The Early Mesozoic volcanic arc of western North America in

Transcription

The Early Mesozoic volcanic arc of western North America in
Journal of South American Earth Sciences 25 (2008) 49–63
www.elsevier.com/locate/jsames
The Early Mesozoic volcanic arc of western North America
in northeastern Mexico
José Rafael Barboza-Gudiño a,*, Marı́a Teresa Orozco-Esquivel b,
Martı́n Gómez-Anguiano c, Aurora Zavala-Monsiváis d
a
Instituto de Geologı́a, Universidad Autónoma de San Luis Potosı́, Manuel Nava No. 5. Zona Universitaria, 78240 San Luis Potosı́, S.L.P., Mexico
b
Centro de Geociencias, Universidad Nacional Autónoma de México, Campus Juriquilla, 76230 Querétaro, Qro., Mexico
c
Universidad Tecnológica de La Mixteca, Carretera a Acatuma Km. 2.5, 69000 Huajuapan de León, Oaxaca, Mexico
d
Posgrado en Geologı́a Aplicada, Universidad Autónoma de San Luis Potosı́, Manuel Nava No. 5, Zona Universitaria,
78240 San Luis Potosı́, S.L.P., Mexico
Abstract
Volcanic successions underlying clastic and carbonate marine rocks of the Oxfordian–Kimmeridgian Zuloaga Group in northeastern
Mexico have been attributed to magmatic arcs of Permo–Triassic and Early Jurassic ages. This work provides stratigraphic, petrographic
geochronological, and geochemical data to characterize pre-Oxfordian volcanic rocks outcropping in seven localities in northeastern
Mexico. Field observations show that the volcanic units overlie Paleozoic metamorphic rocks (Granjeno schist) or Triassic marine strata
(Zacatecas Formation) and intrude Triassic redbeds or are partly interbedded with Lower Jurassic redbeds (Huizachal Group). The volcanic rocks include rhyolitic and rhyodacitic domes and dikes, basaltic to andesitic lava flows and breccias, and andesitic to rhyolitic
pyroclastic rocks, including breccias, lapilli, and ashflow tuffs that range from welded to unwelded. Lower–Middle Jurassic ages (U/
Pb in zircon) have been reported from only two studied localities (Huizachal Valley, Sierra de Catorce), and other reported ages (Ar/
Ar and K–Ar in whole-rock or feldspar) are often reset. This work reports a new U/Pb age in zircon that confirms a Lower Jurassic
(193 Ma) age for volcanic rocks exposed in the Aramberri area. The major and trace element contents of samples from the seven localities
are typical of calc-alkaline, subduction-related rocks. The new geochronological and geochemical data, coupled with the lithological features and stratigraphic positions, indicate volcanic rocks are part of a continental arc, similar to that represented by the Lower–Middle
Jurassic Nazas Formation of Durango and northern Zacatecas. On that basis, the studied volcanic sequences are assigned to the Early
Jurassic volcanic arc of western North America.
Ó 2007 Elsevier Ltd. All rights reserved.
Keywords: Stratigraphy; Volcanic rocks; Arc; Jurassic; Mexico
1. Introduction
Volcanic successions underlie clastic and carbonate marine sequences of the Oxfordian–Kimmeridgian Zuloaga
Group in northeastern Mexico. Studies of the successions
(e.g., Pantoja-Alor, 1972; Blickwede, 2001; López-Infanzón, 1986; Jones et al., 1990, 1995; Bartolini, 1998; Barto-
*
Corresponding author. Fax: +52 444 8111741.
E-mail address: [email protected] (J.R. Barboza-Gudiño).
0895-9811/$ - see front matter Ó 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsames.2007.08.003
lini et al., 2003; Barboza-Gudiño et al., 1998, 1999, 2004)
reveal diverse lithologies and stratigraphic position below
Oxfordian limestones. In northern Durango and Zacatecas, the volcanic pre-Oxfordian rocks have been assigned
to the Jurassic continental volcanic arc, related to the
active continental margin of southwestern North America
(Grajales-Nishimura et al., 1992; Jones et al., 1995; Bartolini, 1998; Bartolini et al., 2003), whereas in areas of San
Luis Potosı́, Nuevo León, and Tamaulipas, they have been
considered products of a Permo–Triassic volcanic arc (Meiburg et al., 1987; Bartolini et al., 1999).
50
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
However, the assignments are uncertain because reliable isotopic data are lacking, and petrographic and
geochemical information is scarce. In this article, we
report new geochemical, petrographic, and stratigraphic
data for pre-Oxfordian rocks exposed in northeastern
Mexico in the Sierra de Salinas, Sierra de Charcas,
and Sierra de Catorce in San Luis Potosı́; the Aramberri and San Marcos areas in Nuevo León; and the
Huizachal Valley in Tamaulipas (Fig. 1). The main purpose of the geochemical analyses is to characterize the
rocks with regard to the tectonic setting in which they
could have originated, rather than providing information leading to a detailed evolutionary model of the
magmas. The new data, supported by information from
the literature, help establish the tectonic setting and the
correlations among pre-Oxfordian volcanic rocks in
northeastern Mexico.
2. The exposed sequences
Localities described in this section are shown in Fig. 1.
Some outcrop aspects and textural or microstructural
details of the pre-Oxfordian volcanic rocks studied in localities from northeastern Mexico are illustrated in Fig. 2.
In northern Durango, intermediate to felsic volcanic
rocks are exposed in the Villa Juárez area. Pantoja-Alor
(1972) defines this unit as the Nazas Formation, with its
type locality at Cerritos Colorados and a reported Pbage of 230 ± 20 Ma from a rhyolitic flow. In the Villa
Juárez area, Bartolini and Spell (1997) obtain a 40Ar/39Ar
age of 195 ± 55 Ma from plagioclase in rhyolitic rocks,
probably comparable to those dated by Pantoja-Alor
(1972). The Nazas Formation is the oldest exposed unit
in the Villa Juárez region, but 200 km northwest of this
locality, in Santa Marı́a del Oro, northern Durango, the
Fig. 1. Geologic map of northeastern Mexico, showing outcrops of pre-Oxfordian volcanic and sedimentary rocks and location of samples.
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
51
Fig. 2. Details of the textures and microstructures observed in outcrops, hand samples, and thin sections of the volcanic rocks. (a) Volcanic breccia in the
pyroclastic deposits of the Aramberri area, Nuevo León. Knife is 11 cm long. (b) General aspect of a rhyolitic dike (R) in the Sierra de Catorce; Juj, Upper
Jurassic La Joya Formation; volc., intermediate pre-Oxfordian volcanic rocks. Note person for scale. (c) Lithophyse contained in pyroclastic deposits
outcropping west of Charcas, San Luis Potosı́ (long edge = 10 cm). (d) Fiamme structures in ignimbrites from Aramberri, Nuevo León; hand-lens
diameter is 2.5 cm. (e) Fragments of partially collapsed pumice in an ignimbrite of Charcas, San Luis Potosı́. (f) Trachytic or pylotaxitic texture in a
basaltic andesite of La Ballena, Zacatecas, San Luis Potosı́.
unit overlies Paleozoic metamorphic rocks (Bartolini, 1998)
and unconformably underlies Upper Jurassic sandy limestone of La Gloria Formation (Imlay, 1936).
