Sunteți pe pagina 1din 12

Downloaded from gsabulletin.gsapubs.

org on May 7, 2012

Geological Society of America Bulletin


Cordilleran deformation along the eastern edge of the VallesSan Luis Potos
carbonate platform, Sierra Madre Oriental fold-thrust belt, east-central Mexico
MAX SUTER
Geological Society of America Bulletin 1984;95, no. 12;1387-1397
doi: 10.1130/0016-7606(1984)95<1387:CDATEE>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 Geological Society of


America Bulletin

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 gsabulletin.gsapubs.org on May 7, 2012

Cordilleran deformation along the eastern edge


of the Valles-San Luis Potos carbonate platform,
Sierra Madre Oriental fold-thrust belt, east-central Mexico
MAX SUTER* Instituto de Geologa, Universidad Nacional Autnoma, de Mxico, Ciudad Universitaria, Delegacin Coyoacn, 04150 Mxico, D.F.

ABSTRACT
The eastern margin of the Cretaceous
Valles-San Luis Potosi carbonate platform
(Hidalgo, Queretaro, and San Luis Potosi
States) became activated during the formation of the Sierra Madre Oriental fold-thrust
belt. A series of thrust sheets, with a minimum linear shortening of 3 to 4 km each and
10 to 12 km in total, formed where the platform edge was approximately normal to the
greatest principal stress trajectories, as deduced from fold axes. A tear fault developed
where the platform edge formed an acute
angle with the same stress trajectories. The
overthrusts cut across the competent layers of
the Lower Cretaceous on tectonic ramps that
dip >10 and became partly steepened by
subsequent imbrication or folding. In the mechanically weak Upper Cretaceous rocks, the
thrust faults are nearly parallel to bedding for
several kilometres. The location of the thrusts
is controlled mainly by the lithological and
thickness change of the middle Cretaceous
rocks from an orthotropically layered basin
facies into a two to five times thicker homogeneous platform-edge assemblage. The bank
margin was a zone of stress concentration.
The magnitude of the horizontal tectonic
components which caused the deformation
may have changed by a factor of two to five
across the platform margin, due to the change
in cross-sectional area. Oblique and layerparallel discrete shears and subordinate tectonic stylolites appear to have been the
dominant deformational mechanisms. The
rocks did not suffer any measurable penetrative ductile deformation; ooids present at the
base of the Xilitla Thrust are unflattened but
are marked by a closely spaced stylolitic
cleavage perpendicular to bedding. The de-

*Present address: Schlumberger Offshore Services,


Interpretation Development, Apartado 53-937,11300
Mxico, D.F.

formations are bracketed by the paleontological age of the youngest strata affected by the
overthrusts (Globotruncana contusa planktonic foraminiferal zone) and by the isotopic
age of a post-tectonic pluton (62.2 Ma) and
are thus of late Maastrichtian/Paleocene age.
INTRODUCTION
The Sierra Madre Oriental is a physiographic
province that is tectonically made up of the outermost outcropping segment of the Cordilleran
fold belt (King, 1969) in central Mexico. The
orogenic belt trends north-northwest and is
bounded on the north by the east-trending
Monterrey-Torren transverse zone (de Cserna,
1956) and on the east by the flysch-filled
Tampico-Misantla foredeep (Busch and Gavela,
1978). It becomes covered by the late MioceneQuaternary Trans-Mexican Neovolcanic belt
(Thorpe, 1977; Demant, 1978; Nixon, 1982) on
the south, and by the basin-range-type, faultbounded, and debris-filled basins of the Central
Mesa and the ignimbrite sheets of the Sierra
Madre Occidental (McDowell and Clabaugh,
1979) on the west.
A structural traverse across the Sierra Madre
Oriental fold-thrust belt at approximately latitude 21N in southern San Luis Potos, northern
Hidalgo, and eastern Quertaro States was made
in a research study by the Mexican National
University. The research revealed that the style
of deformation is principally controlled by the
lithology and thickness of the middle Cretaceous
rocks. Large overthrusts developed along the
edges of massive carbonate platforms, whereas
the synchronal basin facies and the interior of
the platforms are characterized by chiefly upright, open to closed flexural-slip folds (Suter,
1983).
This paper describes the Cordilleran structures on the eastern edge of the Valles-San Luis
Potos (VSP) Platform over a distance of 70 km
between Huichihuayn and Cerro del Aguila
(Fig. 1). The over-all geometry of this major

Geological Society of America Bulletin, v. 95, p. 1387-1397,12 figs., December 1984.


1387

carbonate bank was outlined by Carrillo-Bravo


(1971). The sedimentation and diagenesis of the
platform and slope carbonates were studied
by Griffith and others (1969), Enos (1974),
Carrasco (1977), and Minero (1983a, 1983b).
Problems in interpretation of the depositional
environment of the bank foreslopes have concentrated mainly on the question of whether the
great difference in relief and the common faulting at the bank edges are structural or depositional (Wilson, 1975). Enos (1974) assumed the
prominent escarpments on the eastern edge of
the VSP and El Doctor Platforms to reflect depositional topography but was also aware that
later folds and faults parallel with the bank margins obscured the transition from platform to
basinal rocks. Systematic detailed structural
mapping of the Cordilleran deformation along
the platform margin was undertaken to quantify
the very little-known structural geometry and
deformational style of the Sierra Madre Oriental
fold-thrust belt and to determine whether the
vertical juxtaposition of facies belts of the
platform-basin transition zone is due to progradation of the platform or to imbricate overthrusting from the platform toward the basin.
STRATIGRAPHY
Lithology of the shallow-water VSP platform
consists of an approximately 1,500- to 2,000-mthick pile of carbonates and evaporites (El Abra
and Guaxcamd Formations). The bank-edge
facies belts (also referred to as the "Taninul
facies") are slightly outbuilding (Fig. 2 and see
Fig. 9 below) and include rudist boundstone
and nonparallel, discontinuously bedded bioclastic lime grainstone-rudstone. Dolomitization obscures their distribution in the Xilitla
Mountains (Fig. 3). The deposits of the platform
interior are mainly well-bedded micritic limestones (also known as "El Abra facies"; for the
facies belts and microfacies types of the VSP
Platform, the reader is referred to the review by
Wilson, 1975) and contain evaporites (Guax-

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


1388

M. SUTER

2T30"

