You are on page 1of 22

Significance of Provenance Ages from the Chiapas Massif Complex

(Southeastern Mexico): Redefining the Paleozoic Basement


of the Maya Block and Its Evolution in a
Peri-Gondwanan Realm

Bodo Weber, Victor A. Valencia,1 Peter Schaaf,2 Valerie Pompa-Mera,2


and Joaquin Ruiz1
División Ciencias de la Tierra, Centro de Investigación Científica y de Educación Superior de Ensenada
(CICESE), km. 107 carretera Tijuana-Ensenada, 22860 Ensenada, Baja California, Mexico
(e-mail: bweber@cicese.mx)

ABSTRACT
Medium- to high-grade metasedimentary rocks are exposed as isolated domains in mostly metaigneous crystalline rocks
of the Chiapas Massif Complex (CMC), which forms the basement of the southern Maya Block, southeastern Mexico.
Laser ablation multicollector inductively coupled plasma mass spectrometry U-Pb isotope analyses on inherited detrital
zircon cores show two principal age distributions: (a) Group I, with its highest peak at 500–650 Ma and smaller peaks at
∼380–400 Ma, 1.0–1.2 Ga, 1.5–2.0 Ga, and 2.6–3.1 Ga, and (b) Group II, with its highest peak at 1.0–1.2 Ga, a minor peak
at 1.5–1.6 Ga (or ∼1:0-Ga zircon cores only), and a lack of 500–650-Ma zircons. The cores are commonly surrounded by
metamorphic or anatectic overgrowths, and some zircon ages were reset by a tectonothermal event at ∼250 Ma. The age
distribution of zircon cores from Group I metasediments are similar to detrital zircon ages in Carboniferous-Permian
sediments (Santa Rosa Formation) in southeastern Chiapas, which were shed mostly from continental crust dominated
by Pan-African–Brasiliano orogenic cycles. Group II metasediments have different sources, principally from Grenville
orogens, such as the Oaxacan Complex, and early mid-Proterozoic sources, such as the Rio Negro–Juruena Province of
western Amazonia. The lowest stratigraphic unit in the CMC (the Jocote Unit) is intruded by lower Ordovician S-type
granite and contains only 1.5–1.6-Ga and older detrital zircons. The protoliths of the CMC can be correlated with
Paleozoic strata across the southern Maya Block, and they show parallels with the geologic history and provenance
patterns of protoliths from the Acatlán Complex (Mixteca Terrane) of central southern Mexico. The geologic history of
the CMC and the Maya Block can be interpreted in terms of tectonothermal events accompanying accretion of the CMC
and polarity reversal during formation of the western Pangea margin.

Online enhancements: appendix tables.

Introduction
Analytical improvement of laser ablation multicol- sors (e.g., Fedo et al. 2003). In metamorphosed sedi-
lector inductively coupled plasma mass spectrom- mentary rocks, zircons tend to register complex
etry (LA-MC-ICPMS) has recently facilitated the histories because detrital zircons may either lose
determination of U-Pb ages of individual spots on part of their radiogenic Pb or develop rims of newly
zircon crystals (Gehrels et al. 2006). The statistical formed or recrystallized zircon. Thus, it is still a
analysis of U-Pb ages from individual detrital grains challenge to date tiny detrital zircon cores in meta-
has become a powerful tool for determining possible sedimentary rocks by the laser ablation technique.
source regions of sediments and sedimentary precur- Mexican geology is characterized by a patchwork
of crustal fragments evolved through the Phanero-
Manuscript received April 16, 2008; accepted July 14, 2008. zoic by the assembly and fragmentation of crustal
1
Department of Geosciences, University of Arizona, 1040 blocks (Sedlock et al. 1993; Dickinson and Lawton
East Fourth Street, Tucson, Arizona 85721-0077, U.S.A. 2001). The occurrence of several ∼1:0- Ga granulite
2
Instituto de Geofísica, Universidad Nacional Autónoma exposures in central and southern Mexico has been
de México (UNAM), Ciudad Universitaria, 04510 Coyoacán,
Distrito Federal, Mexico. interpreted in terms of a single microcontinent,

[The Journal of Geology, 2008, volume 116, p. 619–639] © 2008 by The University of Chicago.
All rights reserved. 0022-1376/2008/11606-0006$15.00 DOI: 10.1086/591994

619
620 B. WEBER ET AL.

referred to as Oaxaquia (Ortega-Gutiérrez et al. (Keppie and Ramos 1999). However, there is an on-
1995), that was incorporated into the Rodinia going debate about the location of these land masses
supercontinent during the Grenville orogeny. after the rifting of Rodinia and before the late
Paleozoic sedimentary rocks discordantly overlay Paleozoic assembly of the Pangea supercontinent
the Oaxaquia Terrane in southern Mexico (Pantoja- (Keppie 2004). Existing models suggest that these
Alor and Robinson 1967) and in northeastern landmasses represent peri-Gondwanan blocks or
Mexico (Stewart et al. 1999). Paleozoic igneous outboard terranes related to either the Iapetus
and metamorphic rocks in southern Mexico are (Talavera-Mendoza et al. 2005; Vega-Granillo et al.
exposed in the Mixteca Terrane (Acatlán Complex; 2007) or Rheic oceans (Keppie et al. 2006, 2008a;
fig. 1) and record a complex history of rift magma- Murphy et al. 2006; Nance et al. 2006, 2007). Because
tism, high-pressure metamorphism, decompression of reported U-Pb and Rb-Sr ages in the range of the
melting, low-grade metamorphism, thrusting, and Pan-African–Brasiliano orogenic cycle (Krogh et al.
transtension throughout the Paleozoic (e.g., Keppie 1993; Schaaf et al. 2002; Weber et al. 2006 and
et al. 2008a and references therein). forthcoming), the Maya Block, or at least part of
The general term (cratonic) “Middle America” it, has been interpreted as peri-Gondwanan. On the
was suggested by Keppie and Ortega-Gutiérrez other hand, Grenville-type granulites (Weber and
(1999) and Keppie (2004) for land masses in Mexico, Köhler 1999) link the Maya Block with Oaxaquia.
Central America, and northwestern South America Therefore, the geologic history and basement char-
that have Precambrian or Paleozoic basement and acteristics of the Maya Block are of particular
a likely Gondwanan provenance. Middle America interest.
is inferred to have formed a belt along the In this article, we present new LA-MC-ICPMS
northwestern margin of the Amazon craton of South U-Pb zircon ages from metasedimentary domains in
America in the Late Proterozoic and early Paleozoic the Chiapas Massif Complex (CMC) of southern