In northern Zacatecas, lava flows, airfall and ashflow
tuffs, and lahars correlated with the Nazas Formation have
been described in the Caopas–Rodeo area, including Sierra
de Teyra to the west and Sierra de San Julián to the east
(Blickwede, 1981, 2001). Some units (Caopas schist and
Rodeo Formation) initially were considered pre-Jurassic
because of their very deformed and metamorphosed
aspects (de Cserna, 1956; Córdoba-Méndez, 1964). Subsequent studies indicated that the deformed units are coeval
with the Nazas Formation and belong to the same Jurassic
continental volcanic arc sequence (López-Infanzón, 1986;
Jones et al., 1990, 1995). This hypothesis is supported by
a K–Ar age of 183 Ma determined in hornblende from
the so-called Rodeo Formation (López-Infanzón, 1986)
and an apparent U–Pb age of 158 ± 4 Ma in zircon grains
from the Caopas schist (Jones et al., 1995). In the Sierra de
Teyra, the Nazas Formation overlies the marine siliciclastic
Taray Formation (Córdoba-Méndez, 1964) of Triassic age
(Silva-Romo et al., 2000) and underlies Upper Jurassic
continental deposits of La Joya Formation (Mixon et al.,
1959) and shallow marine limestones of the Zuloaga Formation (Imlay, 1938).
In western San Luis Potosı́, volcanic sequences comparable to those of Durango and Zacatecas rest on Triassic
thin-bedded strata interpreted as part of a turbiditic
sequence known as the Zacatecas Formation (Martı́nezPérez, 1972) or the La Ballena Formation (Silva-Romo,
0.717
0.592
0.690
193.6
193.7
199.4
Pb/206Pb age
(Ma)
207
193.1
191.2
183.0
Pb/235U age
(Ma)
207
0.04997
0.04997
0.05010
0.16
0.22
0.17
0.20943
0.20720
0.19744
0.11
0.12
0.11
0.030397
0.030071
0.028584
%Err
Pb
Pb/
Rad.
%Err
U
Pb/
Rad.
%Err
U
Pb/
2276.1
2019.8
2234.7
0.6
0.6
0.6
12.4
4.4
6.3
1.7
4.4
3.5
Z1
Z2
Z3
402
142
211
/
Pb (corr.)
/
0.133
0.155
0.169
Pb (corr.)
Rad.
0.11
0.18
0.12
193.0
191.0
181.7
Pb/238U age
(Ma)
206
206
207
235
207
238
206
208 206
Corrected atomic ratios
206 204
Com. Pb
(pg)
Pb
(ppm)
U
(ppm)
Weight
( g)
Sample
1993; Centeno-Garcı́a and Silva-Romo, 1997). The volcanic units are unconformably overlain by Upper Jurassic
redbeds of La Joya Formation and shallow marine limestones of the Zuloaga Formation. The volcanic sequences
crop out in the Sierra de Salinas, at the state border
between Zacatecas and San Luis Potosı́ (Silva-Romo,
1993; Barboza-Gudiño et al., 1998, 1999; Zavala-Monsiváis, 2000; Gómez-Anguiano, 2001), in the Sierra de
Charcas, and in the region of Tepozan in western San Luis
Potosı́ state (Tristán-González and Torres-Hernández,
1992, 1994; Tristán-González et al., 1995; Zavala-Monsiváis, 2000), as well as in the Sierra de Catorce (LópezInfanzón, 1986; Barboza-Gudiño et al., 1998, 1999; Zavala-Monsiváis, 2000; Hoppe, 2000; Gómez-Anguiano,
2001). Barboza-Gudiño et al. (2004) report U–Pb isotopic
analyses of zircon for rhyolite of the Sierra de Catorce.
The fraction recording the least inheritance yields an age
of 174.7 ± 1.3 Ma, a maximum age of the rock.
In Nuevo León state, volcanic rocks comparable in
their lithology and stratigraphic position with those
described previously have been observed in a sequence
exposed in Aramberri (Jones et al., 1995). The sequence
consists predominantly of ignimbrites, volcanic breccias,
and tuffs of intermediate to felsic composition, which
overlie Paleozoic schist and unconformably underlie the
transgressive Upper Jurassic strata. Some authors consider these volcanic units related to a Permo–Triassic volcanic arc (Meiburg et al., 1987; Bartolini et al., 1999). We
analyze three single zircon grains from a rhyodacitic
ignimbrite at the Mezquital section, north of Aramberri.
One grain yields essentially concordant Pb/U ratios that
indicate an age of 193.1 ± 0.3 Ma for the rock (Table 1);
an upper intercept at 193.3 ± 1.5 Ma is also shown in a
discordia plot (Fig. 3).
In the redbed sequence exposed near San Marcos,
south of Galeana, Nuevo León, the volume of volcanic
and subvolcanic rocks is small, and the stratigraphic relations are uncertain. At this locality, some dikes and sills of
trachytic composition intrude Upper Triassic redbeds of
the Lower Huizachal Group and are truncated by the
unconformity beneath Upper Jurassic breccias of La Joya
Formation and Oxfordian–Kimmeridgian evaporites of
the Minas Viejas Formation (Gómez-Anguiano, 2001).
Pre-Oxfordian volcanic rocks are also present in the
State of Tamaulipas, at Huizachal Valley (Jones et al.,
1995; Fastowsky et al., 2005), the Miquihuana area (Bartolini et al., 2003), and La Boca Canyon (Fig. 1). Metarhyolite exposed in the nearby Caballeros Canyon are
apparently older and possibly of Paleozoic age (Gursky
and Ramı́rez-Ramı́rez, 1986; Stewart et al., 1999). The
best exposures of pre-Oxfordian volcanic and subvolcanic
rocks in Tamaulipas are found in the Huizachal Valley,
where they represent the basal part of the exposed Mesozoic sequence, underlying and partially intruding Lower
Jurassic redbeds; no older units crop out in this area. Volcanic rocks in this locality consist of ignimbrite, finegrained tuff, breccia, and lava of intermediate to felsic
Corr.
coef.
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
Table 1
U–Pb data for single zircon crystals from a rhyodacitic ashflow tuff, Aramberri area (sample MZQTG1)
52
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
53
Fig. 3. Concordia diagram for U–Pb isotope ratios of zircons from a rhyodacitic ashflow tuff in the Aramberri area (see Table 1). Error ellipses of
individual spots are 2 .
composition. In the Huizachal Valley, zircon grains from a
pyroclastic flow are dated by U–Pb at 189 ± 0.2 Ma
(Fastowsky et al., 2005). At this locality, Fastowsky et al.
(2005) recognize an older volcanic sequence beneath an
angular unconformity that underlies the dated pyroclastic
flow at the base of the Lower Jurassic redbed sequence of
La Boca Formation. All features described by Fastowsky
et al. (2005) and those we observe in the same outcrops
of the ‘‘older’’ sequence are typical textures and structures
of rhyolitic domes: spherulitic structures, steeply dipping
flow-like bands, lithophysae, peripheral lava flows or lobes,
and associated ash flow tuffs. In addition, we observe intrusive relationships to Lower Jurassic redbeds of La Boca
Formation. We interpret the steeply dipping features in this
sequence as subvertical flow bands resulting from magma
injection into a volcanic dome, whose emplacement age is
not necessary much older than the dated Lower Jurassic
pyroclastic flow, and thus, we consider plausible the inclusion of this lower sequence as part of the same Lower
Jurassic volcanic arc.
Isotopic ages reported to date for pre-Oxfordian volcanic rocks in northeastern Mexico are summarized in Table
2. Three Lower–Middle Jurassic ages were obtained by the
U–Pb method in zircons, whereas other included K–Ar and
Ar/Ar ages (whole-rock or feldspar) indicate Cretaceous–
Paleogene ages and probably reflect age resetting during
the Laramide event.