21'00

as "Tamaulipas Superior" and "Tamaulipas


Inferior," respectively. None of these 1; thostratigraphic units has ever been formally introduced, however. By contrast with tie bank
facies, the basinal facies is composed cf formations of similar thicknesses but alternating competences (Fig. 2).
The foreslope deposits are made up of sedimentary breccia and bioclastic lime packstone.
They vary laterally in thickness and bsisinward
extension. (1) In the Xilitla area (Fig. 3), the
foreslope deposits are mapped as an individual
unit (Tamabra Formation), which constitutes a
15-km-wide wedge that measures at leaiit 600 m
at the platform margin and thins to 20C m near
Xilitla. The Tamabra Formation grades into pelagic limestone of Ahuacatlan Formation lithology to the east (Fig. 2). The Tamabra Formation
near Xilitla is typically composed of bioclastic
rocks in the uppermost 50 m, whereas the lower
part consists largely of platform-derived exoclasts in a matrix of lime mudstone-wac.iestone.
The contact with the underlying Ahuacatlan
Formation is defined by morphological contrast
(see Fig. 5 below) rather than by abundance of
exoclasts. The great extension and thickness of
the Tamabra Formation near Xilitla can be explained by the re-entrant geometry of the shelf
edge (Enos, 1983) that changes from north to
northeast orientation. (2) To the south of the
Tancuilin River (Fig. 3), the foreslope deposits
form a much narrower facies belt and, depending on their lithology, are mapped as pait of the
Ahuacatlan or El Abra Formations. Along
Federal Highway 85 (to the north of Cuesta
Colorada) (Fig. 3), for example, the rudist
boundstone is absent at the bank margin, and
the foreslope realm measures only a few
hundred metres.
STRUCTURE
Tlamaya-Huichihuayin Lineament

20"45'
99"00'

Figure 1. Outline map showing the location and generalized structure of the study area.
Dotted: Cretaceous Valles-San Luis Potosi carbonate platform. Framed: Figure 3.

cama Formation). The bank forms a thick, isolated, competent layer that in the study area is
sandwiched by thinner, incompetent layers
composed of argillaceous limestone and shale
(Soyatal and Pimienta Formations) (Fig. 2).
The basinal facies to the east of the platform
measures 300 to 1,200 m and is composed
of parallel-bedded, 30-cm- to 2-m-thick lime
mudstone-wackesione strata with shale partings
and intercalated chert and bentonite layers. Near
the toe of slope, platform-derived, graded,

bioclastic-oolitic-lithoclastic packstones (see Fig.


6 below) alternate with the former lithology.
The basin facies was divided by Bodenlos
(1956a) and by Bonet (1956) into the Ahuacatldn and Chapulhuacan Formations. The Ahuacatl&n Formation is thinner bedded and has
lower relief than the underlying Chapulhuacan
Formation (see Fig. 5 below) in the area of
Highway 120, where the former contains
numerous bentonite layers. The Ahuacatlan and
Chapulhuacan Formations are also referred to

In the northeastern part of the study area, the


eastern edge of the VSP Platform passes below
the surface and reaches the front of the Sierra
Madre Oriental to the southwest of Huichihuayan (Figs. 1 and 3). It then crosses the outermost part of the fold-thrust belt in a N45Etrending lineament. The segment to the northwest dips - 1 5 against the lineament, and its
topographic relief is 300 to 400 m higher than
that of the flat-lying counterpart to the southeast
(foreslope facies).
The lineament very likely was activated as a
strike-slip fault during the formation af the
Huayacocotla Anticline and the San Juanito
and Xilitla Thrusts (Figs. 1 and 3). North of the
lineament, the belt is bounded on the east by the

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


CORDILLERAN DEFORMATION, MEXICO
platform

platform

interior

edge

San Juanito Thrust, which brings El Abra limestone in tectonic contact with Tanlajs flysch of
the foreland near San Juanito (Bonet, 1956) and
with Mndez beds at the Huichihuayn-San
Pedro trail (Fig. 3). South of the lineament, the
shortening is taken up in a box fold (Huayacocotla Anticline), and there is no frontal thrust. The
east limb of the fold is located ~ 1,200 m west of
the front of the San Juanito thrust plate. The
more internally located Xilitla Thrust (Fig. 1)
has a linear shortening of ~10 km and terminates laterally as it reaches the projected western
continuation of the lineament. Across the lineament, the Xilitla Thrust is replaced by folds,
which accounts for only a fraction of the thrust
shortening.
This arrangement of structures with different
geometries and shortening amounts on either
side of the lineament implies its activation as a
tear fault. Dextral relative movement along the
frontal part of the fault presumably equals approximately the unknown magnitude of the
shortening of the San Juanito Thrust to the west
of Huichihuayn. The existence of a dextral
transcurrent fault is also supported by the orientation of the lineament with reference to the
regional compressive stress field causing the
Cordilleran deformations, as deduced from fold
axes (Fig. 3). Although the Cordilleran activation of the lineament is compatible with the
structural evidence on a kilometre scale, it could
not be substantiated by smaller-scale features,
due to karst morphology, tropical vegetation,
and the absence of marker horizons.
Xilitla Thrust
The Xilitla Thrust, first recognized by Heim
(1940), is situated where the orientation of the
VSP bank edge changes gradually from a N45E

foreslope

basin

direction in the Tlamaya area to a north-south


direction south of Federal Highway 120 (Fig. 3).
The thrust forms an angle o f - 6 0 with the platform edge to the northwest. It does not affect the
edge proper, but runs within the foreslope and
basin sediments, which makes it possible to distinguish structural relief from depositional relief.
The geometry of the thrust is well exposed
normal to its trend on the northern side of Arroyo Seco valley (see Fig. 5 below). The mean
inclination of the thrust fault is 7.5 (6.5 in the
eastern part, 9 in the western part). The San
Felipe beds of the footwall are not transected by
the fault plane (Fig. 4a), and they remain at
constant thickness between the innermost and
the outermost outcrop. The mean angle between
the exposed beds of the upper plate and the fault
is <*i = 8 for the interval Chapulhuacn Formation to Tamabra Formation and a^ <2.5 for
the Pimienta Formation. The corresponding
rocks in the lower plate must be truncated in the
subsurface to the west at similar angles. The Xilitla Thrust therefore probably includes a ramp
segment with an inclination cq that connects the
outcropping layer-parallel fault segment with a
lower dcollement running in the Pimienta beds.
Other evidence for the ramp geometry in the
lower plate is the exposed fold that is present
only in the upper plate and does not continue in
the footwall (Figs. 4a and 5). The fold has an
angle 7 = 85.5 between its axial plane and
bedding (Fig. 4a). This rootless ramp fold is
caused by the displacement of the thrust sheet
over the unexposed tectonic ramp (Rich, 1934).
This is in agreement with the geometric relation between the cutoff angle <*i and the axial
angle 7 for a single step in dcollement, cq =
arctan(sin 2 7 / I + 2cos27), as deduced by Suppe
and Namson (1979, equation 4; Suppe, 1983,
equation 12). The rotational component caused