Figure 1. a, Plate tectonic overview of Central America, showing Oaxaquia, Maya, and Chortis blocks. b, Geologic
map showing pre-Mesozoic rocks exposed in southern Mexico and central America (modified after Ortega-Gutiérrez
et al. 1992, 2007 and French and Schenk 1997). BVF ¼ Baja Verapaz Fault, CMC ¼ Chiapas Massif Complex,
CUI ¼ Cuicateco Terrane, MFZ ¼ Motagua Fault Zone, MM ¼ Maya Mountains, PFZ ¼ Polochic Fault Zone, Z ¼
Zapoteco Terrane (Oaxaquia Terrane).
Journal of Geology C H I A PA S MA S S I F AG E S 621

Mexico, together with U-Pb zircon and Sm-Nd the northeast over Paleozoic metasediments
garnet–whole-rock ages of a granite that intrudes (San Gabriel Sequence) that are intruded by an
metasedimentary rocks. New constraints on strati- Ordovician granite (Rabinal Suite; Ortega-Obregón
graphic age and the comparison of detrital zircon et al. 2008). North of the Polochic Fault, in the Sierra
ages with provenance ages from Paleozoic sedimen- Los Cuchumatanes (fig. 1), metasedimentary rocks
tary rocks provide new insights into the Paleozoic intruded by a Devonian (391 ± 7 Ma) granitoid
history of the Maya Block and important constraints (Solari et al., forthcoming) are exposed.
on the tectonic evolution of Middle America. At the southwestern edge of the Maya Block,
west of the Tehuantepec isthmus, ∼1- Ga granulites
of the Guichicovi Complex (fig. 1) are similar to
Geologic Frame and Previous Work ∼1- Ga rocks from the Oaxacan Complex (fig. 1;
The Maya Block. The Maya Block (Dengo et al. Ruiz et al. 1999; Weber and Köhler 1999; Weber and
1985) is the southeasternmost of the Mexican Hecht 2003). The Guichicovi Complex is intruded
terranes and includes the Yucatan Peninsula and by Late Permian granitoids (Damon et al. 1981;
southeastern Mexico from the Tehuantepec isth- Murillo-Muñetón 1994; Weber 1998).
mus to Guatemala (fig. 1). The Yucatan Peninsula is The Chiapas Massif. The Chiapas Massif, which
mainly composed of a Cretaceous carbonate plat- covers an area of more than 20; 000 km2 parallel to
form with an important 200-km-wide impact struc- the Pacific coast in southeastern Mexico, is the
ture, the Chicxulub crater, which was formed at the most voluminous Permian batholithic complex in
Cretaceous-Tertiary boundary (e.g., Alvarez et al. Mexico. This complex forms the crystalline base-
1980; Blum et al. 1993). Zircons found in ejecta ment of the southwestern Maya Block (fig. 1) and is
from the Chicxulub event have been dated at about dominated by strongly deformed orthogneisses,
545 Ma (Krogh et al. 1993), indicating that base- anatexites, and amphibolites intruded by less de-
ment rocks include zircons that were formed during formed granitoids (Schaaf et al. 2002).
the Pan-African–Brasiliano orogenic cycle. Pre- Parts of the magmatic precursors of the ortho-
Mesozoic rocks are exposed only in the southern gneisses are less than 20 m.yr. older (∼272 Ma) than
part of the Maya Block: La Mixtequita (Guichicovi a medium- to high-grade metamorphic event that
Complex), the Chiapas Massif, central Guatemala, culminated in anatexis and the intrusion of the main
and the Maya Mountains of Belize (fig. 1). plutonic masses in the latest Permian (254–252 Ma;
The southern boundary of the Maya Block is Weber et al. 2007). Inherited zircon cores from or-
defined by the roughly east-west-trending Motagua thogneisses with ∼1:0- Ga upper-intercept ages (de-
and Polochic fault systems, which separate the termined by isotope dilution–thermal ionization
North American Plate from the Caribbean Plate mass spectrometry [ID-TIMS] on abraded zircons;
(Chortis Block; fig. 1). The fault zone is a complex Weber et al. 2005), together with T DMðNdÞ model ages
composite structural boundary on which left-lateral between 1.0 and 1.4 Ga (Schaaf et al. 2002), indicate
displacement is accommodated along several faults the possible influence of mid-Proterozoic crust
(Guzmán-Speziale 2000). To the west, in Chiapas, during the evolution of the Chiapas Massif. This
these individual faults splay into diffuse structural makes it likely that the Chiapas Massif is underlain
lineaments that may or may not continue to the by a ∼1- Ga basement similar to the granulitic
Pacific Ocean (Muehlberger and Ritchie 1975; Guichicovi and Oaxacan complexes. Such lower-
Burkart and Self 1985; Keppie and Morán-Zenteno crustal rocks, however, are not exposed east of the
2005) and the southern margin of the Chiapas Tehuantepec isthmus.
Massif. The metaigneous units of the Chiapas Massif
In central Guatemala, the Chuacús “group” are interlayered with metasedimentary rocks
(Dengo 1985) or “complex” (Ortega-Gutiérrez et al. (Weber at al. 2002, 2007; Hiller et al. 2004; Estrada-
2004) is situated between the Polochic and Motagua Carmona et al. 2007) that provide important in-
faults (fig. 1). Dengo (1985) correlated the Chuacús formation about the prebatholithic history and
“group” with metasediments from the Maya Block possible correlation within the southern Maya
in Chiapas. However, recent studies suggest that the Block. The provenance of these metasediments is
Chuacús “complex” is possibly an independent, the main topic of this article. In accord with strati-
fault-bounded terrane between the Maya and graphic nomenclature for high-grade metamorphic
Chortis blocks (Ortega-Gutiérrez et al. 2007) that terranes where primary stratigraphic relations have
extends south of the Polochic Fault as far as the Baja been essentially destroyed (Whitaker et al. 1991),
Verapaz shear zone (fig. 1; Ortega-Obregón et al. we propose the new name “Chiapas Massif Com-
2008). The Chuacús “complex” thrusts toward plex” (CMC) in preference to “Chiapas Batholith”
622 B. WEBER ET AL.

