You are on page 1of 5

Downloaded from geology.gsapubs.

org on August 28, 2014

Geology

Active detachment faulting above the Peruvian flat slab


Brendan McNulty and Daniel Farber

Geology 2002;30;567-570
doi: 10.1130/0091-7613(2002)030<0567:ADFATP>2.0.CO;2

Email alerting services click www.gsapubs.org/cgi/alerts to receive free e-mail alerts when
new articles cite this article
Subscribe click www.gsapubs.org/subscriptions/ to subscribe to Geology
Permission request click http://www.geosociety.org/pubs/copyrt.htm#gsa to contact GSA
Copyright not claimed on content prepared wholly by U.S. government employees within scope of
their employment. Individual scientists are hereby granted permission, without fees or further
requests to GSA, to use a single figure, a single table, and/or a brief paragraph of text in
subsequent works and to make unlimited copies of items in GSA's journals for noncommercial
use in classrooms to further education and science. This file may not be posted to any Web site,
but authors may post the abstracts only of their articles on their own or their organization's Web
site providing the posting includes a reference to the article's full citation. GSA provides this and
other forums for the presentation of diverse opinions and positions by scientists worldwide,
regardless of their race, citizenship, gender, religion, or political viewpoint. Opinions presented in
this publication do not reflect official positions of the Society.

Notes

Geological Society of America


Downloaded from geology.gsapubs.org on August 28, 2014
Active detachment faulting above the Peruvian flat slab
Brendan McNulty Department of Earth Sciences, California State University Dominguez Hills, 1000 East Victoria Street,
Carson, California 90747, USA
Daniel Farber Lawrence Livermore National Laboratory, Institute of Geophysics and Planetary Physics, University of California,
7000 East Avenue, Livermore, California 94550, USA