3. Petrography
The petrography and petrology of the Lower Jurassic
volcanic rocks in the Nazas Formation and comparable
rocks from northeastern Mexico have been described from
different localities by various authors. Blickwede (1981,
2001) provides petrographic and petrologic data on rocks
from Sierra de San Julián; López-Infanzón (1986) offers
data on rocks from the Caopas–Rodeo and Sierra de Teyra
areas and compares them with rocks of the Sierra de Catorce; Jones et al. (1995) describe volcanic rocks of the Caopas–Pico de Teyra region and compare them with rocks
Table 2
Summary of isotopic ages of pre-Oxfordian volcanic rocks from northeastern Mexico, including several Ar/Ar and K–Ar ages determined in whole-rock
or feldspar that are considered reset ages
Location name
State
Rock type
Method
Material dated
Age (Ma)
Error (Ma)
Source
Huizachal Valley
Sierra de Salinas
Aramberri
Huizachal Valley
Miquihuana
San Marcos
Sierra de Salinas
Sierra de Catorce
Sierra de Catorce
Aramberri
Tamaulipas
S. Luis Potosı́
Nuevo León
Tamaulipas
Tamaulipas
Nuevo León
S. Luis Potosı́
S. Luis Potosı́
S. Luis Potosı́
Nuevo León
Pyroclastic flow
Basalt
Rhyolite
Rhyolite
Rhyolite
Andesitic dike
Andesitic basalt
Rhyolite
Rhyolite
Pyroclastic rock
U–Pb
40
Ar/39Ar
K–Ar
K–Ar
K–Ar
K–Ar
K–Ar
K–Ar
U–Pb
U–Pb
Zircon
Whole rock
Whole rock
Whole rock
Whole rock
Whole rock
Whole rock
Feldspar
Zircon
Zircon
189.0
82.9
70.7
52.1
57.8
104
81.9
110.0
174.7
193.1
0.2
0.6
1.8
1.4
1.5
3
4.1
1.9
1.3
0.3
Fastowsky et al. (2005)
Bartolini (1998)
Fastowsky et al. (2005)
Bartolini (1998)
Bartolini (1998)
Bartolini (1998)
Barboza-Gudiño et al. (1999)
This work
Barboza-Gudiño et al. (2004)
This work
54
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
from Torreón, Sierra de Catorce, Charcas, Aramberri,
Miquihuana, and Huizachal Valley; and finally, ZavalaMonsiváis (2000) describes the petrography of pre-Oxfordian volcanic rocks included in this study from western
San Luis Potosı́ State at Sierra de Salinas, Charcas,
Tepozán, and Sierra de Catorce. We compile relevant
petrographic data and add new petrographic descriptions
of the main lithologic features, which we summarize in
Table 3.
3.1. Rhyolite
Porphyritic rhyolite with quartz and sanidine phenocrysts is abundant in the ‘‘Caopas schist,’’ exposed near
the town of same name, and in the Nazas Formation outcrops of Villa Juárez and Sierra de San Julián. In the Sierra
de Catorce, a rhyolitic dike (Fig. 2b) exhibits porphyritic
texture, with phenocrysts of hypidiomorphic quartz in a
groundmass of feldspar and quartz. The rhyolitic dike is
strongly altered, as shown by totally kaolinitized feldspar
phenocrysts and subsequent silicification of the rock.
Locally, foliation with lepidoblastic texture results from
the presence of oriented sericite associated with dynamic
metamorphism. In the Huizachal Valley, rhyolitic domes
are characterized by steeply dipping flow bands and the
development of abundant spherulites, ranging from one
to several centimeters in diameter.
3.2. Basalts and basaltic andesites
In the Caopas–Rodeo area, Jones et al. (1995) describe
sparse basaltic lavas in the Nazas Formation, whereas
andesitic lava flows are more common in the unit known
as the Rodeo Formation. In the Sierra de Salinas, north
of the town La Ballena, basaltic–andesitic lava is the dominant rock type; some lavas display fluidal porphyritic texture with highly altered, probable hornblende phenocrysts,
scarce pyroxene, olivine, and plagioclase in a fine groundmass composed of acicular plagioclase, ferromagnesian,
and opaque grains. At the base of the sequence, the lavas
are brecciated. Similar basaltic–andesitic lavas crop out
at Sierra de Catorce and Charcas. In the latter area, andesitic lava flows contain a brecciated zone that, given the
presence of autoclasts arranged in a ‘‘puzzle structure,’’ is
interpreted as an autoclastic breccia and probable peperite,
associated with a flow front or basal breccia that was
apparently engulfed by igneous material of the same composition, and in part mixing with wet sediment. Finally,
volcanic products in the Huizachal Valley include layers
of cinders, scoreaceous material, and andesitic lava. These
rocks are dark to reddish brown due to oxidation and have
a very compact, deformed aspect, with vestiges of vesicles.
3.3. Pyroclastic deposits
The most common rocks are andesitic to dacitic
and rhyolitic pyroclastic products that show clear flow
indicators. The textures are diverse, including breccia, lapilli tuff, and fine volcanic ash. Some deposits display welding and devitrification, whereas others appear to be
unwelded rocks, compacted by deformational processes,
that record intense foliation and jointing.
In the Sierra de San Julián, Blickwede (1981, 2001)
describes a 1000 m thick volcanic sequence consisting of
65% pyroclastic rocks, 25% sedimentary and volcaniclastic
rocks, and only 10% lavas. Volcanic breccias or tuff-breccias are the predominant rocks in the various small outcrops exposed in the Arroyo El Tepozán, northern Sierra
de Charcas. A volcanic breccia that crops out in Aramberri, Nuevo León (Fig. 1), consists of angular fragments
of rhyodacitic to rhyolitic welded tuffs in a sandy matrix
(Fig. 2a). These deposits are directly overlain by shallow
marine sandstone with calcareous cement of La Joya Formation that represents the base of the Late Jurassic marine
transgression (Mixon et al., 1959).
Typical ignimbrites with notable development of fiamme
structures are identified in outcrops at Charcas, Aramberri,
and Huizachal Valley (Fig. 2d and e). In La Boca Canyon,
the pyroclastic deposits are represented by fine-grained
tuffs that partly contain deformed accretionary lapilli.
4. Geochemistry
Ten samples were collected for geochemical analysis
from six exposures of pre-Oxfordian volcanic units in the
states of San Luis Potosı́, Nuevo León, and Tamaulipas
(Fig. 1). Sampling of the diverse lithologies in the studied
areas was strongly restricted by the availability of acceptably fresh samples. Major and trace-element analysis of
whole-rock samples was performed by ICP-MS at the Centre des Recherches Pétrographiques et Geóchimiques
(CNRS) of Nancy, France (samples: PBLG1 from Sierra
de Salinas, CHRG1 from Sierra de Charcas, RCG1 from
Sierra de Catorce; MZQTG1 collected in the Mezquital
area, north of Aramberri; SM1 from the San Marcos area
south of Galeana, Nuevo León; HZCHG1 and HZCHG2
from Huizachal Valley in Tamaulipas), and Activation
Laboratories Ltd. (samples: TPZ1 from Tepozán outcrops,
northern Sierra de Charcas; RCG3 from north of Real de
Catorce, Sierra de Catorce; SM3 from San Marcos area
in Nuevo León). The composition of analyzed samples is
summarized in Table 4.