(e.g., Morán-Zenteno 1984), which does not include (Hernández-García 1973; López-Ramos 1979).
the metamorphic basement, or “Chiapas Massif” Fossiliferous gray limestones of Leonardian age
(Weber et al. 2005, 2007), which is a geographic (Paso Hondo Formation; fig. 2; Hernández-García
feature. 1973) concordantly overlie the Grupera Formation.
Pre-Mesozoic Stratigraphy of the Maya Block. In On the most recent geologic map (Jiménez-
Chiapas, east of the CMC, Mississippian-Permian Hernández et al. 2005), Paleozoic schists and
sedimentary rocks of the Santa Rosa Formation phyllites are identified in contact with the south-
(SRF) constitute a thick, laminated sequence of eastern part of the CMC (fig. 2), but no details have
mainly siltstone, minor sandstone, and some cal- been reported on the age and contact relationships of
careous horizons (upper SRF) that is best exposed in these sedimentary rocks.
the Chicomuselo area. The Pennsylvanian age of the On the basis of lithological similarities and
upper SRF is documented by fossils (Hernández- stratigraphic relations, the Pennsylvanian-Permian
García 1973). This sequence is underlain by a strata of the SRF in Chiapas have been correlated
monotonous sequence of grayish slates and phyllites with the Santa Rosa Group of central Guatemala
(lower SRF) that crop out south of La Concordia (Bohnenberger 1966; Clemons and Burkart 1971).
(fig. 2; Hernández-García 1973; Weber et al., forth- In the Baja Verapaz area of central Guatemala (south
coming). The SRF is unconformably overlain of the Polochic Fault; fig. 1), conglomerates, shales,
by siliceous shales and limestones (Grupera and limestones with Mississippian conodonts
Formation) containing Early Permian fusulinids (Ortega-Obregón 2005) overlie clastic metasedi-

Figure 2. Simplified geologic map of the study area, showing sample locations (modified from Jiménez-Hernández et al.
2005 and Martínez-Amador et al. 2005).
Journal of Geology C H I A PA S MA S S I F AG E S 623

