- of mantle would have to have been relocated
nation anomalies; the Ibero-Variscan oroclines wana, and the palinspastic restoration of the from the regions underlying the orocline cores.
affect crust characterized by the Armorican in- Ibero-Variscan oroclines can explain some, but Removal of the mantle lithosphere underpin-
clination anomaly (Tait et al., 1997). Additional not all, of the Armorican inclination anomaly. ning the orogen during orocline formation

Geological Society of America Bulletin, May/June 2013 651


Johnston et al.

A >310 Ma D 300–292 Ma
2007; Lorinczi and Houseman, 2009; Fillerup
et al., 2010). Formation of the Carpathians was
outer arc inner arc
Connental Crust related to the Mesozoic and Cenozoic closure
Lithospheric Mantle
of the Tethys Ocean during collision of the Eur-
asian and African plates (Knapp et al., 2005).
outer arc

inner arc

Here, the active Vrancea seismic zone has been


Asthenospheric
Mantle imaged as a steeply dipping volume of inter-
mediate-depth seismicity in the upper mantle
directly beneath the major bend in the Eastern
Carpathians (Fillerup et al., 2010). Today, the
Carpathians are only supported by ~33 km of
crust, whereas the proximal basins have greater
B Kasimovian E 292–285 Ma than 40-km-thick crust (Gvirtzman, 2002;
Knapp et al., 2005), indicating that not only was
lithospheric mantle involved with delamination,
but also the lowermost continental crust. If the
removal of the lithospheric mantle underpin-
ning an orogen is a requirement of secondary
orocline formation, then it follows that a cold,
thick, structurally coherent and therefore strong
lithospheric mantle may preclude secondary
orocline formation or at least limit the amount
of oroclinal buckling. Orogenic thickening of
C 305–300 Ma F the lithosphere can weaken the lithospheric
mantle (Pysklywec et al., 2000); through-going
faults provide planes of weakness along which
brittle failure can occur (e.g., Cook et al., 1998),
and thickened mantle can become gravitation-
ally unstable (e.g., Bird, 1979; Houseman et al.,
1981; Houseman and Molnar, 1997). Over time,
tectonized mantle anneals and attains thermal
and gravitational equilibrium. There may, there-
fore, only be a short interval between the end of
orogenesis and the subsequent stabilization of
the lithospheric mantle available for secondary
Gabbroic rocks orocline formation, which would explain oro-
Outer Arc (mantle derived) cline development soon after the termination of
“Hot” granites orogenesis. Magmatic arcs are underpinned by
Inner Arc (crust derived) a hot and hence weak mantle lithosphere, and
“Cold” granites their arcuate geometry may, therefore, be at-
(crust derived) tributable to the same processes responsible for
secondary oroclines.
Volcanic rocks Secondary oroclines do not develop in re-
(mantle derived) sponse to the same stress field responsible for
orogenesis. Instead, buckling is attributable to a
principal compressive stress oriented strike- or
Figure 7. (A–E) Schematic maps (at left) and lithosphere-scale cross sections (right) showing
orogen-parallel and positioned at a high angle
the sequential development of the Cantabrian orocline, and explaining the character and dis-
to the stress responsible for development of
tribution of synoroclinal magmatism as the result of involvement of the lithospheric mantle
the orogen. The lack of significant extension
in oroclinal buckling. (F) Diagram showing how tangential longitudinal strain associated with
around the outer arcs of either of the coupled
oroclinal buckling explains the development of lithospheric mantle thickness variations dur-
Cantabrian or Central Iberian oroclines im-
ing orocline formation. Figure is modified after Gutiérrez-Alonso et al. (2004, 2011a, 2011b).
plies that buckling is a response to a compres-
sive stress applied parallel to the long axis of
the orogen, and that vertical free surfaces are
negates this space problem, permits orocline de- Earth’s lithosphere (e.g., Bird, 1979; House- available to accommodate large amounts of
velopment to be balanced at depth by the flow of man et al., 1981; Houseman and Molnar, 1997; orogen-parallel displacement. Furthermore,
asthenosphere, and is expected to be a common Ducea, 2011). A modern example of mantle de- a significant component of lithospheric-scale
consequence of orogen-scale orocline forma- lamination likely related to secondary oroclinal flexural slip is required in order to balance the
tion. Besides subduction, mantle delamination processes is found in the Romanian Carpathians difference in line length measured around the
is the most important mechanism for recycling (Knapp et al., 2005; Houseman and Gemmer, outer versus inner boundaries of the buckling

652 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

Siberia
Obstacle-induced progressive oroclines de-
Siberia
A B velop in response to interaction between an
advancing thrust or thrust belt and a footwall
LAURUSSIA LAURUSSIA “obstacle” such as a basement massif or a pre-
Balca Balca
existing structure such as a strike-slip fault (Weil
and Sussman, 2004). These obstacles interfere
Rheic Rheic with the thrust sheet, restraining its movement
within a geographically limited region, and give
Laurena Laurena
Arm Arm rise to divergent transport around the obstacle.
The Wyoming Salient has been interpreted as
Paleo-Tethys an obstacle-induced progressive orocline pro-
Africa
duced by interaction between the leading edge
So South America of the eastward-propagating Sevier thrust belt
uth Africa
Am GONDWANA and the Gros Ventre uplift and Uinta Arch base-
eri GONDWANA
ca ment massifs (Paulsen and Marshak, 1997,
1998, 1999). However, these basement mas-
Figure 8. Paleogeographic models at 395 Ma based on (A) paleomagnetic data (Van der sifs, although faulted and uplifted during Late
Voo, 1982; Tait, 1999) that imply 40 degrees latitude of separation between “Armorican” Cretaceous and Tertiary Laramide orogenesis,
components of the Variscan orogen and Gondwana; and (B) stratigraphic correlations and originated as ancestral basement highs and are
faunal data (Robardet and Gutiérrez-Marco, 1990; Paris and Robardet, 1990) that preclude characterized by Paleozoic cover sequences that
any significant separation of Armorican rocks from Gondwana. Figure is modified after thin and change character toward the basement
Martínez Catalán et al. (2007). massifs (Dickinson, 2004). Hence, it is likely
that the Wyoming Salient is a product of a com-
bination of critical wedge and obstacle-induced
orogen. The distribution and timing of strike- ture; and (4) produce “progressive” oroclines curvature. Where “obstacles” consist of base-
slip faulting in Iberia are consistent with these in which map-view curvature develops dur- ment massifs or preexisting coastline features
faults having developed in response to, and ing thrust belt formation and in response to (promontories) that exerted a control on the
having accommodated, flexural slip during the strike-perpendicular stress (attributable to character of the sedimentary cover sequence, it
orocline formation; sinistral faults characterize the indenting backstop) responsible for orogen is likely that the resulting progressive oroclines
the south limb of the convex-to-the-east Central development. will, like the Wyoming Salient, reflect a com-
Iberian orocline (Quesada, 1991), whereas a Two main types of progressive oroclines bination of critical wedge and obstacle-induced
mix of dextral and sinistral faults characterizes are distinguished: critical wedge and obstacle curvature.
central and northern Iberia (Gutiérrez-Alonso induced (Marshak, 2004; Weil and Sussman,
et al., 2004). Finally, the development of 2004). Critical wedge progressive oroclines de- Secondary Oroclines
coupled oroclines requires an orogen-parallel velop curvature in response to the maintenance
stress; during formation of the Ibero-Variscan of a critical taper within the growing thrust belt. Any model of secondary orocline formation
coupled oroclines, a 2300-km-long orogen was Factors that affect the geometry of a critical must: (1) place orocline formation within a
significantly shortened, with >1100 km of rela- wedge include the thickness of the sedimentary plate-tectonic framework; (2) show how buck-
tive translation of the end points of the buckled succession, depth to the basal décollement, lith- ling of a lithospheric beam (including the man-
orogen toward one another. ology (and hence strength) of the detachment tle lithosphere) is accommodated in plan and
horizon, slope of the basal décollement; and section view; (3) explain how a buckling orogen
DISCUSSION lithology of the rocks carried within the thrust appears to behave as a beam that buckles about
wedge (Davis et al., 1983; Dahlen, 1990). Thick a vertical axis with a vertical axial surface; and
Progressive Oroclines sedimentary sequences consisting of relatively (4) account for the amounts and rapid rates of
weak rock (shale and mudstone) carried along translation during buckling. Because secondary
Progressive oroclines are adequately mod- a low-slope décollement located within a rheo- oroclines commonly form soon after the end of
eled using analogue sandbox experiments (Mar- logically weak unit (e.g., evaporites) will yield orogenesis, it is also necessary to be able to ex-
shak, 1988; Marshak et al., 1992) in which an a low taper critical wedge bound by a leading plain a rapid switch of the principal compressive
indentor or backstop is pushed or impressed thrust that will propagate far into the foreland. stress from orogen normal to orogen parallel
into a sand-filled box. Similar results have been In contrast, a thin sedimentary sequence con- prior to oroclinal buckling.
obtained from viscous fluid analogue modeling sisting of strong rock carried on a steeply slop- The rapid rates at which secondary oroclines
(Marshak et al., 1992) and computer modeling ing décollement located within a rheologically form imply that subduction is the main driving
(Macedo and Marshak, 1999). These models: strong unit will yield a high taper critical wedge force responsible for secondary orocline forma-
(1) are, by definition, thin skinned, restricting with limited foreland propagation of its lead- tion. Formation of the Ibero-Variscan oroclines
map-view bending to the sequence above the ing thrusts. Hence, along-strike changes in the involved sustained relative rates of translation of
basal décollement; (2) yield oroclines that are nature of the pre-orogenic sedimentary basin between 6 and 12 cm/yr over a 10–20 m.y. in-
limited in scale to the size of a thrust sheet or affecting the thickness or nature of the sedimen- terval. The coupled Alaskan oroclines involved
thrust belt; (3) generate oroclines in which the tary strata will affect the geometry of the critical relative rates of translation of 10 cm/yr sus-
degree of secondary rotation about a vertical wedge and give rise to curvature of the thrust tained over 20–30 m.y. The average rate of con-
axis is always less than the measured curva- belt (Marshak, 2004). vergence across subduction zones over the past

Geological Society of America Bulletin, May/June 2013 653


Johnston et al.

3 m.y. is ~7 cm/yr (DeMets et al., 1990; Gordon, cian Grampian orogeny, in which arc-continent metrically elegant, the self-subduction model is
1998; Engebretson et al., 1985; Engebretson collision involving the Laurentian margin was dependent upon the unique geometry of Pangea
and Richards, 1992); maximum convergence followed by a subduction polarity flip (Dewey, and its associated plate boundaries; cannot be
rates of ~11 cm/yr characterize the subduction 2005). Where collision zones abut along strike readily applied to explain the majority of sec-
zones of the southwest Pacific (Gordon, 1998); against an orthogonal subduction zone, the ces- ondary oroclines; and does not address the fate
and relative translation rates of 10 cm/yr for sus- sation of subduction within the collision zone of the lithosphere that formerly lay between the
tained periods of time are limited to plates at- can be followed by an orthogonal switch of the limbs of the Ibero-Variscan oroclines.
tached to subducting slabs (Gordon, 1998). The principal compressive stress. A modern example A set of geometric models for the develop-
implication is that where secondary orocline of a collisional orogen that ends against an or- ment of secondary oroclines is presented in
formation involves sustained relative transla- thogonal subduction zones is Taiwan (Suppe, Figures 10 and 11. Like balanced and palinspas-
tion rates of 7 cm/yr or greater, the oroclines are 1984). Finally, crustal blocks that extrude lat- tically restorable cross sections (Dahlstrom,
forming within and are a characteristic of plates erally out of a collision zone are characterized 1969), such geometric models provide a test of
attached to and continuous with a subducting by an orthogonal stress switch as the collisional the feasibility of an interpretation, and a means
slab. A marginal location with respect to a sub- stress field is replaced during lateral escape by a of quantifying its implications for line length
ducting oceanic slab is also a requirement given stress field related to translation out of the col- and displacement, but they do not necessarily
the space problems and free surface require- lision zone. imply that such an interpretation is correct. The
ments of buckling a lithospheric beam around These two assumptions, (1) that secondary first model (Fig. 10) assumes concentric buck-
a vertical axis. oroclines form associated with a subducting ling, with the buckle radius and amplitude equal
Escape tectonics may provide an additional slab, and (2) that collision is followed by the to the width of the orogen (r = a = w); and a
setting amenable to orocline formation. Sus- linked consumption of oceanic lithosphere in an buckle wavelength equal to twice the width (λ =
tained moderate rates of translation of 2–7 cm/yr orthogonal subduction zone, informed a “self- 2w). The second model (Fig. 11) assumes linear
characterize extrusion zones. Examples include subduction” model for the Ibero-Armorican limbs throughout orocline formation, and that
westward extrusion of Anatolia out of the oroclines by Gutiérrez-Alonso et al. (2008). In orocline amplitude and wavelength are three
Arabian–Eurasian collision zone (Tartar et al., their model, collision of Laurussia with Gond- and four times the width of the buckling orog-
2012); southeastern Asian extrusion eastward wana along a north-south–striking collisional eny, respectively (a = 3w; λ = 4w). Calculations
out of the Himalayan collision zone (Burchfiel, orogen was followed by subduction of Tethyan of line length around secondary oroclines and
2004); and east-directed escape out of the Aus- oceanic lithosphere into an east-west–striking relative translations all assume a pre-orocline
tralian–Indonesian collision zone (Pubellier and subduction zone lying east of and perpendicu- orogen width of 300 km (w = 300 km), similar
Ego, 2002). In addition, extruding blocks are lar to the Ibero-Variscan orogen. The Tethyan to the width of the Variscan orogen. The basic
bound by vertical lithospheric strike-slip shear oceanic lithosphere subducting north beneath elements of these models are: (1) an elongate,
surfaces that provide a means of accommodat- northeastern Pangea was continuous to the linear, collisional orogen that is bounded by
ing orocline formation. The coupled Miocene south with the Ibero-Variscan crust of Pangea, continent on one side and oceanic lithosphere
Carpathian–Balkan oroclines provide an exam- placing it within the lower subducting plate. on the other, and that ends along strike against
ple of escape-tectonic–related oroclines. Their The result was self-subduction, and orocline a subduction zone oriented perpendicular
formation: (1) involved moderate relative rates formation was explained as a response to the re- to the orogen; (2) subduction of the oceanic
of translation (4 cm/yr); (2) was kinematically sulting within-plate torque (Fig. 9). While geo- lithosphere that lies adjacent to the collisional
linked to westward escape of Anatolia out of
the Arabian–Eurasian collision zone (Shaw and
Johnston, 2012); (3) was coeval with abundant A Late Carboniferous (~305 Ma) B Carboniferous - Permian boundary (~299 Ma)
magmatism in and adjacent to the Carpathian–
Peripheral
Per
Balkan belt attesting to synoroclinal subduc- Seaa
Seas
tion as a means of lithospheric removal during Oslo-North Sea
Ris
buckling of the belt (Cavazza et al., 2004); ck-
Ba
l i ata
and (4) saw the delamination of the orogen’s Me Basi
n
lithospheric mantle as imaged by ongoing seis- rasi
a Arc
Lau
TENING

