Download as pdf or txt
Download as pdf or txt
You are on page 1of 31

Downloaded from specialpapers.gsapubs.

org on June 27, 2015

Geological Society of America


Special Paper 393
2005

The Taray Formation: Jurassic(?) mélange in northern Mexico—


Tectonic implications

Thomas H. Anderson
Department of Geology and Planetary Sciences, University of Pittsburgh, Pittsburgh, Pennsylvania 15260, USA

Norris W. Jones
James W. McKee
Department of Geology, University of Wisconsin, Oshkosh, Oshkosh, Wisconsin 54901, USA

ABSTRACT

The Taray Formation is a mudstone-rich body of rock characterized by dismembered


beds of sandstone and shale and fragments of a great range of sizes in a finer matrix. The
formation is known only from exposures in northern Zacatecas, in the southern part of
the San Julian uplift. Structurally, it underlies a thick Jurassic sequence that includes
the volcanogenic Caopas, Rodeo, and Nazas formations and younger rocks of La Joya
(volcaniclastic) and Zuloaga (limestone) formations. The base of the Taray is unknown.
The texture and structures of Taray and its tectonostratigraphic association with Juras-
sic volcanic rocks suggest that it is formed of mélange of probable Jurassic age.
Sections of interbedded sandstone and mudstone commonly show progressive
dismemberment. Initial subtabular beds commonly pinch and swell. Further exten-
sion was accommodated by boudinage or by faults, which cut gently across bedding
and dissect sandy layers into lenses and slivers. Progressive stratal disruption formed
isolated, deformed inclusions of sandstone, commonly mixed with volcanic and cherty
debris, encased in dark-gray, fine-grained, foliated, crenulated matrix. Some large
boulders of quartzose sand that slid downslope, crumpling sediments at their bows,
were subsequently encased by matrix and both were distinctly extended so that the
boulders were fractured, whereas the weaker, pliant matrix responded ductilely.
Taray also contains blocks (as much as hundreds of meters in maximum dimen-
sion) of diverse composition including: laminated and massively bedded, light-colored
chert; fragmented volcanic flow rock; fine, quartzose sandstone; and carbonate beds,
some of which contain fusulinids and crinoid debris. Some of these rocks show tight
folds that formed, in some cases, while they were semiconsolidated; later they were
incorporated in muddy matrix as lithified fragments.
The unexposed contact between mudstone-rich Taray Formation and the volca-
nic and volcaniclastic beds of the Nazas Formation also marks changes in the attitude
of planar structures. We interpret Taray as part of a Jurassic(?) accretionary prism.
The tectonostratigraphic sequence at this latitude (west to east) of accretionary prism
(Taray), arc (Nazas, Rodeo, Caopas Formations), and overlying clastic cover (La Joya
Formation), which developed against and upon a margin of Paleozoic and Precam-

Anderson, T.H., Jones, N.W., and McKee, J.W., 2005, The Taray Formation: Jurassic(?) mélange in northern Mexico—Tectonic implications, in Anderson, T.H.,
Nourse, J.A., McKee, J.W., and Steiner, M.B., eds., The Mojave-Sonora megashear hypothesis: Development, assessment, and alternatives: Geological Society of America
Special Paper 393, p. 427–455. doi: 10.1130/2005.2393(16). For permission to copy, contact editing@geosociety.org. ©2005 Geological Society of America.

427
Downloaded from specialpapers.gsapubs.org on June 27, 2015

428 T.H. Anderson, N.W. Jones, and J.W. McKee

brian crust, cannot be extended northward and must either turn abruptly westward
along the southern edge of the Mesozoic Coahuila Island or be truncated and offset,
probably along the Mojave-Sonora megashear.

Keywords: Taray Formation, mélange, accretionary prism, Jurassic, volcanic arc.

INTRODUCTION AND REGIONAL SETTING ash-flow tuff and tuffaceous siltstone, which are locally strongly
deformed, are included within Caopas, Nazas, and Rodeo Forma-
The Mojave-Sonora megashear is postulated to curve from tions (Jones et al., 1995). Outcrops of these pre-Oxfordian, Meso-
Sonora east-southeastward across northern Mexico (Anderson and zoic, commonly subaerial rocks occur within a discrete region
Schmidt, 1983). The inferred trace is constrained by the distribu- south of the inferred trace of the megashear (McKee et al., 1989).
tion of pre–Late Jurassic rocks, especially those in the vicinity of In the southern part of the San Julian uplift a mudstone-rich
the great Monterrey salient of the Sierra Madre Oriental, which is body of rock (Taray Formation), which is characterized by dis-
defined principally by resistant ridges of Oxfordian and younger membered beds of sandstone and shale and by fragments of a great
carbonates. The stratigraphic discontinuity cited as an indication range of sizes in a finer matrix, underlies the volcanogenic rocks
of displacement along the fault occurs between Paleozoic rocks of the Nazas Formation (Fig. 2). The unexposed contact between
(King, 1934, 1944) that underlie the southern part of the Coahuila these two units is interrupted by low-dipping faults. Southwest-
block (also described as a peninsula or island) and pre-Cretaceous ward-directed tectonic transport of the subaerial rocks of the mag-
Mesozoic rocks exposed southward in the San Julian uplift (Cór- matic arc onto fine-grained marine rocks is compatible with the
doba-M., 1964), across the inferred megashear (Fig. 1). asymmetry of folds recorded by units high in the volcanic section
During the late 1970s Anderson learned of regional map- and overlying Oxfordian(?) strata (Anderson et al., 1991).
ping of the relevant pre-Cretaceous rocks being conducted by Our purpose is to: (1) describe the lithologic, compositional,
McKee and Jones in northeastern Mexico. An outcome of their and structural features of Taray that led us to conclude it is part
work in the state of Coahuila is the recognition that initial activity of an accretionary prism (López-Infanzón, 1986), and (2) relate
along the regional San Marcos fault, a west-northwest–striking Taray rocks within the San Julian uplift to older and younger
structure bounding the northern margin of the Mesozoic Coa- units, as well as to the Mojave-Sonora megashear along which
huila Island, is contemporaneous with the age proposed for the the formation has been displaced (Jones et al., 1995).
megashear (McKee and Jones, 1979; 1990). Soon thereafter we
joined to conduct regional studies with the objective of assessing TARAY FORMATION
the postulated stratigraphic discontinuity between the Coahuila
block and the San Julian uplift to the south (Fig. 1). We interpret the dismembered muddy rocks composing
The initial cooperative study focused upon late Paleozoic, Taray Formation as mélange. They were first mapped as a forma-
arc-proximal marine strata and related intrusive rocks, named the tion by Córdoba-M. (1964) during his geologic investigations of
Las Delicias Formation, that underlie at least 6000 km2 of the the San Julian uplift in northern Zacatecas, Mexico (Figs. 1 and
Coahuila block (McKee et al., 1988, 1999). The folded strati- 2). He proposed the name “Taray Formation” for gray and brown
fied section consists mainly of thousands of meters of sand- to phyllite, graywacke, and conglomerate with scattered, bedded
megaboulder-size volcaniclastic debris (mostly andesitic and masses of dolomite and chert exposed along Arroyo El Taray
dacitic) with conspicuous, although volumetrically minor, thick north of the prominent Pico de Teyra.
beds composed of allochthonous basin-margin limestone that is The principal exposure of the Taray Formation occupies
intercalated with fine-grained marine pelagic sediment. Coherent ~60 km2, west of Sierra El Solitario de Teyra and around Pico
or semicoherent allochthonous bodies of carbonate and volcanic de Teyra (Figs. 2 and 3). Access to the area is by very poor roads
rocks that moved down the flanks of the sedimentary basin under and arroyos and most of our field work was done on foot. Taray-
the influence of gravity may include blocks with dimensions of like rocks also are exposed in an area of a few hundred square
hundreds or thousands of meters. One slab may encompass a few meters in extent along the road just east of Caopas (Fig. 2); these
hundred square kilometers. Neither the bottom nor the top of the outcrops, which are 15 km north of the main exposure, are sur-
section is exposed. The unit includes strata with ages that fall rounded by volcanic rocks of the Jurassic Caopas-Rodeo-Nazas
within the interval from late Mississippian(?) through most of the Formations (Jones et al., 1995). López-Infanzón (1986) reported
Permian (McKee et al., 1999). Taray Formation to the east of Cerro Europa, a hill ~4 km north
South of the Delicias terrane across the inferred megashear, of Caopas and halfway between Caopas and Rodeo, but we were
which is covered by Cretaceous strata preserved in the Parras unable to verify this occurrence.
basin, major exposures of pre-Cretaceous rocks crop out within The Taray Formation is a mudstone-rich body of rock char-
the San Julian uplift. Among the hills of the uplift, thousands acterized by: (1) dismembered beds of sandstone and shale, and
of meters of flows, breccia, laharic conglomerate, air-fall tuff, (2) fragments or blocks up to hundreds of meters long, in a fine-
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico

Figure 1. Map of Mexico showing postulated Jurassic tectonic discontinuities and locations of: (1) towns, (2) exposures of Jurassic, Triassic, and late Paleozoic rocks, (3) terranes, and
429

(4) tectonostratigraphic domains mentioned in text.


Downloaded from specialpapers.gsapubs.org on June 27, 2015

430 T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 2. Geologic map of the San Julian Uplift. The Upper Zuloaga Formation was not mapped in the Sierra Solitario de Las Teyra and is in-
cluded there under the symbol for post-Zuloaga formations, undifferentiated.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 431

grained matrix. Many exposures of matrix composed of dark- base of the Nazas, led us to the conclusion that these parts of the
gray, fine-grained platy rocks are characterized by closely spaced Taray had been within reach of a volcanic source other than that
cleavage that may be planar (poorly to well developed) or phacoi- related to oceanic volcanism.
dal. Locally the foliated rocks record crenulations. In places we We discuss the enigmatic Taray-Nazas contact in a separate
were tempted to label the curved anastomosing cleavage planes section of this report that is devoted to the structural and strati-
in matrix as scaly cleavage (e.g., Lundberg and Moore, 1986; graphic relationship between those units. The base of the Taray
Pini, 1999, fig. 12A therein) equivalent to that which character- is not exposed.
izes the tectonosomes of the argille scagliose of the Bologna The age of the principal components of the Taray Forma-
Apennines of Italy (Pini, 1999). However, we are not confident tion, such as the common gray, fine-grained foliated rocks,
that the cleavage in the Taray is restricted to bodies comparable to is unknown because (1) the stratigraphic position is not well
tectonosomes nor does it exhibit typical splintery character. constrained, (2) the only fossils are in clasts, and (3) no isotopic
Parts of the Taray contain tuffaceous horizons. Although dates are available. Córdoba-M. (1964) tentatively suggested an
these bodies of rock may be compared to progressively extended Early Devonian to Early Pennsylvanian age based on its per-
sequences of mudstone, tuff, chert and sandstone of Cowan’s ceived lithologic similarity to rocks of that age in the Marathon
(1985) type II mélange, their stratigraphic position, close to the region of West Texas. Rare clasts (up to boulder size) of crinoidal

km

Figure 3. Geologic map of the principal area of exposure of the Taray Formation.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

432 T.H. Anderson, N.W. Jones, and J.W. McKee

A B

Figure 4. Conglomerate in Taray Formation. (A) Clasts in near-vertical beds are mainly quartz-rich sandstone and chert; note the gentle dip of
cleavage. Hammer length = 36 cm. (B) Rare conglomerate with rounded cobble of muscovite granite. Scale in cm.

limestone are assumed to be Paleozoic. Some limestone cobbles Blanco (San Luis Potosí and Zacatecas), Zacatecas (Zacatecas)
contain fusulinids that were identified as Pennsylvanian by Mer- and Arteaga (Michoacan) distinguish a domain of marine
lynd Nestell (1983, personal commun.). Detrital zircons from clastic rocks, commonly strongly disrupted by layer-parallel
sandstone in Taray yield ages as young as 270 Ma, suggesting extension. These strata occur beneath Jurassic volcanic units
that the sandstone formed between Late Permian and Nazas time (Nazas Formation and others; Blickwede, 1981) of the Sierra
(Middle Jurassic) (Diaz-Salgado et al., 2003). López-Infanzón Madre terrane, as well as Early Cretaceous volcanic rocks of
(1986) and Barboza-Gudino et al. (1999) assigned a Late Triassic the Guerrero terrane (Centeno-García and Silva-Romo, 1997).
age on the basis of their correlation of Taray with the fossiliferous We propose that the mélange-like structures of Taray formed
Zacatecas Formation (Burckhardt and Scalia, 1906). during east-northeast–directed, Middle Jurassic convergence
The Zacatecas Formation (Carrillo-Bravo, 1968), exposed and that the uppermost units of the Taray may be transitional
in the vicinity of the city of Zacatecas, links Taray Formation into the Nazas Formation.
to other lithologically similar bodies of fine-grained clastic
rocks (e.g., Varales lithofacies of the Arteaga Complex; Cen- LITHOLOGIC COMPONENTS OF THE TARAY
teno-García et al., 2003) and locally mafic marine rocks to the FORMATION
southwest. Together the widespread exposures of comparable
strata at Taray (Zacatecas), Charcas (San Luis Potosí), Peñon Overview