ments of the San Gabriel sequence (Solari et al., is composed of quartz, K-feldspar, plagioclase, and
forthcoming), which are intruded by the S-type clinopyroxene. Granular titanite is an abundant
Rabinal granite of Ordovician age (∼462–453 Ma; accessory mineral, together with zircon, apatite,
Ortega-Obregón et al. 2008). In the Altos Cuchuma- and alteration minerals such as sericite and epidote.
tanes (north of the Polochic Fault; fig. 1), limestones K-feldspar is mostly microcline, but in some cases
of the Santa Rosa Group are underlain by low- to microperthitic exsolutions are still preserved. The
high-grade metamorphic rocks that are intruded metagreywacke is in contact with calcsilicate rocks
by Permian (269 ± 29 Ma) and Early Devonian mainly composed of quartz and zoned garnet with
(391 ± 7 Ma) granites (Solari et al., forthcoming). anomalous interference colors. The contact zone is
In the Maya Mountains of Belize, Dixon (1956) rich in clinozoisite.
suggested subdivision of the Paleozoic sedimentary Sample CB36 (93.6126°W, 16.2748°N) is a two-
rocks into two series: (1) the Macal series, with mica gneiss with reddish-brown biotite, two
abundant Pennsylvanian to middle Permian fossils, generations of white mica (>1- mm laths and fine-
and (2) the mildly deformed Maya series of unknown grained fibrolithic mica), quartz, plagioclase, and
age, which discordantly underlies the Macal series. minor K-feldspar, together with myrmekite.
Bateson and Hall (1977) proposed combining these Anhedral opaque minerals, of which pyrite is the
sequences and considered them to be equivalent to most abundant, are a major constituent. Occasional
the Santa Rosa Group. Steiner and Walker (1996) sillimanite needles (fibrolite) occur within and
obtained Silurian U-Pb crystallization ages (404–418 around white mica crystals, and cordierite was ob-
Ma) from granites that have contact aureoles in the served in quartz-rich bands. Further accessory min-
sediments of the Maya Mountains, indicating that erals are titanite, zircon, and apatite. We interpret
these sequences are older than the fossils described this sample as having a pelitic precursor.
from the Macal series and the Santa Rosa Group. About 15 km south of Los Angeles (Chiapas),
Recently, Martens et al. (2006) demonstrated the south of the main escarpment of the Chiapas Massif
existence of a lower sequence of Silurian or older mountain range, metapelites are more common.
strata underlying the Pennsylvanian to middle Per- Sample CB33 (93.6636°W, 16.1059°N) is from a river
mian strata of the Santa Rosa Group equivalent outcrop east of Tonalá (fig. 2). Sillimanite and Ti-rich
sequence in the Maya Mountains. biotite define a crenulated foliation in these rocks.
Andalusite is the dominant aluminosilicate phase
and forms 2–5-mm anhedral crystals that overgrow
Metasediments of the CMC the foliation. Aluminosilicate, biotite, and quartz
and Sample Descriptions are the major rock-forming minerals. Opaque min-
Samples were collected in areas of the CMC that are erals and large apatite crystals are abundant. Occa-
dominated by metasedimentary rocks. Compared sional plagioclase and K-feldspar with perthitic
with the abundance of igneous and metaigneous lamellae were also observed. Biotite occurs in two
rocks, metasedimentary rocks are extremely rare in generations, of which the later one is contempora-
the CMC, making correlations between units or neous with fine-grained muscovite, crosscutting the
even individual outcrops difficult or impossible. foliation. The mineral assemblages indicate that in
The geology and petrography of the key units or outcrop CB33, the high-T/medium-P dynamic meta-
outcrops and the analyzed samples are briefly morphism (sillimanite þ biotite 1 þ K-feldspar) was
described in this section. followed by a static medium- to high-T/low-P over-
The Sepultura Unit. Metasedimentary rocks in print (andalusite þ biotite 2 þ muscovite).
the Chiapas Massif were first described from the El At the only road outcrop on the federal road
Tablón River valley at Los Angeles and named the between Tonalá and Arriaga (fig. 2), massive
Sepultura Unit (fig. 2; Weber et al. 2002). At its type calcsilicate rocks with a matrix of carbonate, quartz,
locality, the sequence is mainly composed of meta- epidote, and wollastonite contain large (up to
psammites and of metagreywackes with intercalated 10-cm), euhedral vesuvianite crystals and garnet.
calcsilicates and marbles. The rocks have mineral These rocks are intercalated with intensely folded
assemblages that indicate medium- to high-grade quartz-feldspar-(hornblende)-gneisses. Sample CB44
metamorphism with local anatexis (Weber et al. (93.8052°W, 16.1364°N) is from a quartzite inter-
2007), yielding peak P-T conditions at 730°–780°C layered in these calcsilicates. This fine-grained
and ∼5:8 kbar and retrograde thermal overprinting at quartzite contains brown biotite (not Ti-rich) that
∼540°C and ∼4:5 kbar (Hiller et al. 2004). defines the foliation and also small (<1- mm)
Sample CB54 (93.6138°W, 16.2819°N) is a meta- idiomorphic biotite of the same color in the matrix.
greywacke with a fine-grained, granular texture and Minor constituents are muscovite, plagioclase,
624 B. WEBER ET AL.

garnet, secondary actinolite, and Mg-rich chlorite. CMC into discrete blocks, making correlations dif-
Small lithic fragments composed of plagioclase and ficult. In addition, late Miocene to Pliocene
quartz (±muscovite ± biotite) can be identified by magmatism at the southern edge of the CMC
their larger grain size with respect to the matrix. (Damon and Montesinos 1978) has hydrothermally
The absence of high-T minerals, such as Ti-rich altered the host rocks in this area.
reddish-brown biotite, suggests that outcrop CB44 Sample CB64 (92.3808°W, 15.3396°N) is a mus-
represents a higher crustal level with metamor- covite paragneiss from an outcrop along the dirt
phism in the lower amphibolite facies. A secondary road to San José Ixtepec, close to the southern
static overprint is accompanied by metasomatism, highway to Motozintla (fig. 2). This outcrop is
forming monomineralic zones of epidote and sulfide located south of the main trace of the Polochic
aggregates in the calcsilicates. Fault (after Burkart et al. 1987). The sample is
The Custepec Unit. In the central-southeastern mainly composed of muscovite and quartz. Feldspar
part of the CMC, anatectic amphibolites with or was not observed, but abundant secondary epidote
without garnet define the Custepec Unit, with a is probably the result of the retrogression of plagio-
type locality close to “Finca Custepec” (fig. 2; clase, forming pseudomorphs of former porphyro-
Weber et al. 2007). Because of the strong composi- clasts. Garnet, which is partly chloritized, is locally
tional layering in the amphibolites and the presence present. Muscovite, together with abundant opaque
of metapelites, marbles, and calcsilicate layers or minerals, defines an east-west-trending mylonitic
pebbles, the Custepec Unit has been interpreted as foliation containing a steep stretching lineation
being of sedimentary or volcanosedimentary origin (180°=74°). Shear indicators give top-to-north move-
(Estrada-Carmona et al. 2007). However, it cannot ment, indicating that the southern block has been
be completely excluded that the metasedimentary elevated with respect to the northern. These para-
layers and pebbles are deformed xenoliths in an gneisses are either intruded by or thrust over
igneous amphibolite. The garnet-biotite geother- gabbros and weakly serpentinized pyroxenites of
mometer and several geobarometers in metapelite unknown age that crop out about 1.5 km northeast
and amphibolite yielded peak metamorphic condi- of sample site CB64.
tions above 800°C and 9 kbar (Estrada-Carmona Sample SR05 (92.3270°W, 15.3850°N) is a meta-
et al. 2007). These data, together with the presence sandstone (quartz-phyllite) and is the least meta-
of brown Ti-rich hornblende in garnet-amphibolite morphosed of all the analyzed metasediments. It
and sillimanite and rutile in metapelite, place the comes from an outcrop north of the Polochic Fault,
metamorphic peak in the upper amphibolite- to on the road from Motozintla to Siltepec (fig. 2), that
granulite-facies transition at a 25–30-km depth we define as the Jocote Unit. On the geologic map,
(Estrada-Carmona et al. 2007). Hence, the Custepec the Jocote Unit is located within an area mapped as
Unit not only originated from different protoliths Paleozoic sedimentary rocks in direct contact with
but also represents a deeper crustal level than the the CMC igneous and metamorphic rocks (fig. 2).
Sepultura Unit. The Jocote Unit is intruded by grayish-white garnet-
Sample CB47 (92.9627°W, 15.7240°N) is from bearing S-type granite (sample 705). Sillimanite
an anatectic garnet-amphibolite from an outcrop occurs directly at the intrusive contact. The meta-
south of Custepec. This rock has a medium- sandstone is composed of quartz and fine-grained
grained gneissic texture and is mainly composed white mica that, together with opaque minerals,
of brown hornblende, plagioclase, garnet, and quartz define a crenulated foliation dipping steeply to the
(±biotite). The brown hornblende is rimmed by retro- south (180°=65°). Some larger (∼0:5 -mm) laths of
grade green hornblende. Epidote, clinozoisite, seri- white mica crosscut the foliation, and some plagio-
cite, and chlorite were also formed during retrograde clase is also present.
metamorphism. The presence of a relic bluish am-
phibole inclusion in garnet documents the prograde
Analytical Techniques
path at lower temperatures but similar pressures,
relative to peak conditions (Estrada-Carmona et al. Zircons were separated by standard procedures at
2007). Pebbles of marble and calcsilicate at outcrop the Geology Department of the Centro de Inves-
CB47 demonstrate the sedimentary origin of these tigación Científica y de Educación Superior de
rocks (see above). Ensenada Baja California (CICESE) by using a
The Southern CMC and the Jocote Unit. Close to Wilfley table, a Frantz isodynamic separator, heavy
the Guatemalan border west of Motozintla, east- liquids, and handpicking techniques. LA-MC-
west-trending branches of the sinistral Polochic ICPMS was conducted according to the method
Fault dismember the southernmost part of the described by Gehrels et al. (2006). Zircon cores were
Journal of Geology C H I A PA S MA S S I F AG E S 625