micity (Fillerup et al., 2010, and references Paleoteth


Paleotethys
Pa
Paleotet
Paaleotethy
lleot
le
eot
eootet
o tet
eth
eth
hyy Paleotethys
co
SION

therein). Slower relative rates of translation dur- G on d Ocea


Oc
O
Oce
Ocean
ccean
ean
eea
aan
n Ocean
wana C i mm
ing orocline formation may point to additional p assive
OR

erian
TEN

margin R ibb
orocline-forming processes. For example, for- SH Ne on
ote
EX

mation of the Bolivian orocline has involved thy


sR
i
relative rates of translation of the Andes south of
the bend relative to the northern Andes of only
Madagascar Ri
1–2 cm/yr (Allmendinger et al., 2005). Oroclinal
Buckling
It is common for the principal compressive
stress affecting an orogen to change orientation
after collision. Arc-continent collisions are com- Figure 9. Schematic maps of Pangea at (A) 305 Ma and (B) 299 Ma showing how self-
monly followed by a subduction flip, yielding a subduction of Pangea oceanic lithosphere underlying the Tethys beneath the active margin
change in the polarity, but not necessarily orien- bounding the southeast portion of the Laurussian (northern) portion of Pangea leads to
tation of the principal compressive stress (e.g., orocline formation within the core of Pangea and coeval extension around its peripheries.
Clift et al., 2003). Examples include the Ordovi- Figure is modified after Gutiérrez-Alonso et al. (2008, 2012).

654 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

Figure 10. Geometric tectonic


models of secondary orocline A B C
development. Dark brown—
autochthonous continent; red—
continental orogen (300 km
wide); blue—oceanic litho-
sphere. (A) Subduction of the
oceanic lithosphere, which is
fully coupled to the continental
orogen, into a subduction zone λ
oriented perpendicular to the
orogen drives the unsubduc-
a
table orogen into and pins it
against the subduction zone.
Continued subduction leads to
buckling of the continental oro-
gen, forming (B) a concentric,
convex-oceanward orocline;
a—amplitude = w; λ—wave-
length = 2w; Ls—shortening of
orogen in response to orocline
formation. Continued sub-
duction can lead to multiple
concentric, convex-oceanward
Ls = 1150 km
oroclines of this sort. (C) Alter-
natively, continued trenchward
motion of the initial concentric Ls = 350 km
orocline explains the formation
of coupled oroclines. See text w w = 300 km
for complete explanation.

orogen; and (3) coupling of the collisional oro- of time); and the delamination and subsequent clines are required in order to maintain contact
gen with the subducting oceanic lithosphere. foundering of the lithospheric mantle, all of between the buckling orogen and the autoch-
Coupling of the orogen with the subducting which allow the unsubductable linear crustal thon; and (2) the oroclines are characterized
oceanic lithosphere implies that the orogen orogen to buckle. by axial surfaces that are perpendicular to the
moves as part of the oceanic plate, requires the bounding strike-slip faults. The convex-ocean-
presence of a strike-slip fault separating the Concentric Models ward Zagros oroclines are bounded landward by
orogen from the autochthonous continent, and As depicted in Figure 10, orocline forma- major strike-slip faults oriented at a high angle
transports the orogen into the subduction zone. tion is characterized by the counterclockwise to the axial surfaces of the oroclines (Hollings-
Complete coupling between plates is required rotation of the trenchward orocline limb and worth et al., 2010), and may represent an ex-
to explain the observed translation rates. Upon clockwise rotation of the trailing limb. In this ample of such concentric secondary oroclines.
entry into the subduction zone, the continental case, the oceanward migration of the orocline Simple concentric buckling cannot, however,
crust of the linear orogen, being unsubductable, hinge is facilitated by overthrusting of the ad- explain coupled oroclines. Coupled oroclines,
collides with the upper plate, yielding to a jacent oceanic lithosphere. The orocline hinge such as the Ibero-Variscan oroclines, consist
switch of the principal compressive stress, from is required to migrate 300 km over the adjacent of two convex-in-opposite-direction oroclines
orogen normal to orogen parallel. Continued oceanic lithosphere, enough to initiate subduc- linked through a common limb, and hence do
subduction of the oceanic lithosphere leads to tion of the overthrust oceanic lithosphere, sug- not consist of two complete (trough to trough)
buckling of the coupled orogen, and formation gesting that oroclinal buckling may provide a buckles (Fig. 1). In addition, the axial sur-
of a secondary orocline. means of initiating localized subduction. The faces of the coupled oroclines are commonly
In these models, it is assumed that buck- length of a medial line around a single concen- oriented subparallel with adjacent, bounding
ling is accommodated by delamination of the tric orocline affecting a 300-km-wide orogen strike-slip faults. A modified concentric buckle
lithospheric mantle. Hence, the behavior of the is 950 km (l = 950 km). Since the straight-line orocline model, based on the assumption that
orogen as a beam that buckles about an axis distance measured from the fore- to backlimb of an initial concentric orocline will continue to
perpendicular to the plane of the plate is a the orocline is 600 km (lo = 600 km), a rela- move trenchward with the adjacent oceanic
product of the high strain rates; the significant tive translation, which can also be thought of as lithosphere, addresses these problems. Having
shortening required by the ongoing transla- the amount of shortening accommodated during an initial concentric orocline move trenchward
tion of the linear orogen toward the subduction orocline formation, of 350 km (ls = 350 km) is with the adjacent oceanic lithosphere requires
zone (shortening that could only be accommo- implied. In a model of simple concentric buck- that the initial orocline rotates 90° around the
dated by crustal thickening for a short period les: (1) complete convex-toward-the-ocean oro- fixed trenchward forelimb of the buckling

Geological Society of America Bulletin, May/June 2013 655


Johnston et al.

geometry. Any oceanic lithosphere generated


within the interlimb basin likely subducts be-
neath the trailing limb of the orocline, and there
may be little record of the ephemeral interlimb
basin. The Tyrrhenian Sea provides an exam-
ple of a potentially ephemeral basin; the basin
opened in response to buckling of the Apennine
chain of Italy, forming the Calabrian orocline,
and is now beginning to close by subduction be-
neath the north margin of Sicily as the Calabrian
λ = 4w orocline continues to tighten (Fig. 12; Johnston
w a = 3w and Mazzoli, 2009).
These models demonstrate that secondary
oroclines can be placed within a plate-tectonic
framework, and are geometrically viable, and
they show how secondary orocline formation
can be explained as a product of slab pull. In
addition, they account for the fate of lithosphere
that formerly lay adjacent to the buckled orogen,
Figure 11. Geometric block diagram showing fixed limb secondary orocline development. as well as the basic characteristics of secondary
Colors and labels are as in Figure 10. Here, the orocline amplitude (a) is three times the oroclines, including stress switching and mantle
width (w) of the buckling orogen, and the wavelength (λ) is four times. Linear (fixed) limbs lithosphere removal. While the models pre-
are employed. See text for discussion. sented here are driven by subduction, extrusion
or escape tectonics could equally be called upon
to drive orocline formation. However, coupled
orogen. Rotation of the original orocline into amplitude is twice the width of the buckling oroclines require >1000 km of translation of
a position oceanward of the forelimb is ac- orogen (a = 3w) (Fig. 11). the trailing and leading orocline limbs relative
commodated by overthrusting of the forelimb As in the concentric orocline models, fixed to one another, and in even simple concentric
onto the autochthon. The result is two coupled limb orocline formation, as depicted in Figure oroclines, the orocline hinge line has to migrate
oroclines, including an oceanward orocline 11, is characterized by the counterclockwise ro- >300 km out over adjacent lithosphere. The only
that is convex toward the adjacent trench, and tation of the subduction-side orocline limb and way to accommodate these substantial transla-
a landward orocline convex in the opposite di- clockwise rotation of the trailing limb relative tions is by subduction.
rection (Fig. 10). The line length around the to the leading, trenchward end of the orogen.
coupled oroclines is >2350 km (l ~2350 km); The oceanward migration of the orocline hinge Coupled Andean Oroclines
the straight-line distance from back to forelimb is facilitated by overthrusting of the adjacent
is 1200 km (lo = 1250); and the relative transla- oceanic lithosphere; with a 300-km-wide orogen, How broadly applicable are our models for
tion of the trailing limb relative to the forelimb the orocline hinge migrates 900 km oceanward secondary oroclines? To address that question,
is >1150 km (ls ~1150 km). The axial surfaces and requires substantial subduction beneath the we turn to perhaps the best known “buckled”
of the coupled oroclines are curviplanar, with growing buckle. Unlike the restricted amplitude orogen: the Andes. The Andean orogen is char-
the oceanward orocline axial surface curving (a = w) of the concentric orocline models, fixed acterized by two coupled oroclines: a northern
landward from the core of the orocline, and the limb orocline growth is facilitated by develop- convex-oceanward Peruvian orocline and a
landward orocline exhibiting the opposite sense ment of an interlimb basin during oceanward mi- more southerly convex-landward Arica orocline,
of curvature. The orocline axial surfaces strike gration of the orocline hinge. The interlimb basin classically called the “Bolivian orocline” (Fig.
subparallel to the bounding strike-slip fault. The develops landward of the oceanward migrating 13). The buckles of the Andean orogeny are sec-
modified concentric buckle model yields oro- orocline hinge and is initially bound by normal ondary oroclines: They buckled a preexisting,
clines that: (1) closely resemble, in map pattern, faults; both orocline limbs move away from the previously linear orogen. Structures and fabrics
documented coupled oroclines, including their continent, and the strike-slip fault separating attributable to Jurassic and Cretaceous terrane
curviplanar axial surfaces; (2) are characterized the translating orogen from the continent con- accretions and the pan-Andean Eocene Incaic
by line lengths >2000 km and that involve rela- verts to a zone of extension. Continued growth orogeny, which may be related to subduction of
tive translations of >1000 km; and (3) are bound of the orocline expands the interlimb basin and the Nazca plate underneath South America, are
by strike-slip faults oriented sub-parallel to their leads to oceanic lithosphere formation that is continuous around both oroclines (Isacks, 1988;
axial surfaces. produced along a complicated set of at least two Allmendinger et al., 1997) and limit orocline
spreading centers. While the trailing edge of the formation to the late Eocene or later. Paleomag-
Fixed Limb Models leading limb of the orocline remains a passive netic, geodetic, and geomorphic studies show
Geometric models for fixed (linear) limb margin once the interlimb basin is established, that vertical-axis rotations are occurring at pres-
oroclines, in which buckle amplitude is greater the leading edge of the trailing limb converts to ent (e.g., Norabuena et al., 1998; Lamb, 2000;
than the width of the buckling orogen, provide a thrust fault, and, as the orocline tightens, the Beck, 2004; Allmendinger et al., 2005, 2007;
a means of accommodating great line lengths trailing limb overthrusts the basin (Fig. 11). Maffione et al., 2009), indicating that oroclinal
around individual oroclines. Here, a fixed limb Complete closure of the interlimb basin coin- bending is ongoing. Geodetic studies constrain
model is presented in which the orocline buckle cides with the orocline attaining an isoclinal the rate of bending around the Arica orocline to