The analyzed volcanic rocks are classified as intermediate to felsic after the silica content limits proposed by Peccerillo and Taylor (1976). Following the total alkalis vs.
silica (TAS) classification of Le Bas et al. (1986), the samples are classified as trachyandesite, andesite, dacite, and
rhyolite. The use of the TAS diagram to classify volcanic
rocks is restricted to fresh rocks because of the high mobility of alkaline elements during secondary processes. The
analyzed samples have high volatile contents (LOI) and
show evidence of oxidation, which basically makes them
unsuitable for TAS classification. Nevertheless, the
obtained chemical classification is supported by its agree-
Table 3
Petrographic features observed in the studied pre-Oxfordian volcanic rocks of northeastern Mexico
Feature
RhyoliteRhyodacite
Macrostructure
La Ballena
Charcas
Texture
Components
Alteration
minerals
Basalt–
Andesite
Macrostructure
Texture
Components
Alteration
minerals
Pyroclastic
rocks
Lava flow, basal
flow breccia
Airfall deposits,
laminated ash,
porphyritic,
trachytic, pilotaxitic
Plagioclase, olivine,
pyroxene,
hornblende
Chlorite, epidote,
oxide
Lava flow, basal flow breccia
Porphyritic, trachytic,
pilotaxitic
Plagioclase, olivine,
Pyroxene,
Chlorite, epidote, oxide
Sierra de Catorce
Aramberri
Huizachal Valley
Lava dome, dikes, flow
banding
Porphyritic
Quartz, feldspar
Kaolinite, sericite
opaque minerals
Lava dome, lava flow
flow banding
Glassy
Quartz, feldspar
Kaolinite, opaque
minerals (oxide)
Lava flow, basal flow
breccia
Porphyritic, trachytic,
pilotaxitic
Lava flow basal flow
breccia
Vesicular
Plagioclase, olivine,
pyroxene, biotite,
hornblende
Chlorite, sericite,
epidote, oxide
Plagioclase
La Boca
Canyon
Opaque minerals (oxide)
Macrostructure
Airfall deposits,
laminated ash
Ash flow tuff (ignimbrites),
airfall deposits, welded basal
vitrophyre, spherulites
vapour zone, massive welded
zone, eutaxitic structure,
fiammen, volcaniclastic
breccia
Airfall deposits,
laminated ash
Ash flow tuff (ignimbrites),
airfall deposits, laminated
ash, massive welded zone,
eutaxitic structure, fiammen,
volcaniclastic breccia
Texture
Unwelded
Unwelded
Components
Quartz, lithic
fragments, pumice
fragments
Porphyritic, glassy, welded,
unwelded
Quartz, lithic fragments,
pumice fragments
Quartz, lithic fragments,
pumice fragments
Porphyritic, glassy, welded,
unwelded
Quartz, lithic fragments,
pumice fragments
Alteration
minerals
Kaolinite, epidote,
opaque minerals
(oxide)
Kaolinite, sericite, epidote,
opaque minerals (oxide)
Sericite, opaque minerals
(oxide)
Kaolinite, opaque minerals
(oxide)
Ash flow tuff
(ignimbrites), airfall
deposits, spherulites
vapor zone, massive
welded zone, eutaxitic
structure, fiammen,
rheomorphic folding,
ash flows, flow banding
Glassy welded, unwelded
Quartz, lithic fragments,
pumice fragments
Kaolinite, opaque
minerals (oxide)
Airfall deposits,
laminated ash
Unwelded
Quartz plagioclase,
lithic fragments,
pumice fragments,
distorted lapillus
Kaolinite,opaque
minerals (oxide)
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
Rock
55
56
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
Table 4
Major- and trace-element composition of pre-Oxfordian volcanic rocks, northeastern Mexico
CHRG1a
SCH
Dacite
TPZ1b
SCH
Andesite
RCG1a
SC
Dacite
RCG3b
SC
Rhyolite
MZQTG1a
A
Rhyolite
SM1a
SM
Trachyandesite
SM3 b
SM
Trachyandesite
HZCHG1a
HV
Rhyolite
HZCHG2a
HV
Rhyolite
Major elements (wt%)
SiO2
53.16
TiO2
1.25
Al2O3
18.79
4.54
Fe2O3
FeO
1.92
MgO
5.21
MnO
0.11
CaO
2.81
Na2O
6.2
K2O
1.82
P2O5
0.32
LOI
3.55
Total
99.68
65.07
0.48
15.23
5.6
0.79
1.53
Traces
0.67
2.21
3.72
0.16
3.02
98.48
54.50
0.954
16.19
6.37
1.12
6.06
0.081
5.38
3.83
0.76
0.19
4.17
99.60
62.19
2.11
11.85
11.97
0.66
0.75
Traces
2.19
0.08
4.24
1.34
2.35
99.73
78.75
0.235
14.10
0.69
0.13
0.09
0.005
Traces
0.24
3.57
0.05
2.19
100.05
71.97
0.15
13.25
2.06
0.55
1.09
0.02
1.74
0.53
4.59
0.04
3.83
99.82
52.53
0.78
18.34
2.94
4.59
2.23
0.12
4.58
5.52
1.64
0.3
6.42
99.99
54.68
0.638
18.0
6.33
1.12
1.86
0.098
4.03
5.52
2.07
0.32
4.78
99.44
77.34
0.24
12.51
3.16
0.06
0.26
Traces
0.31
0.25
3.17
Traces
2.49
99.79
76.89
0.27
14.24
0.57
0.22
0.36
Traces
0.11
0.21
3.88
0.06
2.74
99.51
Trace elements (ppm)
Rb
57.82
Sr
387
Y
28.6
Zr
212
Nb
12.93
Cs
5.1
Ba
538
La
20.41
Ce
47.53
Pr
5.93
Nd
23.5
Sm
5.12
Eu
1.57
Gd
5.16
Tb
0.76
Dy
4.52
Ho
1.02
Er
2.56
Tm
0.4
Yb
2.42
Lu
0.4
Hf
4.96
Ta
0.943
Th
4.58
U
1.26
154.0
156
36.1
338
17.51
12.96
1987
38.73
81.14
9.85
38.77
7.92
2.12
6.75
1.04
6.18
1.25
3.46
0.53
3.42
0.54
8.62
1.33
10.07
1.82
14
480
18.8
129
5.7
6.6
501
15.1
33.3
4.10
17.9
4.20
1.33
4.04
0.65
3.78
0.77
2.26
0.331
2.03
0.325
3.2
0.35
2.10
0.63
139.3
107
32.9
804
27.74
10.65
1069
128.8
331.3
35.48
129.4
20.68
8.04
13.27
1.51
6.95
1.06
2.73
0.34
2.09
0.29
20.4
1.5
28.55
5.05
90
56
14.4
113
8.9
8.4
318
34.8
68.3
7.43
27.8
5.58
1.59
4.65
0.62
2.94
0.54
1.63
0.245
1.62
0.253
3.4
1.42
8.44
1.15
166.6
30
16.9
142
7.27
7.38
709
25.55
46.74
4.92
16.9
3.17
0.72
2.71
0.41
2.48
0.542
1.54
0.249
1.66
0.27
3.94
0.749
12.54
2.97
51.09
320
20.7
126
5.78
2.16
455
38.68
76.64
8.82
34.83
6.93
2.18
5.25
0.76
3.99
0.66
1.97
0.29
1.98
0.3
3.32
0.424
9.25
2.28
54
251
27.9
120
4.2
4.0
227
39.0
79.9
8.94
35.5
7.0
1.65
5.83
0.88
4.91
1.03
3.18
0.475
3.06
0.502
3.0
0.18
13.8
3.69
94.91
29.1
31.4
230
9.87
2.99
799
27.81
60.28
7.05
27.8
5.69
1.08
4.52
0.77
4.86
1.01
3.04
0.51
3.3
0.52
5.93
0.872
10.44
1.91
94.21
30.5
23.7
219
11.01
3.22
607
41.55
82.97
9.43
35.11
5.9
0.95
4.02
0.65
4.29
0.82
2.51
0.4
2.5
0.43
6.02
0.974
11.56
3.75
11.8
–
16.3
–
43.2
6.6
6.5
–
65.0
10.1
0.1
–
48.3
4.5
3.0
–
1.4
–
8.9
–
3.6
0.1
8.1
–
64.3
8.3
1.5
–
62.5
10.0
0.6
–
Sample
Locality
Type
PBLG1a
SS
Trachyandesite
CIPW normative minerals (wt.%)
qz
–
35.9
c
2.4
7.1
hy
9.9
6.7
ol
2.3
–
Notes: SS, Sierra de Salinas; SCH, Sierra de Charcas; SC, Sierra de Catorce; A, Aramberri; SM, San Marcos; HV, Huizachal Valley. For sample locations,
see Fig. 1.
a
Analyses performed by ICP-MS at Centre des Recherches Pétrographiques et Geóchimiques (CNRS) of Nancy, France.
b
Analysis performed by ICP-MS at Activation Laboratories, Ltd., Ancaster, Ontario, Canada.
ment with the petrographic classification based on the mineralogical composition of the samples.