analyzed with a Micromass Isoprobe multicollector


ICPMS linked to a New Wave ArF Excimer laser
ablation system, which has an emission wavelength
of 193 nm. The collector configuration allows mea-
surement of 204 Pb with an ion-counting channel,
whereas 206 Pb, 207 Pb, 208 Pb, 232 Th, and 238 U are
measured simultaneously with Faraday detectors.
All analyses were conducted in static mode with a
laser beam diameter of 35 or 25 μm, operated with
output energy of ∼32 mJ (at 23 kV) and a pulse rate of
8 Hz. Each analysis consisted of one 20-s integration
on peaks with no laser firing for backgrounds and
12 1-s integrations on peaks with the laser firing.
Hg contribution to the 204 Pb mass position was
removed by subtracting on-peak background values.
Interelement fractionation was monitored by
analyzing an in-house zircon standard that has a
concordant ID-TIMS age of 563:5 ± 3:2 Ma (2σ;
Gehrels et al. 2008). This standard was analyzed
once for every five unknown zircon grains. Uranium
and thorium concentrations were monitored by an-
alyzing a standard (NIST 610 Glass) with ∼500 ppm
Th and U. The lead isotopic ratios were corrected for
common Pb by using the measured 204 Pb, assuming
an initial Pb composition according to Stacey and
Kramers (1975) and respective uncertainties of
1.0, 0.3, and 2.0 for 206 Pb=204 Pb, 207 Pb=204 Pb,
and 208 Pb=204 Pb.
Systematic errors were propagated separately and
include the age of the standard, calibration correc-
tion from standard analyses, composition of
common Pb, and U decay constant uncertainties.
For these samples, the systematic errors were 0.9%
for 206 Pb=238 U and 1.3% for 207 Pb=206 Pb. The age
probability plots (Ludwig 2003) and the “best ages”
were based on the 206 Pb=238 U age for young
(<0:9- Ga) zircons and the 207 Pb=206 Pb age for older
(>0:9- Ga) grains. This division at 0.9 Ga was chosen
because of the increasing uncertainty of 206 Pb=238 U
ages and the decreasing uncertainty of 207 Pb=206 Pb
ages as a function of age. All U-Th-Pb data and
apparent ages are shown in table A1, available online
or from the Journal of Geology office.
About 200 mg of garnet was separated from
granite sample 705 at Universidad Nacional
Figure 3. Cathodoluminescence images of zircons from
Autónoma de México (UNAM). The dissolved samples CB36 (a), CB44 (b; numbers are best apparent
garnet sample was split into two aliquots. The ages for individual spots), and CB54 (c), showing laser pits
smaller (1=3) aliquot and two whole-rock aliquots of 35-μm diameter on zircons with inherited cores and
were spiked for isotope dilution analysis (ID run) for metamorphic/anatectic rims.
both Sm and Nd; the larger dissolved garnet aliquot
was for a Nd isotope composition run, to achieve
precise isotope ratios. Sm and Nd isotopic ratios
were measured in static mode on a Finnigan MAT Nd isotopic ratios were corrected for mass frac-
262 mass spectrometer at the Laboratorio Univer- tionation by normalizing to 152 Sm=147 Sm ¼
sitario de Geoquímica Isotópica (UNAM). Sm and 1:78308 and 146 Nd=144 Nd ¼ 0:7219. Sm-Nd isotope
626 B. WEBER ET AL.

data are shown in supplementary table A2, avail-


able online or from the Journal of Geology office.