656 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

stress fields is likely responsible for the devia-


A tions in observed rotation magnitude away from
the orocline hinges. Reviews of paleomagnetic
data indicate that the observed rotation patterns
are not systematic along orogenic strike (Prezzi
and Alonso, 2002; Beck, 2004; Roperch et al.,
LI 2006; Arriagada et al., 2006; Maffione et al.,
2010), and thus have led to models that rely
on along-strike differential displacements that
200 km
incorporate localized modification and progres-
Figure 12. Sequential paleo- sive rotations produced by distributed shear,
geographic maps showing the wrenching, and fault-bound block rotations
development of the Calabrian (e.g., Kley and Monaldi, 1998; Kley, 1999;
secondary orocline and related Hindle et al., 2000, 2002; McQuarrie, 2002;
opening and incipient closure of B Oncken et al., 2006; Arriagada et al., 2008).
the ephemeral Tyrrhenian Sea A model of secondary orocline formation ac-
(TS) at 10 Ma (A), 6 Ma (B), companied by orogen-perpendicular differential
3 Ma (C), and at present (D). shortening requires significant strike-slip defor-
Green—present-day exposed TS mation (either transport parallel or orogen paral-
continent; brown—full extent lel), as well as orogen-parallel extension (e.g.,
of continental lithosphere; Gephart, 1994; Müller et al., 2002; Arriagada
gray—transitional to oceanic et al., 2008). Such is the case in the Andean sys-
lithosphere. The secondary oro- tem, where the presence of extensive orogen-
200 km
cline is characterized by fixed parallel and high-angle strike-slip fault systems
linear limbs; orocline amplitude are found throughout the Andean oroclines (e.g.,
is twice the width of the buck- Allmendinger et al., 1983; Strecker et al., 1989;
ling Apennine-Sicilian orogen. C Hérail et al., 1996; Riller et al., 2001; Müller
The Tyrrhenian Sea reached its et al., 2002; Riller and Oncken, 2003; Maffione
greatest extent between 3 and et al., 2009).
1 Ma and has begun to shrink As with other secondary oroclines, the
by clockwise rotation of the TS Andean secondary oroclines are plate-scale fea-
southern (Sicilian) limb of the tures that involve the lithospheric mantle (e.g.,
orocline over the southern mar- Gutiérrez-Alonso et al., 2004, 2011a, 2011b;
gin of the Tyrrhenian Sea. LI— Fillerup et al., 2010). Recent work investigating
Lagonegro-Imerese-Sicanian the rapid uplift of the Altiplano plateau has re-
basin; I—Ionian basin. Figure 200 km sulted in models that connect lithospheric thick-
is modified after Johnston and ening and ultimately delamination near the core
Mazzoli (2009). of the Bolivian orocline (below the Altiplano
D Plateau) to oroclinal buckling (e.g., Garzione
et al., 2006; Ghosh et al., 2006; Molnar and Gar-
zione, 2007). Like the Iberian and Carpathian
oroclines, delamination is likely related to over-
TS thickening of the crust and mantle lithosphere
within the orocline hinge regions (Pastor-Galán
et al., 2012b).
I A key characteristic of secondary oroclines
200 km is their expansive (orogen to continental)
scale; the Andean oroclines are expansive. The
buckles affect a 400-km-wide orogen (w =
400 km) yielding two open oroclines with inter-
1.5–2°/m.y. (Allmendinger et al., 2005, 2007), formation and require significant mobility of limb angles of 125°; a wavelength of ~3500 km
consistent with the paleomagnetic data from the Andean orogen relative to the autochthon (λ ~3500 km = 9w); and an amplitude of
Miocene and younger rocks. Assuming that (Lamb, 2001; MacFadden et al., 1995; Gilder ~400 km (a = 400 km = w). The long wave-
the current rate of buckling is representative et al., 2003; Beck, 2004). length relative to width relationship (λ ~ 9w)
of the rate of buckling through time, construc- Unlike the models for secondary oroclines of the Andean oroclines implies a component
tion of the oroclines could have been accom- developed here, the formation of the Andean of fixed limb buckling (as opposed to simple
plished in 35–50 m.y., consistent with oroclines secondary oroclines has been at least partly concentric oroclinal buckling). Cenozoic ex-
being post-Incaic structures. Paleomagnetic contemporaneous with continuous shortening tension east of the Peruvian orocline within the
data showing significant Cretaceous and ear- perpendicular to the original Andean orogenic Amazonian Basin (Costa et al., 2001) may be
lier rotations of Andean rocks predate orocline grain. Such a combination of regional tectonic in part a product of the fixed limb buckling. The

Geological Society of America Bulletin, May/June 2013 657


Johnston et al.

CaP (Fig. 14). As in our model for secondary orocline


N10 o formation, the source of the orogen-parallel
AP stress in the Andes can be explained as a result
of subduction and collision along a convergent
plate boundary oriented nearly perpendicular
CoP Equa to the buckling orogen—in this case, the east-
tor
Peruvian west–striking South American–southern Carib-
Orocline bean plate margin, where continental South
SAP America is being dragged into a collision with
the Caribbean plate by northward subduction of
NP
the South American plate in the northeast and
S10 o
Arica southward subduction of the Caribbean plate in
ge

(Bolivian) the northwest (Avé Lallemant, 1997; Audemard


Rid

Orocline
a

and Audemard, 2002; Miller et al., 2009) (Fig.


zc
Na

14). The coupled secondary oroclines that char-


acterize the Panamanian isthmus are interpreted
corn
Capri to be forming in response to the same north- to
rop ic of
T northwestward motion of the South American
plate relative to the Caribbean and North Ameri-
e
r Ridg can plates (Silver et al., 1990; Farris et al., 2011;
Easte
S30 o Montes et al., 2012). The state of stress for the
Andes, as indicated by earthquake focal mecha-
nisms, indicates primarily east-west compres-
sion (Assumpcao, 1992), reflecting the ongoing
W70o W50o subduction of oceanic lithosphere of the Nazca
W30o plate beneath western South America. However,
east of the Andes, continental South America
is characterized by north-south compression,
AnP
W90o probably reflecting collision with the Caribbean
plate and consistent with orocline formation in
Figure 13. Satellite image of South America showing the coupled Peruvian and Arica- response to a north-south compressive stress
Bolivian oroclines of the Andean orogen. AnP—Antarctic plate; AP—African plate; CaP— (Assumpcao, 1992).
Caribbean plate; Cop—Cocos plate; NP—Nazca plate; SAP—South American plate. Red The rate of northward migration, and as a
line with teeth (teeth point in to upper plate)—convergent plate boundaries; red line— consequence the rate of buckling, is slow (1–2
divergent plate boundaries. cm/yr compared to 6–10 cm/yr for the Ibero-
Variscan and Alaskan oroclines), probably due
to the entry of continental lithosphere of the
arc length of the buckled portion of the Andes The significant north-south shortening South American plate into the collision zone
is ~3800 km, indicating that orocline formation (~300 km) during orocline formation, the (Miller et al., 2009). An additional factor that
accommodated ~300 km of north-south short- coupled geometry of the two Andean oroclines, may have limited the rate of orocline formation
ening. Given the 35–50 m.y. time span avail- and the synoroclinal thickening and delamina- pertains to the geometry of the Andean oro-
able for orocline development, a relative rate of tion of the mantle lithosphere around the Arica- clines. Unlike our model for coupled oroclines
translation of the ends of the Andes toward one Bolivian orocline point to the Andean oroclines (Fig. 11) in which the trailing orocline relative
another is a modest 6–9 mm/yr. as being the result of an orogen-parallel stress to the subduction zone is convex oceanward, the

Figure 14. Model of the development of


N
the coupled secondary Andean oroclines. CP
(A) Formation of the Eocene Incaic orogeny; NP
red arrows indicate orogen-normal princi- CP
pal compressive stress responsible for oro-
genesis. (B) Secondary orocline formation. NP
Blue arrows indicate orogen-parallel prin- SAP
cipal compressive stress responsible for sec- A
ondary orocline development. Red—orogen;
coral—autochthonous continental crust; B SAP
blue—oceanic crust; gray—lithospheric
mantle; CP—Caribbean plate; NP—Nazca
plate; SAP—South American plate.