Sample PBLG1 from Sierra de Salinas is a silica undersaturated trachyandesite with 2.3% normative olivine and
9.9% normative hypersthene (hy). Samples SM1 and SM3
from the San Marcos area in Nuevo León are slightly oversaturated trachyandesite with low normative quartz (qz)
content (1.4 and 3.6%, respectively) and 8.1–8.9% hy. A
sample collected in the northern Sierra de Charcas
(TPZ1) is a silica saturated andesite with 11.8% qz and
16.3% hy. Samples classified as dacites were collected at
Sierra de Charcas (CHRG1) and Sierra de Catorce
(RCG1); they have qz contents of 35.9% and 43.2% and
hy contents of 6.7% and 6.5%, respectively. Samples RC3
(Sierra de Catorce), MZQTG1 (Aramberri area),
HZCHG1 and HZCHG2 (Huizachal Valley) are rhyolite,
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
with variable qz contents between 48.3% and 65.0% and hy
contents of 0.1–3.0%; these rocks are peraluminous with
normative corundum contents varying from 4.5% to 10.1%.
The analyzed samples trend toward lower contents of
most oxides (e.g., CaO, Al2O3, MgO, NaO, TiO2, and
P2O5) as silica increases, whereas K2O tends to be enriched
as silica increases and then diminish at higher silica contents, as is characteristic of calc-alkaline magmas evolving
through the fractional crystallization of plagioclase and
ferromagnesian minerals. The late potassium depletion
could indicate K-feldspar crystallization in the most
evolved rocks. The samples represent a broad region, so
they cannot be treated as comagmatic, though the observed
tendencies prompt us to interpret them as originating in a
common tectonic setting.
Trace-element abundances are shown in a multi-element
diagram (Fig. 4), normalized against the primordial mantle
composition of Sun and McDonough (1989). Although the
57
samples have different trace-element enrichments, their patterns are similar and show common features. The most relevant feature observed in this diagram is a well-developed
negative anomaly in the elements Nb and Ta, considered
one of the most distinctive characteristics of rocks generated by subduction processes (e.g., Hawkesworth et al.,
1993), and generally related to the enrichment of largeion lithophile elements (e.g., Rb, Ba, K) and light rare
earth elements (e.g., La, Ce) in fluids and melts released
from the subducting plate to the overlying mantle. The negative anomalies in Sr, P, and Ti are most likely controlled
by the fractionation of individual minerals, in that they
become more pronounced in the most differentiated rocks.
Because Sr is a fluid-mobile element, the Sr anomalies
could result partly from element mobilization during alteration, though the development of similar anomalies for
relatively immobile elements such as P and Ti as differentiation proceeds indicate that mineral fractionation is the
Fig. 4. Normalized multielement diagrams for samples collected from six different outcrops of pre-Oxfordian volcanic rocks in northeastern Mexico (for
sample location, see Fig. 1). Sample elemental abundances are normalized to the primitive mantle values of Sun and McDonough (1989). A, andesite; TA,
trachyandesite; D, dacite; R, rhyolite.
58
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
dominant process of depletion. In the analyzed samples,
concentrations of P could have been controlled by apatite
fractionation; Sr by plagioclase, and Ti by ilmenite, rutile,
or sphene (Rollinson, 1993).
The patterns in this diagram do not show significant
deviations caused by element mobilization, despite of the
observed alteration in the rocks. The trace-element patterns in the normalized multi-element diagram are strong
evidence of an origin of pre-Oxfordian rocks in a continental arc setting.
Rare earth element (REE) abundances normalized
against chondrite values from Anders and Grevesse
(1989) are shown in Fig. 5. The REE can be used reliably
in this type of rock because they are considered to remain
immobile during alteration. All samples are enriched in
light REE relative to heavy REE. With the exception of
sample RCG1 from Sierra de Catorce, all samples have a
relatively steep slopes for the LREE (La-Eu) and a flat pattern for the HREE (Gd-Lu). The most evolved samples
(northeast area: MZQTG1, HZCHG1, and HZCHG2)
show weak negative Eu-anomalies, indicating plagioclase
fractionation. The REE pattern of sample RCG1 is characterized by a stronger enrichment in LREE relative to
HREE and a continuous decrease of element abundances
between La and Lu.
The REE diagrams show that the pre-Oxfordian volcanic rocks from broadly distributed localities have similar
compositions and probably originated in similar conditions. The differences observed in sample RCG1 could be
attributed to differences in source composition or melting
process, but the available data do not allow a more definitive interpretation.
Discrimination among tectonic settings on the basis of
geochemical data has been proposed in several works
(e.g., Pearce and Cann, 1971, 1973; Wood, 1980; Pearce,
Fig. 6. Tectonomagmatic discrimination diagrams showing the composition of analyzed samples. (a) Th–Ta–Hf/3 ternary diagram of Wood (1980).
(b) Rb vs. Y + Nb diagram after Pearce et al. (1984). Symbols as in Fig. 5.
Fig. 5. Rare earth element diagrams of the pre-Oxfordian volcanic units from northeastern Mexico, normalized to chondrites values of Anders and
Grevesse (1989). A, andesite; TA, trachyandesite; D, dacite; R, rhyolite.
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
1982; Shervais, 1982; Meschede, 1986). The diagram Hf/3–
Th–Ta of Wood (1980) is particularly useful for discriminating altered and metamorphosed rocks because of the
relative immobility of these elements during secondary processes. This diagram was originally proposed for basic
rocks but can be satisfactorily applied to rocks of intermediate to felsic composition (Wood, 1980). The pre-Oxfordian volcanic rocks analyzed in this work plot within the
field of calc-alkaline continental volcanic arcs (Fig. 6A).
To confirm the tectonomagmatic discrimination, the data
were plotted in the Rb vs. Y + Nb diagram (Fig. 6B) proposed by Pearce et al. (1984), which applies to rocks of granitic composition and is thus more indicative for the
evolved rocks. Again, all pre-Oxfordian volcanic rocks plot
in the field of volcanic arc granites (VAG), independently
of their composition. It is noteworthy that the samples plot
in the same field in both diagrams, though Rb is considered
a mobile element during secondary processes.
The general features of the exposed volcanic sequences,
the petrography of the diverse materials, and the geochemical data support the conclusion that all the studied preOxfordian volcanic rocks originated in a continental arc.
Our analyses provide a general idea of the compositional
variations among the sequences or localities, which, however, are common in volcanic arcs composed of different volcanic centers. These variations also document the changes in
composition of all volcanic products during magmatic evolution in space and time. The following observations suggest
an origin of these rocks in a continental volcanic arc:
59
1. Pyroclastic volcanic products, in the form of ashflows
and ignimbrites, airfall tuffs, breccias, and probable lahars and avalanches predominate at all localities, whereas
lava flows, rhyolitic domes, and dikes are present in lesser proportion.
2. The observed volcanism is of eminently subaerial character. The existence of large volcanoes is suggested by
the presence of lahars and possible avalanche deposists.