Much of the matrix of Taray Formation is gray to dark-


gray, light-gray-weathering argillite, slate, or phyllite that
shows scant evidence of bedding. However, in most outcrops
some sandstone or metasandstone is present, either in lenses
or in rare continuous beds. Distinctly bedded sections of sand-
stone, mudstone, conglomerate, and pebbly mudstone that
occur within the Taray are generally within blocks.
Conglomerate, though a relatively minor component, is
widespread both as beds and within blocks. Some conglomer-
ate consists of lenses of granules or pebbles of sandstone or
slate that were probably derived internally. Other conglomerate
bodies contain subrounded to rounded clasts that suggest a his-
tory of subaerial transport. Among these latter conglomerates,
components include: (1) subangular to rounded granules of
metaquartzite or vein quartz, (2) rounded pebbles or cobbles of
chert, (3) subrounded cobbles to boulders of sandstone (Fig. 4A),
Figure 5. Quartzite boulder in Taray Formation with person for scale. (4) pebbles and cobbles of carbonate, and (5) rounded cobbles
Note crumpled sediments to right of boulder. of granitoids (one exposure only) (Fig. 4B). A conglomerate
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 433

of carbonate are of several types: some are crinoid-bearing


limestone; some contain dolomite intimately mixed with vol-
canic material (near La Cejita mine, Fig. 3); and, on Cerro El
Pedernal, ankerite occurs interbedded with magnetite in a lami-
nated to thin-bedded metasediment. Some masses of intermedi-
ate to mafic volcanic rock occur near the southeastern edge of
the Taray exposure, but they may be fault-bounded slivers of the
overlying Nazas Formation.

Matrix

Exposures of massive, fine-grained matrix of the Taray


consist of black, greenish-black, gray or silvery gray phyllitic
metashale or metasiltstone that is commonly foliated, locally
Figure 6. Thin-bedded chert in block in Taray Formation. Note isocli- crenulated, and metamorphosed to low greenschist facies. Bed-
nal fold near centimeter scale. ding is difficult to discern in many outcrops.
Thin sections show that grain size is typically coarse silt to
clay, but scattered sand-size grains or lensoid granule-size intra-
clasts are common (Fig. 7). Optical microscopy and X-ray dif-
exposed near the western side of the Taray exposures contains fraction (XRD) show that subangular quartz is invariably present,
cobbles of crinoid- and fusulinid-bearing limestone. 2MIIb clinochlore occurs in nearly all samples, and fine-grained,
Blocks up to several hundred meters long, characterized 2M1 white mica (illite, sericite, or muscovite) is a common, but
by coherent bedded strata, and other large clasts are distinctive not invariable, component. Albite is less common, and biotite is
components of the Taray. Clasts, commonly lensoid, of quartz- generally absent or a minor component. Rounded zircon was
ose sandstone that range from pebble to house-size boulders are identified in some.
most common (Fig. 5). The blocks or inclusions up to 300 m Foliation, generally evident in thin section, is shown (Fig. 8)
long consist of thin- to medium-bedded, light-gray to black by: (1) common extinction of sheet silicates, (2) the presence of
chert, some of which was isoclinally folded before incorpora- wisps of white mica and chlorite, (3) the presence of aligned iron
tion in the Taray (Fig. 6). A complex block, or blocks, contain- oxide, or (4) the orientation of lensoid intraclasts. Thin seams
ing chert, diabase, scoriaceous material, and turbiditic wacke of opaque minerals (Fig. 7) that separate lensoid domains are
forms Cerro El Pedernal (Fig. 3; Klein et al., 1990). Inclusions interpreted as the residuum of dissolution and define foliation.

Figure 7. Photomicrograph (plane light) of matrix and a few clasts of a Figure 8. Photomicrograph (crossed polars) of metasiltstone composed
conglomeratic phyllite. Matrix and siltstone and sandstone clasts are lens- of quartz, muscovite, albite, and clinochlore. Extension resulting in the
oid as a result of the development of slip surfaces, commonly parallel to formation of slip surfaces parallel to compositional layers and small
compositional layers, and small normal faults that transect the layers and normal faults that transect layers has juxtaposed slightly different
intersect the slip surfaces at acute angles; opaques accentuate foliation. lithologies and produced lenses. Lenses, common orientation of sheet
Minerals present include quartz, albite, muscovite, and clinochlore. silicates, and iron oxides on cleavage surfaces define the foliation.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

434 T.H. Anderson, N.W. Jones, and J.W. McKee

Layers, or zones that resemble beds, are evident in some samples.


Dark beds contain more iron oxide and/or chlorite with a strong
preferred orientation. Lens-shaped fragments also define crude
bedding. Siliceous lenses that resemble chert in hand specimen
contain very fine grains of quartz. Some grains may be derived
from metamorphosed rock as suggested by irregular domains
with sutured boundaries and elongate crystal aggregates. In some
samples, fractures filled with iron oxide cut the foliation and bed-
ding. Quartz veins both cut and parallel foliation. Quartz veins
in some samples have very irregular shapes and appear to have
been deformed.
Locally in the topographically higher parts of the Taray,
light-colored clasts and matrix appear to be tuffaceous. We tried
to separate these rocks from the rest of the Taray, but the transi-
tions are so subtle that they are not mappable. A thin section Figure 9. Photomicrograph (crossed polars) of fine-grained meta-ash-
of what we interpreted to be light-gray, very fine-grained tuff tuff, which consists of quartz, sericite, plagioclase, and particles of
(Fig. 9) reveals aligned folia of sericite (composing ~50% of probable devitrified volcanic glass. Some quartz fragments are elon-
the rock), quartz and plagioclase crystals, less than ~0.03 mm gate, curved, and angular.
in diameter, and chert-like material. Some quartz has shard-like
shapes; fragments tend to be elongate, curved and angular. The
chert-like material does not clearly occur as grains, but seems to
be matrix; it could be devitrified glass. (Fig. 11). Unfortunately, we were unable to map these sandstones
as a distinct unit.
Sandstone Texturally, all of the sandstones are wackes, but the grain
mineralogy is variable (Table 1, Figs. 11, 12, 13, 14). The princi-
The most prominent masses of sandstone occur as lensoid pal grains are quartz (composing from ~18–70% of the sample),
masses with visible dimensions up to 5 m thick and 10 m long. plagioclase (trace to 18%), and lithic fragments (0–25%); matrix
The sandstone beds vary in composition, but most are quartz- ranges from 18% to 65%. Sedimentary chert, volcanic chert, and
rich. Color varies from dark gray to buff to nearly white, but fine-grained polycrystalline quartz are difficult to subdivide.
most are gray. Grain size varies from fine- to coarse-sand and Differences among individual quartz grains suggest the
angularity from angular to rounded. Slightly flattened grains and existence of multiple provenances. Many grains show undula-
wisps of dark material, probably iron oxides, define weak folia- tory extinction or polygonization, and some show deformation
tion in some sandstone beds (Fig. 10). Sorting ranges from poor lamellae typical of highly sheared rocks suggestive of deriva-
to moderately well, and sandstones are both grain and matrix tion from metamorphic rocks (Fig. 12). Some grains show
supported. very straight extinction, are slightly embayed (Fig. 11), or
Distinct beds (versus lenses) of sandstone are rare. Where have inclusions indicating a volcanic source. In one wacke,
present, they show some graded bedding and channeling and may quartz is angular and has pressure fringes of sericite and chlo-
be part of an exotic block. Beds are mostly thin, 2–5 cm, but there rite that parallel the plane of foliation. Plagioclase varies from
are rare beds of sandstone up to 1 m thick. fresh grains with good sharp twins to grains that are slightly to
Some sandstone bodies contain inclusions. In one exposure, strongly altered (Fig. 13). In a few samples, twin lamellae are
massive sandstone contains clasts of quartz sandstone ~0.6 m bent, but most are undistorted. Much feldspar is untwinned, but
long, a bed or elongate fragment of chert, and small, 1–10 cm, staining for K indicated there is very little K-feldspar. Other
variously colored wisps of fine sandstone that probably were minerals include clinochlore (which may have formed from
formed before lithification. detrital Fe-Mg minerals), muscovite, and iron oxides. Zircon
In the topographically upper part of the Taray Formation, and rare hornblende are accessory minerals. Rock fragments
some of the sandstone becomes feldspathic. These beds are include volcanic rocks (ranging from very fine-grained frag-
dark-gray to moderate brown, rusty weathering, fine- to coarse- ments that are unidentifiable to chert-like siliceous rocks to pla-
grained quartz wackes. Some possess a slight foliation and may gioclase-phyric andesite), shale (some with fine quartz grains),
even be somewhat schistose. Muscovite occurs in some as both chert, chlorite (probably derived from clasts), sandstone, and
detrital and metamorphic grains; biotite is rare. Some have scat- quartzite (or recrystallized portions of vein quartz).
tered clasts of shale fragments, milky quartz, feldspars, and even The matrix of the sandstones is generally composed of very
carbonate, both as grains and as matrix. In thin section the princi- fine-grained quartz and a fine-grained white mica and clino-
pal difference from the other sandstones is that lithic fragments, chlore. In some rocks it is difficult to distinguish matrix from
especially volcanic rock fragments, are somewhat more abundant distorted rock fragments.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 435

0.5mm
1mm

Figure 10. Photomicrograph (plane light) of fine-grained sandstone from Figure 11. Photomicrograph (crossed polars) of medium-grained,
lensoid body in Taray Formation. Angular to subangular quartz, plagio- grain-supported wacke with grains of quartz, plagioclase, and volca-
clase, and detrital muscovite (lower left) are contained in a matrix of very nic rock fragments. Note embayed volcanic quartz in rock fragment.
small crystals of silicates, including quartz and clinochlore, that may be Sample from lensoid body in Taray.
the remains of highly altered volcanic rock fragments. Note the foliation.

range from tabular blocks, up to 2 m by 30 cm, to 1 cm lenticular


clasts. In some outcrops, beds pinch and swell in a manner sug-
Broken Formation and Pebbly Mudstone gesting soft-sediment deformation. In thin section, most clasts are
lensoid (Fig. 16); in some, fine-grained phyllitic material interfin-
Discontinuous layers of sandstone, recording layer-parallel gers with sandstone so that the term clast is inappropriate. Entirely
extension, that occur in a matrix of metashale/siltstone are compa- disrupted sand bodies that are mixed in an irregular fashion within
rable to broken formation (Hsü, 1974; Kusky and Bradley, 1999) argillaceous matrix are similar to pebbly mudstone as proposed
(Fig. 15). Sand horizons in less-disrupted zones are segmented by Hsü (1974) and discussed by Kusky and Bradley (1999). Most
and the broken aspect is clear. Sand bodies in some of these zones common are pebble- to cobble-size, rounded or elongate lenses of