658 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

“trailing” Arica orocline (relative to the South tation of a rigid backstop into a sedimentary Armstrong, R.L., 1968, Sevier orogenic belt in Nevada and
Utah: Geological Society of America Bulletin, v. 79,
American–Caribbean plate boundary) is convex sequence, while instrumental in understanding p. 429–458, doi:10.1130/0016-7606(1968)79[429:
landward (Figs. 13 and 14). The result is that progressive oroclines, can tell us nothing about SOBINA]2.0.CO;2.
the Arica-Bolivian orocline has overthrust con- secondary oroclines. Similarly, a common and Arriagada, C., Roperch, P., Mpodozis, C., and Fernández,
R., 2006, Paleomagnetism and tectonics of the south-
tinental autochthonous South America during unjustifiable assumption in paleogeographic ern Atacama Desert (25–28°S), northern Chile: Tec-
buckling, a geometry that has likely contributed and tectonic modeling is that all oroclines are tonics, v. 25, p. TC4001, doi:10.1029/2005TC001923.
to the slow rate of buckling. progressive. Such models fail to adequately ad- Arriagada, C., Roperch, P., Mpodozis, C., and Cobbold, P.R.,
2008, Paleogene building of the Bolivian orocline: Tec-
To summarize, interpretation of the buckles dress the significance of secondary oroclines. tonic restoration of the central Andes in 2-D map view:
of the Andean orogen as secondary oroclines For example, we have yet to develop a fully bal- Tectonics, v. 27, TC6014, doi:10.1029/2008TC002269.
Assumpcao, M., 1992, The regional intraplate stress field in
is consistent with their geometry and with the anced palinspastic restoration of the secondary South America: Journal of Geophysical Research, v. 97,
plate-tectonic setting in which they developed. oroclines of the Variscan orogen, and hence our no. B8, p. 11,889–11,903, doi:10.1029/91JB01590.
The oroclines postdate and buckled a preexist- understanding of the tectonic and paleogeo- Audemard, F.E., and Audemard, F.A., 2002, Structure of the
Merida Andes, Venezuela: Relations with the South
ing orogen; involve the entire plate, including graphic evolution of Pangea remains incom- America–Caribbean geodynamic interaction: Tectono-
the lithospheric mantle underpinning the Andes; plete. Determinations of how to palinspastically physics, v. 345, p. 299–327, doi:10.1016/S0040-1951
are expansive of scale, affecting a 400-km- restore secondary oroclines and evaluations (01)00218-9.
Avé Lallemant, H.G., 1997, Transpression, displacement
wide orogen, with an amplitude of 400 km of their tectonic significance require that we partitioning, and exhumation in the eastern Caribbean/
and a wavelength of 3500 km; and developed understand the relationship between such huge South American plate boundary zone: Tectonics, v. 16,
in response to an orogen-parallel compressive structures and subduction, and this understand- p. 272–289, doi:10.1029/96TC03725.
Beck, M.E., 2004, The central Andean rotation pattern: An-
stress that has resulted in ~300 km of north- ing awaits successful analogue and computer other look: Geophysical Journal International, v. 157,
south shortening of the Andes. Three compli- modeling. p. 1348–1358, doi:10.1111/j.1365-246X.2004.02266.x.
Beck, M.E., and Noson, L., 1972, Anomalous paleolatitudes
cating factors are that (1) the oroclines have in Cretaceous granitic rocks: Nature, v. 235, p. 11–13.
ACKNOWLEDGMENTS
developed contemporaneously with ongoing Bird, P., 1979, Continental delamination and the Colorado
orogen-perpendicular compression (attributable Plateau: Journal of Geophysical Research, v. 84,
This paper is a contribution to the International p. 7561–7571, doi:10.1029/JB084iB13p07561.
to subduction of the Nazca plate); (2) the rate Geoscience Program Project 574: Buckling and Bent Boyer, S.E., and Elliott, D., 1982, Thrust systems: The Amer-
of orocline formation is slow relative to other Orogens and Continental Ribbons. Financial support ican Association of Petroleum Geologists Bulletin,
coupled oroclines; and (3) the geometry of the was provided by the Spanish Ministry of Science and v. 66, p. 1196–1230.
Innovation to Gutiérrez-Alonso for the Research Proj- Braid, J.A., Murphy, J.B., Quesada, C., Bickerton, L., and
trailing orocline is convex landward. Mortensen, J.K., 2012, Probing the composition of unex-
ect ODRE II (Oroclines and Delamination: Relations
posed basement, South Portuguese zone, southern Iberia:
and Effects) CGL2009–1367; and the National Sci-
Implications for the connections between the Appala-
CONCLUSIONS ences and Engineering Research Council (NSERC) chian and Variscan orogens: Canadian Journal of Earth
to Johnston through a Discovery Grant. Comments, Sciences, v. 49, no. 4, p. 591–613, doi:10.1139/e11-071.
There are two fundamentally different types corrections, and suggested revisions provided by two Burchfiel, B.C., 2004, New technology: New geological
anonymous reviewers and Editor Brendan Murphy challenges: GSA Today, v. 14, no. 2, p. 4–9, doi:10.1130
of oroclines: progressive and secondary. Progres- significantly improved the manuscript. We would like /1052-5173(2004)014<4:PANNGC>2.0.CO;2.
sive oroclines characterize thrust belts, develop to thank Jess Shaw and Dani Pastor-Galán for sharing Burchfiel, B.C., and Davis, G.A., 1972, Structural framework
in response to the same orogen-normal stress their insights regarding the Ibero-Variscan oroclines; and evolution of the southern part of the Cordilleran
orogen, western United States: American Journal of Sci-
responsible for thrust development, and can be and Rob Van der Voo for his support, interest, and
ence, v. 272, p. 97–118, doi:10.2475/ajs.272.2.97.
explained as a result of along-strike changes in many discussions. Burtman, V.S., 1986, Origin of structural arcs of the Car-
pathian-Balkan region: Tectonophysics, v. 127, p. 245–
the character of the deforming sequence (criti- REFERENCES CITED 260, doi:10.1016/0040-1951(86)90063-6.
cal taper progressive oroclines) or to interaction Carey, S.W., 1955, The orocline concept in geotectonics:
with structural features that characterize the Aerden, D.G.A.M., 2004, Correlating deformation in Vari- Proceedings of the Royal Society of Tasmania, v. 89,
scan NW-Iberia using porphyroblasts; implications for p. 255–288.
footwall to the thrust belt (obstacle-induced pro- the Ibero-Armorican arc: Journal of Structural Geology, Cavazza, W., Roure, F., and Ziegler, P.A., 2004, The Medi-
gressive oroclines). Because basement structures v. 26, p. 177–196, doi:10.1016/S0191-8141(03)00070-1. terranean area and the surrounding regions; active
and paleogeographic features such as promon- Allmendinger, R.W., Ramos, V.A., Jordan, T.E., Palma, M., processes, remnants of former Tethyan oceans and
and Isacks, B.L., 1983, Paleogeography and Andean related thrust belts, in Cavazza, W., Roure, F., Spak-
tories control the character of the sedimentary structural geometry, northwest Argentina: Tectonics, man, W., Stampfli, G.M., and Ziegler, P.A., eds., The
cover sequence, it is likely that most progres- v. 2, p. 1–16, doi:10.1029/TC002i001p00001. TRANSMED Atlas; the Mediterranean Region from
sive oroclines, like the Wyoming Salient, are Allmendinger, R.W., Jordan, T.E., Kay, S.M., and Isacks, Crust to Mantle; Geological and Geophysical Frame-
B.L., 1997, The evolution of the Altiplano-Puna Pla- work of the Mediterranean and the Surrounding Areas:
products of a combination of critical taper and teau of the Central Andes: Annual Review of Earth Berlin, Springer-Verlag, p. 1–29.
obstacle-induced curvature. Secondary oroclines and Planetary Sciences, v. 25, p. 139–174, doi:10.1146 Channell, J.E.T., Lowrie, W., Medizza, F., and Alvarez, W.,
/annurev.earth.25.1.139. 1978, Palaeomagnetism and tectonics in Umbria, Italy:
are extra-orogenic, developing in response to an Allmendinger, R.W., Smalley, R., Bevis, M., Caprio, H., and Earth and Planetary Science Letters, v. 39, p. 199–210,
orogen-parallel stress applied to a preexisting Brooks, B., 2005, Bending the Bolivian orocline in real doi:10.1016/0012-821X(78)90196-6.
orogen; involve relative translations measured time: Geology, v. 33, p. 905–908, doi:10.1130/G21779.1. Cifelli, F., Mattei, M., and Rossetti, F., 2007, Tectonics evo-
Allmendinger, R.W., Reilinger, R., and Loveless, J.P., 2007, lution of arcuate mountain belts on top of a retreating
in the hundreds to thousands of kilometers; and Strain and rotation rate from GPS in Tibet, Anatolia, subduction slab: The example of the Calabrian arc:
are lithospheric scale, affecting crust and litho- and the Altiplano: Tectonics, v. 26, TC3013, doi: Journal of Geophysical Research, v. 112, B09101,
spheric mantle. The only way to accommodate 10.1029/2006TC002030. doi:10.1029/2006JB004848.
Arenas, R., and Martínez-Catalán, J.R., 2003, Low-P Cifelli, F., Mattei, M., and Della Seta, M., 2008, Calabrian
the scale of deformation, the involvement of the metamorphism following a Barrovian-type evolution: arc oroclinal bending: The role of subduction: Tectonics,
lithospheric mantle, and the significant transla- Complex tectonic controls for a common transition, as v. 27, TC5001, doi:10.1029/2008TC002272.
deduced in the Mondoñedo thrust sheet (NW Iberian Clift, P.D., Schouten, H., and Draut, A.E., 2003, A general
tions required during secondary orocline forma- Massif ): Tectonophysics, v. 365, p. 143–164, doi: model of arc-continent collision and subduction polar-
tion is through subduction. 10.1016/S0040-1951(03)00020-9. ity reversal from Taiwan and the Irish Caledonides, in
Failure to distinguish between progressive Armstrong, F.C., and Oriel, S.S., 1965, Tectonic develop- Larter, R.D., and Leat, P.T., eds., Intra-Oceanic Sub-
ment of the Idaho-Wyoming thrust belt: The American duction Systems: Tectonic and Magmatic Processes:
and secondary oroclines has resulted in much Association of Petroleum Geologists Bulletin, v. 49, Geological Society of London Special Publication 219,
confusion. Analogue models involving inden- p. 1847–1866. p. 81–98, doi:10.1144/GSL.SP.2003.219.01.04.

Geological Society of America Bulletin, May/June 2013 659


Johnston et al.