Furthermore, there is evidence of separate volcanic centers or volcanic fields in the region.
3. The rocks in this study have intermediate to felsic compositions and calc-alkaline characters, as found in volcanic arcs worldwide. The abundance of silica-rich
magmas, partly K-rich, is also typical for continental
arcs. Moreover, the patterns observed in normalized
multi-element diagrams are characteristic of subduction-related rocks.
4. Two different discrimination diagrams that show the
behavior of some trace elements for different rock compositions support an origin of these rocks in a continental volcanic arc.
5. Correlation
The stratigraphic correlation of pre-Oxfordian units in
northeastern Mexico is shown in Fig. 7. In most outcrops
of Zacatecas and San Luis Potosı́ (southwestern part of
the studied region), the volcanic rock sequences overlie
Upper Triassic siliciclastic marine sequences of the Zacate-
Fig. 7. Stratigraphic correlation of the pre-Oxfordian units of northeastern Mexico. The studied volcanic sequences belong to the Nazas Formation,
which overlies Triassic units of marine origin in central Mexico and sequences of continental origin in northeastern Mexico, and underlies or is partly
interlayered with the basal parts of the Middle–Upper Jurassic redbed sequences (La Boca Formation).
60
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
Fig. 8. Tectonic setting of pre-Oxfordian volcanic rocks in a model of continental volcanic arc associated with the development of the active continental
margin of southwestern North America during Late Triassic–Middle Jurassic time.
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
cas and Taray formations (Cantú-Chapa, 1969; GalloPadilla et al., 1993; Gómez-Luna et al., 1998), whereas in
the northeastern part of the region, the underlying rocks
are continental redbed sequences of Late Triassic age
(Weber, 1997; Silva-Pineda and Buitrón-Sánchez, 1999).
Triassic rocks are absent in localities where the volcanic
rocks directly overlie older sedimentary or metamorphic
rocks. In turn, the volcanic rocks underlie or are interbedded in the basal part of redbeds of either La Boca Formation (Mixon et al., 1959) – dated as Sinemurian by RuedaGaxiola et al. (1993) and as Early–Middle Jurassic by
Fastowsky et al. (2005) – or the Upper Jurassic La Joya
Formation (Mixon et al., 1959). Both Jurassic redbed
sequences frequently contain clasts of pre-Oxfordian volcanic rocks and, in some localities, are almost solely constituted by these rocks. La Joya Formation, in turn, is
overlain by the transgressive carbonate sequences of the
Oxfordian–Kimmeridgian Zuloaga Group.
The lithological similarities, stratigraphic position, and
available absolute ages of the localities studied in this work
lead us to conclude that all the volcanic sequences
described here are part of a Jurassic volcanic arc, related
to the active margin of southwestern North America. The
available ages (Tables 1 and 2) indicate that the volcanic
arc was probably active for a period of 40 Ma during the
Jurassic. The volcanic rocks of the described localities correlate with the Nazas Formation of northern Durango and
Zacatecas and therefore represent a key unit for the stratigraphic subdivision, as well as paleogeographic and paleotectonic interpretations of north and northeastern Mexico.
6. Conclusions
The evidence presented herein indicates that the preOxfordian volcanic rocks of Tamaulipas and Nuevo León
are similar in composition, age, and tectonic setting to
those of Durango and Zacatecas and that these sequences
continue toward the western part of San Luis Potosı́. The
tectonic setting, defined on the basis of petrographic and
geochemical studies of the volcanic rocks, is a continental
arc. Fig. 8 summarizes the tectonic evolution of pre-Oxfordian units from northeastern Mexico according to their
stratigraphic positions, geochemical–petrological character, and available paleontological and isotopic age
determinations.
The Upper Triassic and Lower Jurassic rocks from
northeastern México may be interpreted as related to the
evolution of the Cordilleran system (Jones et al., 1995;
Bartolini et al., 2003). During the Middle–Late Triassic,
turbiditic sequences (Zacatecas Formation) related to submarine fans formed at the western margin of North America; subsequent eastward subduction of the Kula plate
caused the first deformation of the turbiditic sequences.
In the Early Jurassic, the development of a continental volcanic arc (recorded by the Nazas Formation) began as part
of the Cordilleran magmatic arc. Finally, the volcanic arc
was eroded and buried beneath the Upper Huizachal
61
Group and La Joya Formation, followed by the eastward
displacement of Mexico along the Mojave–Sonora megashear (Silver and Anderson, 1974; Anderson and Silver,
2005) and/or other Late Jurassic sinistral strike-slip faults.
Roughly coeval to sinistral transcurrent faulting, eastward
subduction of a North American segment of oceanic lithosphere took place under the Pacific plate (Barboza-Gudiño
et al., 1998; Dickinson and Lawton, 2001), resulting in the
development of an intraoceanic volcanic arc complex
(Guerrero Terrane). Since Late Jurassic time, the evolution
of northeastern Mexico has been associated with the evolution of the Gulf of Mexico Basin, and the ‘‘Cordilleran
terranes’’ in this region were covered by the Upper Jurassic–Cretaceous Gulf sequences (e.g., Wilson and Ward,
1993).
The stratigraphic position and existing absolute ages of
the studied volcanic sequences indicate they belong to a
Jurassic arc, related to the active Pacific continental margin
of southwestern North America, rather than to the Permo–
Triassic arc. This older arc has been inferred from isolated
localities east and north of the studied area under the Gulf
of Mexico coastal plain in Tamaulipas and near Coahuila
and Chihuahua (Bartolini et al., 1999; Torres et al., 1999).
The assignment of the studied volcanic sequences to the
Jurassic arc provides an excellent guide for discriminating
between Triassic and Jurassic redbeds, which commonly
have been considered a single unit (La Boca Formation
[Mixon et al., 1959]; Huizachal Formation [Carrillo-Bravo,
1961]; or Los San Pedros Alogroup [Rueda-Gaxiola et al.,
1993), which blurs stratigraphic details and leads to a misunderstanding of the tectonic evolution and main processes
acting in these periods.
Acknowledgments
We acknowledge support from SEP/CONACYT (project 485100-5-25400T and 2002-CO2-41239) and FAI/
UASLP (project C02-FAI-11-27.88) and thank J. Blickwede and C. Bartolini for their revisions and suggestions.
Detailed reviews by the JSAES-designed reviewers, T.H.
Anderson and T.F. Lawton, greatly helped improve this
work.
References
Anders, E., Grevesse, N., 1989. Abundances of the elements: meteoritic
and solar. Geochimica et Cosmochimica Acta 53, 197–214.
Anderson, T.H., Silver, L.T., 2005. The Mojave-Sonora Megashear – field
and analytical studies leading to the conception and evolution of the
hypothesis. In: Anderson, T.H., Nourse, J.A., McKee, J.W., Steiner,
M.B. (Eds.), The Mojave-Sonora Megashear Hypothesis: Development, Assessment and Alternatives. Geological Society of America
Special Paper, vol. 393, pp. 1–50.
Barboza-Gudiño, J.R., Tristán-González, M., Torres-Hernández, J.R.,
1998. The Late Triassic–Early Jurassic active continental margin of
western North America in northeastern México. Geofı́sica Internacional 37 (4), 283–292.
Barboza-Gudiño, J.R., Tristán-González, M., Torres-Hernández, J.R.,
1999. Tectonic setting of pre-Oxfordian units from central and
62
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
northeastern Mexico: A review. In: Bartolini, C., Wilson, J.L.,
Lawton, T.F. (Eds.), Mesozoic Sedimentary and Tectonic History of
North-Central Mexico. Geological Society of America Special Paper,
vol. 340, pp. 197–210.