ABSTRACT scarp fronting a $1-km-thick mylonite belt (Figs. 1 and 2). The foot-
The Cordillera Blanca detachment fault in the Peruvian Andes wall block, which forms the highest peaks in Peru (Fig. 1), is primarily
is, to our knowledge, the first active detachment to be documented constructed of the ca. 8 Ma (;3 kbar) Cordillera Blanca batholith
above a modern flat slab. Crustal detachment has unroofed the ca. (McNulty et al., 1998). The supradetachment basin (Callejon de Huay-
8 Ma Cordillera Blanca batholith, now the backbone of the highest las: Fig. 1) contains 7.8–4.65 Ma ignimbrites (Cobbing et al., 1981;
mountain range in Peru. Large-magnitude slip along the fault was Bonnot, 1984), glacial deposits, and mass-wasting–dominated alluvial
thermally enhanced by emplacement of the batholith, the penulti- fans shed off the rapidly elevating arc massif.
mate magmatic event prior to flattening of the Nazca slab. However, Kinematic indicators and mineral assemblages in the Cordillera
extensional models based on arc magmatism and crustal thickening Blanca mylonite document oblique (sinistral) down-to-the-west ductile
alone do not adequately explain the scale or structural asymmetry shear under conditions of the upper to lower greenschist facies. Stretch-
of a series of young, deep-seated, west-dipping normal faults across ing lineations in the mylonite and slip directions along brittle Pliocene–
Peru. Here we show that the onset of detachment faulting coincided Quaternary faults are oblique (trend, ;2208; pitch, ;708; Bonnot, 1984;
with subduction of the aseismic Nazca Ridge and consequent flat- this study). Kinematic indicators (S-C fabric, oblique quartz fabrics, mica
tening of the Nazca slab. We propose that slab buoyancy from ridge fish, asymmetric tails on feldspar porphyroclasts) perpendicular to foli-
subduction triggered extensional collapse of the prethickened con- ation and parallel to lineation invariably exhibit a normal-shear sense;
tinental crust, and that this buoyancy drove footwall uplift that ex- those perpendicular to foliation and lineation show a predominantly si-
ceeds basin subsidence. The west-dipping asymmetry of late Ceno- nistral shear sense. Mineral assemblages are quartz 1 biotite 1 mus-
zoic extensional faults in Peru may be controlled by a preexisting covite 6 actinolite 6 epidote 6 titanite; retrograde reactions include
crustal anisotropy (older thrusts), and/or formation of Riedel-like hornblende → actinolite 6 biotite, amphibole 6 biotite → chlorite, feld-
shears kinematically linked to the flat Nazca slab. spar → white mica. Upsection, the mylonite is progressively overprinted
by cataclasite and chlorite breccia, recording a retrograde ductile through
Keywords: Nazca Ridge, detachment, flat subduction, Peru, Andes, brittle history of deformation (Petford and Atherton, 1992). To the west,
extension. seismogenic, west-dipping normal faults cut Quaternary glacial moraines
and alluvial fans (Schwartz, 1988; Deverchere et al., 1989). The 10Be
INTRODUCTION and 26Al model ages of offset moraines yield vertical uplift rates of ;4
Detachment faults—regional low- to moderate-angle normal faults mm/yr (cf. .1 mm/yr long-term rate from exhumed batholith: DeSmedt
rooted in the ductile middle crust—are primary structures that allow et al., 1999). Figure 3 shows the fault to be listric, as inferred from
extension of Earth’s crust (Davis and Lister, 1988). Here we identify numerous back-rotated fault blocks in the hanging wall, and deep seated