1389

Figure 2. Sketch of the platformbasin transition zone on the eastern


edge of the Cretaceous Valles-San
Luis Potos Bank; after Wilson (1975),
modified. 1 = Mndez Formation. 2 =
Soyatal Formation. 3 = San Felipe
Formation. 4 = Agua Nueva Formation. 5 = Ahuacatln Formation. 6 =
Chapulhuacn Formation. 7 = Tamabra Formation. 8 = El Abra Formation,
platform-edge member (also referred
to as "Taninul fades"). 9 = El Abra
Formation, platform-interior member
(also referred to as "El Abra fades").
10 = Guaxcam Formation (evaporites
of the platform interior). 11 = Pimienta
Formation. 12 = Tamn Formation. 13
= Santiago Formation. Lithology description provided in the text.

by frictional drag (Berger and Johnson, 1980) is


zero for the upper and lower plates of the Xilitla
Thrust. A value for the ramp height h (Fig. 4a) is
given by the greatest measurable structural relief
of the fault, which is 1,050 m. The lower limit
Ax of the horizontal slip component is calculated by measuring the distance between the erosional front of the upper plate east of Miramar
and the innermost outcrop of the footwall near
Xilitlilla. This value calculated normal to the
structural trend is 3,600 m. It has to be taken as
a minimum amount, because the distance the
thrust originally traveled horizontally in the San
Felipe beds is not known, and the shortening
above the San Felipe Formation and below the
Pimienta Formation has not been considered.
The strata of the hanging wall to the west of
the innermost outcrop of the thrust fault do not
necessarily parallel the inclination of the faultramp segment. These rocks are at the platform
edge, and so they may not have been deposited
horizontally. Another possibility is that the incompetent Pimienta beds of the upper plate
might have thickened tectonically in the lower
ramp area, in a fashion similar to that of smallscale overthrust faults described by Serra (1977,
1978) and as shown from geometrical models of
Jura ramp anticlines (Suter, 1981).
The thrust displacement tapers toward the
south and terminates in the Tancuilin valley at a
steep dip. The shortening increase between the
Tancuilin River and the Arroyo Seco section
(Fig. 4a) must be gradual; no strike-slip fault can
be recognized where a discontinuous shortening
change would be possible. Only a minor amount
of the thrust displacement can be transferred to
the fold that develops in the upper plate, where
the thrust shortening terminates (Fig. 3). Most of
the shortening is probably transferred to the

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


1390

M. SUTER

Figure 3. Tectonic map of the study area.


Note that the Lower Cretaceous formations
are lateral equivalents. For a more detailed
map of the northeastern part, see Figure 2 in
Suter (1980).

folds that develop east of Xilitlilla (Fig. 3) in


front of the Xilitla Thrust.
Deformation Mechanism. Small-scale features can be studied in the lowermost part of the
upper plate, where the Xilitla Thrust is crossed
by Federal Highway 120. The well-beddei limestones were deformed under a minimum of
1,500 m of cover. They show extension
microfractures normal to bedding and, less
frequently, concordant with bedding (Fig. 6a).
The microfractures are filled with eutiedral,
equigranular to bladed calcite crystals. The veins
normal to bedding often begin on bidding
planes and become gradually thinner toward the
middle of the layers. Tensile stresses must have
developed locally during the movement of the
thrust plate, causing these extensional cracks, although the regional stress field was compressive.
The extension microfractures could have formed
when the upper-plate rocks passed from the
ramp through the upper hinge onto the Upper
Cretaceous bedding-plane segment. Shear fractures, on the other hand, are less frequent than
extension fractures, which can be explained by
the low tensile strength of brittle rocks, by the
relatively low overburden, or by high por;-fluid
pressures. Ooids, present close to the thrust fault,
make excellent strain markers and are pervasively diagenetically deformed at grain tc grain
contacts (Fig. 6a) but are unflattened. They are
affected by a closely spaced stylolitic ckavage
(Fig. 6b). The stylolite seams generally follow
grain boundaries and only rarely pass through
grains (Fig. 6a). The random orientation of stylolites in Figure 6b can be caused by the grain
array or by progressive rotation during movement along the thrust fault. The pressuresolution surfaces are likely to be the source for
the material precipitated in the extension veins
described above (Alvarez and others, 1976;
Fletcher and Pollard, 1981).
Thrust Sheets between Federal Highway 120
and the Amajac River
Cordilleran deformations to the south o:' Federal Highway 120 are characterized by a series
of thrust plates (Fig. 7) that have a slightly
oblique trend to the edge of the VSP carbonate
bank between the highway and La Misin and a
parallel trend to the bank edge between Li Misin and the Amajac River (Fig. 3).
The geometry of the Lobo-Cinega Thrust
(Fig. 1) is well exposed along Federal Highway

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


^ ^

Pea de S a n A n t o n i o

Xilitla

Xalcuayo

A'

2000

1000

Om

C e r r o de S a n A g u s t n

g '

2000

Om

Cerro

Macangui

2000

1000

Om
O i -

M n d e z and
Soyatal Formations

V V

I J L I

San Felipe F o r m a t i o n

A g u a Nueva

Formation

l y r
aTA
r-rr

-t 2 k m
El A b r a F o r m a t i o n
platform-interior member

J,I A h u a c a t l n F o r m a t i o n

El A b r a F o r m a t i o n
platform-edge member

Chapulhuacn F o r m a t i o n

Tamabra Formation

Pimienta

Formation

Figure 4. Sections across the eastern margin of the Valles-San Luis Potos carbonate bank. The traces are marked on Figure 3. 1 = Xilitla
Thrust. 2 = Misin Thrust. 3 = Lobo-Cinega Thrust. 4 = Agua Zarca Thrust. Symbols used for the geometry description of the Xilitla Thrust:
a = cutoff angles; y = axial angle; h = ramp height; Ax = outcropping length of the layer-parallel fault segment.