Cobbold, P.R., and Davy, P., 1988, Indentation tectonics in tional, v. 101, p. 425–478, doi:10.1111/j.1365-246X of paleomagnetic rotations in Upper Permian to Lower
nature and experiments: II. Central Asia: Bulletin of .1990.tb06579.x. Jurassic rocks from northern and southern Peru: Earth
the Geological Institutions of the University of Upp- Dewey, J.F., 2005, Orogeny can be very short: Proceedings and Planetary Science Letters, v. 210, p. 233–248,
sala, v. 14, p. 143–162. of the National Academy of Sciences of the United doi:10.1016/S0012-821X(03)00102-X.
Colmenero, J.R., and Prado, J.G., 1993, Coal basins in the States of America, v. 102, no. 43, p. 15,286–15,293, Gómez-Fernández, F., Both, R.A., Mangas, J., and Arribas,
Cantabrian Mountains, northwestern Spain: Interna- doi:10.1073/pnas.0505516102. A., 2000, Metallogenesis of Zn-Pb carbonate-hosted
tional Journal of Coal Geology, v. 23, p. 215–229, doi: Dickinson, W.R., 2004, Evolution of the North American mineralization in the southeastern region of the Picos de
10.1016/0166-5162(93)90049-G. Cordillera: Annual Review of Earth and Planetary Europa (central northern Spain) Province; geologic, fluid
Cook, F.A., van der Velden, A.J., Hall, K.W., and Roberts, Sciences, v. 32, p. 13–45, doi:10.1146/annurev.earth inclusion, and stable isotope studies: Economic Geology
B.J., 1998, Tectonic delamination and subcrustal im- .32.101802.120257. and the Bulletin of the Society of Economic Geologists,
brication of the Precambrian lithosphere in northwest- Díez Balda, M.A., Vegas, R., and González Lodeiro, F., v. 95, p. 19–40, doi:10.2113/gsecongeo.95.1.19.
ern Canada mapped by LITHOPROBE: Geology, v. 26, 1990. Structure, in Dallmeyer, R.D., and Martínez Gordon, R.G., 1998, The plate tectonic approximation: Plate
p. 839–842, doi:10.1130/0091-7613(1998)026<0839: García, E., eds., Pre-Mesozoic Geology of Iberia: Berlin, nonrigidity, diffuse plate boundaries, and global plate
TDASIO>2.3.CO;2. Springer-Verlag, p. 172–188. reconstructions: Annual Review of Earth and Plan-
Costa, J.B.S., Léa Bemerguy, R., Hasui, Y., and da Silva Ducea, M.N., 2011, Fingerprinting orogenic delamination: etary Sciences, v. 26, p. 615–642, doi:10.1146/annurev
Borges, M., 2001, Tectonics and paleogeography Geology, v. 39, p. 191–192, doi:10.1130/focus022011.1. .earth.26.1.615.
along the Amazon River: Journal of South Ameri- Eldredge, S., and Van der Voo, R., 1988, Paleomagnetic Gray, M.B., and Stamatakos, J., 1997, New model for evolu-
can Earth Sciences, v. 14, p. 335–347, doi:10.1016 study of thrust sheet rotations in the Helena and Wyo- tion of fold and thrust belt curvature based on integrated
/S0895-9811(01)00025-6. ming Salients of the Northern Rocky Mountains, in structural and paleomagnetic results from the Pennsyl-
Craddock, J.P., 1992, Transpression during tectonic evolu- Schmidt, C.J., ed., Interaction of the Rocky Mountain vania Salient: Geology, v. 25, no. 12, p. 1067–1070,
tion of the Idaho-Wyoming fold-and-thrust belt, in Foreland and the Cordilleran Thrust Belt: Geological doi:10.1130/0091-7613(1997)025<1067:NMFEOF>2.3
Link, P., Duntz, M.A., and Platt, L.B., eds., Regional Society of America Memoir 171, p. 319–332. .CO;2.
Geology of Eastern Idaho and Western Wyoming: Geo- Eldredge, S., Bachtadse, V., and Van der Voo, R., 1985, Grubbs, K.L., and Van der Voo, R., 1976, Structural deforma-
logical Society of America Memoir 179, p. 125–139. Paleomagnetism and the orocline hypothesis: Tecto- tion of the Idaho-Wyoming overthrust belt (USA), as de-
Craddock, J.P., Kopania, A.A., and Wiltschko, D.V., 1988, nophysics, v. 119, p. 153–179, doi:10.1016/0040-1951 termined by Triassic paleomagnetism: Tectonophysics,
Interaction between the northern Idaho-Wyoming (85)90037-X. v. 33, p. 321–336, doi:10.1016/0040-1951(76)90151-7.
thrust belt and bounding basement blocks, central Engebretson, D.C., and Richards, M.A., 1992, 180 million Gutiérrez-Alonso, G., Fernández-Suárez, J., and Weil, A.B.,
western Wyoming, in Schmidt, C.J., and Perry, W.J., years of subduction: GSA Today, v. 2, no. 5, p. 93–100. 2004, Orocline triggered lithospheric delamination, in
Jr., eds., Interaction of the Rocky Mountain Foreland Engebretson, D.C., Cox, A., and Gordon, R.G., 1985, Rela- Sussman, A.J., and Weil, A.B., eds., Orogenic Curva-
and the Cordilleran Thrust Belt: Geological Society of tive Motions between Oceanic and Continental Plates ture: Integrating Paleomagnetic and Structural Analy-
America Memoir 171, p. 333–351. in the Pacific Basin: Geological Society of America ses: Geological Society of America Special Paper 383,
Crosby, G.W., 1969, Radial movements in the western Wyo- Special Paper 206, 59 p. p. 121–131.
ming Salient of the Cordilleran overthrust belt: Geologi- Farris, D.W., Jaramillo, C., Bayona, G., Restrepo-Moreno, Gutiérrez-Alonso, G., Fernández-Suárez, J., Weil, A.B., Mur-
cal Society of America Bulletin, v. 80, p. 1061–1077, S.A., Montes, C., Cardona, A., Mora, A., Speakman, phy, B.J., Nance, R.D., Corfu, F., and Johnston, S.T.,
doi:10.1130/0016-7606(1969)80[1061:RMITWW]2.0 R.J., Glascock, M.D., and Valencia, V., 2011, Fractur- 2008, Self-subduction of the Pangean global plate: Na-
.CO;2. ing of the Panamanian Isthmus during initial collision ture Geoscience, v. 1, p. 549–553, doi:10.1038/ngeo250.
Dahlen, F.A., 1990, Critical taper model of fold-and-thrust with South America: Geology, v. 39, p. 1007–1010, Gutiérrez-Alonso, G., Murphy, B., Fernández-Suárez, J.,
belts and accretionary wedges: Annual Review of Earth doi:10.1130/G32237.1. Weil, A.B., Franco, M.P., and Gonzalo, J.C., 2011a,
and Planetary Sciences, v. 18, p. 55–99, doi:10.1146 Fernández-Suárez, J., Dunning, G.R., Jenner, G.A., and Lithospheric delamination in the core of Pangea:
/annurev.ea.18.050190.000415. Gutiérrez-Alonso, G., 2000, Variscan collisional mag- Sm-Nd insights from the Iberian mantle: Geology,
Dahlstrom, C.D.A., 1969, Balanced cross sections: Cana- matism and deformation in NW Iberia: Constraints v. 39, p. 155–158, doi:10.1130/G31468.1.
dian Journal of Earth Sciences, v. 6, p. 743–757, doi: from U-Pb geochronology of granitoids: Journal of Gutiérrez-Alonso, G., Fernández-Suárez, J., Jeffries, T.E.,
10.1139/e69-069. the Geological Society of London, v. 157, p. 565–576, Johnston, S.T., Pastor-Galán, D., Murphy, J.B., Franco,
Dallmeyer, R.D., Martínez-Catalán, J.R., Arenas, R., Gil Ibar- doi:10.1144/jgs.157.3.565. M.P., and Gonzalo, J.C., 2011b, Diachronous post-
guchi, J.I., Gutiérrez-Alonso, G., Farias, P., Bastida, F., Fillerup, M.A., Knapp, J.H., Knapp, C.C., and Raileanu, V., orogenic magmatism within a developing orocline in
and Aller, J., 1997, Diachronous Variscan tectonother- 2010, Mantle earthquakes in the absence of subduction? Iberia, European Variscides: Tectonics, v. 30, TC5008,
mal activity in the NW Iberian Massif: Evidence from Continental delamination in the Romanian Carpathians: doi:10.1029/2010TC002845.
40
Ar/39Ar dating of regional fabrics: Tectonophysics, Lithosphere, v. 2, p. 333–340, doi:10.1130/L102.1. Gutiérrez-Alonso, G., Johnston, S.T., Weil, A.B., Pastor-
v. 277, p. 307–337, doi:10.1016/S0040-1951(97)00035-8. Franke, W., Matte, P., and Tait, J., 2005, Variscan Orog- Galán, D., and Fernández-Suárez, J., 2012, Buckling
Davis, D., Suppe, J., and Dahlen, F.A., 1983, Mechanics of eny: Encyclopedia of Geology: Amsterdam, Elsevier, an orogen: The Cantabrian Orocline: GSA Today,
fold-and-thrust belts and accretionary wedges: Jour- p. 75–85. v. 22, no.7, doi:10.1130/GSATG141A.1.
nal of Geophysical Research, v. 88, p. 1153–1172, Garzione, C.N., Molnar, P., Libarkin, J.C., and MacFadden, Gutiérrez-Marco, J.C., Robardet, M., Rábano, I., Sarmiento,
doi:10.1029/JB088iB02p01153. B., 2006, Rapid late Miocene rise of the Bolivian Alti- G.N., Sán Jose Lancha, M.Á., and Herranz, P.A., 2002,
Davy, P.H., and Cobbold, P.R., 1991, Experiments on plano: Evidence for removal of mantle lithosphere: Ordovician, in Gibbons, W., and Moreno, T., eds., The
shortening of a 4-layer model of the continental litho- Earth and Planetary Science Letters, v. 241, p. 543– Geology of Spain: Bath, UK, Geological Society of
sphere: Tectonophysics, v. 188, p. 1–25, doi:10.1016 556, doi:10.1016/j.epsl.2005.11.026. London, p. 31–49.
/0040-1951(91)90311-F. Gasparrini, M., Bakker, R.J., Bechstädt, T., and Boni, M., Gvirtzman, Z., 2002, Partial detachment of a lithospheric
DeCelles, P.G., 1994, Late Cretaceous–Paleocene syn- 2003, Hot dolomites in a Variscan foreland belt: Hydro- root under the southeast Carpathians: Towards a better
orogenic sedimentation and kinematic history of the thermal flow in the Cantabrian zone (NW Spain): Jour- definition of the detachment concept: Geology, v. 30,
Sevier thrust belt, northeast Utah and southwest Wyo- nal of Geochemical Exploration, v. 78–79, p. 501–507, p. 51–54, doi:10.1130/0091-7613(2002)030<0051:
ming: Geological Society of America Bulletin, v. 106, doi:10.1016/S0375-6742(03)00115-8. PDOALR>2.0.CO;2.
p. 32–56, doi:10.1130/0016-7606(1994)106<0032: Gasparrini, M., Bechstädt, T., and Boni, M., 2006a, Massive Hatcher, R.D., Jr., 1989, Tectonic synthesis of the U.S. Appa-
LCPSSA>2.3.CO;2. hydrothermal dolomites in the southwestern Canta- lachians, in Hatcher, R.D., Jr., Thomas, W.A., and Viele,
DeCelles, P.G., 2004, Late Jurassic to Eocene evolution of brian zone (Spain) and their relation to the Late Vari- G.W., eds., The Appalachian-Ouachita Orogen in the
the Cordilleran thrust belt and foreland basin system, scan evolution: Marine and Petroleum Geology, v. 23, United States: Boulder, Colorado, Geological Society of
western U.S.A.: American Journal of Science, v. 304, p. 543–568, doi:10.1016/j.marpetgeo.2006.05.003. America, Geology of North America, v. F-2, p. 511–535.
p. 105–168, doi:10.2475/ajs.304.2.105. Gasparrini, M., Bakker, R.J., and Bechstädt, T., 2006b, Char- Hatcher, R.D., Jr., 2002, Alleghanian (Appalachian) orogeny,
DeCelles, P.G., and Coogan, J.C., 2006, Regional structure acterisation of dolomitising fluids in the Carboniferous a product of zipper tectonics: Rotational, transpressive
and kinematic history of the Sevier fold-and-thrust of the Cantabrian zone (NW Spain): A fluid inclusion continent–continent collision and closing of ancient
belt, central Utah: Geological Society of America Bul- study with cryo-Raman spectroscopy: Journal of Sedi- oceans along irregular margins, in Martínez Catalán,
letin, v. 118, p. 841–864, doi:10.1130/B25759.1. mentary Research, v. 76, p. 1304–1322, doi:10.2110 J.R., Hatcher, R.D., Jr., Arenas, R., and Díaz García,
DeCelles, P.G., Ducea, M.N., Kapp, P., and Zandt, G., 2009, /jsr.2006.106. F., eds., Variscan-Appalachian Dynamics: The Building
Cyclicity in Cordilleran orogenic systems: Nature Geo- Gephart, J.W., 1994, Topography and subduction geometry of the Late Paleozoic Basement: Geological Society of
science, v. 2, p. 251–257, doi:10.1038/ngeo469. in the central Andes: Clues to the mechanics of a non- America Special Paper 364, p. 199–208.
de la Rosa, J., Jenner, J., and Castro, A., 2002, A study of collisional orogen: Journal of Geophysical Research, Hérail, G., Oller, J., Baby, P., Blanco, J., Bonhomme, M.G.,
inherited zircons in granitoid rocks from the South v. 99, p. 12,279–12,288, doi:10.1029/94JB00129. and Soler, P., 1996, The Tupiza, Nazareno and Estarca
Portuguese and Ossa Morena zones, Iberian Massif: Ghosh, P., Garzione, C.N., and Eiler, J.M., 2006, Rapid basins (Bolivia): Strike-slip faulting and related basins
Support for the exotic origin of the South Portuguese uplift of the Altiplano revealed through 13C-18O bonds in the Cenozoic evolution of the southern branch of the
zone: Tectonophysics, v. 352, p. 245–256, doi:10.1016 in paleosol carbonates: Science, v. 311, p. 511–515, Bolivian orocline: Tectonophysics, v. 259, p. 201–212,
/S0040-1951(02)00199-3. doi:10.1126/science.1119365. doi:10.1016/0040-1951(95)00108-5.
DeMets, C., Gordon, R.G., Argus, D.F., and Stein, S., 1990, Gilder, S., Rousse, S., Farber, D., Sempere, T., Torres, V., Hindle, H., and Burkhard, M., 1999, Strain, displacement and
Current plate motion: Geophysical Journal Interna- and Palacios, O., 2003, Post–middle Oligocene origin rotation associated with the formation of curvature on