the Cordillera Blanca detachment fault, an example of a seismogenic, (microseismicity to 10 km depth; Deverchere et al., 1989). The Cordil-
low-angle extensional shear zone in the Andean continental margin lera Blanca probably marks the juvenile phase of metamorphic core com-
(Fig. 1). We show that detachment faulting began just after the final plex formation (Chery, 2001).
phase of continental-arc magmatism in northern Peru, at a time coin- The following heat-flow data, temperature versus time data from
cident with subduction of the Nazca Ridge and flattening of the un- the batholith, and stratigraphic data from the Callejon de Huaylas basin
derthrusting Nazca slab. We propose that detachment faulting was trig- indicate that active detachment faulting began between ca. 8 Ma and
gered by buoyant forces associated with aseismic (hotspot) ridge ca. 5 Ma. (1) Present-day heat flow in the region is low (;20mW/m22;
subduction, superposed upon prethickened continental crust locally Henry and Pollack, 1988), indicating a low regional geothermal gra-
weakened by arc magmatism. dient in the latest Miocene. The lithospheric heat necessary to initiate
The Cordillera Blanca offers an ideal geological setting to study detachment faulting (Buck, 1991) thus had to have been derived from,
the evolution of detachment systems. Subduction and mountain building and localized to, the ca. 8 Ma Cordillera Blanca batholith. Moreover,
in Peru are ongoing, and both the lithospheric structure and late Miocene mylonite sequences are confined to the batholith. (2) The 40Ar/39Ar
to present plate kinematic framework are well defined. For example, slip muscovite ages show that the Cordillera Blanca mylonite cooled
vectors for the Nazca and South American plates have been quantified through ;350 8C by ca. 3.6 Ma (Petford and Atherton, 1992). The
over a length of geologic time, including those for today (via the Global simplest interpretation is that the rapid cooling was facilitated by uplift
Positioning System [GPS]: Norabuena et al., 1998) and for the late Ce- and/or exhumation along the fault prior to this time. (3) In the supra-
nozoic (NUVEL-1A: DeMets et al., 1994; Somoza, 1998). Other well- detachment basin, Pliocene strata above a ca. 5 Ma basal ignimbrite
defined parameters include the geometry, age, and structure of the sub- coarsen upward, from fine-grained sediment to fanglomerate, and thick-
ducting slab, downdip location of subducted hotspot ridges, crustal and en upward (tenfold increases in bed thickness; Bonnot, 1984), indicat-
lithospheric thicknesses of both plates, stress fields in the overriding ing substantial fault-generated relief by the earliest Pliocene.
plate, heat flow, and Andean topography (e.g., Suarez et al., 1983; James
and Sacks, 1999; Gutscher et al., 2000). OTHER NORMAL FAULTS
The Quiches fault, located northeast of the Cordillera Blanca (Fig.
CORDILLERA BLANCA DETACHMENT FAULT 3), is an active, N308W-trending, west-dipping normal fault with
The Cordillera Blanca fault has been studied previously (Wilson oblique sinistral offset (i.e., geometry and kinematics similar to Cor-
et al., 1967; Sebrier et al., 1988; Schwartz, 1988; Petford and Atherton, dillera Blanca detachment). In 1946, a major earthquake (M 5 6.3–
1992), but had not been recognized as a detachment fault. This 6.9; Doser, 1987) nucleated on the Quiches fault. The fault-plane so-
;N308W-striking fault is exposed as a spectacular 208–458W-dipping lution for the event, the largest well-documented normal-fault