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


1392

M. SUTER

Figure 5. Xilitla Thrust and ramp anticline in the Arroyo Seco valley; view from the south. 1 = Chapulhuac&i Formation. 2 = Ahuacatl&n
Formation. 3 = Tamabra Formation. 4 = El Abra Formation, platform-edge member. 5 = Agua Nueva Formation. 6 = San Felipe Formation. * = Highway 120. The fault parallels beds in the lower plate, whereas bedding in the upper plate is truncated at an angle of 8 in the interval
Tamabra-Ahuacatl&i Formations and at an angle of <2.5 below. The upper-plate layers form a fold with an angle of 85.5 between the axial
plane and bedding. The fold is probably caused by the transport of the hanging wall over a tectonic ramp, the inclination of which is given by the
cutoff angle of the frontal part of the thrust sheet. The visible horizontal displacement component measures 3,600 m.

120, where a near-surface dcollement developed in the Soyatal Formation. A layer-parallel


thrust segment, with a mean dip of 2 and a
length of 3,500 m, is connected with a steeperdipping thrust segment by a hinge (Figs. 4b and
8). The steep-dipping segment between the
highway and the Moctezuma River has a mean
inclination of 30, exposed over an altitude difference of nearly 1,000 m. The angle between
the fault and both hanging wall and footwall
bedding is 7 (Figs. 4b and 9). There is a partial
rotation, subsequent to thrusting, of both the
upper and the lower plates. This is in contrast to
the Xilitla Thrust, where upper-plate rotation is
caused primarily by a stepped geometry of the

Figure 6. Small-scale deformations in Lime


packstone-grainstone, with concavo-convex
and sutured grain boundaries, located close
to the Xilitla thrust fault in the upper plate of
the thrust, a. The visible diagenetic indentation is approximately vertically oriented (frpendicular to bedding), as are an observable
tectonic stylolite and an extension crack. The
stylolite follows grain boundaries but cuts
across a peloid in the center-left part of the
figure, b. The components are in other places
entirely framed by a dense net of tectonic
stylolites, whereas no penetrative ductile deformation can be observed (the cryptocrystalline ooid in the upper left part of the figure is
still circular in shape). The irregular array of
the stylolites can be explained by progressive
rotation during transport of the thrust sheet.

fault. The trace of the thrust is difficult to pinpoint north of Federal Highway 120, where the
fault plane has a dip component toward the
north. Here, the thrust surface crosses the El
Abra-Soyatal contact and passes into bedding
within the El Abra Formation. A thrust fault
mapped in the western flank of the; Cerro
Grande Anticline (Suter, 1980, Figs. 2 and 3)
could be the northern continuation of the LoboCinega Thrust.
Between the Moctezuma River and Federal
Highway 85 (Figs. 3 and 4c), the Lobo-Cinega
Thrust has a throw of 1,300 m, dips 26W,
and is subparallel to bedding. Given the shear
angle of 7 and the vertical displacement
component of 1,300 m, a linear shortening of
10,590 m results. The fault was mapped by
Bodenlos (1956b) where it is crossed by Federal
Highway 85.
In the Upper Cretaceous rocks, twe repetitions of platform-edge fades can be found (Fig.
10). The lower one follows the Lobo-Cinega
thrust fault in the El Lobo half-window (Fig. 3)
along the entire outcropping layer-parallel
thrust segment. It is not clear whether the repetition is caused by migration of the facies belts, or
whether it is of tectonic nature, which would
increase the shortening considerably. Microfossils sampled directly below (SM 74 and 75) and
above (SM 76 to 78) the repetition of platform
sediments inside the Soyatal Formation along
Highway 120 (Fig. 10, section A) belong to the
Globotruncana elevata planktonic foraminiferal
zone s.s. (early Campanian), but only SM 75
contains the zone index fossil.

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


CORDILLERAN DEFORMATION, MEXICO

Pena P r i e t a

Cerro Macangu Moctezuma canyon

1393

C e r r o de la Yesca W

Figure 7. Stack of imbrcate thrust slices on the eastern edge of the Valles-San Luis Potos Platform, view from the El Lobo-Agua Zarca
road, looking south. Agua Zarca is located in the saddle east of the Agua Zarca Thrust, and the visible road connects Agua Zarca with El Lobo
and Highway 120. A = Agua Zarca Thrust. B = Lobo-Cinega Thrust. C = Misin Thrust. 1 = Ahuacatln Formation. 2 =Tamabra Formation. 3 = El Abra Formation, platform-edge member. 4 = Agua Nueva Formation. 5 = San Felipe Formation. 6 = Mndez Formation.
The outcrops to the south of Federal Highway 120 reveal two more thrusts (Figs. 3 and 7)
on the edge of the VSP Platform: (1) the more
internal and steeper Misin Thrust (Fig. 1),
which is characterized by a linear shortening of
at least 1,300 m, a mean dip of 41, and a mean
angle of the thrust fault of 6 with regard to
bedding of the upper plate (Fig. 4b); its throw is
indeterminate, because no correlative strata crop
out on either side of the thrust, due to the facies
change of the middle Cretaceous; and (2) the
Agua Zarca Thrust (Fig. 1), which has, between
the Moctezuma River and Federal Highway 85, a
structural relief of somewhat more than 1,000 m
(Fig. 4c), dips ~30W, and also forms a low
angle with bedding.

W C e r r o de la Yesca

Farther south, the Amajac canyon exposes


outcrops with a relief of as much as 2,000 m.
The Lobo-Cinega and Misin thrust sheets that
follow the platform edge dip subvertically but
still have a shallow angle with bedding (Figs. 11
and 12). They are located in the western limb of
a major anticlinorium (Fig. 3), which explains
the steep inclination of the thrust faults. The
Lobo-Cinega Thrust has been mapped to date
as far as its intersect with a radial fault of the
Cerro del Aguila stock. It very likely continues
along the platform edge beyond the Cerro del
Aguila mountain, whereas the other two described thrust faults die out in the basin (Agua
Zarca Thrust) and in the platform interior
(Misin Thrust) facies.