660 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

fold belts; the example of the Jura arc: Journal of Struc- Julivert, M., and Marcos, A., 1973, Superimposed folding Arcs, Developments in Geotectonics 21: Amsterdam,
tural Geology, v. 21, p. 1089–1101, doi:10.1016/S0191 under flexural conditions in the Cantabrian zone (Her- Netherlands, Elsevier, p. 141–158.
-8141(99)00021-8. cynian Cordillera, northwest Spain): American Journal Lucente, F.P., and Speranza, F., 2001, Belt bending driven
Hindle, D., Besson, O., and Burkhard, M., 2000, A model of Science, v. 273, p. 353–375, doi:10.2475/ajs.273 by lateral bending of subducting lithospheric slab:
of displacement and strain for arc-shaped moun- .5.353. Geophysical evidences from the northern Apennines
tain belts applied to the Jura arc: Journal of Struc- Keep, M., 2000, Models of lithospheric-scale deformation (Italy): Tectonophysics, v. 337, p. 53–64, doi:10.1016
tural Geology, v. 22, p. 1285–1296, doi:10.1016 during plate collision: Effects of indentor shape and /S0040-1951(00)00286-9.
/S0191-8141(00)00038-9. lithospheric thickness: Tectonophysics, v. 326, p. 203– Macedo, J., and Marshak, S., 1999, Controls on the geometry
Hindle, D., Kley, J., Klosko, E., Stein, S., Dixon, T., and 216, doi:10.1016/S0040-1951(00)00123-2. of fold-thrust belt salients: Geological Society America
Norabuena, E., 2002, Consistency of geologic and geo- Kent, D., 1988, Further paleomagnetic evidence for oroclinal Bulletin, v. 111, p. 1808–1822, doi:10.1130/0016-7606
detic displacements during Andean orogenesis: Geo- rotation in the central folded Appalachians from the (1999)111<1808:COTGOF>2.3.CO;2.
physical Research Letters, v. 29, 1188, doi:10.1029 Bloomsburg and the Mauch Chunk formations: Tectonics, MacFadden, B., Anaya, F., and Swisher, C., III, 1995, Neo-
/2001GL013757. v. 7, p. 749–759, doi:10.1029/TC007i004p00749. gene paleomagnetism and oroclinal bending of the
Hobbs, W.H., 1914, Mechanics or formation of arcuate Kent, D., and Opdyke, N.D., 1985, Multicomponent mag- central Andes of Bolivia: Journal of Geophysical Re-
mountains: The Journal of Geology, v. 22, p. 71–90, netization from the Mississippian Mauch Chunk For- search, v. 100, p. 8153–8167, doi:10.1029/95JB00149.
doi:10.1086/622134. mation of the central Appalachians and their tectonic Maffione, M., Speranza, F., and Faccenna, C., 2009, Bending
Hollingsworth, J., Fattahi, M., Walker, R., Talebian, M., Bah- implications: Journal of Geophysical Research, v. 90, of the Bolivian orocline and growth of the Central Andean
roudi, A., Bolourchi, M.J., Jackson, J., and Copley, A., p. 5371–5383, doi:10.1029/JB090iB07p05371. plateau: Paleomagnetic and structural constraints from
2010, Oroclinal bending, distributed thrust and strike- Kley, J., 1999, Geologic and geometric constraints on a the Eastern Cordillera (22–24°S, NW Argentina): Tec-
slip faulting, and the accommodation of Arabia–Eurasia kinematic model of the Bolivian orocline: Journal of tonics, v. 28, TC4006, doi:10.1029/2008TC002402.
convergence in NE Iran since the Oligocene: Geophysi- South American Earth Sciences, v. 12, p. 221–235, Maffione, M., Speranza, F., Faccenna, C., and Rossello, E.,
cal Journal International, v. 181, no. 3, p. 1214–1246. doi:10.1016/S0895-9811(99)00015-2. 2010, Paleomagnetic evidence for a pre–early Eocene
Houseman, G.A., and Gemmer, L., 2007, Intra-orogenic ex- Kley, J., and Monaldi, C.R., 1998, Tectonic shortening and (~50 Ma) bending of the Patagonian orocline (Tierra
tension driven by gravitational instability: Carpathian- crustal thickening in the Central Andes: How good is the del Fuego, Argentina): Paleogeographic and tectonic
Pannonian orogeny: Geology, v. 35, p. 1135–1138, doi: correlation?: Geology, v. 26, p. 723–726, doi:10.1130 implications: Earth and Planetary Science Letters,
10.1130/G23993A.1. /0091-7613(1998)026<0723:TSACTI>2.3.CO;2. v. 289, p. 273–286, doi:10.1016/j.epsl.2009.11.015.
Houseman, G.A., and Molnar, P., 1997, Gravitational Knapp, J.H., Kapp, C.C., Raileanu, V., Matenco, L., Marshak, S., 1988, Kinematics of orocline and arc forma-
(Rayleigh-Taylor) instability of a layer with non-lin- Mocanu, V., and Dinu, C., 2005, Crustal constraints tion in thin-skinned orogens: Tectonics, v. 7, no. 1,
ear viscosity and convective thinning of continental on the origin of mantle seismicity in the Vrancea zone, p. 73–86, doi:10.1029/TC007i001p00073.
lithosphere: Geophysical Journal International, v. 128, Romania: The case for active continental lithospheric Marshak, S., 2004, Salients, recesses, arcs, oroclines, and
p. 125–150, doi:10.1111/j.1365-246X.1997.tb04075.x. delamination: Tectonophysics, v. 410, p. 311–323, syntaxes; a review of ideas concerning the formation
Houseman, G.A., McKenzie, D.P., and Molnar, P., 1981, doi:10.1016/j.tecto.2005.02.020. of map-view curves in fold-thrust belts, in McClay,
Convection instability of a thickened boundary-layer Kollmeier, J.M., van der Pluijm, B.A., and Van der Voo, R., K.R., ed., Thrust Tectonics and Hydrocarbon Systems:
and its relevance for the thermal evolution of continen- 2000, Analysis of Variscan dynamics; early bending American Association of Petroleum Geologists Mem-
tal convergent belts: Journal of Geophysical Research, of the Cantabria-Asturias arc, northern Spain: Earth oir 82, p. 131–156.
v. 86, p. 6115–6132, doi:10.1029/JB086iB07p06115. and Planetary Science Letters, v. 181, p. 203–216, Marshak, S., Wilkerson, M.S., and Hsui, A.T., 1992, Gen-
Irving, E., and Opdyke, N.D., 1965, The paleomagnetism of doi:10.1016/S0012-821X(00)00203-X. eration of curved fold-thrust belts: Insight from simple
the Bloomsburg redbeds and its possible application to Kotasek, J., and Krs, M., 1965, Palaeomagnetic study of tectonic physical and analytical models, in McClay, K.R., ed.,
the tectonic history of the Appalachians: Geophysical rotation in the Carpathian Mountains of Czechoslovakia: Thrust Tectonics: London, Chapman & Hall, p. 83–92.
Journal of the Royal Astronomical Society, v. 9, p. 153– Palaeogeography, Palaeoclimatology, Palaeoecology, Martínez-Catalán, J.R., 1990, A non-cylindrical model for
167, doi:10.1111/j.1365-246X.1965.tb02067.x. v. 1, p. 39–49, doi:10.1016/0031-0182(65)90005-2. the northwest Iberian allochthonous terranes and their
Irving, E., Wynne, P.J., Thorkelson, D.J., and Schiarizza, Kraig, D.H., Wiltschko, D.V., and Spang, J.H., 1987, Interaction equivalents in the Hercynian belt of western Europe:
P., 1996, Large (1000–4000 km) northward move- of basement uplift and thin-skinned thrusting, Moza Tectonophysics, v. 179, p. 253–272, doi:10.1016/0040
ments of tectonic domains in the northern Cordillera, arch and the Western Overthrust Belt, Wyoming: a hy- -1951(90)90293-H.
83 to 45 Ma: Journal of Geophysical Research, v. 101, pothesis: Geological Society of America Bulletin, v. 99, Martínez-Catalán, J.R., Arenas, R., Díaz García, F., and Abati,
p. 17,901–17,916, doi:10.1029/96JB01181. p. 654–662, doi:10.1130/0016-7606(1987)99<654: J., 1997, Variscan accretionary complex of NW Iberia:
Isacks, B.L., 1988, Uplift of the central Andean plateau IOBUAT>2.0.CO;2. Terrane correlation and succession of tectonothermal
and bending of the Bolivian orocline: Journal of Geo- Kraig, D.H., Wiltschko, D.V., and Spang, J.H., 1988, The events: Geology, v. 25, p. 1103–1106, doi:10.1130
physical Research, v. 93, p. 3211–3231, doi:10.1029 interaction of the Moxa Arch (La Barge Platform) /0091-7613(1997)025<1103:VACONI>2.3.CO;2.
/JB093iB04p03211. with the Cordilleran thrust belt, south of Snider Basin, Martínez-Catalán, J.R., Arenas, R., Díaz García, F., Gomez-
Johnston, S.T., 2001, The Great Alaskan terrane wreck: southwestern Wyoming, in Schmidt, C.J., and Perry, Barreiro, J., Gonzalez Cuadra, P., Abati, J., Castineiras,
Reconciliation of paleomagnetic and geological data W.J., eds., Interaction of the Rocky Mountain Foreland P., Fernández-Suárez, J., Sanchez Martínez, S., Ando-
in the northern Cordillera: Earth and Planetary Sci- and the Cordilleran Thrust Belt: Geological Society of naegui, P., Gonzalez Clavijo, E., Diez Montes, F.A.,
ence Letters, v. 193, p. 259–272, doi:10.1016/S0012 America Memoir 171, p. 395–410. Rubio Pascual, J., and Valle Aguado, B., 2007, Space
-821X(01)00516-7. Kwon, S., and Mitra, G., 2004, Strain distribution, strain and time in the tectonic evolution of the northwestern
Johnston, S.T., 2004, The d’Entrecasteaux orocline and rota- history, and kinematic evolution associated with the Iberian massif: Implications for the Variscan belt, in
tion of the Vanuatu–New Hebrides arc: An oroclinal orgy formation of arcuate salients in fold-thrust belts; Hatcher, R.D., Jr., Carlson, M.P., McBride, J.H., and
and analogue for Archean craton formation, in Sussman, the example of the Provo Salient, Sevier Orogen, Utah Martínez Catalán, J.R., eds., 4-D Framework of Con-
A., and Weil, A.B., eds., Orogenic Curvature: Integrating 2004, in Sussman, A.J., and Weil, A.B., eds., Orogenic tinental Crust: Geological Society of America Memoir
Paleomagnetic and Structural Analyses: Geological So- Curvature: Integrating Paleomagnetic and Structural 200, p. 403–423.
ciety of America Special Paper 383, p. 225–236. Analyses: Geological Society of America Special Martín-Izard, A., Fuertes, M., Cepedal, A., Moreiras, D.,
Johnston, S.T., 2008, The Cordilleran ribbon continent of Paper 383, p. 205–223. Nieto, J.G., Maldonado, C., and Pevida, L., 2000,
North America: Annual Review of Earth and Plan- Lamb, S.H., 2000, Active deformation in the Bolivian The Rio Narcea gold belt intrusions; geology, petrol-
etary Sciences, v. 36, p. 495–530, doi:10.1146/annurev Andes, South America: Geophysical Research v. 105, ogy, geochemistry and timing: Journal of Geochemi-
.earth.36.031207.124331. p. 25,627–25,653, doi:10.1029/2000JB900187. cal Exploration, v. 71, p. 103–117, doi:10.1016/S0375
Johnston, S.T., and Gutierrez-Alonso, G., 2010, The North Lamb, S.H., 2001, Vertical axis rotation in the Bolivian -6742(00)00148-5.
American Cordillera and West European Variscides: orocline, South America: 1. Paleomagnetic analysis of Matte, P., 2001, The Variscan collage and orogeny (480–
Contrasting interpretations of similar mountain sys- Cretaceous and Cenozoic rocks: Journal of Geophysi- 290 Ma) and the tectonic definition of the Armorica
tems: Gondwana Research, v. 17, no. 2–3, p. 516–525, cal Research, v. 106, p. 26,605–26,632, doi:10.1029 microplate: A review: Terra Nova, v. 13, p. 122–128, doi:
doi:10.1016/j.gr.2009.11.006. /2001JB900012. 10.1046/j.1365-3121.2001.00327.x.
Johnston, S.T., and Mazzoli, S., 2009, The Calabrian oro- Lefort, J.P., 1988, Basement Correlation across the North McCaig, A.M., and McClelland, E., 1992, Palaeomagnetic
cline: Buckling of a previously more linear orogen, in Atlantic: Berlin, Germany, Springer, 148 p. techniques applied to thrust belts, in McClay, K.R.,
Murphy, J.B., Keppie, J.D., and Hynes, A.J., eds., An- Linzer, H.G., 1996, Kinematics of retreating subduction ed., Thrust Tectonics: London, Chapman and Hal,
cient Orogens and Modern Analogues: Geological So- along the Carpathian arc, Romania: Geology, v. 24, p. 209–216.
ciety of London Special Publication 327, p. 113–125. p. 167–170, doi:10.1130/0091-7613(1996)024<0167: McQuarrie, N., 2002, The kinematic history of the central
Jordan, T.E., 1981, Thrust loads and foreland basin evolu- KORSAT>2.3.CO;2. Andean fold-thrust belt, Bolivia: Implications for
tion, Cretaceous, western United States: The American Lorinczi, P., and Houseman, G.A., 2009, Lithospheric gravi- building a high plateau: Geological Society of Amer-
Association of Petroleum Geologists Bulletin, v. 65, tational instability beneath the southeast Carpathians: ica Bulletin, v. 114, no. 8, p. 950–963, doi:10.1130
p. 2506–2520. Tectonophysics, v. 474, p. 322–336, doi:10.1016/j.tecto /0016-7606(2002)114<0950:TKHOTC>2.0.CO;2.
Julivert, M., 1971, Décollement tectonics in the Hercynian .2008.05.024. Miller, J.D., and Kent, D.V., 1986a, Paleomagnetic of the
Cordillera of NW Spain: American Journal of Science, Lowrie, W., and Hirt, A.M., 1986, Paleomagnetism in arcu- Upper Devonian Catskill Formation from the southern
v. 270, p. 1–29, doi:10.2475/ajs.270.1.1. ate mountain belts, in Wezel, F.C., ed., The Origin of limb of the Pennsylvania Salient: Possible evidence of

Geological Society of America Bulletin, May/June 2013 661


Johnston et al.