q 2002 Geological Society of America. For permission to copy, contact Copyright Permissions, GSA, or editing@geosociety.org.
Geology; June 2002; v. 30; no. 6; p. 567–570; 4 figures. 567
Downloaded from geology.gsapubs.org on August 28, 2014

Figure 1. Composite Landsat Thematic Mapper image


(bands 1, 2, 3) of central Peru. Map at lower left shows lo-
cation of Landsat image (red inset box), location of Nazca
Ridge (thick lines) at 8, 5, and 0 Ma, extent of Peruvian flat
slab slip (thin lines), and vector (arrow) of Nazca plate (N).
B is snow-capped Cordillera Blanca batholith, CB is Cordil-
lera Blanca detachment fault, CH is Callejon de Huaylas ba-
sin, CN is Cordillera Negra, Q is Quiches fault, M is Maranon
fault, H is Huascaran (highest peak in Peru, 6768 m). Also
shown are locations of Figure 2 (A–C) photos, cross-
sectional line A-A9 (Fig. 3), extension directions at CB and
Q, and Global Positioning System data at Peruvian coast at
98309 (projected 15 km eastward onto image). Sinistral com-
ponent of these faults is attributed to oblique subduction
beneath Peru (McNulty et al., 1998).

earthquake in the Andes, defines a plane dipping 308 6 108W at a


hypocentral depth of 15–17 km (cf. 588W-dipping surface rupture).
Both the Cordillera Blanca and Quiches faults thus appear to be listric
and extend deep into the upper crust. Farther east, in the Maranon fold-
and-thrust belt, mylonitized Precambrian-Paleozoic rocks are exposed
in the footwall of a west-dipping normal fault (Maranon fault; Figs. 1
and 3). These rocks are interpreted to be a metamorphic core complex. Figure 2. Field photos of Cordillera Blanca detachment fault. A:
Mylonitic scarp; for scale, visible topographic relief is 1600 m;
Three similar, large-scale, west-dipping normal faults occur farther 408W-dipping mylonites can be seen in shaded canyon walls.
east, but the age of faulting in this remote region is unknown. B: View looking along strike of 208W-dipping mylonitic belt;
GPS data along the Peruvian coast (98309S) document an anom- road at bottom right for scale. C: Active frontal fault scarp with
alous motion of the upper plate toward (;2058) the Peru trench (Nor- down-dropped block in foreground; scarp is ~10–20 m high.
abuena et al., 1998; Fig. 1). A simple explanation is that westward slip
along Cordillera Blanca detachment extends under the coast (i.e., ex-
tension is not restricted to high Andes). The lateral extent, depth of tachment. The batholithic heat source, present from ca. 8 Ma to ca. 3 Ma,
faulting, oblique slip sense, and predominance of one dip direction of facilitated ductile extensional strain within the $1-km-thick mylonitic belt,
these normal faults (Fig. 3) suggest forcing factors other than the large cumulative slip along the fault, and rapid tectonic and erosional
spreading of orogenic relief (cf. Dalmayrac and Molnar, 1981; Suarez exhumation of the ;3 kbar batholith. It is evident that this fault has un-
et al., 1983; Sebrier et al., 1988; Mercier et al., 1992). dergone a longer period of displacement, and/or faster slip rates, than other
normal faults in the Peruvian Andes, most of which lack a heat source.
DRIVING FORCES FOR EXTENSION
Thermal Forcing: Batholith Emplacement Ridge Subduction, Slab Flattening, and Lithospheric Buoyancy
Thermal effects associated with emplacement of the Cordillera Blan- In Peru, active detachment faulting occurs above the modern flat
ca batholith played a key role in initiation of the Cordillera Blanca de- slab. Flat subduction here is primarily caused by the subducted part of