The geometry of the thrusts (Fig. 4b) suggests


that the plates were rotated by the subsequent
formation of more external overthrusts and by
formation of the west limb of the Pisaflores
Anticlinorium (Fig. 1). Deformation thus probably proceeded from west to east, similar to the
Idaho-Wyoming segment (Rubey and Hubbert,
1959; Armstrong and Oriel, 1965; Wiltschko
and Dorr, 1983) and the southern Canadian
segment (Bally and others, 1966) of the Cordilleran fold-and-thrust belt.
Structural versus Depositional Relief
The edge of the VSP bank is shown here to be
affected by Cordilleran faulting. The structural

Valle de G u a d a l u p e

Figure 8. Thrust plates along the eastern margin of the Valles-San Luis Potos Platform, to the north of the Moctezuma River. View from
Cerro Fro, looking north-northwest. 1 = Ahuacatln Formation. 2 = Tamabra Formation. 3 = El Abra Formation, platform-edge member. 4 =
Agua Nueva Formation. 5 = San Felipe Formation. 6 = Mndez Formation. A = Aqua Zarca Thrust. B = Lobo-Cinega Thrust. C = Misin
Thrust. The Lobo-Cinega thrust fault is composed of a 3,500-m-long near-surface dcollement and an inclined segment that is caused by the
rotation, subsequent to folding, of both the upper and the lower plates. This is in contrast to the Xilitla Thrust, the style of which is caused by a
stepped geometry of the fault.

Downloaded from gsabulletin.gsapubs.org on May 7, 2012

sw

C e r r o Sapo

NE

Figure 9. Lobo-Cinega (A) and Misin (B) Thrusts in the


Moctezuma canyon. 1 = Ahuacatln Formation. 2 = Tamabra
Formation. 3 = El Abra Formation, platform-edge member. 4 =
El Abra Formation, platform-interior member. 5 = San Felipe
Formation. 6 = Mndez Formation. The outward progradation
of the facies belts causes the lateral transition platform interiorplatform edge-foreslope-basin to be seen here as a vertical sequence. The relief of 1,500 m between Cerro Sapo and the
Lobo-Cinega thrust fault contains both a part of the structural
relief of the two marked overthrusts and a component of the
depositional topography between the Valles-San Luis Potos
Platform and the adjacent basin.

Argillaceous limestone
shale and sandstone

N
UJ
Q
UJ

5
o

P a r t i a l Bouma
cycles with graded and
convolute bedding

Sandstone, shale and


pelagic lime
wacke stone

Bioclastic
grainstone

Platformfacies
exoclasts

Figure 10. Lithology of the Upper Cretaceous across the eastern edge of the Valles-San
Luis Potos Platform, in the lower plate of the
Lobo-Cinega Thiust. Section A is located on
the platform edge (El Abra Formation) and
consists of the pelagic Soyatal Formation.
Section B is located 1.5 km farther southeast o n foreslope deposits (Tamabra
Formation) and is made up of a pelagic
sequence ( A g u a Nueva, San Felipe, and
M n d e z Formaitions) that b e c o m e s increasingly clastic t o w a r d its upper part.
Repetitions of platform-edge facies in the
Soyatal, A g u a Nueva, and San Felipe
Formations could be due to faulting, or
they may be sedimentary repetitions due
to migrating facies boundaries.

?
i

i:

i ir
i

Shale

and

subordinate
pelagic lime
mudstone

1
1
l-l

M
l-l
1 1 1
l-l
1.

Bioclastic
grainstone
Pelagic lime mudstone,
wackestone,
subordinate
shale and chert

_[_' i l l 1 i l l
1 1 1 1

Bioclastic

cot< <

->

p a r t l y dolomitized

LOBO
A) EL
VILLAGE

t O
et p
03

<

: *

oc
<t o

_ tc
bJ o
li.

microbreccia

and grainstone

V- li.
LOBO-AGUA
B) EL
ZARCA ROAD

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


CORDILLERAN DEFORMATION, MEXICO

NE Puerto del Baile

Piedra Blanca

Figure 11. Lobo-Cinega Thrust in the Amajac canyon. 1 = Chapulhuacn Formation. 2 = Ahuacatln Formation. 3 = Soyatal Formation (Upper Cretaceous). The thrust fault is located in the western
limb of a major anticlinorium. It nearly parallels bedding and dips
subvertically, due to subsequent rotation.

relief of these faults measures 1,050 m for the


Xilitla Thrust and approximately 3,000 m for
the stack of imbricates to the south of Federal
Highway 120. An exception is made by the area
to the northeast of El Lobo (Fig. 3), which appears to be the only locality of the study area
where the platform-slope-basin transition is exposed and not overprinted by Cordilleran deformations. The platform-edge facies is slightly
outbuilding and shows a basinward inclination
of - 2 0 in the southern wall of the Pea Prieta
mountain. No synsedimentary faults could be
observed along the escarpment. A minimum
value for the depositional relief is 1,070 m,
which is given by the altitude difference between
the erosion surface of the platform (Pea Prieta)
and the top of the Ahuacatln Formation to the
east. A comparable value was obtained for the

1395

W S W

Figure 12. Misin Thrust in the Amajac canyon (southwestern continuation of Fig. 11). 1 = Chapulhuacn Formation. 2 = Ahuacatln
Formation. 3 = El Abra Formation, platform-edge member. 4 =
Soyatal Formation (Upper Cretaceous).

more easterly situated Golden Lane Platform


by Enos (1977). The value is also in accord
with the depositional relief between the Great
Bahama Bank and Tongue of the Ocean
(Ball, 1967).
Timing of the Deformation
A maximum age limit for the regional compressive deformations was biostratigraphically
determined from the flysch-type Chicontepec
Formation from outcrops immediately below
the San Juanito thrust fault (sample SM 65) and
~80 m above the contact between the Velasco
and Chicontepec Formations, on the slope behind the San Juanito ranch (2.5 km to the
northwest of Huichihuayan). The sample contains fauna of the Globotruncana contusa plank-

tonic foraminiferal zone of late Maastrichtian


age (van Hinte, 1976), as does sample SM 62,
taken from the top of the Velasco Formation
directly behind the farmhouse. This is surprising,
as the Globotruncana contusa zone is also found
in the upper part of the Mndez Formation (50
to 100 m downsection) at its type locality (Pessagno, 1969) farther northeast in the flysch
basin, and because a Paleocene age is generally
assigned to the lower part of the Chicontepec
Formation (Lpez-Ramos, 1956). The discrepancy may exist because the type locality of the
Chicontepec Formation is farther southeast in
the Tampico-Misantle foredeep and because of
the well-known diachroneity of clastic wedges,
due to lateral accumulation. It cannot be totally
excluded, however, that the samples contain reworked Mndez fauna.