oroclinal rotation: Geophysical Research Letters, v. 13, Pastor-Galán, D., Gutiérrez-Alonso, G., Zulauf, G., and the Iapetus Ocean: Journal of Geophysical Research,
p. 1173–1176, doi:10.1029/GL013i011p01173. Zanella, F., 2012b, Analogue modeling of lithospheric- v. 81, p. 5605–5619, doi:10.1029/JB081i032p05605.
Miller, J.D., and Kent, D.V., 1986b, Synfolding and pre- scale orocline buckling: Constraints on the evolution Rast, N., O’Brien, B.H., and Wardle, R.G., 1976, Relation-
folding magnetizations in the upper Devonian Catskill of the Iberian-Armorican Arc: Geological Society of ships between Precambrian and Paleozoic rocks of the
formation of eastern Pennsylvania: Implication for America Bulletin, v. 124, p. 1293–1309, doi:10.1130 ‘Avalon Platform’ in New Brunswick, the northeast
the tectonic history of Acadia: Journal of Geophysi- /B30640.1. Appalachians and the British Isles: Tectonophysics,
cal Research, v. 91, p. 12,791–12,803, doi:10.1029 Paulsen, T., and Marshak, S., 1997, Structure of the Mount v. 30, p. 315–338, doi:10.1016/0040-1951(76)90192-X.
/JB091iB12p12791. Raymond transverse zone at the southern margin of the Ries, A.C., and Shackleton, R.M., 1976, Patterns of strain varia-
Miller, M.S., Levander, A., Niu, F., and Li, A., 2009, Upper Wyoming Salient in the Sevier fold-thrust belt, Utah: tion in arcuate fold belts: Philosophical Transactions of
mantle structure beneath the Caribbean–South Ameri- Tectonophysics, v. 280, p. 199–211, doi:10.1016/S0040 the Royal Society of London, v. 283, p. 281–288.
can plate boundary from surface wave tomography: -1951(97)00205-9. Riller, U., and Oncken, O., 2003, Growth of the central
Journal of Geophysical Research, v. 114, B01312, doi: Paulsen, T., and Marshak, S., 1998, Structure of the Andean plateau by tectonic segmentation is controlled
10.1029/2007JB005507. Charleston transverse zone, Wasatch Mountains, Utah: by the gradient in crustal shortening: The Journal of
Mitra, G., 1994, Strain variation in thrust sheets across the Tectonic evolution of the northern margin of the Provo Geology, v. 111, p. 367–384, doi:10.1086/373974.
Sevier fold-and-thrust belt (Idaho-Utah-Wyoming); salient in the Sevier fold-thrust belt: Geological Society Riller, U., Petrinovic, I., Ramelow, J., Strecker, M., and
implications for section restoration and wedge taper of America Bulletin, v. 110, p. 512–522, doi:10.1130 Oncken, O., 2001, Late Cenozoic tectonism, collapse
evolution: Journal of Structural Geology, v. 16, p. 585– /0016-7606(1998)110<0512:CTZWMU>2.3.CO;2. caldera, and plateau formation in the central Andes:
602, doi:10.1016/0191-8141(94)90099-X. Paulsen, T., and Marshak, S., 1999, Origin of the Uinta recess, Earth and Planetary Science Letters, v. 188, p. 299–311,
Mitra, G., and Yonkee, W.A., 1985, Spaced cleavage and Sevier fold-thrust belt, Utah: Influence of basin architec- doi:10.1016/S0012-821X(01)00333-8.
its relationship to folds and thrusts in the Idaho-Utah- ture on fold-thrust belt geometry: Tectonophysics, v. 312, Robardet, M., 2002, Alternative approach to the Variscan
Wyoming thrust belt of the Rocky Mountain Cordi- p. 203–216, doi:10.1016/S0040-1951(99)00182-1. belt in southwestern Europe: Preorogenic paleobio-
lleras: Journal of Structural Geology, v. 7, p. 361–373, Pereira, M.F., and Silva, J.B., 2001, The northeast Alentejo geographical constraints, in Martinez Catalan, J.R.,
doi:10.1016/0191-8141(85)90041-0. Neoproterozoic–Lower Cambrian succession (Portu- Hatcher, R.D., Jr., Arenas, R., and Diaz Garcia, F.,
Molnar, P., and Garzione, C.N., 2007, Bounds on the vis- gal): Implications for regional correlations in the Ossa eds., Variscan-Appalachian Dynamics: The Building
cosity coefficient of continental lithosphere from Morena zone (Iberian Massif): Sociedad Geologica de of the Late Paleozoic Basement: Geological Society of
removal of mantle lithosphere beneath the Altiplano Espana: Geogaceta, v. 30, p. 106–111. America Special Paper 364, p. 1–15.
and Eastern Cordillera: Tectonics, v. 26, TC2013, doi: Pérez-Estaún, A., Bastida, F., Alonso, J.L., Marquinez, J., Robardet, M., 2003, The Armorica ‘microplate’: Fact or fic-
10.1029/2006TC001964. Aller, J., Álvarez-Marron, J., Marcos, A., and Pulgar, tion? Critical review of the concept and contradictory
Montes, C., Bayona, G., Cardona, A., Busch, D.M., Silva, J.A., 1988, A thin-skinned tectonics model for an arcu- palaeobiogeographical data: Palaeogeography, Palaeo-
C.A., Morón, S.A., Hoyos, N., Ramirez, D.A., Jara- ate fold and thrust belt: The Cantabrian zone (Variscan climatology, Palaeoecology, v. 195, no. 1–2, p. 125–
millo, C., and Valencia, V., 2012, Arc-continent col- Armorican Arc): Tectonics, v. 7, p. 517–537. 148, doi:10.1016/S0031-0182(03)00305-5.
lision and orocline formation: Closing of the Central Pérez-Estaún, A., Martinez-Catalán, J.R., and Bastida, F., Robardet, M., and Gutiérrez-Marco, J. C., 1990, Sedimen-
America Seaway: Journal of Geophysical Research, 1991, Crustal thickening and deformation sequence in tary and faunal domains in the Iberian Peninsula dur-
v. 117, B04105, doi:10.1029/2011JB008959. the footwall to the suture of the Variscan belt of north- ing Lower Paleozoic times: Pre-Mesozoic Geology of
Müller, J., Kley, J., and Jacobshagen, V., 2002, Structure and west Spain: Tectonophysics, v. 191, no. 3–4, p. 243– Iberia, IGCP-Project 233, p. 383–395.
Cenozoic kinematics of the Eastern Cordillera, south- 253, doi:10.1016/0040-1951(91)90060-6. Roperch, P., Sempere, T., Macedo, O., Arriagada, C., For-
ern Bolivia (21°S): Tectonics, v. 21, 1037, doi:10.1029 Perroud, H., Bonhommet, N., and Ribeiro, A., 1985, Paleo- nari, M., Tapia, C., García, M., and Laj, C., 2006,
/2001TC001340. magnetism of late Paleozoic igneous rocks from south- Counterclockwise rotation of late Eocene–Oligocene
Muttoni, G., Argnani, A., Kent, D.V., Abrahamsen, N., and ern Portugal: Geophysical Research Letters, v. 12, fore-arc deposits in southern Peru and its significance
Cibin, U., 1998, Paleomagnetic evidence for Neogene p. 45–48, doi:10.1029/GL012i001p00045. for oroclinal bending in the central Andes: Tectonics,
tectonic rotations in the northern Apennines, Italy: Earth Perroud, H., Calza, F., and Khattach, D., 1991, Paleomag- v. 25, TC3010, doi:10.1029/2005TC001882.
and Planetary Science Letters, v. 154, p. 25–40, doi: netism of the Silurian volcanism at Almaden, south- Rosenbaum, G., and Lister, G.S., 2004, Neogene and Qua-
10.1016/S0012-821X(97)00183-0. ern Spain: Journal of Geophysical Research, v. 96, ternary rollback evolution of the Tyrrhenian Sea, the
Muttoni, G., Lanci, L., Argnani, A., Hirt, A.M., Cibin, U., p. 1949–1962, doi:10.1029/90JB02226. Apennines and the Sicilian Maghrebides: Tectonics,
Abrahamsen, N., and Lowrie, W., 2000, Paleomagnetic Prezzi, C., and Alonso, R., 2002, New paleomagnetic data v. 23, TC1013, doi:10.1029/2003TC001518.
evidence for a Neogene two-phase counterclockwise from the northern Argentine Puna: Central Andes ro- Royse, F., 1993, An overview of the geologic structure of
tectonic rotation in the Northern Apennines (Italy): tation pattern reanalyzed: Journal of Geophysical Re- the thrust belt in Wyoming, northern Utah, and eastern
Tectonophysics, v. 326, p. 241–253, doi:10.1016 search, v. 107, no. B2, doi:10.1029/2001JB000225. Idaho, in Snoke, A.W., Steidtmann, J.R., and Roberts,
/S0040-1951(00)00140-2. Protzman, G.M., and Mitra, G., 1990, Strain fabric associ- S.M. eds., Geology of Wyoming: Geological Survey of
Norabuena, E., Leffler-Griffin, L., Mao, A., Dixon, T., Stein, ated with the Meade thrust sheet: Implications for Wyoming Memoir 5, p. 272–311.
S., Sacks, S., Ocola, L., and Ellis, M., 1998, Space cross-section balancing: Journal of Structural Geology, Royse, F., Warner, M.A., and Reese, D.L., 1975, Thrust belt
geodetic observations of the Nazca–South American v. 12, p. 403–417, doi:10.1016/0191-8141(90)90030-3. structural geometry and related stratigraphic problems,
convergence across the Central Andes: Science, v. 279, Pubellier, M., and Ego, F., 2002, Anatomy of an escape tec- Wyoming-Idaho-northern Utah, in Bolyard, D.W.,
p. 358–362, doi:10.1126/science.279.5349.358. tonic zone: Western Irian Jaya (Indonesia): Tectonics, ed., Rocky Mountain Association of Geologists Sym-
Oncken, O., Hindle, D., Kley, J., Elger, K., Victor, P., and v. 21, no. 4, p. 1019, doi:10.1029/2001TC901038. posium on Deep Drilling Frontiers in Central Rocky
Schemmann, K., 2006, Deformation of the central Pueyo, E.L., Pares, J.M., Millán, H., and Pocoví, A., 2003, Mountains: Denver, Colorado, Rocky Mountain Asso-
Andean upper plate system—Facts, fiction, and con- Conical folds and apparent rotations in paleomagnetism ciation of Geologists, p. 41–45.
straints for plateau models, in the Andes, in Oncken, (a case study in the southern Pyrenees): Tectonophysics, Rubey, W.W., and Hubbert, M.K., 1959, Role of fluid pres-
O., et al., eds., Active Subduction Orogeny: Berlin, v. 362, p. 345–366, doi:10.1016/S0040-1951(02)00645-5. sure in mechanics of overthrust faulting: Part II. Over-
Springer, p. 3–28. Pueyo, E.L., Pocovi, A., Millan, H., and Sussman, A.J., thrust belt in geosynclinal area of western Wyoming in
Parés, J.M., and Van der Voo, R., 1992, Paleozoic paleomag- 2004, Map-view models for correcting and calculat- light of fluid pressure hypothesis: Geological Society
netism of Almaden, Spain: A cautionary note: Journal ing shortening estimates in rotated thrust fronts using of America Bulletin, v. 70, p. 115–206.
of Geophysical Research, v. 97, p. 9353–9356, doi: paleomagnetic data, in Sussman, A.J., and Weil, A.B., Rubio, F.J., Arenas, R., Martínez Catalán, J.R., Rodríguez
10.1029/91JB03073. eds., Orogenic Curvature: Integrating Paleomagnetic Fernández, L.R., and Wijbrans, J.R., 2013, Thickening
Parés, J.M., Van der Voo, R., Stamatakos, J.A., and Pérez- and Structural Analyses: Geological Society of Amer- and exhumation of the Variscan roots in the Iberian Cen-
Estaún, A., 1994, Remagnetization and postfolding oro- ica Special Paper 383, p. 57–72. tral System: Tectonothermal processes and 40Ar/39Ar
clinal rotations in the Cantabrian/Asturian Arc, northern Pysklywec, R.N., Beaumont, C., and Fullsack, P., 2000, ages: Tectonophysics, doi:10.1016/j.tecto.2012.10.005
Spain: Tectonics, v. 13, p. 1461–1471, doi:10.1029 Modeling the behaviour of the continental mantle (in press).
/94TC01871. lithosphere during plate convergence: Geology, v. 28, Ruffett, G., 1990, 40Ar-39Ar dating of the Beja Gabbro: Tim-
Paris, F., and Robardet, M., 1990, Early Paleozoic paleo- p. 655–658, doi:10.1130/0091-7613(2000)28<655: ing of the accretion of southern Portugal: Geophysi-
biogeography of the Variscan regions: Tectonophysics, MTBOTC>2.0.CO;2. cal Research Letters, v. 17, p. 2121–2124, doi:10.1029
v. 177, p. 193–213, doi:10.1016/0040-1951(90)90281-C. Quesada, C., 1991, Geological constraints on the Paleozoic /GL017i012p02121.
Pastor-Galán, D., Gutierréz-Alonso, G., and Weil, A.B., tectonic evolution of tectonostratigraphic terranes in Russo, R., and Silver, P.G., 1996, Cordillera formation,
2011, Orocline timing through joint analysis: Insights the Iberian Massif: Tectonophysics, v. 185, p. 225–245, mantle dynamics, and the Wilson cycle: Geology, v. 24,
from the Ibero-Armorican Arc: Tectonophysics, v. 507, doi:10.1016/0040-1951(91)90446-Y. p. 511–514, doi:10.1130/0091-7613(1996)024<0511:
no. 1–4, p. 31–46. Quesada, C., and Dallmeyer, R.D., 1994, Tectonothermal CFMDAT>2.3.CO;2.
Pastor-Galán, D., Gutiérrez-Alonso, G., Mulchrone, K.F., evolution of the Badajoz-Cordoba shear zone (SW Schellart, W.P., and Lister, G.S., 2004, Tectonic models
and Huerta, P., 2012a, Conical folding in the core of Iberia): Characteristics and 40Ar/ 39Ar mineral age for the formation of arc-shaped convergent zones and
an orocline. A geometric analysis from the Cantabrian constraints: Tectonophysics, v. 231, p. 195–213, doi: backarc basins, in Sussman, A.J., and Weil, A.B., eds.,
Arc (Variscan Belt of NW Iberia): Journal of Struc- 10.1016/0040-1951(94)90130-9. Orogenic Curvature: Integrating Paleomagnetic and
tural Geology, v. 39, p. 210–223, doi:10.1016/j.jsg Rankin, D.W., 1976, Appalachian salients and recesses: Late Structural Analyses: Geological Society of America
.2012.02.010. Precambrian continental breakup and the opening of Special Paper 383, p. 237–258.