568 GEOLOGY, June 2002


Downloaded from geology.gsapubs.org on August 28, 2014

Figure 3. Cross-section A-A9 (no vertical exaggeration) showing west-dipping structural asymmetry across Peru (88S–108S)
from coast to Maranon fold-and-thrust belt (location, symbols, and abbreviations in Fig. 1; Quiches fault is projected into
section). Also shown are hypocenter and microseismicity locations (red ellipses—Quiches and Cordillera Blanca faults), in-
ferred brittle-ductile transition (BDT), Huancaspata (f6) and Huachay (f7) faults, and Eocene thrusts (green lines). Exhumed
metamorphic rocks are Precambrian (Complejo Maranon, P – C —purple waves) and Paleozoic-Mesozoic (Pz—orange).

the aseismic Nazca Ridge (Sacks, 1983; Gutscher et al., 2000). The
orientation of this hotspot track, which is ;2200 km long by ;200
km wide, is oblique (0408) to the vergence direction of the Nazca plate
(0828); the collision zone of the ridge with the Peru margin has thus
moved southward from 88S to 158S during the past 8 m.y. (Fig. 1,
inset). From plate reconstructions (DeMets et al., 1994; Somoza, 1998),
we estimate that the leading (southern) flank of the Nazca Ridge at
088S entered the Peru trench ca. 8 Ma (Figs. 1 and 4A). During this
same time period, the submerged Peruvian margin rapidly thinned by
subduction erosion (von Huene et al., 1996), whereas the calc-alkalic
Cordillera Blanca arc was active 300 km inboard of the trench. If we
assume a plate-margin geometry similar to that of the present and a
plate-convergence rate of 11 cm/yr (Somoza, 1998), the 308-dipping
slab would have brought the southward-migrating edge of the Nazca
Ridge to a depth of 100 km by ca. 6 Ma. The slab then began to pass
subhorizontally, bringing the aseismic ridge beneath the Cordillera
Blanca by ca. 5 Ma (Figs. 1 and 4B). This reconstruction suggests that
arrival of the Nazca Ridge, the leading edge of the flat slab, under the
Cordillera Blanca coincided with the cessation of arc magmatism and
onset of detachment faulting. The time-space correspondence between
subduction of the Nazca Ridge and detachment faulting suggests to us
a fundamental link between ridge subduction and crustal extension.
We propose that subduction of the ;17-km-thick Nazca Ridge
(Bialis et al., 2000) beneath the arc provided sufficient lithospheric
buoyancy to drive uplift in the overriding plate. Furthermore, we sug-
gest that this combined lithospheric buoyancy (thick subducted oceanic
crust plus thick Andean crust) drove footwall uplift in the Cordillera
Blanca that greatly exceeds hanging-wall subsidence. The high topo-
graphic relief of the Cordillera Blanca footwall and the low cumulative
Figure 4. Model for formation of detachments above Pe-
amount of sedimentation in the supradetachment Callejon de Huaylas
ruvian flat slab. Features drawn to scale: yellow—sub- basin both support a model of absolute footwall uplift (as opposed to
ducting Nazca slab (N), thick black line on top of slab— basin subsidence; Friedmann and Burbank, 1995). Buoyancy effects of
Nazca Ridge (NR), blue—South American lithosphere the subducted Nazca Ridge, which were superposed upon preexisting
(SA), magenta—asthenosphere, light pink—Cordillera
Blanca batholith (B). A: At 8 Ma, Cordillera Blanca arc is
thermal and gravitational conditions, were at a maximum ca. 3.5 Ma,
active above normal-dipping slab; Nazca Ridge enters Pe- when the entire ;300 km width (oblique section) of the ridge was
ruvian trench. B: At 5 Ma, leading flank of Nazca Ridge passing subhorizontally beneath the Cordillera Blanca (Fig. 4C).
passes under Cordillera Blanca (also see Fig. 1 inset),
calc-alkalic magmatism ceases, and detachment faulting
is underway. C: At 3.5 Ma, entire Nazca Ridge subducts Asymmetrical Extension Above the Flat Slab
under Peruvian Andes, driving tectonic and erosional ex- A question remains as to why the extensional structure in the
humation of batholith; up-to-east movement of footwall Peruvian Andes is dominated by large-scale, west-dipping normal
blocks is driven by buoyancy (due to ridge subduction)
and basal traction (due to flat slab). D: Riedel shear mod-
faults (e.g., Cordillera Blanca, Quiches, Maranon; Fig. 3). One possi-
el, Y—interplate shear, R—detachment; geometry favored bility is that this asymmetry is controlled by the reactivation of older
by preexisting thrust fault anisotropy. west-dipping thrusts (Fig. 3) formed during Andean shortening (Eocene