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


1396

A minimum age limit is put on the regional


compressive deformations by the radiometric
age of the Cerrc del Aguila stock (Fig. 3) of
62.2 1.5 Ma (early Paleocene: Ness and
others, 1980; late Paleocene: 1983 DNAG time
scale), as determined by conventional potassium-argon dating methods from separated
hornblende (P. Damon, sample PED-HI18-81).
Both the Cerro del Arguila pluton and another
major stock to the south of Jacala (Fig. 3; radiometrically dated at 41.5 0.9 Ma from separated biotite; P. Damon, sample PED-HI 17-81)
superimpose the Cordilleran structures. The
decay constants are: Xp = 4.963 10~10 a - 1 ; Xe =
0.581 10" 10 a" 1 ; k = 5.544 10" 10 a" 1 ;
K/K = 1.167 10~4 atom/atom. The two
intrusive ages fit ;:nto the Cenozoic magma distribution documented by Damon and others
(1981) and by Clark and others (1982) for
northern Mexico as a function of time and distance from a paleatrench on the Pacific margin.
The sample PED-HI 18-81 is located near the
upper boundary of the distribution, which can
be explained by the high crystallization temperature of hornblende and by an early cooling due
to the location of the sampling site at the very
margin of the intrusive.
The observed deformations occurred during
the Hidalgoan orogenic phase (de Cserna, 1960,
1975; Guzmn and de Cserna, 1963). They are
of the same age as the Late Cretaceous to early
Eocene compressional Laramide structures of
the external Rocky Mountains (Burchfiel and
Davis, 1975; Dickinson and Snyder, 1978), and
the same time span of deformation is also occupied by thin-skinned thrusting in the IdahoWyoming segmen t of the Cordillera (Royse and
others, 1975; Wiltschko and Dorr, 1983). The
structural fronts of the external Rocky Mountains and of the Sierra Madre Oriental are probably linked through the Monterrey-Torren
transverse zone and the Chihuahua-Coahuila
foreland fold-thrust belt, despite the different
structural style of the external Rocky Mountains
(thick-skinned and large vertical displacement
components). The probable connection is obscured, however, by the postorogenic tectonic
development of the transition zone (Rio Grande
Rift) and by the younger basin and range normal displacements that have taken place along
some of the Rocky Mountain thrust faults
(Young, 1979). The fold-thrust belt branches in
northern Chihuahua. From there, a western belt
follows the southern limit of the essentially unshortened Colorado Plateau in a zone of sinistral
transpression through southern New Mexico

M. SUTER

and southern Arizona (Drewes, 1981) and connects along the western edge of the Colorado
Plateau with the Idaho-Wyoming fold-thrust
belt.

elastic experiments, applying an external load in


the west, will clarify the stress distribution that
caused the observed deformations.
ACKNOWLEDGMENTS

DISCUSSION AND CONCLUSIONS


The eastern edge of the VSP carbonate platform is affected by large, low-angle Cordilleran
overthrusts that are at places steepened by subsequent imbrication or folding. Similar deformations exist along the western limit (Suter, 1982a,
1982b), and along the northeastern limit of the
platform near Aramberri (Tardy and others,
1976). Overthrusts have also deformed the eastern margin of the El Doctor Bank (Carrill.oMartinez and Suter, 1982) and the northeastern edge of the Cordoba Platform (Gonzalez
Alvarado, 1976; Mossman and Viniegra, 1976),
where the platform margins run parallel to the
Cordilleran structural trend. The thrust sheets
show a shortening of at least 3 to 4 km each and
10 to 12 km in total, and the actual horizontal
displacement amount may be greater, due to the
low angles between the thrust faults and
bedding.
On a regional scale, the thrusts are discrete
segments of a right-lateral shear zone with an en
echelon array, localized by a change in trend of
the platform margin. Displacement transfer ttetween the segments is partly discrete along a
transcurrent fault (Tlamaya-Huichihuayan lineament) and partly diffuse. Shortening of the individual segments is greatest where they coincide
with the platform edge, whereas fault slip decreases as the thrusts pass inside the platform or
basin facies.
The style of deformation of the study area is
controlled mainly by the lithologic change of the
middle Cretaceous rocks from an orthotropically layered basin facies in the east to a two to
five times thicker, homogeneous platform-edge
assemblage in the west (Fig. 2). The bank margin has been a zone of shear instability.
The magnitude of the horizontal tectonic components of the stress field which caused the
deformation may have changed by a factor of
two to five across the margin of the platform,
due to the difference in cross-sectional area. Additionally, a horizontal gravitational stress component was caused by the inclination of the
foreslope in the direction of tectonic transport
(Elliott, 1976; Chappie, 1978). Scaled modeling
of the deformations in the platform-basin
transition-zone, either numerically, using finiteelement techniques, or by mechanical and photo-

This report is a result of a co-operative project


between the University of Basel (Switzerland)
and the Mexican National Autonomous University (UNAM). Financial and logistic support
was provided by UNAM and by the Swiss Ministry of Foreign Affairs (contract t. 541) a nd has
been appreciated.
I am greatly indebted to Zoltan de Cserna for
his advice during my stay at UNAM; to Paul
Damon for sampling and determination of radiometric ages; to Mrs. Caron, Hans Bolli, and
Mario Wannier for the determination of microfossils; to W. R. Muehlberger, Zoltan de (-serna,
Jay Namson, Dietrich Roeder, and Frank Royse
for review of the manuscript; and to Pau 1 Enos
and Jos Longoria for written comments. Discussion in the field with Maria Fernanda
Campa, Miguel Carrillo, Zoltan de (-serna,
Stanislaw Kwiatkowski, Hans Laubscher, Jay
Namson, and Fernando Ortega was helpful.
Paul Damon acknowledges his support from the
National Science Foundation and Consejo Nacional de Ciencia y Tecnologa.