TABLE 1. POINTS COUNTED ON TARAY SANDS

A. No associated volcanic rocks*


Sample: 3 4 5 6 7A 7B 10A 10B 11A 11B 19B 29 30B 38 40B 40C
Qm 19.7 22.7 26.7 46.7 42.0 39.7 53.7 59.7 18.3 26.0 33.3 50.7 19.7 34.0 45.3 33.3
Qp – 0.7 24.7 6.0 23.0 27.3 4.7 10.3 1.0 2.3 1.3 5.7 2.0 3.7 5.0 25.7
P 18.3 18.0 13.3 12.7 0.3 9.0 4.3 1.7 13.7 10.3 17.3 15.0 7.3 18.0 16.7 8.3
K 4.7 2.3 1.7 – – – – – 1.7 – –
Lv 7.3 8.7 – 3.3 1.0 1.7 – – 4.3 5.3 12.3 5.3 2.7 18.3 1.3 3.3
Ls 1.3 – – 0.7 – 0.7 – – – – 6.3 0.3 3.3 7.7 0.3 10.7
M 48.7 47.7 33.7 30.7 33.7 21.7 37.3 28.3 61.0 56.0 29.3 22.3 65.0 18.3 31.3 18.7
B. Volcanic rocks associated*
Sample: 3 8A 8B 8C 8D 9C 9D 9E 10A 10B 10C 11A 11C 14B
Qm 40.3 28.0 25.3 24.0 32.3 44.3 27.0 18.3 23.0 24.0 20.3 17.3 16.7 28.0
Qp 13.7 2.0 1.3 3.3 8.7 13.0 4.3 5.7 2.3 3.7 1.0 42.0 18.0
P 7.3 7.7 13.7 13.0 9.0 2.0 4.3 10.7 11.0 18.7 10.3 18.3 10.7 7.0
K 0.3
Lv 4.0 9.7 6.3 4.7 8.3 3.7 2.3 5.7 7.0 11.0 22.3 3.0 11.7 1.7
Ls 4.0 8.3 6.3 0.3 0.7 2.0 4.0 3.0 3.0 7.7 2.0 2.7
M 30.3 44.3 47.3 58.0 46.3 39.3 53.3 53.3 50.3 37.7 35.7 58.3
16.3 45.3
Note: 300 points counted, values in percent. Explanation of symbols: Qm—monocrystalline quartz grains; Qp—polycrystalline quartz grains;
P—plagioclase grains; K—K–feldspar grains; Lv—volcanic lithic fragments; Ls—sedimentary lithic fragments; M—matrix and other grains.
*No rocks associated with the sandstones in Table 1A are thought to have a volcanic component vs. those in Table 1B. Comparison of Tables 1A
and 1B shows that the sands themselves do not reflect this difference. Rocks associated with the sandstones in Table 1B are thought to have a
volcanic component vs. those in Table 1A.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

436 T.H. Anderson, N.W. Jones, and J.W. McKee

1mm

Figure 12. Photomicrograph (crossed polars) of a coarser-grained Figure 15. Example of “broken formation” in which what probably were
sandstone lense. Quartz in this sample is variable; extinction varies once-continuous layers of fine-grained sandstone in a shale matrix have
from undulatory to straight, and some lithic grains are polygonized. been dismembered and pinched out. Note overturned slump(?) fold in
Plagioclase, sandstone, and volcanic fragments also are present. detached sandstone layer, upper left corner. Field of view ~3.5 m.

Figure 13. Photomicrograph (crossed polars) of wacke with grains of Figure 16. Photomicrograph (crossed polars) of “broken formation”
plagioclase (note albite twinning), quartz, both volcanic and sedimen- with grains and lenses of siltstone and sandstone contained within a
tary lithic fragments, and a few grains of K-feldspar. foliated, fine-grained matrix.

Figure 14. Photomicrograph (crossed polars) of wacke with abundant Figure 17. Photomicrograph (plane light) of black chert cut by quartz
sedimentary and volcanic(?) lithic fragments. veins, some of which display ptygmatic folds. Chert contains small
amount of clinochlore.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 437

sandstone that occur as inclusions within finer-grained matrix. The iron oxides, though concentrated relative to the rest of the rock, are
clasts of sandstone typically are composed of quartz with some more disseminated near gradational contacts.
trace amounts of clinochlore and white mica between grains. Gen- Clinochlore and muscovite are detectable with XRD in all
erally the fragments are grain supported although some are matrix chert samples, and finely disseminated iron oxide is common in
supported. Grain shapes range from angular (most common) to some chert samples. Small, somewhat rounded particles of iron
subrounded. Almost all quartz grains show undulatory extinction. oxides are less common, and fractures may contain iron oxide.
Also present are irregular-shaped, angular grains of plagioclase Some chert has a volcanic affinity as indicated by possible
commonly with bent twin lamellae. One clast has layers of fine- relic shards and rare angular fragments of volcanic rock. Textural
and very fine-grained sand. Many sandstone fragments have a properties suggestive of clastic origin in some cherts are consis-
slight foliation indicated by common extinction of sheet silicates tent with the volcanic affinity. For example, lamination in at least
and by thin seams of iron oxide (Fig. 10). one sample is due to variation in grain size. Some laminae show
Sandstone and shale clasts clearly were derived from the fine cross-bed–like texture as though some laminae were eroded
rock that contains them, but other rare clasts of quartzite, granite, before the next was deposited. This is especially noticeable where
and chert are exotic. Exotic clasts include quartzite with coarse there are small concentrations of slightly coarser chert contained
quartz grains showing undulatory extinction and sutured bound- within a finer-grained matrix. In some cases, these coarse ovoids
aries and a clast consisting of quartz and perthitic feldspar (gran- are truncated by the next bed. In some layers, shreds of white
ite). There are a few fragments of very irregularly shaped chert mica are abundant and define a bedding-parallel foliation. We
that are difficult to distinguish from the matrix. found probable radiolaria in only one thin section, but did not
Separating discontinuous layers of sandstone and other search exhaustively.
disrupted sand bodies is matrix of metashale/siltstone. Typically
the foliated fine-grained matrix contains a very fine-grained mix- Composite Blocks
ture of aligned, finely crystalline flakes of white mica and very Some large, matrix-poor blocks are lithologically and struc-
fine grains of quartz. Individual grains are very poorly defined turally heterogeneous, containing volcanic, carbonate, chert, or
with indistinct borders. Aligned iron oxides or concentrations of siliciclastic sedimentary rocks that bear no structural relation
organic material also help to define the foliation or cleavage. As to one another. The most conspicuous of these forms Cerro El
in some of the other rock types, there are microcrenulations that Pedernal (Fig. 3; Klein et al., 1990), which is capped by a block
provide evidence for a second deformation. of chert several hundred meters long. On the north side of that
Shale clasts also occur within sand, sometimes as pebble- hill, there are a few poor outcrops of highly altered volcanic and
like fragments, sometimes as irregular sheets. Some shale con- volcaniclastic rocks including a black, laminated, water-laid (?)
tains irregular fragments and stringers of sandstone, as though tuff, a coarse clastic rock that we interpret to be a submarine flow
they had been mixed and smeared, probably during soft-sediment of debris, and an enigmatic laminated ankerite-magnetite rock.
deformation of sand beds. Layering in the tuff is due to differences in grain size and
relative amounts of sheet silicates and quartz. The sheet silicates
Blocks (Or Inclusions) are clinochlore and a yellowish biotite; they occur as masses of
very small, diversely oriented, indistinct crystals (Fig. 9).
Chert The debris-rich unit is a dark-gray to black rock consisting
Blocks of well-bedded chert may be more than 100 m in of a variety of dark fragments. The mass is so altered that it is
maximum exposed dimension. Such masses of chert may form difficult to identify rock types, even in thin section. Zones or
isolated hills; chert is the most prominent and resistant element domains with different textures are evident; we infer the presence
of Cerro El Pedernal (Fig. 3). Isoclinal and chevron folds (Fig. 6) of sedimentary rocks, including chert and mafic slate, based upon
are conspicuous in these exotic cherts. These folds were formed grain size and the abundance of mafic material. Chlorite is locally
prior to deposition in the Taray, as shown by the strong discor- abundant. Some clasts are replaced by very small, poorly defined,
dance among bedding attitudes. Bedded chert is white to light gray yellowish-green grains of a sheet silicate. In these same zones,
to black and weathers brownish-gray to pale buff. Beds typically there are curved, angular pieces of quartz that are reminiscent of
are 1–5 cm thick but may be up to 1 m. Black chert occurs as thin volcanic shards.
beds in folded metashale at a few locations. At one, chert occurs Another rock in the Pedernal assemblage is a black, thin-
as discontinuous, 10–200-cm-long, 1–3-cm-thick bed-like clasts bedded magnetite-ankerite rock. Beds are 0.5–2 cm thick and are
that pinch out or end abruptly. Some of these clasts are isoclinally graded in terms of mineral percentage. In thin section, the contact
folded. In thin section, chert shows typical texture with zones or between nearly pure ankerite and nearly pure magnetite is sharp.
veins of quartz (some of which are ptygmatically folded and/or off- The magnetite near the contact consists of poorly defined grains
set on small faults; Fig. 17). Thin beds of slaty shale that may occur that are ~0.02 mm in diameter (the same size as the ankerite in
between chert beds range from ~0.5–5 cm thick. The contacts this iron-rich part of the sample). In reflected light the iron oxide
between the shale and chert range from very sharp to gradational. is milky, suggesting the presence of titanium; XRD indicates it
Sharp contacts have iron oxide concentrated along them, whereas is magnetite. Upward(?) there is a gradational change with a
Downloaded from specialpapers.gsapubs.org on June 27, 2015

438 T.H. Anderson, N.W. Jones, and J.W. McKee

decrease in magnetite and an increase in amount and grain size


of carbonate. Near the presumed top of a bed, ankerite comprises
~95% of the rock and consists of crystals or zones of relatively
homogenous carbonate in which the individual grains or particles
are ~0.2–0.3 mm across, and in which there are some magnetite
grains between the carbonates. In some parts of the sample there
are also small, poorly defined grains of quartz.
On the south side of Cerro El Pedernal, there are a few small
outcrops of diabase, which we interpret as another block in the
mélange. The diabase is highly altered, but the original diabasic
texture is easily recognizable. Plagioclase is altered to sericite
and some calcite and Fe-Mg silicate minerals have been com-
pletely replaced by chlorite and calcite. On the southeast corner
of the hill, there is an outcrop of proximal turbidite wacke with
15–30-cm-thick beds of sand interbedded with 2-cm-thick beds Figure 18. Photomicrograph (crossed polars) of dolomite-magnesite-
of shale. The beds are overturned, as evidenced by burrows and talc rock from the La Cejita talc mine. Photo shows subhedral crystals
sole markings. Because it is unlike other parts of the Taray, we of dolomite in matrix of finely crystalline magnesite and talc. Also
consider it to be part of an exotic block. present are quartz and clinochlore.
Another large composite block contains La Cejita talc mine
(Fig. 3). The mine consists of a number of small pits dug into
light greenish-yellow to dark greenish-gray zones. Optical and
XRD analysis of five samples from these pits indicate that clino-
chlore is abundant and talc is not; presumably the talc has been
mined out. Also in the mine areas are scattered, rusty, siliceous
(veined) dolomite blocks up to 10 m or so in exposed diameter.
In thin section, abundant dolomite occurs as closely packed
crystals or as crystals that are completely surrounded by mag-
nesite (Fig. 18). Dolomite crystals are randomly distributed or
concentrated along certain zones. The magnesite is finely crystal-
line and massive and was identified by XRD methods. Scattered
throughout are grains of quartz that appears to be detrital. Talc
and clinochlore are minor constituents.
Northeast of the mine and exposed along a ridge, dolomitic
rocks have a variety of textures, some of which, mesoscopically,
appear as though the dolomite contains volcanic clasts. In these Figure 19. Photomicrograph (crossed polars) of mixed volcanic and
rocks, dolomite and magnesite (not distinguishable in thin sec- carbonate rock from ~150 m east of the La Cejita talc mine. Dark areas
tion) occur as very irregular-shaped patches and aggregates of are mainly clinochlore, light areas are dolomite-magnesite; scattered
anhedral grains, intimately mixed with small amounts of chert- quartz grains are also present.
like material of variable grain size (Fig. 19). There are also less
common patches of clinochlore and fractured, equant grains up
to 0.5 mm across of an opaque mineral. The texture and mineral-
ogy of this rock are consistent with our field interpretation of a association of chert with volcanic rocks and a laminated ankerite-
mixture of volcanic material and carbonate, and it is possible that magnetite rock suggests a submarine environment that may have
metasomatism of a volcanogenic parent produced this rock. included active hydrothermal vents. The turbidite near Cerro El
Blueschist and other rocks containing other high-pressure Pedernal likely accumulated offshore, below wave base, perhaps
minerals, and peridotite and related ultramafic rocks, are not in deep water. The mixed volcanic-carbonate rock near La Cejita
known from Taray Formation. Despite the absence of these mine may have been deposited in shallow water with an active
lithologies common in occurrences of mélange, some blocks volcano nearby.
in the Taray are exotic and indicate contributions from mul-
tiple sources. One source contains fossiliferous limestone of METAMORPHISM
late Paleozoic age; another provided chert that probably was
deformed as soft sediment prior to inclusion in the Taray. The Several small plutons of Tertiary(?) granite intrude the
few blocks containing volcanic rocks among siliceous strata are Taray and nearby Nazas Formations near the southwest limit of
consistent with submarine volcanism. At Cerro El Pedernal, the the San Julian uplift, including the one that forms Pico de Teyra
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 439