GEOLOGY, June 2002 569


Downloaded from geology.gsapubs.org on August 28, 2014

Incaic orogeny; Megard, 1984). We suggest an alternate, yet specula- Snelling, N.J., 1981, The geology of the western Cordillera of northern Peru: London,
Institute of Geological Sciences Overseas Memoir 5, 143 p.
tive, model where crustal detachments are kinematically linked to the Dalmayrac, B., and Molnar, P., 1981, Parallel thrusts and normal faulting in Peru and
subducting Nazca slab. Notably, fault geometry and kinematics relative constraints on the state of stress: Earth and Planetary Science Letters, v. 55,
p. 473–481.
to the flat slab resemble Riedel shear systems, where the sense of slip Davis, G.A., and Lister, G.A., 1988, Detachment faulting in continental extension: Per-
on detachments (R shears) is the same as that on the plate interface (Y spectives from the southwestern U.S. Cordillera, in Clark, S.P., et al., eds., Processes
shear), with an intervening angle of ;108–308 (Fig. 4D; also see Yin in continental lithospheric deformation: Geological Society of America Special Paper
218, p. 133–159.
and Kelty, 1991). The model is based on stress reorientation (west- DeMets, C., Gordon, R.G., Argus, D.F., and Stein, S., 1994, Effect of recent revisions to
plunging s1) associated with increased basal traction on the plate in- the geomagnetic reversal timescale on estimates of current plate motions: Geophys-
terface (Bird, 1998; Gutscher et al., 2000). An increase in interplate ical Research Letters, v. 21, p. 2191–2194.
DeSmedt, M., Farber, D.L., Hancock, G., Finkel, R., McNulty, B., and Torres, V., 1999,
shear might also relax normal compressional stress at the western edge Uplift and extension rates of the central Peruvian Andes: Slip rates along the Cor-
of the South American plate, thereby enabling extension. The model dillera Blanca detachment fault deduced from in-situ produced 10Be and 26Al cos-
mogenic radionuclide model ages: Eos (Transactions, American Geophysical Union),
also requires that middle to upper crustal strain be coupled to the flat v. 80, p. F1037.
slab through a strong lower lithosphere. This may be the situation in Deverchere, J., Dorbath, C., and Dorbath, L., 1989, Extension related to a high topography:
Peru, where flat-slab subduction has refrigerated the overriding litho- Results from a microseismic survey in the Andes of Peru and tectonic implications:
Geophysical Journal International, v. 98, p. 281–292.
sphere and increased its strength and seismicity (Suarez et al., 1983; Doser, D.I., 1987, The Ancash, Peru, earthquake of 1946 November 10: Evidence for low-
Henry and Pollack, 1988; Gutscher et al., 2000). angle normal faulting in the high Andes of northern Peru: Royal Astronomical So-
ciety Geophysical Journal, v. 91, p. 57–71.
SUMMARY AND IMPLICATIONS Friedmann, S.J., and Burbank, D.W., 1995, Rift basins and supradetachment basins: Intra-
continental end members: Basin Research, v. 7, p. 109–127.
This study provides one of the few examples of an active low- Gutscher, M., Spakman, W., Bijwaard, H., and Engdahl, E.R., 2000, Geodynamics of flat
angle detachment fault (Wernicke, 1995). The study also documents subduction: Seismicity and tomographic constraints from the Andean margin: Tec-
tonics, v. 19, p. 814–833.
detachment faulting in the Andean convergent margin, a tectonic re- Henry, S.G., and Pollack, H.N., 1988, Terrestrial heat flow above the Andean subduction
gime where these low-angle faults have not been widely recognized. zone in Bolivia and Peru: Journal of Geophysical Research, v. 93, p. 15 153–15 162.
Moreover, because it is widely held that flat-slab subduction drives James, D.E., and Sacks, I.S., 1999, Cenozoic formation of the central Andes: A geophysical
perspective, in Skinner, B.J., ed., Geology and ore deposits of the Central Andes:
contraction in the upper plate (e.g., Bird, 1998), our study presents an Society of Economic Geologists Special Publication 7, p. 1–25.
example of extensional strain above a modern-day flat slab. Our model McNulty, B.A., Farber, D.L., Wallace, G., and Lopez, R., 1998, Role of plate kinematics
predicts that detachment faults should occur elsewhere above the Pe- and plate-slip-vector partitioning in continental magmatic arcs: Evidence from the
Cordillera Blanca, Peru: Geology, v. 26, p. 827–830.
ruvian flat slab and that they should young southward in concert with Megard, F., 1984, The Andean orogenic period and its major structures in central and
southward migration of the Nazca Ridge. Similarly, we would expect northern Peru: Geological Society [London] Journal, v. 141, p. 893–900.
Mercier, J.L., Sebrier, M., Lavenu, A., Cabrera, J., Bellier, O., Dumont, J.-L., and Machare,