REFERENCES CITED
Alvarez, Walter, Engelder, Terry, and Lowrie, William, 1976, Formation
of spaced cleavage and folds in brittle limestone by dissolution: Geology,
v. 4, p. 698-701.
Armstrong, F, C., and Oriel, S. S., 1965, Tectonic development of IdahoWyoming thrust-belt: American Association of Petroleum Geologists
Bulletin, v. 49, p. 1849-1866.
Ball, M. M., 1967, Carbonate sand bodies of Florida and the Bahamas: Journal
of Sedimentary Petrology, v. 37, p. 556-591.
Bally, A. W., Gordy, P. L., and Stewart, G. A., 1966, Structure, seismic data
and orogenic evolution of southern Canadian Rocky Mountains: Canadian Petroleum Geology Bulletin, v. 14, p. 337-381.
Berger, Philip, and Johnson, A. M., 1980, First-order analysis of defcrmation of
a thrust sheet moving over a ramp: Tectonophysics, v. 70, p. T9-T24.
Bodenlos, A. J., 1956a, Notas sobre la geologa de la Sierra Madre en la seccin
Zimapn-Tamazunchale: International Geological Congress, 20th, Mexico, Excursions A-14 and C-6, Guidebook, p. 293-309.
1956b, Itinerario Zimapn, Hgo.-Tamazunchale, San Luis Potos: International Geological Congress, 20th, Mexico, Excursions A-14 and
C-6, Guidebook, p. 179-215.
Bonet, Federico, 1956, Tamazunchale-Taninul: International Geokgical Con.
gress, 20th, Mexico, Excursions A-14 and C-6, Guidebook, p. 217-240.
Burchfiel, B. C., and Davis, G. A., 1975, Nature and controls of (Cordilleran
orogenesis, western United States: Extensions of an earliei synthesis:
American Journal of Science, v. 275-A, p. 363-396.
Busch, D. A., and Gavela, Amado, 1978, Stratigraphy and structure of Chicontepec turbidites, southeastern Tampico-Misantle Basin, Mexico: American Association of Petroleum Geologists Bulletin, v. 62, p. 2)5-246.
Carrasco, Baldomero, 1977, Albian sedimentation of submarine autcchthonous
and allochthonous carbonates, east edge of the Valles-San Luis Potos
platform, Mexico, in Cook, H. E., and Enos, Paul, eds.. Deep-water
carbonate environments: Society of Economic Paleontologists and Mineralogists Special Publication 25, p. 263-272.
Carrillo-Bravo, Jos, 1971, La Plataforma Valles-San Luis Potos: Asociacin
Mexicana de Geologos Petroleros Boletn, v. 23, p. 1-102.
Carrillo-Martnez, Miguel, and Suter, Max, 1982, Tectnica de los r (rededores
de Zimapn, Hidalgo: Sociedad Geolgica Mexicana Convencin geolgica nacional, 6th, Excursin a la regin de Zimapn y r a s circundantes, Guidebook, p. 1-20.
Chappie, W. M., 1978, Mechanics of thin-skinned fold-and-thrast "lelts: Geo-

Downloaded from gsabulletin.gsapubs.org on May 7, 2012


1397

CORDILLERAN DEFORMATION, MEXICO


logical Society of America Bulletin, v. 89, p. 1189-1198.
Clark, K. F., Foster, C. T., and Damon, P. E., 1982, Cenozoic mineral deposits
and subduction-related magmatic arcs in Mexico: Geological Society of
America Bulletin, v. 93, p. 533-544.
Damon, P. E., Shaftqullab, Muhammad, and Clark, K. F., 1981, Age trends of
igneous activity in relation to metallogenesis in the southern Cordillera:
Arizona Geological Society Digest, v. 14, p. 137-154.
de Cserna, Zoltan, 1956, Tectnica de la Sierra Madre Oriental de Mxico,
entre Torren y Monterrey: International Geological Congress, 20th,
Mxico, 87 p.
1960, Orogenesis in time and space in Mexico: Geologische Rundschau,
v. 50, p. 595-605.
1975, Mexico, in Fairbridge, R. W., ed., Encyclopedia of World Regional GeologyPart 1, Western Hemisphere: Stroudsburg, Pennsylvania, Dowden, Hutchinson and Ross, p. 348-360.
Demant, Alain, 1978, Caractersticas del Eje Neovolcnico Transmexicano y
sus problemas de interpretacin: Universidad Nacional Autnoma de
Mxico, Instituto de Geologa, Revista, v. 2, p. 172-187.
Dickinson, W. R., and Snyder, W. S., 1978, Plate tectonics of the Laramide
orogeny: Geological Society of America Memoir 151, p. 355-366.
Drewes, Harald, 1981, Tectonics of southeastern Arizona: U.S. Geological
Survey, Professional Paper 1144.
Elliott, David, 1976, The motion of thrust sheets: Journal of Geophysical
Research, v. 81, p. 949-963.
Enos, Paul, 1974, Reefs, platforms and basins of middle Cretaceous in northeast
Mexico: American Association of Petroleum Geologists Bulletin, v. 58,
p. 800-809.
1977, Tamabra limestone of the Poza Rica trend. Cretaceous, Mexico,
in Cook, H. E., and Enos, Paul, eds., Deep-water carbonate environments: Society of Economic Paleontologists and Mineralogists Special
Publication 25, p. 273-314.
1983, Sierra Madre Oriental, in Eby, D. E., and Clarke, R. T., eds.,
Sedimentation and diagenesis of mid-Cretaceous platform margin eastcentral Mexico: Dallas Geological Society, Guidebook, p. 78-94.
Fletcher, R. C., and Pollard, D. D., 1981, Anticrack model for pressure solution
surfaces: Geology, v. 9, p. 419-424.
Gonzlez-Al varado, Jorge, 1976, Resultados obtenidos en la exploracin de la
plataforma de Crdoba y principales campos productores: Sociedad
Geolgica Mexicana Boletn, v. 37, p. 53-59.
Griffith, L. S., Pitcher, M. G., and Rice, G. W., 1969, Quantitative environmental analysis of a Lower Cretaceous reef complex, in Friedman,
G. M., ed., Depositional environments in carbonate rocks: Society of'
Economic Paleontologists and Mineralogists Special Publication 14,
p. 120-138.
Guzmn, E. J., and de Csema, Zoltan, 1963, Tectonic history of Mexico:
American Association of Petroleum Geologists Memoir 2, p. 113-129.