(Fig. 3). Contact metamorphism in the vicinity of Pico de Teyra Fabric elements measured in blocks of cherty rock that are
has formed fine-grained spotted schist; the contact metamorphic encased in matrix are generally distinguished on the data plots,
aureole is tens to hundreds of meters wide. The schist has a lepi- because we were unable to determine unambiguously how they
doblastic texture due to white mica and porphyroblastic texture related to deformation common to matrix and aligned fragments.
due to small (0.03 mm) porphyroblasts of andalusite and biotitic Structures, such as isoclinal folds, within blocks of cherty strata
mica. In places, as the contact is approached, quartz veins and are generally discordant with foliation and folds in matrix, how-
dikes are present and the Taray appears to be somewhat silicified ever, exceptions exist.
and hydrothermally altered. Elsewhere, the degree of metamor-
phism is less and does not seem to be related to known plutons. Planar Structures
Regionally, the shale-siltstone protolith is of low green-
schist facies. Away from intrusive rocks, the only metamorphic Bedding
minerals noted in thin section are chlorite and white mica. Most In general, the measurements of bedding are orientations of
chlorite is clinochlore IIb, as indicated by brown anomalous the bounding surfaces of tabular fragments of sandy beds against
interference colors, length fast character, and XRD, but in a few matrix (Fig. 21A). Some beds clearly occur within blocks sur-
thin sections, two chlorites were noted, one with brown and the rounded by matrix; other beds form seemingly coherent sections
other with blue anomalous colors. The fine-grained white mica is but are very likely part of a large block, the boundaries of which
muscovite, based on XRD. Some larger grains of muscovite and are not exposed or recognized. Most sections of interbedded
biotite probably are detrital. sandstone and mudstone record progressive dismemberment and
Metamorphic grade near the La Cejita mine is somewhat are probably not part of a block (Fig. 15). Initial attenuation is
higher than in surrounding rocks, and it is assumed that the min- suggested by subtabular beds that pinch and swell or by those
eral assemblages are products of contact metamorphism (there that show boudinage. Numerous small faults that generally show
are small intrusions in the area, although no plutonic rocks are normal separation further dissect the beds into lenses and slivers
exposed at the mine). Microscopic and X-ray studies reveal the (Fig. 22). Stratal disruption yields an end product of isolated,
following mineral assemblages: clinochlore-Fe-oxide, quartz- aligned inclusions of sandstone, commonly mixed with volcanic
clinochlore, quartz-clinochlore-albite, dolomite-talc-Fe-oxide- and other clasts that are encased in a matrix of somewhat scaly,
clinochlore, and dolomite-magnesite-quartz-talc-(clinochlore). cleaved, crenulated mudstone (Fig. 23). The oriented slabs of
The talc-bearing rocks are interpreted as siliceous magnesite- bedding define a “fragment foliation” (Cowan, 1985) consisting
dolomite rocks that are partially replaced by talc. of subparallel surfaces, although they are not as well oriented as
foliation in matrix. In the northernmost part of the principal Taray
STRUCTURE exposure, where foliation is not pronounced, well-developed
fragment foliation strikes northeast (Fig. 21A).
Overview Some features of deformed beds suggest penecontempo-
raneous deformation. This type of deformation is recorded by
Planar structures are the principal structural elements rare, isolated, large boulders of well-cemented sandstone that
recorded in outcrops of Taray Formation (Fig. 20). Planar slid downslope, crumpling sediments at their bows (Fig. 4). An
features include the following: (1) S0 surfaces of sandy layers exposure of one of these large blocks, ~20 m long, clearly shows
(Fig. 21A). (The term “surfaces” is used rather than bedding to a convex upward, striated base distinguished by sheared and
indicate our uncertainty about them as primary depositional fea- broken shale with thin interbeds of sandstone. Above the striated
tures; many parallel planes [see C foliation below and Fig. 21C] surface, the base of the coherent block is underplated by pieces
record movement.) (2) S1 and C, principal foliations in argilla- broken and corraded from the block. Beneath the underplate, a
ceous matrix (Fig. 21B–C). (Indications of slippage distinguish mixture of pebbles and cobbles of sandstone and black shale with
C planes that may be parallel to fragment surfaces or that steepen lenticular inclusions of thin-bedded gray shale were overridden.
and ramp gently across layers, which commonly show offset.) Some of the thin beds are crumpled, and crenulations with hinges
(3) Axial surfaces of folds (AP1 and AP2). (4) Locally developed parallel to striations at the base of the slide block are observable.
cleavage (S2) that is axial planar to folds defined by S1 (Fig. 21D). Irregular fractures in this horizon are filled with fine-grained
(5) Faults along which gouge and/or breccia occur. (These faults sediment. At the southwest bow of the boulder, a sandstone bed
record brittle deformation after lithification and are separated is imbricated and folded. A set of quartz-filled fractures in the
from structures that may have formed when the rocks were duc- sandstone boulder is more-or-less orthogonal to slickenlines on
tile, perhaps not fully lithified.) (Fig. 21E). (6) Steeply dipping the slide surfaces and to the direction of transport (Fig. 21F).
fractures within sandy rocks, commonly filled with quartz. Most The fractures indicate that the boulders were sufficiently coher-
planar features strike easterly to northeasterly and dip south. ent during or after sliding for brittle deformation in response to
Filled fractures (tension gashes) strike northwesterly. extension or stretching to occur .
Linear structures include fold hinges (F1, F2), slickenlines (Lf), Penecontemporaneous intraformational deformation is also
and intersection lineations (L1x0) (Fig. 21F). inferred by the small, strongly overturned folded layer detached
Downloaded from specialpapers.gsapubs.org on June 27, 2015

440 T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 20. Map of Taray Formation showing bedding and foliation.


Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 441

A N B N C N

So – Surfaces of beds S1 – Foliation C – Planes


Tabular layers - N = 54 Foliation – N = 93 Planes that cut layers and
Semi-lithified (?) - N = 16 Foliation; evidence for show normal offset – N = 40
Chert layers – N = 15 penecontemporaneous Planes parallel to fragments
deformation preserved – N = 25 and layers that show evidence
of slip – N = 33

D N E N F N

Axial surfaces and foliation (S2) Brittle Fractures Linear elements


parallel to axial surface Brittle faults F1 - Hinge
AP1 - N = 24 Tension Gashes F1 - Hinge recorded in
AP1 evidence for soft sediment
penecontemporaneous F1 - Chert
deformation preserved - N = 3 F2 - Hinge
AP1 in chert - N = 11 Lf - Slickenline
AP2 - N = 6 L1Xo
S2 - N = 27 L1Xo - Recorded in
soft sediment
L1Xo - Chert
Figure 21. Stereoplots for planar and linear structural elements recorded by Taray Formation. These and all subsequent stereoplots are equal-area,
lower-hemisphere projection plots.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

442 T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 22. Low angle fault cutting layers below at a high angle and Figure 23. Dismembered, attenuated sandstone in dark, muddy matrix
nearly parallel to bedding above. Rock appears to be altered ash-tuff exposed in the crest of a small antiform. Scale length = 10 cm.
and chert. Hammer length = 36 cm.

from unfolded lower beds (Fig. 15). The thinner, underlying (Fig. 21E). In general they strike somewhat more easterly and are
beds are segmented at ramps where faults parallel to layering cut slightly steeper. However, some faults fall within the loose cluster
across them. The exposures suggest unconstrained downslope of foliation planes and shear surfaces. Where faults cut sandy lay-
movement followed by similarly directed segmentation during ers, normal displacements are recorded.
subsequent stretching. Fractures, commonly lenticular and filled with quartz,
are best developed in resistant sandstone blocks and boulders,
Foliation and Other Spaced Planar Features although they may occur in fine-grained matrix. They are orthog-
Foliation is most conspicuous in the fine-grained matrix but is onal to the trend of striations, slickenlines, and hinges of crenula-
also evident in some coarser sandy layers that may record spaced tions, and to the dip of bedding and S1 and C foliations.
cleavage with dips steeper than typical S1 (Fig. 21B). Strike is
parallel to that of bedding. The general dip is gentler than that of Linear Features
bedding; however the dip may vary from nearly vertical to 15°SE
(Fig. 21B). Within the matrix, S1 fabric is accentuated by strong Linear features include fold hinges recorded by beds (F1),
alignment of tabular, lenticular and oblate fragments (Fig. 15). hinges of crenulations (F2), intersection lineation (Ls1xso), stria-
Although prominent, S1 foliation is not strongly planar and may tions and slickenlines (Lf) (Fig. 21F). Most lineations and fold
exhibit common deflections around clasts and inclusions (Fig. 4). hinges are co-linear with gentle plunge.
S1 is also commonly tangential or subparallel to gently dipping
surfaces that accommodate slip (C planes) as shown by slicken- Mesoscopic and Macroscopic Folds
lines or offsets, principally of bed surfaces. Together the foliations Folds are evident in matrix (Figs. 23 and 24) and large
resemble phacoidal foliation (Figs. 15, 23; Kusky and Bradley, blocks; folds in the latter are commonly isoclinal (Fig. 6). Rare
1999) composed of cleavage (S1) and slip surfaces (C planes) exposures show folds in detached layers (Fig. 15). The distinc-
(Fig. 21C). Where beds are preserved, C planes commonly accom- tion between structures formed before sediment was lithified
modate normal displacements of individual layers that indicate and those formed later is commonly not clear. Confusion arises
extension in the matrix. In places, gash fractures filled with quartz because many planar structures, i.e., bedding and foliations, are
cut the foliated matrix. Distinguishing among anastomosing planar subparallel, and large blocks of fractured (lithified?) material are
elements with respect to their origins is difficult. entrained in foliated shale that we interpret to have been rela-
The planarity of foliation may be further modified and tively soft, pliable and wet (?) during some part of deformation.
obscured by folds (F2) (Fig. 23) and/or crenulations defined by We conclude that some folds formed in unlithified sediment.
S1 foliation. An axial planar S2 cleavage is developed in the hinge
regions of some F2 folds. Environment of Deformation as Suggested by Structural Style

Brittle Faults and Fractures Rocks of Taray Formation that are: (1) disrupted, frag-
Brittle faults, marked by breccia and/or gouge, are less mented bodies of interbedded sandstone and shale that lack
well oriented than striated surfaces grouped with the C surfaces coherent stratigraphy, and (2) argillaceous masses containing
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 443

within a shallow, nonmetamorphic part of an accretionary prism;


we have no evidence of the latter.
Taray contains blocks of many sizes in fine-grained matrix.
The most conspicuous are exotic blocks of chert, several hundred
meters long that form isolated hills. One of these hills, Cerro El
Pedernal, is shown in Figure 3. Other exotic blocks are composed
of carbonate, some of which is fossiliferous, quartzose sandstone,
and volcanogenic rocks. Sandstone, siltstone, and calcareous lay-
ers interbedded with shale also form blocks, as do meter- to centi-
meter-sized fragments of the layers, which float in matrix. These
bodies of disordered rock composed of diverse blocks with definite
shapes and sharp outlines randomly distributed in shaly matrix also
fit well with Pini’s (1999) description of olistostromes.
We interpret the mesoscopic fabric and lithologic composi-
Figure 24. Chevron folds with gently dipping axial planes. Plier length tion of Taray Formation as recording progressive layer-parallel
= 20 cm. extension of interbedded shale and sandstone leading to disrup-
tion and fragmentation compatible with sliding and fragmenta-
tion during gravity-driven, downslope movement that culminated
in mixing of muddy debris. The processes responsible for the
formation of the rocks probably occurred within the high part
exotic blocks are comparable to types I and III mélange (Cowan, of an accretionary wedge at a convergent margin (Fig. 25, after
1985). Type I mélange is distinguished by sequences of sand- Cowan, 1985), or near the thrust front (e.g., Pini, 1999, their fig.
stone and mudstone that show variable disruption and fragmenta- 48), although deeper settings are not precluded. We were unable
tion, principally in response to layer-parallel extension. Although to test for the record of coaxial strain, however, some sliding
interbedded mudstone and sandstone are the dominant lithologies clearly occurred at depths sufficiently shallow so that strongly
of Taray Formation, continuous layers of sandstone are restricted noncoaxial shear was unlikely.
to outcrop-scale occurrences and are not preserved as mappable The processes that we infer to be recorded by structures
units. Sandstone beds show progressive dismemberment, rang- contained within Taray are comparable to those occurring within
ing from tabular layers preserved on the scale of an outcrop, to actively developing accretionary prisms such as the Nankai
beds that pinch and swell, to beds dismembered by boudinage (Chamot-Rooke et al., 1992). Near the toe of the eastern part
or low-angle faults. Some isolated lenses of sand 10 m or more of the Nankai wedge, layers above gentle thrusts bend down-
long are interpreted to record boudinage during gravity-driven, ward conformably across topographic slopes as steep as 20° to
downslope sliding of partially lithified sediment followed by fur- 30°. Locally, bed dips are steeper than the slope and may reach
ther stretching during tectonic extension. These rock masses con- 50–60°. In the same area, a slump scar, the face of which extends
tain mesoscopic features comparable to those of tectonosomes as for more than 1.5 km, cuts steeply (in places 45°) across almost
recently described by Pini (1999). Similarities include boudinage 1000 m of section composed of subhorizontal mudstone, inter-
and disruption of competent beds of sandstone, siltstone, and cal- preted as silty turbidite with interbeds of conglomerate contained
careous layers by fragmentation into blocks with sharp outlines. within stacked thrusts (Lallemand et al., 1992). Fluid is expelled
The fragments show a common trend and strong preferred orien- along shallow detachments that are commonly parallel to bed-
tation parallel to the plane of major extension, which coincides ding, and possibly along steep joints. We believe that these condi-
with the surfaces of beds. tions must be conducive to the formation of slides, slumps, and
According to Cowan (1985), masses of rock that contain debris-rich matrix characteristic of Taray.
inclusions: (1) the origins of which are obscure, (2) the structures
of which are discordant and the lithologies of which are exotic STRUCTURAL AND STRATIGRAPHIC RELATIONSHIP
to the basin, and (3) that are entrained in foliated matrix, are BETWEEN TARAY AND JURASSIC VOLCANIC
interpreted to belong to Cowan’s (1985) type III mélange. The FORMATIONS
formation of this type requires a multistage process involving:
(1) acquisition of rock fragments from diverse, possibly widely The principal planar structures (beds and foliation) in the
separated sources, (2) dispersion of the inclusions during mixing Taray Formation strike northeast. Beds and foliation in the over-
with matrix, and (3) common bulk deformation involving flat- lying Jurassic volcanic section, including Nazas, Rodeo, and
tening and development of scaly foliation. The common charac- Caopas Formations, and higher Zuloaga carbonate strike north-
teristics of type III mélange suggest emplacement as mud-rich, west, placing the strikes of the Taray and overlying Jurassic units
submarine debris flows or olistostromes coupled with migration at right angles to one another. Foliation and bedding in the Juras-
and intrusion of these materials during upwelling or diapirism sic rocks generally conform to the northwest trend of the Late
Downloaded from specialpapers.gsapubs.org on June 27, 2015