Barboza-Gudiño, J.R., Hoppe, M., Gómez-Anguiano, M., Martı́nezMacı́as, P.R., 2004. Aportaciones para la interpretación estratigráfica
y estructural de la porción noroccidental de la Sierra de Catorce, San
Luis Potosı́, México. Revista Mexicana de Ciencias Geológicas 21 (3),
299–319.
Bartolini, C., 1998. Stratigraphy, geochronology, geochemistry and
tectonic setting of the Mesozoic Nazas Formation, north-central
México. Unpublished Ph.D. thesis, University of Texas, El Paso, 558
pp.
Bartolini, C., Spell, T., 1997. An Early Jurassic age (40Ar/39Ar) for the
Nazas Formation at the Cañada Villa Juárez, northeastern Durango,
México. Geological Society of America, Abstracts with Programs, vol.
29 (2), p. 3.
Bartolini, C., Lang, H., Stinnesbeck, W., 1999. Volcanic rock outcrops in
Nuevo León, Tamaulipas and San Luis Potosı́, Mexico: Remnants of
the Permian–Early Triassic magmatic arc? In: Bartolini, C., Wilson,
J.L., Lawton, T.F. (Eds.), Mesozoic Sedimentary and Tectonic History
of North-Central Mexico. Geological Society of America Special
Paper, vol. 340, pp. 347–356.
Bartolini, C., Lang, H., Spell, T., 2003. Geochronology, geochemistry, and
tectonic setting of the Mesozoic Nazas arc in north-central Mexico,
and its continuation to north South America. In: Bartolini, C., Buffler,
R.T., Blickwede, J.F. (Eds.), The Circum Golf of Mexico and the
Caribbean: Hydrocarbon Habitats, Basin Formation and Plate Tectonics. American Association of Petroleum Geologists Memoir, vol.
79, pp. 427–461.
Blickwede, J.F., 1981. Stratigraphy and petrology of Triassic(?) ‘‘Nazas
Formation’’, Sierra de San Julian, Zacatecas, Mexico. Unpublished
M.S. thesis, New Orleans, Louisiana, University of New Orleans, 100
pp.
Blickwede, J.F., 2001. The Nazas Formation: A detailed look at the early
Mesozoic convergent margin along the western rim of the Gulf of
Mexico Basin. In: Bartolini C., Buffler R.T., Cantú-Chapa, A. (Eds.),
The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins,
and Petroleum Systems. American Association of Petroleum Geologists Memoir, vol. 75, pp. 317–342.
Cantú-Chapa, A., 1969. Una nueva localidad del Triásico Superior
marino en México. Instituto Mexicano del Petróleo, Revista, vol. 1,
pp. 71–72.
Carrillo-Bravo, J., 1961. Geologı́a del Anticlinorio Huizachal-Peregrina,
al NW de Ciudad Victoria, Tamaulipas. Boletin de la Asociación
Mexicana de Geólogos Petroleros 13 (1-2), 1–98.
Centeno-Garcı́a, E., Silva-Romo, G., 1997. Petrogenesis and tectonic
evolution of central Mexico during Triassic-Jurassic time. Revista
Mexicana de Ciencias Geológicas 14 (2), 244–260.
Córdoba-Méndez, D.A., 1964. Geology of Apizolaya quadrangle (east
half), northern Zacatecas, Mexico. Unpublished M.A. thesis, Austin,
The University of Texas, 111 p.
de Cserna, Z., 1956. Tectónica de la Sierra Madre Oriental de México,
entre Torreón y Monterrey. XX. Congreso Geológico Internacional,
México, 87 p.
Dickinson, W.R., Lawton, T.F., 2001. Carboniferous to Cretaceous
assembly and fragmentation of Mexico. Geological Society of America
Bulletin 113, 1142–1160.
Fastowsky, D.E., Hermes, O.D., Strater, N.H., Bowring, S.A.,
Clark, J.M., Montellano, M., Hernández, R.R., 2005. Pre-Late
Jurassic, fossil-bearing volcanic and sedimentary red beds of
Huizachal Canyon, Tamaulipas, México. In: Anderson, T.H.,
Nourse, J.A., McKee, J.W., Steiner, M.B. (Eds.), The MojaveSonora Megashear Hypothesis: Development, Assessment and
Alternatives. Geological Society of America Special Paper, vol.
393, pp. 401–426.
Gallo-Padilla, I., Gómez-Luna, M.E., Contreras y Montero, B., CedilloPardo, E., 1993. Hallazgos paleontológicos del Triásico marino en la
región central de México. Revista de la Sociedad Mexicana de
Paleontologı́a 6 (1), 1–9.
Gómez-Anguiano, M., 2001. Cartografı́a geológica del área de San Juan
de Matanzas en la porción noroccidental de la Sierra de Catorce, San
Luis Potosı́ y correlación estratigráfica y geoquı́mica de rocas
volcanogénicas pre-Cretácicas del noreste de México. Unpublished
M.S. thesis, Facultad de Ciencias de la Tierra, Universidad Autónoma
de Nuevo León, 164 p.
Gómez-Luna, M.E., Cedillo-Pardo, E., Contreras y Montero, B., GalloPadilla, I., Martı́nez-Cortés, A., 1998. Un nuevo perfil del LadinianoCárnico Inferior con fauna de amonoideos en La Ballena, Zacatecas,
México. Revista Mexicana de Ciencias Geológicas 15 (1), 38–45.
Grajales-Nishimura, J.M., Terrell, D., Damon, P., 1992. Evidencias de la
prolongación del arco volcánico cordillerano del Triásico tardı́oJurásico en Chihuahua, Durango y Coahuila. Boletı́n de la Asociación
Mexicana de Geólogos Petroleros 42, 1–18.
Gursky, H.J., Ramı́rez-Ramı́rez, C., 1986. Notas preliminares sobre el
descubrimiento de volcanitas acidas en el Cañón de Caballeros (núcleo
del Anticlinorio Huizachal-Peregrina, Tamaulipas, México): Actas de
la Facultad de Ciencias de la Tierra, Universidad Autónoma de Nuevo
León, 1 (1), 1–22.
Hawkesworth, C.J., Gallagher, K., Hergt, J.M., McDermott, F., 1993.
Mantle and slab contributions in arc magmas. Annual Review of Earth
and Planetary Sciences 21, 175–204.
Hoppe, M., 2000. Geologische Kartierung (1:10 000) im Gebiet Ojo de
Agua, nordwestliche Sierra de Catorce und sedimentpetrologische
Untersuchungen an prä-oberjurassischen Sedimenten (‘‘Zacatecas
Formation’’). Unpublished M.S. thesis, Technische Universität Clausthal, Germany, 235 pp.
Imlay, R.W., 1936. Evolution of the Coahuila Peninsula, Mexico, part IV,
Geology of the western part of the Sierra de Parras. Geological Society
of America Bulletin 47, 1091–1152.
Imlay, R.W., 1938. Studies of the Mexican Geosyncline. Geological
Society of America Bulletin 49, 1657–1959.
Jones, N.W., McKee, J.W., Anderson, T.H., Silver, L.T., 1990. Nazas
Formation: a remnant of the Jurassic arc of western North America in
north-central Mexico. Geological Society of America, Abstracts with
Programs 22 (7), A327.
Jones, N.W., McKee, J.W., Anderson, T.H., Silver, L. T., 1995. Jurassic
volcanic rocks in northeastern Mexico: a posible remnant of a
Cordilleran magmatic arc. In: Jaques-Ayala, C., Gonzalez-León,
C.M., Roldán Quintana, J. (Eds.), Studies on the Mesozoic of Sonora
and Adjacent Areas. Geological Society of America Special Paper, vol.
301, pp. 179–190.
Le Bas, M.J., Le Maitre, R.W., Streckeisen, A., Zanettin, B., 1986. A
chemical classification of volcanic rocks based on the total alkali-silica
diagram. Journal of Petrology 27, 745–750.