Heim, Arnold, 1940, The front ranges of Sierra Madre Oriental, Mexico, from
Ciudad Victoria to Tamazunchale: Eclogae Geologicae Helvetiae, v. 33,
p. 313-352.
King, P. B,. 1969, Tectonic map of North America (1:5,000,000): U.S. Geological Survey.
Lpez-Ramos, Ernesto, 1956, Terciario marino del este de Mxico: International Geological Congress, 20th, Mexico, Excursion C-16, Guidebook,
94 p.
McDowell, F. W., and Clabaugh, S. E., 1979, Ignimbrites of the Sierra Madre
Occidental and their relation to che tectonic history of western Mexico:
Geological Society of America Special Paper 180, p. 113-124.
Minero, C. J., 1983a, Discussion of back-reef and platform-interior and sedimentation in the El Abra Formation, in Eby, D. E., and Clarke R. T.,
eds., Sedimentation and diagenesis of mid-Cretaceous platform margin
east-central Mexico: Dallas Geological Society Guidebook, p. 110-124.
1983b, Sedimentary environments and diagenesis of the El Abra Formation (Cretaceous), Mexico [Ph.D. thesis]: Binghamton, New York,
State University of New York, 367 p.
Mossman, R. W., and Viniegra, Francisco, 1976, Complex fault structures in
Veracruz province of Mexico: American Association of Petroleum
Geologists Bulletin, v. 60, p. 379-388.
Ness, G., Levi, S., and Couch, R., 1980, Marine magnetic anomaly time-scales
for the Cenozoic and Late Cretaceous: A precis, critique and synthesis:
Reviews of Geophysics and Space Physics, v. 18, p. 753-770.
Nixon, G. T., 1982, The relationship between Quaternary volcanism in central
Mexico and the seismicity and structure of subducted ocean lithosphere:
Geological Society of America Bulletin, v. 93, p. 514-523.
Pessagno, E. A., 1969, Upper Cretaceous stratigraphy of the western Gulf coast
area of Mexico, Texas and Arkansas: Geological Society of America
Memoir 111.
Rich, J. L., 1934, Mechanics of low-angle overthrust faulting illustrated by
Cumberland thrust block, Virginia, Kentucky, and Tennessee: American Association of Petroleum Geologists Bulletin, v. 18, p. 1584-1596.
Royse, F., Warner, M. A., and Reese, D. L., 1975, Thrust belt structural
geometry and related stratigraphie problems: Rocky Mountain Association of Geologists Symposium Volume, p. 41-54.
Rubey, W. W.,and Hubbert, M. K., 1959, Role of fluid pressure in mechanics
of overthrust faultingII, Overthrust belt in gosynclinal area of western Wyoming in light of fluid-pressure hypothesis: Geological Society of
America Bulletin, v. 70, p. 167-206.
Serra, Sandro, 1977, Styles of deformation in the ramp regions of overthrust
faults: Wyoming Geological Association, Annual Field Conference,
29th, Guidebook, p. 487-498.
1978, Styles of deformation in the ramp regions of overthrust faults
[Ph.D. thesis]: College Station, Texas, Texas A&M University, 83 p.
Suppe, John, 1983, Geometry and kinematics of fault-bend folding: American

Journal of Science, v. 283, p. 684-721.


Suppe, John, and Namson, Jay, 1979, Fault-bend origin of frontal folds of
the western Taiwan fold-and-thrust belt: Petroleum Geology of Taiwan,
v. 16, p. 1-18.
Suter, Max, 1980, Tectonics of the external part of the Sierra Madre Oriental
foreland thrust-and-fold belt between Xilitla and the Moctezuma river
(Hidalgo and San Luis Potos States): Universidad Nacional Autnoma
de Mxico, Instituto de Geologa, Revista, v. 4, p. 19-31.
1981, Strukturelles Querprofil durch den nordwestlichen Faltenjura,
Mt-Tem-Randberschiebung-Freiberge: Eclogae Geologicae Helvetiae,
v. 74, p. 255-275.
1982a, Itinerario Zimapn-Jiliapan: Sociedad Geolgica Mexicana
Convencin Geolgica Nacional, 6th, Excursin a la regin de Zimapn
y reas circundantes, Guidebook, p. 43-51.
1982b, Itinerario Puerto de la Estancia-Jagey Colorado-Zimapn: Sociedad Geolgica Mexicana Convencin Geolgica Nacional, 6th Excursin a la regin de Zimapn y reas circundantes, Guidebook,
p56-64.
1983, Structural traverse across the Sierra Madre Oriental fold-thrust
belt in east-central Mexico [abs.]: Geological Society of America Abstracts with Programs, v. 15, p. 702.
Tardy, Marc, Ramirez, Calixto, and Patio, Manuel, 1976, El frente de la Napa
de Parras (Conjunto Cadena alta-Altiplano central) en el rea de
Aramberri, N. L., Sierra Madre Oriental, Mxico: Universidad Nacional
Autnoma de Mxico, Instituto de Geologa, Revista, v. 2 (old series),
p. 1-12.
Thorpe, R. S., 1977, Tectonic significance of alkaline volcanism in eastern
Mexico: Tectonophysics, v. 40, p. T19-T26.
van Hinte, J. E,, 1976, A Cretaceous time scale: American Association of
Petroleum Geologists Bulletin, v. 60, p. 489-497.
Wilson, J. L., 1975, Carbonate fades in geologic history: New York, SpringerVerlag, 471 p.
Wiltschko, D. V., and Dorr, J. A., 1983, Timing of deformation in overthrust
belt and foreland of Idaho, Wyoming, and Utah: American Association
of Petroleum Geologists Bulletin, v. 67, p. 1304-1322.
Young, R. A., 1979, Laramide deformation, erosion and plutonism along the
southwestern margin of the Colorado Plateau: Tectonophysics, v. 61,
p. 25-47.

MANUSCRIPT RECEIVED BY THE SOCIETY SEPTEMBER 1 4 , 1 9 8 2


REVISED MANUSCRIPT RECEIVED M A R C H 7 , 1 9 8 4
MANUSCRIPT ACCEPTED A P R I L 3 0 , 1 9 8 4

Printed in U.S.A.

S-ar putea să vă placă și