444 T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 25. Generalized cross section of


an accretionary prism showing possible
locations and geologic settings com-
patible with the interpreted origins of
rocks mapped as Taray Formation (after
Cowan, 1985).

Jurassic fold that forms the principal structure within the San Metamorphism is higher grade in the Taray, but the differ-
Julian uplift (Anderson et al., 1991). Well-developed multiple ence may be more apparent than real, because Taray offered a
foliations in the Taray indicate a more complex structural history more reactive protolith and it underwent local Tertiary contact
than that recorded by the overlying volcanic rocks. metamorphism. Further, the Taray metamorphism is no higher
Despite wide and diligent search, we never found an than chlorite-grade (except in the vicinity of Pico de Teyra plu-
observable contact between Taray Formation and overlying ton) and Nazas rocks contain chlorite and white mica in their
volcanogenic rock, only covered zones. Córdoba-M. (1964) matrix, suggesting that Nazas also is low greenschist facies.
and López-Infanzón (1986) described the contact as an angular
unconformity and our regional structural data (Fig. 21) appear to RESTORATION OF SOME EFFECTS OF SUPERPOSED
agree with that interpretation. But other observations indicate to DEFORMATION: IMPLICATIONS FOR THE PROPOSED
us that the relationship cannot be so succinctly characterized. TECTONIC SETTING OF TARAY FORMATION
The San Julian uplift contains both an early to middle
Mesozoic mélange and the rocks of a Jurassic arc. This tempo- Deformations that occurred after the formation of Taray
ral compatibility of arc and mélange and their geographical and syndepositional structures include: (1) Late Jurassic folding
stratigraphic proximity suggest that the two were contemporary about northwest-trending hinges during southwest-directed
members of the same stratigraphic regime. The composition of contraction, and accompanying detachment, recorded by beds
the mélange adds support to the idea that arc and mélange are of the Oxfordian(?) lower Zuloaga limestone and facilitated by
related; in the various arroyos that cross the Taray-Nazas contact interbeds of gypsum (Anderson et al., 1991); (2) north-directed,
we saw repeatedly that the parts of the Taray closest to the contact Late Cretaceous folding above a second detachment also within
showed an increase in tuffaceous volcanic component, thereby evaporite-rich strata of the Zuloaga (Tardy, 1980); and (3) Ter-
predicting a gradational contact. Because of these relationships tiary normal faulting and associated tilting related to northeast-
(compositional facies, temporal compatibility, and both geo- southwest extension (Anonymous, 1981) (Figs. 1 and 2). The
graphical and stratigraphical proximity), we hesitate to consider faults and detachments and the directions of tectonic transport
the contact to be an unconformity that represents a significant that we infer from the mapped geological relationships described
temporal hiatus. We hypothesize: (1) that the zone of contact above are shown on schematic structure sections (Fig. 26). Ter-
between forearc and accretionary prism rocks was complex and tiary tilting is recorded by the 20° eastward plunge of fold hinges
may have involved accumulation of forearc strata on deformed in Cretaceous strata along the northeast flank of the San Julian
rock of the prism as portrayed by Cowan (1985) (Fig. 25), uplift (Fig. 2). The relationships are shown by Tardy (1980,
(2) that shortening during formation of a Jurassic fold (Anderson Fig. I3–9), who notes, as have others, the pronounced detachment
et al., 1991) accommodated by low-angle faults resulted in trans- of Late Jurassic and Cretaceous strata above Oxfordian gypsum.
port of the arc rocks across the forearc and onto the accretionary Regionally, east-trending Cretaceous folds with subhorizontal
prism (Figs. 2 and 3), and (3) that the mélange in the footwall of hinges that formed above the detachment are conspicuous topo-
the thrust(s) was largely unaffected by the Jurassic deformation. graphic and structural features along the great westward bend of
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 445

Figure 26. Schematic structure sections


(along lines shown in Fig. 2) that illus-
trate the faults and detachments and the
directions of tectonic transport that we
infer from the mapped geological rela-
tionships. Tertiary tilting is recorded by
the 20° dip of postulated planar surfaces
toward the inferred Cedros fault. Three
deformational events post-formation of
the Taray mélange are emphasized in
the sections: (1) Late Jurassic folding
about northwest-trending hinges dur-
ing southwest-directed contraction, and
accompanying detachment, recorded by
beds of the Oxfordian(?) Lower Zuloaga
limestone and facilitated by interbeds
of gypsum (Anderson et al., 1991);
(2) north-directed, Late Cretaceous
folding above a second detachment
also within evaporite-rich strata of the
Zuloaga (Tardy, 1980); and (3) Tertiary
normal faulting and associated tilting
related to northeast-southwest extension
accommodated by the inferred Cedros
fault (Anonymous, 1981) (Fig. 2).

the Sierra Madre Oriental. Tardy attributes the eastward plunge 1995) and the underlying Taray Formation, and to better interpret
of the folds exposed along the northeast flank of the San Julian the kinematic significance of the structures recorded by Taray
uplift to tilting during northeast-directed Miocene contraction. Formation, we restore the rotation of the hanging wall above the
Tilting also produced the large northwest-trending anticlinorium inferred southwest-dipping Tertiary normal fault (Fig. 27). We
manifest as the core of the uplift. assume that the plunge shown by Cretaceous folds is the result
We have argued that deformation of the northwest-trending of Tertiary tilting, and measure the plunge angle of 20° as shown
core of the San Julian uplift is pre-Cretaceous, as indicated by by Tardy (1980). The plunge is restored to horizontal by counter-
the pronounced deformation of Late Jurassic strata that is not clockwise rotation of 22° about an axis plunging 0°, 330°, paral-
recorded in Cretaceous rocks (Anderson et al., 1991). We agree lel to the inferred fault trace. The sense and magnitude of rotation
that the eastward tilt likely occurred during the Tertiary but attri- are based upon: (1) down-to-the-southwest normal offset along
bute it to rotation in the hanging wall above a major northwest- the fault northeast of the San Julian uplift and (2) sufficient rota-
striking listric normal fault that occurs in the valley northeast of tion (22°) about the axis trending 330°, which is not orthogonal
the uplift (Tardy, 1980; Anonymous, 1981; Fig. 27). The fault to the easterly trending Cretaceous fold hinges, to restore the 20°
disrupts the continuity of the Cretaceous and younger folds as plunge to horizontal.
shown on the regional geologic map. The effect of the rotation upon the structural elements of the
So that we may further assess the relationship between Taray Formation is marked (Fig. 28). Bedding, S0, and foliations,
rocks of the Nazas-Rodeo-Caopas volcanic suite (Jones et al., S1 and C, are parallel and strike northeast. Each group of planes
Downloaded from specialpapers.gsapubs.org on June 27, 2015

446 T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 27. Regional map showing the relationship of San Julian Uplift to Cedros normal fault.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 447

(A) N (B) N (C) N

So – Surfaces of beds S1 – Foliation C – Planes


Tabular layers - N = 54 Foliation – n = 93 Planes that cut layers and
Semi-lithified (?) - N = 16 Foliation; evidence for show normal offset – N = 40
Chert layers – N = 15 penecontemporaneous Planes parallel to fragments
deformation preserved – N = 25 and layers that show evidence
of slip – N = 33
(D) N (E) N (F) N

Axial surfaces and foliation (S2) Brittle Fractures Linear elements


parallel to axial surface
Brittle faults F1 - Hinge
AP1 - N = 24 Tension Gashes F1 - Hinge recorded in
AP1 evidence for soft sediment
penecontemporaneous F1 - Chert
deformation preserved - N = 3 F2 - Hinge
AP1 in chert - N = 11 Lf - Slickenline
AP2 - N = 6 L1Xo
S2 - N = 27 L1Xo - Recorded in
soft sediment
L1Xo - Chert
Figure 28. Stereo plots of restored planar and linear structural elements recorded by Taray Formation. All structures have been rotated 22° coun-
terclockwise, as viewed northwest, about an axis plunging 0°, 330°.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

448 T.H. Anderson, N.W. Jones, and J.W. McKee

dips southeast: S0, 58°; S1, 44°; and C, 37°. Many axial planes lie formation (Diaz-Salgado et al., 2003), and the age of the Nazas,
among these planes and they dip the same direction although only which is probably no younger than 158 ± 4 Ma (Oxfordian),
26°. Most linear elements, including fold hinges, intersection linea- the age of the Caopas Formation (Jones et al., 1995). This idea
tions, and fault-related striations plunge shallowly southwestward remains as a hypothesis in need of further testing.
downdip, orthogonal to the strike of planar elements. Extensional
fractures (tension gashes) filled with quartz are almost vertical and CORRELATION
are mutually perpendicular to the strike of bedding and foliations
and the subhorizontal linear structures. The orientations of planar Taray Formation shares similarities of composition, lithol-
and linear structural elements, restored to positions that we specu- ogy, and structure with other northern Mexico, early Mesozoic,
late may be close to initial strikes and dips, could accommodate marine clastic units: (1) the La Ballena Formation, exposed at
downslope movement of semilithified sediment. Peñon Blanco, San Luis Potosí and Zacatecas (Fig. 1; Labarthe-
We also corrected northwest-trending folds recorded by Hernández et al., 1982), and at Charcas, San Luis Potosí (Fig. 1;
overlying Nazas and lower Zuloaga Formation for the effects of Tristán-González and Torres-Hernández, 1994); and (2) the
Tertiary tilting (Fig. 29). The modest effect shows that the Late Zacatecas Formation, which is exposed at Zacatecas, Zacatecas
Jurassic folds, beneath the detached, east-trending Cretaceous, (Fig. 1; Centeno-García and Silva-Romo, 1997, and references
were asymmetric and verged southwest. We did not restore the therein). The La Ballena consists of quartzose sandstone, black
effects of Late Jurassic folding upon Taray Formation. We suspect shale and conglomerate interpreted as parts of deep marine fans
that low-angle faults at the contact between Taray and overlying formed from deposits carried by turbidity currents (Silva-Romo,
volcanic rocks accommodated Late Jurassic detachment, related 1993). At Charcas, La Ballena beds record slump-folds and brec-
telescoping, and emplacement of the volcanic section across an ciation suggestive of downslope gravity-driven movement. Folia-
assumed forearc basin and onto the accretionary prism. Our inter- tion is commonly parallel to bedding. Late Carnian ammonites
pretation also assumes that: (1) the texture and structure of Taray in the La Ballena (Cantú-Chapa, 1969; Silva-Romo, 1993) are
developed as part of a mélange within an accretionary prism, and correlative with those of the Zacatecas Formation at Zacatecas
(2) the mélange most likely formed during Middle Jurassic time (Burckhardt and Scalia, 1906) that occur within a section of
(Anderson et al., 1990). Unfortunately the age of Taray may be black shale interbedded with quartzose sandstone as well as less
bracketed only between the late Paleozoic, based upon the age common pillow lava and limestone (Ranson et al., 1982; Cuevas-
of the youngest detrital zircons reported from sandstone in the Pérez, 1983; Monod and Calvet, 1991).