that subduction of the ‘‘Lost Inca plateau’’ to the north (Gutscher et J., 1992, Changes in the tectonic regime above a subduction zone of Andean type:
al., 2000) would have generated crustal extension in that region. The Andes of Peru and Bolivia during the Pliocene-Pleistocene: Journal of Geo-
Our consideration differs from previous studies in that it explores physical Research, v. 97, p. 11 945–11 982.
Norabuena, E., Leffler-Griffin, L., Mao, A., Dixon, T., Stein, S., Sacks, I.S., Ocola, L., and
a plate-kinematic–based explanation for detachment faulting. We pro- Ellis, M., 1998, Space geodetic observations of Nazca–South America convergence
pose that buoyancy associated with the subduction of aseismic (hot- across the central Andes: Science, v. 279, p. 358–362.
spot) ridges may drive extension in the overriding plate, and that these Petford, N., and Atherton, M.P., 1992, Granitoid emplacement and deformation along a
major crustal lineament: The Cordillera Blanca, Peru: Tectonophysics, v. 205,
regional effects augment more localized effects from magmatism and p. 171–185.
thickened continental crust. The example of detachment tectonics in Sacks, I.S., 1983, The subduction of young lithosphere: Journal of Geophysical Research,
v. 88, p. 3355–3366.
Peru may provide a perspective for development of post-Laramide Schwartz, D.P., 1988, Paleoseismicity and neotectonics of the Cordillera Blanca fault zone,
metamorphic core complexes in western North America, where sub- northern Peruvian Andes: Journal of Geophysical Research, v. 93, p. 4712–4730.
sequent Basin and Range faulting, isostatic-flexural rebound, and for- Sebrier, M., Mercier, J.L., Machare, J., Bonnot, D., Cabrera, J., and Blanc, J.L., 1988, The
state of stress in an overriding plate situated above a flat slab: The Andes of central
mation of the Pacific–North American transform plate boundary often Peru: Tectonics, v. 7, p. 895–928.
mask the driving forces for detachment faulting. Somoza, R., 1998, Updated Nazca (Farallon)–South America relative plate motions during
the last 40 My: Implications for mountain building in the central Andean region:
ACKNOWLEDGMENTS Journal of South American Earth Sciences, v. 11, p. 211–215.
We thank Greg Davis and Dave Scholl for reviews, and Jim Kellogg, Mian Liu, and Suarez, G., Molnar, P., and Burchfiel, B.C., 1983, Seismicity, fault plane solutions, depth
An Yin for comments. This work was supported by the National Science Foundation (grant of faulting, and active tectonics of the Andes of Peru, Ecuador, and southern Colum-
EAR-9802825), Institute of Geophysics and Planetary Physics at Lawrence Livermore, and bia: Journal of Geophysical Research, v. 88, p. 10 403–10 428.
Instituto de Geológico Minero y Metalurgico in Lima, Peru. von Huene, R., Pecher, I.A., and Gutscher, M.-A., 1996, Development of the accretionary
prism along Peru and material flux after subduction of the Nazca Ridge: Tectonics,
v. 15, p. 19–33.
REFERENCES CITED Wernicke, B., 1995, Low-angle normal faults and seismicity: A review: Journal of Geo-
Bialas, J., Kukowski, N., and GEOPECO Scientific Team, 2000, FS SONNE Cruise Report
physical Research, v. 100, p. 20 159–20 174.
SO146/1 and 2, GEOPECO (Geophysical experiments at the Peruvian Continental
Wilson, J.J., Reyes, L., and Garayer, J., 1967, Geologia de los cuadrangulos de Mollebam-
Margin: Investigations of tectonics, mechanics, gas hydrates, and fluid transport):
ba, Tayabamba, Huaylas, Pomabamba, Carhuaz y Huari: Boletı ´n del Servicio Geo-
Kiel, Germany, University of Kiel, GEOMAR report 96, 490 p.
lógico y Minero de Lima, v. 16, p. 95.
Bird, P., 1998, Kinematic history of the Laramide orogeny in latitudes 35–498N, western
Yin, A., and Kelty, T.K., 1991, Development of normal faults during emplacement of a
United States: Tectonics, v. 17, p. 780–801.
thrust sheet: An example from the Lewis allochthon, Glacier National Park, Montana
Bonnot, D., 1984, Neotectonique et tectonique active de la Cordillere Blanche et du Cal-
(U.S.A.): Journal of Structural Geology, v. 13, p. 37–47.
lejon de Huaylas, Andes Nord-Peruviennes [Ph.D. thesis]: Orsay, University of Paris,
130 p.
Buck, R., 1991, Modes of continental lithospheric extension: Journal of Geophysical Re- Manuscript received September 24, 2001
search, v. 96, p. 20 161–20 178. Revised manuscript received February 20, 2002
Chery, J., 2001, Core complex mechanics: From the Gulf of Corinth to the Snake Range: Manuscript accepted February 26, 2002
Geology, v. 29, p. 439–442.
Cobbing, E.J., Pitcher, W.S., Wilson, J.J., Baldock, J.W., Taylor, W.P., McCourt, W.J., and Printed in USA

570 GEOLOGY, June 2002

You might also like