Results
Metapelite Samples CB36 and CB33 (Sepultura Unit).
Zircons from sample CB36 are mostly rounded or
short and prismatic, having xenocrystic cores with
variable size and appearance (fig. 3a). Zircons from
sample CB33 are similar but considerably smaller
(100–50 μm). We measured 55 spots on zircons from
sample CB36, of which 47 have 207 Pb=206 Pb ages
greater than 900 Ma. The 207 Pb=206 Pb ages of U-Pb
isotope data with less than 10% discordance range
from 959 to 1078 Ma. The concordia plot (fig. 4a)
shows that the data for several individual spots are
discordant but plot on a discordia line with an upper
intercept at 1045 ± 20 Ma, which we consider the
best estimate for the age of most detrital cores from
sample CB36. Only one detrital core has a Silurian
age (206 Pb=238 U age ¼ 428 ± 15 Ma ½2σ), but this is
not a reliable age group (minimum of three zircons
required; Gehrels et al. 2006). The other spots with
Paleozoic ages either are more than 20% discordant
or reflect the age of metamorphism (241–252 Ma),
which is also defined by the lower-intercept age,
calculated at 225 ± 54 Ma. The metamorphic
zircons from sample CB36 can be identified by
their high U=Th (35.5–48.1) and high U contents
(894–1137 ppm). By contrast, the Proterozoic detri-
tal cores have U contents ranging from 50 to 900
ppm (average 200 ppm) and low U=Th (0.6–4.3),
indicating an igneous origin (fig. 4c).
Unfortunately, the cores of zircons from meta-
pelite sample CB33 are too small to have conserved
the age information of their provenance. All but one
of 20 analyzed zircons were formed during high-
grade metamorphism, having 206 Pb=238 U ages
between 213 and 252 Ma, high U=Th (5.7–50.4,
average 29.4), and high U contents (477–1553
ppm, average 933 ppm). The isotope data from spots
that yield the youngest ages are slightly discordant
(fig. 4b), indicating open-system behavior and pos-
sible lead loss. This interpretation is consistent
Figure 4. Concordia diagrams for U-Pb isotope ratios of with the petrographic observations on sample
zircons from the Sepultura Unit samples CB36 (a) and CB33. Only one grain, with a 206 Pb=238 U age of
CB33 (b). Inset, weighted average calculated with the 292 ± 18 Ma (2σ), reflects the age of an inherited
TuffZirc algorithm (Ludwig 2003) yields a coherent group
component or a mixed age of inheritance and meta-
of 13 spots (black bars). The asterisk indicates that propa-
gated age errors include 0.9% of systematic error for
morphism (fig. 4b). In view of the negative (from
206 Pb=238 U ages. Error ellipses and bars of individual spots lead loss) and positive (from inherited cores) age
are 1σ. c, U=Th versus best age of data from both samples, biases of isotope data from sample CB33, we con-
discriminating between igneous and metamorphic zircon. sider an age calculation using “TuffZirc Age” algo-
For best ages, 206 Pb=238 U apparent ages were used for rithm (Isoplot 3.0; Ludwig 2003) to provide the
average apparent ages <900 Ma, and 207 Pb=206 Pb apparent most reliable estimate for the age of high-grade
ages were used for average apparent ages >900 Ma. metamorphism, at 248:5þ2:7 7:5 Ma (2σ). This age is
Journal of Geology C H I A PA S MA S S I F AG E S 627