Figure 29. Stereo plots for planar and


linear structural elements recorded by
Nazas (left) and Zuloaga (right) Forma-
tions. All structures have been rotated
22° counterclockwise, as viewed north-
west, about an axis plunging 0°, 330°.
Foliation (S surface) Foliation
Foliation (C surface) Bedding surface
Bedding surface Slickenline
Brittle fault Brittle fault
Tension gash Fold hinge
Mineral smear Axial surface

Nazas Lower Zuloaga Formation


Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 449

Most of the sand in Taray beds is recycled from an orogen, dark beds of the Varales lithofacies may be interbedded with
although some sand beds record input from cratonic or dis- fine-grained, light-green sandstone and shale of the Jaltomate
sected arc sources (Table 1; Fig. 30). In the field, we thought we lithofacies and light-green chert of the Bocana lithofacies. Within
could see an increase in volcanic aspect of the Taray as a whole the Arteaga Complex, isolated bodies of mafic volcanic rocks
in some of the topographically higher exposures. However, with geochemical signatures similar to mid-ocean-ridge basalt
increased volcanic input is not reflected in the compositions (Charapo lithofacies), foliated diorite, gabbro and plagiogran-
of the sandstones. They indicate a mixed source without a sig- ite (Las Juntas lithofacies), and limestone containing chert and
nificant component reflecting an active, undissected arc, though crinoid fragments, form blocks and lenses incorporated within
we did not sample systematically. Similar to Taray (Fig. 30), the complex during thrusting and shearing. The deformation
La Ballena and Zacatecas Formations (Centeno-García and suggested to Centeno-García et al. (2003) that Arteaga complex
Silva-Romo, 1997, their fig. 4) plot within the recycled orogen may be characterized by broken formation and block-in-matrix
provenance of Dickinson (1985). texture—deformation styles associated with mélange.
These three units (Taray, La Ballena, and Zacatecas) all bear In addition to lithological similarity, age constraints at
resemblance to certain facies of the Arteaga Complex, which Arteaga, based upon the presence of Late Triassic radiolaria
comprises a thick (>1 km) sequence of Triassic-Jurassic rocks (Campa et al., 1982) and detrital zircons (G. Gehrels, personal
exposed near Arteaga, Michoacan (Fig. 1; Centeno-García et commun. to E. Centeno-García, as cited in Centeno-García et al.,
al., 2003). In particular, the Varales lithofacies includes silici- 2003), are also comparable to Taray. Stratigraphic and structural
clastic units composed of black shale, quartzose sandstone and characteristics of the Arteaga Complex are sufficiently similar to
less common black chert and rare conglomerate, interpreted as those of Taray that it may be interpreted as part of a correlative
channel fill. Predominantly black, dark-gray or dark bluish-gray, accretionary prism, although somewhat more oceanward.
2-cm- to 2-m-thick beds of shale and fine- to medium-grained Centeno-García and Silva-Romo (1997) recognized the
(rarely coarse-grained) sandstone may be interbedded with similarities among the mélange-like units but chose to place
thin-bedded black chert or thicker sections of green chert. The them in separate terranes: Taray is included with La Ballena
sandstone and shale beds are commonly broken and disrupted by Formation as part of Sierra Madre terrane of northern Mexico,
the effects of boudinage and may be imbricated with lens-shaped distinguished in part by the postulated presence of sialic base-
masses, tens to hundreds of meters long in dimension, of pillow ment; Zacatecas Formation and rocks of the Arteaga Complex
lava, green chert, limestone blocks and volcaniclastic rocks. The are grouped together as part of the extensive Guerrero compos-
ite terrane, distinguished by oceanic lithosphere, that underlies
much of western Mexico.
Recently reported studies are the basis for a less straightfor-
ward interpretation by Elias-Herrera et al. (2003). They consider
Arteaga to have formed as fill in a basin behind an east-facing
offshore arc with sialic roots that was intruded by plutons during
the late Early Jurassic ~187 m.y. ago. The arc collided against
southern Mexico during the Middle Jurassic.

PLATE TECTONIC IMPLICATIONS

The widely separated exposures of Late Triassic(?) marine


strata composed of dark shale, quartzose sandstone, chert and
matrix-rich debris enclosing blocks of fossiliferous Paleozoic
carbonate and mafic, oceanic igneous rocks, suggest that parts of
a Jurassic(?) accretionary wedge underlie more than one region
of Mexico.
Jones et al. (1995) proposed that Jurassic volcanic rocks,
and accompanying Taray mélange, in northeastern Mexico are
displaced from correlative units in northwestern Mexico (Fig. 1).
We assess the effect of this proposed displacement upon Taray
Formation and its probable equivalents, La Ballena, Zacatecas,
Figure 30. QFL (Quartz-Feldspar-Lithic fragments) diagram after and Varales, by means of restoration of displacement along the
Dickinson (1985) showing provenance categories for sandstones. See hypothetical Mojave-Sonora megashear, as well as other parallel
Table 1 for meanings of Qm, Qp, P, K, Lv, Ls, and M. Open triangles
are samples from Table 1A, those that are not associated with rocks faults to the south as shown by Anderson and Schmidt (1983).
with a volcanic component; solid triangles are samples from Table 1B, The maps (Figs. 31 and 32) employ oblique Mercator projec-
those that are associated with rocks with a volcanic component. tions derived from the pole of rotation of the Mojave-Sonora
450
Downloaded from specialpapers.gsapubs.org on June 27, 2015

T.H. Anderson, N.W. Jones, and J.W. McKee

Figure 31. Oblique Mercator projection of map of Mexico showing postulated Jurassic tectonic discontinuities and location of exposures of Jurassic, Triassic, and late Paleozoic rocks.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico

Figure 32. Oblique Mercator projection of Mexico for Middle Jurassic time showing exposures of Jurassic, Triassic, and late Paleozoic rocks restored along principal, inferred Jurassic
tectonic discontinuities. See text for discussion.
451
Downloaded from specialpapers.gsapubs.org on June 27, 2015

452 T.H. Anderson, N.W. Jones, and J.W. McKee

megashear. As pointed out by Silver (Silver and Anderson, 1974), scheme of Sedlock et al. (1993), who recognized five terranes
restoration removes late Paleozoic overlap of South America in southwestern Mexico. The terranes roughly correspond to
across much of Mesoamerica. principal exposures of the Arteaga Complex (Nahuatl terrane),
Restoration of Jurassic igneous rocks of northeastern Mex- metamorphosed Paleozoic rocks, e.g., Acatlan Complex (Mix-
ico (Grajales-Nishimura et al., 1992) to their pre-fault position teco terrane), Precambrian Oaxaca Complex (Zapoteco terrane),
opposite correlative units in Sonora maintains the oceanward Jurassic arc (Xolapa Complex within Chatino terrane), and
position of the Late Triassic marine rocks relative to the subduc- mainly unmetamorphosed, fossiliferous, Upper Paleozoic strata
tion-related, Andean-style, Jurassic igneous province. The Juras- (Maya terrane). The Nahuatl terrane is the southeastern part of
sic igneous belt may extend eastward from the San Julian Uplift the composite Guerrero terrane of Campa and Coney (1983).
to Aramberi, Nuevo Leon (Fig. 1), where Middle Jurassic(?) The previously discussed Arteaga Complex (Fig. 1; Cen-
volcanic rocks pinch out against underlying paragneiss. From teno-García and Silva-Romo, 1997; Campa and Ramirez, 1979;
this point the belt of “arc” rocks must either bend southward or Campa et al., 1982; Campa and Coney, 1983; Centeno-García
step to the south along a trench-trench transform, perhaps related and Silva-Romo, 1997) is cut by granite plutons, the ages of
to the Oaxaca fault (Alaniz-Alvarez et al., 1996) (Figs. 1, 31, and which fall between ca. 163 Ma (Bathonian) and ca. 158 Ma
32). In either case the eastern limit of the Middle Jurassic “arc” (Callovian) (Centeno-García et al., 2003, and references therein).
likely coincides with exposures of Paleozoic strata above older Other exposures of Triassic-Jurassic rocks have been discovered
crystalline rocks exposed in the uplifted core of the Sierra Madre northwest of Arteaga, where aluminous schist and chlorite schist,
Oriental (e.g., Sedlock et al., 1993). which are probably metamorphosed Late Triassic rocks, comprise
Exposures of Jurassic igneous rocks that may mark the the basement at the Cuale mining district (Schaff et al., 2003).
inboard edge of the accretionary wedge (opposite the Middle There, north-striking faults bound a graben containing rhyolite
Jurassic “arc”) are sparse in central Mexico. Chert, rhyolitic and flows, pyroclastic units, breccia, and conglomerate. Higher in the
rhyodacitic tuff, conglomerate, breccia, sandstone, and siltstone section, the volcanic units are interbedded with and ultimately
crop out in the core of El Chilar anticline, east of Guanajuato buried by black shale, fine sandstone, granule conglomerate, and
(Carrillo-Martinez, 1997). Interbeds of siliceous shale contain coarser conglomerate. Dikes and sills that intrude the basin fill
radiolarians of Middle(?) to Late Jurassic age (Martinez-Hernán- are the youngest units. U-Pb interpreted ages from rhyolite low in
dez, 1979; Chauve et al., 1985). Near Guanajuato (Fig. 1), some the section and the intrusive bodies indicate that the basin prob-
samples from mafic and ultramafic rocks (Quintero-Legorreta, ably formed shortly before ca. 162 Ma and received sediments
1992) of the basal sequence comprising the Guanajuato arc and volcanogenic products until at least ca. 156 Ma (Mortensen
(Ortiz-Hernández and Martinez-Reyes, 1993) yield Late Jurassic et al., 2003). Bissig et al. (2003) interpreted the results of trace
K-Ar dates interpreted as cooling ages of early Mesozoic oce- element geochemistry to indicate that the rhyolites have an “oce-
anic lithosphere. However, these results are considered spurious, anic arc” origin. We offer an alternative interpretation to the idea
because excess argon masks the true Early Cretaceous age of this that the rhyolitic rocks are arc related. The ages and tectonic set-
crust (Lapierre et al., 1992). ting of the Jurassic igneous rocks within the Arteaga Complex
Farther south is the trans-Mexico volcanic belt, a linear and equivalent units are the same as those for pull-apart basins
array of active volcanoes that may reflect the presence of a major along the Mojave-Sonora megashear (Anderson and Nourse, this
fault, as suggested by Gastil and Jensky (1973). The volcanic volume). Anderson speculates that some of the Jurassic magma-
belt and fault (Mexican volcanic belt megashear, Figs. 1, 31, tism may not be arc related but records igneous activity through
and 32; Anderson and Schmidt, 1983) mark a fundamental ter- Arteaga crust thinned and broken during transtension related to
rane boundary. Anderson and Schmidt (1983) treated the rocks displacement along the nearby Acapulco-Guatemala megashear
south of the Mexican volcanic belt as a single block (Maya West) (Figs. 1, 31, and 32; Anderson and Schmidt, 1983).
that they restored northwestward based solely upon geometric Samples of granitic orthogneiss from the Xolapa Complex
constraints stemming from the Bullard reconstruction. Since within the Chatino terrane yield U-Pb ages between ca. 170 and
1983 few additional correlations have been identified that better 158 Ma (Guerrero-García et al., 1978; Ducea et al., 2004). The
constrain the position of the Maya West block. Detrital zircons granitic plutons were emplaced into “basement” composed of
from Cambrian quartzose sandstone north of the Mojave-Sonora Permian and Grenvillian crystalline rocks (Ducea et al., 2003,
megashear in Sonora show a strong peak (14 of 22 grains) at 2004) and possibly Middle Jurassic sedimentary rocks (Ortega-
1.1 Ga (Stewart et al., 2001). The suggested proximity to Late Gutiérrez and Elias-Herrera, 2003).
Proterozoic rocks, such as those that probably form the basement A strong tectono-thermal event, recorded in some deep
of the Maya West block, may imply further northwestward resto- parts of the Acatlan and Oaxacan Complexes at ca. 175 Ma may
ration than inferred by Anderson and Schmidt (1983). reflect the effects of an underlying hot spot (Keppie et al., 2003)
Additional work since 1983 has resulted in recognition or be related to arc-related volcanic units that crop out northwest
of multiple terranes within the Maya West block (Fig. 1; e.g., of Oaxaca (Ortega-Gutiérrez et al., 1997). Unmetamorphosed
Campa and Coney, 1983; Sedlock et al., 1993; Centeno-García arc-related Jurassic rocks resting on continental crust occur
et al., 2003). On Figure 1, we employ the broader classification northwest of Oaxaca, where mafic and intermediate tuff and lava
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 453