López-Infanzón, M., 1986. Estudio petrogenético de las rocas ı́gneas en las
Formaciones Huizachal y Nazas. Boletı́n de la Sociedad Geológica
Mexicana 47 (2), 1–42.
Martı́nez-Pérez, J., 1972. Exploración geológica del área El Estribo-San
Francisco, San Luis Potosı́. Boletı́n de la Asociación Mexicana de
Geólogos Petroleros 24 (7-9), 327–402.
Meiburg, P., Chapa-Guerrero, J.R., Grotehusmann, I., Kustusch, T.,
Lentzy, P., De León-Gómez, H., Mansilla-Terán, M., 1987. El
basamento precretácico de Aramberri, estructura clave para comprender el ‘‘décollement’’ de la cubierta jurásica-cretácica de la Sierra
Madre Oriental, México. Actas de la Facultad de Ciencias de la Tierra,
Universidad Autónoma de Nuevo León 2, 15–22.
Meschede, M., 1986. A method of discriminating between different types
of mid-ocean ridge basalts and continental tholeites with the Nb–Zr–Y
diagram. Chemical Geology 56, 207–218.
Mixon, R.B., Murray, G.E., Diaz, G.T., 1959. Age and Correlation of
Huizachal Group (Mesozoic), State of Tamaulipas, México. American
Association of Petroleum Geologists Bulletin 23, 757–771.
Pantoja-Alor, J., 1972. La Formación Nazas del Levantamiento de Villa
Juárez, Estado de Durango. Segunda Convención Nacional de la
Sociedad Geológica Mexicana, Memorias, pp. 25–31.
J.R. Barboza-Gudiño et al. / Journal of South American Earth Sciences 25 (2008) 49–63
Pearce, J.A., 1982. Trace element characteristics of lavas from destructive
plate boundaries. In: Andesites: Orogenic andesites and Related
Rocks. Wiley Interscience, New York, pp. 525–548.
Pearce, J.A., Cann, J.R., 1971. Ophiolite origin investigated by discriminant analysis using Ti, Zr and Y. Earth and Planetary Science Letters
12, 339–349.
Pearce, J.A., Cann, J.R., 1973. Tectonic setting of basic volcanic rocks
determined using trace element analyses. Earth and Planetary Science
Letters 19, 290–300.
Pearce, J.A., Harris, N.B.W., Tindle, A.G., 1984. Trace element discrimination diagrams for the tectonic interpretation of granitic rocks.
Journal of Petrology 25, 956–983.
Peccerillo, R., Taylor, S.R., 1976. Geochemistry of Eocene calc-alkaline
volcanic rocks from the Kastamonu area, northern Turkey. Contributions to Mineralogy and Petrology 58, 63–81.
Rollinson, H.R., 1993. Using Geochemical Data: Evaluation, Presentation, Interpretation. Longman Group Limited, UK, 352 pp.
Rueda-Gaxiola, J., López-Ocampo, E., Dueñas, M.A., Rodrı́guez, J.L.,
Minero, M., Uribe, G., 1993. Los anticlinorios de Huizachal-Peregrina
y de Huayacocotla, dos partes de la fosa de Huayacocotla–El Alamar.
II Bioestratigrafı́a, cronoestratigrafı́a y paleoecologı́a del Alogrupo
Los San Pedros. Boletı́n de la Asociación Mexicana de Geólogos
Petroleros 53, 1–29.
Shervais, J.W., 1982. Ti–V plots and the petrogenesis of modern and
ophiolitic lavas. Earth and Planetary Science Letters 59, 110–118.
Silva-Pineda, A., Buitrón-Sánchez, B.E., 1999. Mesozoic redbed floras in
east-central Mexico and their stratigraphic relationships with marine
beds. Geological Society of America Special Paper 340, 151–160.
Silva-Romo, G., 1993. Estudio de la Estratigrafı́a y Estructuras
Tectónicas de la Sierra de Salinas, Estados de San Luis Potosı́ y
Zacatecas. Unpublished M.Sc. thesis, Facultad de Ciencias, División
de Estudios de Postgrado, Universidad Nacional Autónoma de
México, 139 pp.
Silva-Romo, G., Arellano-Gil, J., Mendoza-Rosales, C., Nieto-Obregón,
J., 2000. A submarine fan in the Mesa Central, Mexico. Journal of
South American Herat Sciences 13, 429–442.
Silver, L.T., Anderson, T.H., 1974. Possible left-lateral early to middle
Mesozoic disruption of the southwestern North American craton
margin. Geological Society of America, Abstracts with Programs 6 (7),
955–956.
63
Stewart, J.H., Blodgett, R.B., Boucot, A.J., Carter, J.L., López, R., 1999.
Exotic Paleozoic strata of Gondwanan provenance near Ciudad
Victoria, Tamaulipas, México. In: Ramos, V.A., Keppie, J. D.
(Eds.), Laurentia-Gondwana Connections before Pangea. Geological
Society of America Special Paper, vol. 336, pp. 227–252.
Sun, S., McDonough, W., 1989. Chemical and isotopic systematics of
oceanic basalts: Implications for mantle compositions and processes.
In: Saunders, A., Norry, M. (Eds.), Magmatism in the Ocean Basins.
Geological Society Special Publication, vol. 42, pp. 313–345.
Torres, R., Ruiz, J., Patchett, P.J., Grajales, J.M., 1999. Permo-Triassic
continental arc in eastern México: Tectonic implications, for reconstructions of southern North America. In: Bartolini, C., Wilson, J.L.,
Lawton, T.F. (Eds.), Mesozoic Sedimentary and Tectonic History of
North-Central Mexico. Geological Society of America Special Paper,
vol. 340, pp. 191–196.
Tristán-González, M., Torres-Hernández, J.R., 1992. Cartografı́a Geológica 1:50 000 de la Hoja Charcas, Estado de San Luis Potosı́.
Universidad Autonóma de San Luis Potosı́, Instituto de Geologı́a,
Folleto Técnico, vol. 115, 94 pp.
Tristán-González, M., Torres-Hernández, J.R., 1994. Geologı́a de la
Sierra de Charcas, Estado de San Luis Potosı́, México. Revista
Mexicana de Ciencias Geológicas 11 (2), 117–138.
Tristán-González, M., Torres-Hernández, J.R., Mata-Segura, J.L., 1995.
Geologı́a de la Hoja Presa Santa Gertrudis, San Luis Potosı́:
Universidad Autónoma de San Luis Potosı́, Instituto de Geologı́a,
Folleto Técnico, vol. 122, 50 pp.
Weber, R., 1997. How old is the Triassic Flora of Sonora and Tamaulipas
and news on Leonardian floras in Puebla and Hidalgo, México.
Revista Mexicana de Ciencias Geológicas 14 (2), 225–243.
Wilson, J.L., Ward, W.C., 1993. Early Cretaceous carbonate platforms of
northeastern and east-central Mexico. In: Simo, J.A.T., Scott, R.W.,
Masse, J.P. (Eds.), Cretaceous Carbonate Platforms. American Association of Petroleum Geologists Memoir, vol. 56, pp. 35–50.
Wood, D.A., 1980. The application of a Th-Hf-Ta diagram to problems of
tectonomagmatic classification and to establishing the nature of crustal
contamination of basaltic lavas of the British Tertiary volcanic
province. Earth and Planetary Science Letters 50, 11–30.
Zavala-Monsiváis, A., 2000. Petrografı́a de rocas volcánicas pre-Cretácicas en la parte occidental del Estado de San Luis Potosı́. Trabajo
recepcional, Universidad Autónoma de San Luis Potosı́, 112 pp.