identical, within errors, to ages for the high-grade minescence imaging and not the metamorphic rims
metamorphism reported from the CMC (Weber (fig. 3c), we suggest that, whereas the detrital (igne-
et al. 2007). ous) cores lost their radiogenic lead during meta-
Quartzite Sample CB44 (Sepultura Unit). Zircons morphism, the Th and U contents remained almost
from sample CB44 are of variable shape and size unchanged.
(fig. 3b). Apparent ages of 100 analyzed grains range The U-Pb isotope data for detrital zircon cores
from 284 to 3004 Ma. Most of the U-Pb isotope from sample CB54 are often discordant, and a total
ratios are concordant (fig. 5a, 5b). A group of six of eight zircon cores are most probably of Archean
zircons have Archean 207 Pb=206 Pb ages (2.65– and Early Proterozoic age (fig. 5d). Twelve zircon
3.00 Ga). Another seven spots yielded Early Prote- cores have 207 Pb=206 Pb ages in the range from 968 to
rozoic ages from 1540 to 2325 Ma. Twenty-nine 1254 Ma, but several apparently younger zircon
grains have 207 Pb=206 Pb ages in the range from 943 cores are discordant and plot on a discordia line
to 1290 Ma. The highest probability peak of this with an upper-intercept age at 1019 ± 40 Ma (fig. 5e).
provenance group is at 1067 Ma (fig. 5c). A small Thus, we consider most of the 206 Pb=238 U ages
fraction of two grains is ∼850 Ma old. between 700 and 900 Ma (fig. 5f) to date zircon
The most prominent provenance group of 25 cores of Grenville age that lost part of their radio-
zircons from sample CB44 has apparent ages genic lead during Permian metamorphism. The
between 500 and 650 Ma, with the highest proba- most prominent provenance group (n ¼ 25) are
bility peak at 543 Ma (fig. 5c). A smaller group of eight zircon cores from Pan-African–Brasiliano orogenic
grains has ages between 704 and 781 Ma. A total of belts in the range from 517 to 669 Ma, with major
eight grains have Ordovician to Devonian apparent probability peaks at 605 and 560 Ma (fig. 5f).
ages, but this group of zircons does not define any age Paleozoic ages are represented by Silurian,
peak. The youngest apparent age of 284 ± 19 Ma (2σ) Devonian, and two Carboniferous zircon cores.
may reflect the maximum age of deposition rather Provided that no lead loss affected the latter
than the age of metamorphism, because the spot was zircons, they define the maximum age of deposition
measured in the center of a zircon grain showing at ∼308 Ma. However, since lead loss also can be
igneous zoning (fig. 3b), no visible metamorphic expected for these grains, an older depositional age
overgrowth, and a low U=Th of 1.1. In general, all may be plausible.
but eight spots yielded U=Th < 7, indicating prov- Amphibolite Sample CB47 (Custepec Unit). Only a
enance from igneous rocks. few analyses of laser spots on zircon cores from this
Metagreywacke Sample CB54 (Sepultura Unit). high-grade garnet amphibolite yielded reliable data
The provenance groups for 100 analyzed spots from (fig. 6). Many laser experiments on these cores
sample CB54 are similar to those from sample started with isotope ratios corresponding to the
CB44, but their isotope data are more complex age of the core, but as the laser burned deeper into
(fig. 5d–5f). This is due to the high-grade metamor- the zircon, it passed through the cores and sampled
phism at site CB54, which produced metamorphic younger zones, making such analyses useless. Five
overgrowth on these zircons (fig. 3c). For the same analyzed spots yielded reliable Grenville-type ages.
reason, some inherited detrital cores have partly or One discordant grain is clearly older, having a
207
completely lost their radiogenic lead. Pb=206 Pb age of 1441 ± 50 Ma (2σ; fig. 6a). An-
Twenty analyzed cores range from 224 to 299 other analysis yielded concordant 206 Pb=238 U and
207
Ma, and we suggest that the isotope ratios of these Pb=235 U isotope ratios with a 206 Pb=238 U age of
analyses reflect the time of lead loss and resetting 445 ± 57 Ma (2σ). This is the youngest detrital core,
during metamorphism. Age calculation using the and it may indicate the maximum age of deposition
“TuffZirc age” algorithm yields a coherent group of if the rock is sedimentary not igneous; however, it
12 analyses with an age of 260 ± 8 Ma (2σ; fig. 5e). is not a reliable group of at least three ages (Gehrels
The group of analyses giving ages ranging from 284 et al. 2006).
to 299 Ma may date the youngest detrital cores, Most analyzed spots from sample CB47 have
206
but it is also probable that these zircon cores lost Pb=238 U ages ranging from 210 to 280 Ma and
much but not all of their radiogenic lead during have age biases due to either lead loss or inherited
metamorphism. Thus, we omitted the entire group components (fig. 6b). A coherent group of 47 spots
of the 20 youngest zircons from further provenance yielded an age of 250:2þ4:9
3:8 Ma (2σ; fig. 6b), which is
considerations. It is noteworthy, however, that the in agreement with the age of metamorphism and
U=Th ratios of these analyses range from 0.4 to 5.7, anatexis reported by Weber et al. (2007).
indicating igneous and not metamorphic origin. Paragneiss Sample CB64. Twenty-two out of 38
Because we analyzed cores visible with cathodolu- concordant isotope ratios analyzed from sample
Figure 5. a, b, Concordia diagrams for U-Pb isotope ratios of zircons from sample CB44. c, Relative age probability and
histogram plot of best ages for zircons from sample CB44. d, e, Concordia diagrams for U-Pb isotope ratios of zircons
from sample CB54; upper-intercept age for CB54 calculated from isotope ratios of gray dots. e, inset, Weighted average
calculated with TuffZirc algorithm (Ludwig 2003), yielding a coherent group of 12 spots (black bars). f, Relative age
probability and histogram plot of best ages for zircon cores from sample CB54 (only ages older than Permian are shown).
Errors and definitions as for figure 4.
Journal of Geology C H I A PA S MA S S I F AG E S 629

Metasandstone Sample SR05 (Jocote Unit) and S-Type


Granite Sample 705. Metasandstone sample SR05
contained only a few small (40–80 μm) rounded or
elongate zircons, from which 34 grains could be
separated. A group of 23 of those grains, of which
nine are more than 10% discordant, define a regres-
sion line with an upper-intercept age at 1546 ±
42 Ma and a Late Proterozoic to early Paleozoic
lower intercept (fig. 7b). These results indicate that
two-thirds of the analyzed grains have 1.5–1.6-Ga
provenance ages and that no younger zircons are
present. The 206 Pb=238 U ages of the remaining 11
grains range from 1602 to 1822 Ma.
The garnet-bearing S-type granite (sample 705)
that intrudes the metasediments of the Jocote Unit
has few zircons, most of which preserve inherited
ages. Twenty-three (out of 27) analyzed spots yield
mid-Proterozoic ages, with peaks at 931, 1229, and
1491 Ma. The oldest grain has an age of ∼1:88 Ga
(fig. 7c). Only three spots on zircon rims yielded
concordant isotope ratios that reflect the time
of intrusion, and these give a concordia age of
482 ± 5 Ma (2σ; fig. 7c).
To further document the crystallization age of
the S-type granite, we analyzed garnet and whole
rock from the same sample (705) for Sm-Nd. The
Early Ordovician age obtained from the U-Pb iso-
tope ratios of the zircon rims is consistent with the
Sm-Nd garnet–whole-rock isochron, which yielded
an age of 473 ± 39 Ma (fig. 7d).