interbedded with limestone, conglomeratic sandstone, and lime- is the result of shortening, accommodated by low-angle faults
stone conglomerate comprise a section designated as the Ixcatlán concentrated along the enigmatic contact zone, which brought
arc, which is interpreted as near-shore (Ortega-Gutiérrez et al., the volcanic section onto the mélange.
1997). A probable Middle Jurassic age is assigned based upon The Taray is lithostratigraphically correlative with the
plant fossils. Zacatecas and La Ballena Formations, and with the Varales litho-
In summary, no simple relationship between oceanic rocks, facies of the Arteaga Complex. These Late Triassic deep-marine
such as Taray, and a Middle Jurassic arc formed in response to units, which may be contemporary with the Taray, are composed
subduction of the oceanic rocks is recognized south of the Mexi- of dark shale, chert, and quartzose wacke that may enclose blocks
can volcanic belt. Furthermore the Maya East block, composed of Paleozoic carbonate and mafic igneous rocks. Each contains
of Yucatan Peninsula and Guatemala, is awkwardly positioned some structures commonly associated with rocks formed within
with respect to Coahuila Island even after restoration of displace- accretionary prisms.
ments along the inferred Late Jurassic sinistral faults. A determi- Restoration of the mélange-arc suite along the Mojave-
nation of a position for this block away from the late Paleozoic Sonora megashear sufficient to remove ~1000 km of left-lateral
collision zone between North and South America awaits further displacement places it in northwestern Mexico against correla-
study (cf. Steiner et al., this volume). tive igneous rocks (Jones et al., 1995). Restoration of ~300 km
of left-lateral displacement along a fault coincident with the
CONCLUSIONS trans-Mexico volcanic belt (Anderson and Schmidt, 1983) places
the Arteaga Complex and associated Precambrian and Paleozoic
The Taray Formation consists principally of marine sand- sialic basement rocks west of postulated Grenvillian rocks of
stone and mudstone or shale that record progressive disruption by Oaxaquia (Ortega-Gutiérrez et al., 1995).
layer-parallel extension. Locally, disruption is pervasive. Abun-
dant, large, exotic, predeformed inclusions of diverse shapes, ACKNOWLEDGMENTS
sizes, and compositions contained in argillaceous matrix are
an additional characteristic feature. Multiple sources must exist This work was supported by National Science Foundation
for the blocks, many of which are up to several hundred meters grant EAR-8705717 (JWM, NWJ, THA) and by the Faculty
long and are apparently of different ages. The commonly folded Development Fund of the University of Wisconsin–Oshkosh
inclusions, enclosed in slightly to complexly foliated matrix, (JWM). R.T. Klein and S.R. Hannula assisted during study
cannot have been derived by layer-parallel disruption of bedded of Cerro El Pedernal. Discussions with Jon Blickwede, Fer-
sandstone that forms lens-shaped inclusions generally parallel nando Ortega-Gutiérrez, and Elena Centeno-García provided
to foliation. Taray lacks inclusions of blueschist facies rocks or a regional perspective. Scott Davidson carefully drafted the
ophiolitic rocks. figures and Paul Coyle collated and presented the data shown
The lithology, texture, and deformation of these rocks are on the stereo plots. The paper benefited greatly from reviews by
similar to Cowan’s (1985) type I and III mélanges. The principal Darrel Cowan and Mariano Elias-Herrera.
(penetrative) deformation of the Taray is of the kind most likely
to have been imposed on sediment of this kind at a convergent REFERENCES CITED
plate boundary along the line of subduction. We interpret the
Taray as a mélange and part of an accretionary prism. Alaniz-Alvarez, S.A., van der Heyden, P., Nieto-Samaniego, A.F., and Ortega-
Gutiérrez, F., 1996, Radiometric and kinematic evidence for Middle
The maximum age of the Taray, based on fossils in clasts and Jurassic strike-slip faulting in southern Mexico related to the opening
on detrital zircons in sandstone, is late Paleozoic. It is either older of the Gulf of Mexico: Geology, v. 24, p. 443–446, doi: 10.1130/0091-
than the volcanogenic suite (Nazas, Rodeo, and Caopas Forma- 7613(1996)024<0443:RAKEFM>2.3.CO;2.
Anderson, T.H., and Schmidt, V.A., 1983, The evolution of middle America
tions) that lies above it, or a contemporary of that suite. Thus the and the Gulf of Mexico–Caribbean Sea region during Mesozoic time:
Taray may be as young as 158 Ma (latest Middle Jurassic), which Geological Society of America Bulletin, v. 94, p. 941–966, doi: 10.1130/
is an interpreted U-Pb date obtained on zircons from the Caopas 0016-7606(1983)94<941:TEOMAA>2.0.CO;2.
Anderson, T.H., McKee, J.W., and Jones, N.W., 1990, Jurassic (?) mélange
(Jones et al., 1995). in north-central Mexico: Geological Society of America Abstracts with
The present contact of Taray mélange with the volcanic Programs, v. 22, no. 3, p. 3.
rocks of the Nazas Formation marks the discordance of structures Anderson, T.H., McKee, J.W., and Jones, N.W., 1991, A northwest-trending,
Jurassic fold nappe, northernmost Zacatecas, Mexico: Tectonics, v. 10,
within the adjacent units. Nevertheless, the juxtaposition of Taray no. 2, p. 383–401.
beneath the volcanic pile permits the hypothesis that this mélange Anonymous, 1981, Hoja Monterrey carta geológica: Secretaria de Programa-
and arc are contemporary facies within the same convergent plate cion y Presupuesto, Direccion General de Geografía del Territorio Nacio-
nal, México, escala 1:1,000,000.
boundary system. The age of the Taray allows that hypothesis. Barboza-Gudino, J.R., Tristán-González, M., and Torres-Hernández, J.R.,
The relatively scarce evidence of volcanism found in the Taray 1999, Tectonic setting of pre-Oxfordian units from central and northeast-
does not support the hypothesis, but does not disallow it. We ern Mexico: A review, in Bartolini, C., Wilson, J.L., and Lawton, T.F.,
eds., Mesozoic sedimentary and tectonic history of north-central Mexico:
postulate that the absence of an intervening arc-trench gap, or Geological Society of America Special Paper 340, p. 197–220.
arc-proximal facies, richer in volcanic rocks and volcaniclastics,
Downloaded from specialpapers.gsapubs.org on June 27, 2015