662 Geological Society of America Bulletin, May/June 2013


Oroclines: Thick and thin

Schellart, W.P., Freeman, J., Stegman, D.R., Moresi, L., and Tartar, O., Poyraz, F., Gürsoy, H., Cakir, Z., Ergintav, S., nian carbonates from northern Spain: Journal of Geo-
May, D., 2007, Evolution and diversity of subduction Akpinar, Z., Koçbulet, F., Sezen, F., and Türk, T., physical Research, v. 107, B4, 2001JB000200.
zones controlled by slab width: Nature, v. 446, p. 308– Hastaoğ, A.Ö., Polat, A., Mesci, L., Gürsoy, Ö., Ercü- Weil, A.B., and Yonkee, W.A., 2009, Anisotropy of mag-
311, doi:10.1038/nature05615. ment, A., Çakmak, R., Belgen, A., and Yavaşoğlu, H., netic susceptibility in weakly deformed red beds from
Schneider, J., Bakker, R.J., Bechstädt, T., and Littke, R., 2012, Crustal deformation and kinematics of the eastern the Wyoming Salient, Sevier thrust belt: Relations to
2008, Fluid evolution during burial diagenesis and part of the North Anatolian fault zone (Turkey) from GPS layer-parallel shortening and orogenic curvature: Litho-
subsequent orogenetic uplift: The La Vid Group (Can- measurements: Tectonophysics, v. 518–521, p. 55–62. sphere, v. 1, p. 235–256, doi:10.1130/L42.1.
tabrian zone, northern Spain): Journal of Sedimentary Thomas, W.A., 1977, Evolution of Appalachian-Ouachita Weil, A.B., Van der Voo, R., van der Pluijm, B.A., and Parés,
Research, v. 78, p. 282–300, doi:10.2110/jsr.2008.032. salients and recesses from reentrants and promontories J.M., 2000, Unraveling the timing and geometric
Schwartz, S.Y., and Van der Voo, R., 1984, Paleomagnetic in the continental margin: American Journal of Science, characteristics of the Cantabria-Asturias arc (north-
study of thrust sheet rotation during foreland impinge- v. 277, p. 1233–1278, doi:10.2475/ajs.277.10.1233. ern Spain) through paleomagnetic analysis: Journal
ment in the Wyoming-Idaho Overthrust belt: Journal Thomas, W.A., and Whiting, B.M., 1995, The Alabama of Structural Geology, v. 22, p. 735–756, doi:10.1016
of Geophysical Research, v. 89, p. 10,071–10,086, Promontory; an example of the evolution of an Appa- /S0191-8141(99)00188-1.
doi:10.1029/JB089iB12p10077. lachian-Ouachita thrust belt recess at a promontory of Weil, A.B., Van der Voo, R., and van der Pluijm, B.A., 2001,
Shaw, J., and Johnston, S.T., 2012. The Carpathian-Balkan the rifted continental margin, in Hibbard, J.P., van Staal Oroclinal bending and evidence against the Pangea
bends: An oroclinal record of ongoing Arabian–Eur- C.R., and Cawood, P.A., eds., Current Perspectives in megashear: The Cantabria-Asturias arc (northern
asian collision, in Johnston, S.T., and Rosenbaum, G., the Appalachian-Caledonian Orogen: Geological Asso- Spain): Geology, v. 29, p. 991–994, doi:10.1130/0091
eds., Oroclines: Journal of the Virtual Explorer, v. 43, ciation of Canada Special Paper 41, p. 1–20. -7613(2001)029<0991:OBAEAT>2.0.CO;2.
paper 4, doi:10.3809/jvirtex.2012.00310. Trench, A., and Torsvik, T.H., 1992, The closure of the Iapetus Weil, A.B., Yonkee, W.A., and Sussman, A.J., 2010a, Recon-
Shaw, J., Johnston, S.T., Gutierrez-Alonso, G., and Weil, Ocean and Tornquist Sea: New palaeomagnetic con- structing the kinematic evolution of curved mountain
A.B., 2012, Oroclines of the Variscan orogen of Iberia: straints: Journal of the Geological Society of London, belts: A paleomagnetic study of Triassic redbeds from
Paleocurrent analysis and paleogeographic implica- v. 149, no. 6, p. 867–870, doi:10.1144/gsjgs.149.6.0867. the Wyoming Salient, Sevier thrust belt, U.S.A.: Geo-
tions: Earth and Planetary Science Letters, v. 329–330, Valle Aguado, B., Azevedo, M.R., Schaltegger, U., Martínez logical Society of America Bulletin, v. 122, p. 3–23,
p. 60–70, doi:10.1016/j.epsl.2012.02.014. Catalán, J.R., and Nolan, J., 2005, U-Pb zircon and doi:10.1130/B26483.1.
Silver, E.A., Reed, D.L., Tagudin, J.E., and Heil, D.J., 1990, monazite geochronology of the Variscan magmatism Weil, A.B., Gutiérrez-Alonso, G., and Conan, J., 2010b,
Implications of the north and south Panama thrust belts related to syn-convergence extension in central north- New time constraints on lithospheric-scale oroclinal
for the origin of the Panama orocline: Tectonics, v. 9, ern Portugal: Lithos, v. 82, p. 169–184, doi:10.1016 bending of the Ibero-Armorican arc: A palaeomagnetic
p. 261–281, doi:10.1029/TC009i002p00261. /j.lithos.2004.12.012. study of earliest Permian rocks from Iberia: Journal of
Solá, A.R., Pereira, E., Williams, I.S., Ribiero, M.L., Nieva, Van der Voo, R., 1982, Pre-Mesozoic paleomagnetism and the Geological Society of London, v. 167, p. 127–143,
A.M.R., Montero, P., Bea, F., and Zinger, T., 2008, New plate tectonics: Annual Review of Earth and Planetary doi:10.1144/0016-76492009-002.
insights from U-Pb zircon dating of Early Ordovician Sciences, v. 10, p. 191–220, doi:10.1146/annurev Weil, A.B., Gutiérrez-Alonso, G., and Wicks, D., 2012a, In-
magmatism on the northern Gondwanan margin: The .ea.10.050182.001203. vestigating the kinematics of local thrust sheet rotation
Urra Formation (SW Iberian Massif, Portugal): Tec- Van der Voo, R., 2004, Paleomagnetism, oroclines, and in the limb of an orocline: A paleomagnetic and struc-
tonophysics, v. 461, p. 114–129, doi:10.1016/j.tecto growth of the continental crust: GSA Today, v. 14, tural analysis of the Esla tectonic unit, Cantabrian-
.2008.01.011. no. 12, p. 4–9, doi:10.1130/1052-5173(2004)014<4: Asturian arc, NW Iberia: International Journal of Earth
Stamatakos, J.A., and Hirt, A.M., 1994, Paleomagnetic consid- POAGOT>2.0.CO;2. Sciences, p. 1–18, doi:10.1007/s00531-012-0790-3.
erations of the development of the Pennsylvania salient Van der Voo, R., and Channell, J.E.T., 1980, Paleomag- Weil, A., Gutiérrez-Alonso, G., Johnston, S.T., and Pastor-
in the central Appalachians: Tectonophysics, v. 231, netism in orogenic belts: Reviews of Geophysics Galán, D., 2012b, Kinematic constraints on buckling
p. 237–255, doi:10.1016/0040-1951(94)90037-X. and Space Physics, v. 18, p. 455–481, doi:10.1029 a lithospheric-scale orocline along the northern margin
Stegman, D.R., Freeman, J., Schellart, W.P., Moresi, L., and /RG018i002p00455. of Gondwana: A geologic synthesis: Tectonophysics,
May, D., 2006, Influence of trench width on subduc- Van der Voo, R., Stamatakos, J.A., and Pares, J.M., 1997, v. 582, p. 25–49, doi:10.1016/j.tecto.2012.10.006.
tion hinge retreat rates in 3-D models of slab rollback: Kinematic constraints on thrust-belt curvature from Winchester, J.A., Pharaoh, T.C., and Verniers, J., 2002,
Geochemistry Geophysics Geosystems, v. 7, Q03012, syndeformational magnetizations in the Lagos del Palaeozoic amalgamation of Central Europe: An intro-
doi:10.1029/2005GC001056. Valle Syncline in the Cantabrian Arc, Spain: Journal duction and synthesis of new results from recent geo-
Strecker, M.R., Cerveny, P., Bloom, A.L., and Malizia, D., of Geophysical Research–Solid Earth, v. 102, no. B5, logical and geophysical investigations, in Winchester,
1989, Late Cenozoic tectonism and landscape devel- p. 10,105–10,119, doi:10.1029/97JB00263. J.A., Pharaoh, T.C., and Verniers, J., eds., Palaeozoic
opment in the foreland of the Andes: Northern Sierras Van Staal, C.R., Dewey, J.F., MacNiocaill, C., and McKerrow, Amalgamation of Central Europe: Geological Society
Pampeanas, Argentina: Tectonics, v. 8, p. 517–534, doi: W.S., 1998, The Cambrian-Silurian tectonic evolution of London Special Publication 201, p. 1–18.
10.1029/TC008i003p00517. of the northern Appalachians and British Caledonides: Yonkee, W.A., 1997, Kinematics and mechanics of the Wil-
Suppe, J., 1984, Kinematics of arc-continent collision, flip- History of a complex, west and southwest Pacific-type lard thrust sheet, central part of the Sevier orogenic
ping of subduction, and back-arc spreading near Tai- segment of Iapetus, in Blundell, D.J., and Scott, A.C., wedge, north-central Utah: Brigham Young University
wan: Memoir of the Geological Society of China, v. 6, eds., Lyell: The Past Is the Key to the Present: Geological Studies, v. 42, no. part 1, p. 341–354.
p. 21–33. Society of London Special Publication 143, p. 199–242. Yonkee, W.A., and Weil, A.B., 2010a, Reconstructing the
Sussman, A.J., Butler, R.F., Dinares-Turell, J., and Verges, Weil, A.B., 2006, Kinematics of orocline tightening in the kinematics of curved mountain belts: Internal strain
J., 2004, Vertical-axis rotation of a foreland fold: An core of an arc: Paleomagnetic analysis of the Ponga patterns in the Wyoming Salient, Sevier thrust belt,
example from the southern Pyrenees: Earth and Plan- Unit, Cantabrian Arc, northern Spain: Tectonics, v. 25, U.S.A.: Geological Society of America Bulletin, v. 122,
etary Science Letters, v. 218, p. 435–449, doi:10.1016 TC3012, p. 1–23, doi:10.1029/2005TC001861. p. 24–49, doi:10.1130/B26484.1.
/S0012-821X(03)00644-7. Weil, A.B., and Sussman, A., 2004, Classification of curved Yonkee, W.A., and Weil, A.B., 2010b, Quantifying vertical-
Tait, J.A., Bachtadse, V., Franke, W., and Soffel, H.C., orogens based on the timing relationships between axis rotation in curved orogens: Integrating multiple data
1997, Geodynamic evolution of the European Variscan structural development and vertical-axis rotations, in sets with a refined weighted least-squares strike test:
fold belt: Palaeomagnetic and geological constraints: Sussman, A.J., and Weil, A.B., eds., Orogenic Curva- Tectonics, v. 29, TC3012, doi:10.1029/2008TC002312.
Geologische Rundschau, v. 86, p. 585–598, doi:10.1007 ture: Integrating Paleomagnetic and Structural Analy-
/s005310050165. ses: Geological Society of America Special Paper 383, SCIENCE EDITOR: J. BRENDAN MURPHY
Tapponnier, P., Peltzer, G., LeDain, A.Y., Armijo, R., and p. 1–17.
MANUSCRIPT RECEIVED 18 JULY 2012
Cobbold, P., 1982, Propagating extrusion tectonics in Weil, A.B., and Van der Voo, R., 2002, Insights into the REVISED MANUSCRIPT RECEIVED 30 NOVEMBER 2012
Asia: New insights from experiments with plasticine: mechanism for orogen related carbonate remagnetiza- MANUSCRIPT ACCEPTED 6 DECEMBER 2012
Geology, v. 10, p. 611–616, doi:10.1130/0091-7613 tion from growth of authigenic Fe-oxide: A scanning
(1982)10<611:PETIAN>2.0.CO;2. electron microscopy and rock magnetic study of Devo- Printed in the USA

Geological Society of America Bulletin, May/June 2013 663

View publication stats

You might also like