454 T.H. Anderson, N.W. Jones, and J.W. McKee

Bissig, T., Mortensen, J.K., and Hall, B.V., 2003, The volcano-sedimentary Ducea, M., Gehrels, G.F., Shoemaker, S., Ruiz, J., and Valencia, V.A., 2004,
setting of the Kuroko type VHMS district of Cuale, Jalisco, Mexico: Geo- Geologic evolution of the Xolapa Complex, southern Mexico: Evidence
logical Society of America Abstracts with Programs, v. 99, no. 5, p. 61. from U-Pb zircon geochronology: Geological Society of America Bul-
Blickwede, J.F., 1981, Stratigraphy and petrology of Triassic (?) “Nazas Forma- letin, v. 116, no. 7/8, p. 1016–1025.
tion,” Sierra de San Julián, Mexico [M.S. thesis]: New Orleans, Univer- Elias-Herrera, M., Ortega-Gutiérrez, F., Sánchez-Zavala, J.L., and Macías-
sity of New Orleans, 100 p. Romo, C., 2003, The real Guerrero terrane, southern Mexico: New
Burckhardt, C., and Scalia, S., 1906, Geologie des environs de Zacatecas: insights from recent studies: Geological Society of America Abstracts
Mexico, D.F., Instituto Geológica de México, 10th International Con- with Programs, v. 99, no. 5, p. 66.
gress, Excursion Guidebook 16, 26 p. Gastil, R.G., and Jensky, W.A., 1973, Evidence for strike-slip displacement
Campa, M.F., and Ramirez, J., 1979, La evolución geológica y la metalogénesis beneath the trans-Mexican volcanic belt, in San Andreas fault sympo-
de Guerrero: Universidad Autónoma de Guerrero, Serie Técnico-Cienti- sium: Stanford University Publications in the Geological Sciences, v. 13,
fica, v. 1, 84 p. p. 171–180.
Campa, M.F., and Coney, P.J., 1983, Tectono-stratigraphic terranes and mineral Grajales-Nishimura, J.M., Terrell, D.J., and Damon, P.E., 1992, Evidencias
resource distributions in Mexico: Canadian Journal of Earth Sciences, de la prolongación del arco magmático Cordillerano de Triásico Tardío-
v. 20, p. 1040–1051. Jurásico en Chihuahua, Durango y Coahuila: Asociación Mexicana de
Campa, M.F., Ramirez, J., and Bloome, C., 1982, La secuencia volcánica-sedi- Geólogos Petroleros Boletín, v. 42, no. 2, p. 1–18.
mentaria metamorfizada del Triásico (Ladiniano-Cárnico) de la region de Guerrero-García, J.C., Silver, L.T., and Anderson, T.H., 1978, Estudios geo-
Tumbiscatio, Michoacán: Sociedad Geológica Mexicana, 6a Convención cronológicos en el Complejo Xolapa: Boletín de la Sociedad Geológia
Geológica Nacional, Resúmenes, 48 p. Mexicana, v. 39, resúmenes, p. 22–23.
Cantú-Chapa, A., 1969, Una nueva localidad Triásico Superior en México: Hsü, K.J., 1974, Mélanges and their distinction from olistostromes, in Dott,
Revista del Instituto Mexicano del Petroleo, v. 1, no. 2, p. 71–72. R.H., Jr., and Shaver, R.H., eds., Modern and ancient geosynclinal
Carrillo-Bravo, J., 1968, Reconocimiento geológico preliminar de la porción sedimentation: Society of Economic Paleontologists and Mineralogists
central del Altiplano Mexicano: Informe 637-ZN, Pemex, México, Special Publication 19, p. 321–333.
Inédito, 31 p. Jones, N.W., McKee, J.W., Anderson, T.H., and Silver, L.T., 1995, Jurassic vol-
Carrillo-Martinez, M., 1997, Structural geometry of the Sierra Gorda between canic rocks in northeastern Mexico: A possible remnant of a Cordilleran
Jalpan and Tequisquiapan, State of Queretero, Mexico: Libro-guia de las magmatic arc, in Jacques-Ayala, C., González-León, C.M., and Roldán-
excursiones geológicas, II Convención sobre la Evolución Geológica de Quintana, J., eds., Studies on the Mesozoic of Sonora and adjacent areas:
México y Recursos Asociados, Pachuca, Hidalgo, p. 65–71. Geological Society of America Special Paper 301, p. 179–190.
Centeno-García, E., and Silva-Romo, G., 1997, Petrogenesis and tectonic Keppie, J.D., Solari, L.A., Ortega-Gutiérrez, F., Elias-Herrera, M., and Nance,
evolution of central Mexico during Triassic-Jurassic time, in Stanley, R.D., 2003, Paleozoic and Precambrian rocks of southern Mexico—Acat-
G.D., Jr., and González-León, C.M., eds., Special issue dedicated to the lan and Oaxacan Complexes, in Geological transects across Cordilleran
International Workshop on the Geology of Northwestern Sonora, Mexico: Mexico: 99th Geological Society of America Cordilleran Section Annual
Revista Mexicana de Ciencias Geológicas, v. 14, no. 2, p. 244–260. Meeting Field Trip Guidebook, Instituto de Geología, Universidad Nacio-
Centeno-García, E., Corona-Chávez, P., Talavera-Mendoza, Ó., and Iriondo, A., nal Autónoma de México, Special Publication 1, p. 281–314.
2003, Geology and tectonic evolution of the western Guerrero terrane—A King, R.E., 1934, The Permian of southwestern Coahuila, Mexico: American
transect from Puerto Vallarta to Zihuatanejo, Mexico, in Geological tran- Journal of Science, v. 27, p. 98–112.
sects across Cordilleran Mexico: 99th Geological Society of America King, R.E., 1944, Geology and paleontology of the Permian area northwest of
Cordilleran Section Annual Meeting, Field Trip Guidebook, Instituto de Las Delicias, southwestern Coahuila, Mexico: Part 1. Geology: Geologi-
Geología, Universidad Nacional Autónoma de México, Special Publica- cal Society of America Special Paper 52, p. 3–35.
tion 1, p. 201–228. Klein, R.T., Hannula, S.R., Jones, N.W., McKee, J.W., and Anderson, T.H.,
Chamot-Rooke, N., Lallemand, S.J., Le Pichon, X., Henry, P., Sibuet, M., 1990, Cerro El Pedernal: Block-in-mélange features in northern Zacate-
Boulègue, J., Foucher, J.-P., Furuta, T., Gamo, T., Glaçon, G., Kobayashi, cas, Mexico: Geological Society of America Abstracts with Programs,
K., Kuramoto, S., Ogawa, Y., Schultheiss, P., Segawa, J., Takeuchi, A., v. 22, no. 3, p. 35.
Tarits, P., and Tokuyama, H., 1992, Tectonic context of fluid venting at Kusky, T.M., and Bradley, D.C., 1999, Kinematic analysis of mélange fabrics:
the toe of the eastern Nankai accretionary prism: Evidence for a shallow Examples and applications from the McHugh Complex, Kenai Peninsula,
detachment fault: Earth and Planetary Science Letters, v. 109, p. 319–332, Alaska: Journal of Structural Geology, v. 21, p. 1773–1796, doi: 10.1016/
doi: 10.1016/0012-821X(92)90095-D. S0191-8141(99)00105-4.
Chauve, P., Fourcade, E., and Carrillo-Martinez, M., 1985, Les rapports struc- Labarthe-Hernández, G., Tristán-González, M., and Aguillon-Robles, A., 1982,
turaux entre les domaines cordillérain y mésogéen dans la partie centrale Estudio geológico-minero del area de Peñon Blanco estados de San Luis
du Mexique: Comptes Rendus de l’Académie des Sciences, sér. 2., v. 301, Potosi y Zacatecas: Universidad Autónoma de San Luis Potosí, Instituto
p. 335–340. de Geología y Metalurgia Folleto Tecnica 76, 80 p.
Córdoba-M., D.A., 1964, Geology of Apizolaya Quadrangle (east half), northern Lallemand, S.E., Glaçon, G., Lauriat-Rage, A., Fiala-Médioni, A., Cadet, J.-P.,
Zacatecas, Mexico [M.A. thesis]: Austin, The University of Texas, 111 p. Beck, C., Sibuet, M., Iiyama, J.T., Sakai, H., and Taira, A., 1992, Seafloor
Cowan, D.S., 1985, Structural styles in Mesozoic and Cenozoic mélanges in manifestations of fluid seepage at the top of a 2000-metre-deep ridge in
the western Cordillera of North America: Geological Society of America the eastern Nankai accretionary wedge: Long-lived venting and tectonic
Bulletin, v. 96, p. 451–462, doi: 10.1130/0016-7606(1985)96<451: implications: Earth and Planetary Science Letters, v. 109, p. 333–346,
SSIMAC>2.0.CO;2. doi: 10.1016/0012-821X(92)90096-E.
Cuevas-Pérez, E., 1983, Evolución geológica Mesozoic del Estrada de Zacate- Lapierre, H., Ortiz, L.E., Abouchami, W., Monod, O., Coulon, C., and Zim-
cas, México: Zentralblatt für Geologie und Paleontologie, Teil I (3/4), mermann, J.-L., 1992, A crustal section of an intra-oceanic island arc:
p. 190–201. The Late Jurassic–Early Cretaceous Guanajuato magmatic sequence,
Diaz-Salgado, C., Centeno-García, E., and Gehrels, G., 2003, Stratigraphy, central Mexico: Earth and Planetary Science Letters, v. 108, p. 61–77,
depositional environments and tectonic significance of the Taray Forma- doi: 10.1016/0012-821X(92)90060-9.
tion, northern Zacatecas state, Mexico: Geological Society of America López-Infanzón, M., 1986, Estudio petrogenético de las rocas igneas en las
Abstracts with Programs, v. 35, no. 4, p. 71. formaciones Huizachal y Nazas: Sociedad Geológica Mexicana Boletín,
Dickinson, W.R., 1985, Interpreting provenance relations from detrital modes v. 47, no. 2, p. 1–42.
of sandstones, in Zuffa, G.G., ed., Provenance of arenites: Proceedings of Lundberg, N., and Moore, J.C., 1986, Macroscopic structural features in Deep
the NATO Advanced Study Institute on Reading Provenance from Aren- Sea Drilling Project cores from forearc regions: Geological Society of
ites: Bingham, Maryland, D. Reidel Publishing, p. 333–361. America Memoir 166, p. 13–44.
Ducea, M., Shoemaker, S., Gehrels, G., Vervoort, J., and Ruiz, J., 2003, Zircon Martinez-Hernández, S., 1979, Contribución al estudio geológico de una por-
U-Pb geochronology constraints on the magmatic and tectonic evolution ción del sector Vizarrón-Tolimán, Estado de Queretero, México [Bachelor
of the Xolapa Complex, southern Mexico: Geological Society of America thesis]: Mexico, D.F., Universidad Nacional Autónoma de México, Fac-
Abstracts with Programs, v. 99, no. 5, p. 66. ultad de Ingenieria, 82 p.
Downloaded from specialpapers.gsapubs.org on June 27, 2015

The Taray Formation: Jurassic(?) mélange in northern Mexico 455

McKee, J.W., and Jones, N.W., 1979, A large Mesozoic fault in Coahuila, Pacific and Circum-Atlantic Terrane Conference, Guanajuato, Mexico,
Mexico: Geological Society of America Abstracts with Programs, v. 11, November 5–22, Guidebook of Field Trip C, 25 p.
no. 9, p. 476. Pini, G.A., 1999, Tectonosomes and olistostromes in the argille scagliose of the
McKee, J.W., Jones, N.W., and Anderson, T.H., 1988, Las Delicias basin: A record northern Apennines, Italy: Geological Society of America Special Paper
of late Paleozoic are volcanism in northeastern Mexico: Geology, v. 16, 335, 70 p.
p. 37–40, doi: 10.1130/0091-7613(1988)016<0037:LDBARO>2.3.CO;2. Quintero-Legorreta, O., 1992, Geología de la region de Comanja, estados de
McKee, J.W., Jones, N.W., and Anderson, T.H., 1989, The Jurassic margin of Guanajuato y Jalisco: Universidad Nacional Autónoma de México, Insti-
southern North America: Geological Society of America Abstracts with tuto de Geología, Revista, v. 10, no. 1, p. 6–25.
Programs, v. 21, no. 1, p. 34. Ranson, W.A., Fernandez, L.A., Simmons, W.B., Jr., and Enciso de la Vega,
McKee, J.W., Jones, N.W., and Long, L.L., 1990, Stratigraphy and provenance S.E., 1982, Petrology of the metamorphic rocks of Zacatecas, Zac.,
of strata along the San Marcos fault, central Coahuila, Mexico: Geologi- México: Boletín de La Sociedad Geológica Mexicana, v. 43, no. 1,
cal Society of America Bulletin, v. 102, p. 593–614, doi: 10.1130/0016- p. 37–59.
7606(1990)102<0593:SAPOSA>2.3.CO;2. Schaff, P., Hall, B.V., and Bissig, T., 2003, The Puerto Vallarta batholith and
McKee, J.W., Jones, N.W., and Anderson, T.H., 1999, Late Paleozoic and early Cuale mining district, Jalisco, Mexico—High diversity parenthood of
Mesozoic history of the Las Delicias terrane, Coahuila, Mexico, in Bar- continental arc magmas and Kuroko-type volcanogenic massive sulphide
tolini, C., Wilson, J.L., and Lawton, T.F., eds., Mesozoic sedimentary and deposits, in Geological transects across Cordilleran Mexico: 99th Geo-
tectonic history of north-central Mexico: Geological Society of America logical Society of America Cordilleran Section Annual Meeting Field
Special Paper 340, p. 161–189. Trip Guidebook, Instituto de Geología, Universidad Nacional Autónoma
Monod, O., and Calvet, P., 1991, Structural and stratigraphic re-interpretation de México, Special Publication 1, p. 183–199.
of the Triassic units near Zacatecas (Zac.), central Mexico—Evidence of Sedlock, R.L., Ortega-Gutiérrez, F., and Speed, R.C., 1993, Tectonostrati-
a Laramide nappe pile, in Miller, H., Rosenfeld, U., and Weber-Diefen- graphic terranes and tectonic evolution of Mexico: Geological Society of
bach, K., eds., Symposium on Latin American Geosciences: Zentralblatt America Special Paper 278, 153 p.
für Geologie und Palaontologie, Teil 1, p. 1533–1542. Silva-Romo, G., 1993, Estudio de la estratigrafia y estructuras tectónicas de la
Mortensen, J.K., Hall, B.V., Bissig, T., Friedman, R.M., Danielson, T., Oliver, Sierra de Salinas, Estados de San Luis Potosí y Zacatecas [M.Sc. thesis]:
J., Rhys, D.A., and Ross, K.V., 2003, U-Pb zircon age and Pb constraints México, D.F., Universidad Nacional Autónoma de México, Facultad de
on the age and origin of volcanogenic massive sulfide deposits in the Ciencias, 111 p.
Guerrero terrane of central Mexico: Geological Society of America Silver, L.T., and Anderson, T.H., 1974, Possible left-lateral early to middle
Abstracts with Programs, v. 99, no. 5, p. 61–62. Mesozoic disruption of the southwestern North American craton margin:
Ortega-Gutiérrez, F., and Elias-Herrera, M., 2003, Wholesale melting of the Geological Society of America Abstracts with Programs, v. 6, no. 7,
southern Mixteco terrane and origin of the Xolapa Complex: Geological p. 955–956.
Society of America Abstracts with Programs, v. 99, no. 5, p. 66. Stewart, J.H., Gehrels, G.E., Barth, A.P., Link, P.K., Christie-Blick, N., and
Ortega-Gutiérrez, F., Ruiz, J., and Centeno-García, E., 1995, Oaxaquia, Wrucke, C.T., 2001, Detrital zircon provenance of Mesoproterozoic to
a Proterozoic microcontinent accreted to North America during the Cambrian arenites in the western United States and northwestern Mexico:
late Paleozoic: Geology, v. 23, p. 1127–1130, doi: 10.1130/0091- Geological Society of America Bulletin, v. 113, p. 1343–1356, doi:
7613(1995)023<1127:OAPMAT>2.3.CO;2. 10.1130/0016-7606(2001)113<1343:DZPOMT>2.0.CO;2.
Ortega-Gutiérrez, F., Elias-Herrera, M., and Vega-Carillo, J. de J., 1997, A Tardy, M., 1980, Contribution a L’etude geologique de La Sierra Madre Orien-
new Middle Jurassic-Triassic (?) arc-related stratigraphic succession (the tale du Mexique: [Ph.D. thesis]: Paris, L’Universite Pierre et Marie Curie,
Ixcatlán arc) in the southern Mixteco terrane: II Convención Sobre la Evo- 459 p., 6 plates.
lución Geológica de México Y Recursos Asociados, Pachuca, Hidalgo, Tristán-González, M., and Torres-Hernández, J.R., 1994, Geología de la sierra
Universidad Nacional Autónoma de México, Resúmenes, p. 51–52. de Charcas, Estado de San Luis Potosí, México: Revista Mexicana de
Ortiz-Hernández, L.E., and Martinez-Reyes, J., 1993, Geological structure, pet- Ciencias Geológicas, v. 11, no. 2, p. 117–138.
rological and geochemical constraints for the centralmost segment of the
Guerrero terrane (Sierra de Guanajuato, central Mexico): First Circum- MANUSCRIPT ACCEPTED BY THE SOCIETY 24 MARCH 2005

Printed in the USA


Downloaded from specialpapers.gsapubs.org on June 27, 2015
Downloaded from specialpapers.gsapubs.org on June 27, 2015

Geological Society of America Special Papers


The Taray Formation: Jurassic (?) mélange in northern Mexico−−Tectonic
implications
Thomas H. Anderson, Norris W. Jones and James W. McKee

Geological Society of America Special Papers 2005;393; 427-455


doi:10.1130/0-8137-2393-0.427

E-mail 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


Special Papers
Permission request click 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

You might also like