Figure 6. Concordia diagrams for U-Pb isotope ratios of


zircons from sample CB47. a, Inset, Relative age proba- Discussion
bility plot for best ages of inherited cores. b, Inset, Provenance and Depositional Ages of CMC Group I
weighted average calculated with the TuffZirc algorithm Metasediments. We define the metagreywacke
(Ludwig 2003), yielding a coherent group of 47 spots
(CB54) and quartzite (CB44) samples from the
(black bars). Errors and definitions as for figure 4.
Sepultura Unit as CMC Group I metasediments.
Both samples are from outcrops with interlayered
calcsilicates, and both contain detrital zircon cores
CB64 yielded 207 Pb=206 Pb ages from 968 to 1238 Ma, that clearly differ from those of the metapelites (and
with the highest probability peak at 1017 Ma structurally related psammites) in the same area.
(fig. 7a). Another six spots yielding ages between Detrital zircons from CMC Group I metasediments
827 and 942 Ma are slightly (2.9%–5.5%) discordant were derived from regions dominated by rocks that
but of the same provenance. A group of five zircons formed during the Pan-African–Brasiliano orogenic
range in age from 1344 to 1648 Ma, with a proba- cycle, together with Grenville, mid-Proterozoic,
bility peak at 1628 Ma. A third group of five small Early Proterozoic, Archean, and early Paleozoic
zircons have early Paleozoic 206 Pb=238 U ages. Two sources (fig. 8a). The age distributions are similar
of these (∼407 and ∼426 Ma) are 8% and 4% dis- to those reported from the Carboniferous to Early
cordant, with U=Th of 3.0 and 1.5, respectively. The Permian SRF of southeastern Chiapas (fig. 8a; Weber
other three grains (435–461 Ma) are 10%–31% dis- et al. 2006 and forthcoming). Alternating quartzites,
cordant and have high U=Th (32–93). It remains quartz-feldspar-rich layers with lithic fragments,
unclear, however, whether the early Paleozoic zir- calcsilicates, and marbles observed in the Sepultura
cons are detrital grains that were affected by some Unit are also consistent with the stratigraphic suc-
form of U gain or metamorphic zircons possibly cession reported from the SRF (Hernández-García
linked to neighboring mafic intrusives. 1973).
630 B. WEBER ET AL.

Figure 7. a–c, Concordia diagrams for U-Pb isotope ratios, relative age probability plots, and histograms of zircons from
sample CB64 (a), sample SR05 (b; intercept ages calculated from isotope ratios indicated by black dots), and S-type
granite sample 705 (c). Errors and definitions as for figure 4. d, Sm-Nd isochron plot for two whole-rock analyses and
garnet from S-type granite sample 705.

The depositional age of these rocks is difficult to data from other Paleozoic sedimentary rocks and
estimate because the youngest zircons reflect the protoliths of southern Mexico reveals an outstand-
time of metamorphism and lead loss in the zircon ing similarity in provenance age patterns of the
cores. But one igneous zircon core from the CMC Group I metasediments, the SRF (fig. 8a),
quartzite CB44 (286 ± 20 Ma ½2σ), together with and Carboniferous to Permian metasedimentary
two cores with similar ages from metagreywacke rocks of the Acatlán Complex in the Mixteca
CB54 (∼305 Ma), may tentatively define the maxi- Terrane (fig. 8b; Talavera-Mendoza et al. 2005;
mum age of deposition in the Early Permian, re- Keppie et al. 2006, 2008a; Grodzicki et al. 2008;
affirming the possible correlation with the upper Morales-Gámez et al. 2008). The comparable dis-
SRF, from which two similar but discordant detrital tributions of zircon ages suggest that at least part of
zircon ages have been reported (fig. 8b; Weber et al. the Carboniferous to Permian sedimentary rocks
2006). These data suggest that late Paleozoic strata constituting the Acatlán Complex may correlate
correlative with the SRF extend into the CMC as far with CMC Group I metasediments. On the other
as the Arriaga-Tonalá area southwest of the main hand, the Acatlán Complex also contains older (pre-
mountain range escarpment of the Chiapas Massif. Silurian) sedimentary protoliths with age patterns
Comparison of the detrital zircon age data from similar to those of CMC Group I (fig. 8b; Keppie
the CMC Group I metasediments with published et al. 2008b).
Journal of Geology C H I A PA S MA S S I F AG E S 631

Provenance and Depositional Ages of the CMC Group ∼392–413 Ma reported for white micas from the
II Metasediments. We define the second group of Jocote Unit (Pompa-Mera et al. 2007).
CMC metasedimentary rocks by the absence of Late The metasandstone from the Jocote Unit (fig. 8c)
Proterozoic (Pan-African–Brasiliano) zircon cores in contains zircons derived almost exclusively from a
their age patterns. Accordingly, the remaining meta- ∼1:5–1:6-Ga provenance, indicating that such a
sedimentary rocks sampled in the study are included source, which is atypical for Oaxaquia (fig. 8f), plays
in CMC Group II. The metapelite samples from the an important role in the sedimentary precursors of
Sepultura Unit have only small mid-Proterozoic the CMC. The Jocote Unit is intruded by garnet-
detrital zircon cores. However, the results are con- bearing S-type granite of Early Ordovician age that
sistent with SHRIMP data from an anatectic meta- constrains the minimum age of deposition and
psammite interbedded with metapelites from the documents the first evidence for sedimentary pre-
same unit (fig. 8c; Weber et al. 2007) and with the cursors of pre-Ordovician age in the CMC, which is
average upper-intercept age of CMC othogneisses the lowest stratigraphic unit recognized in the
analyzed by ID-TIMS (Weber et al. 2005). Since they CMC and which probably underlies the other
have only zircons of a single provenance, the meta- CMC Group II metasedimentary rocks.
pelites may have been deposited in an isolated basin The stratigraphic relationship between the indi-
or in a distal environment that received detritus vidual outcrops is unknown, but we propose that
principally from a Grenville-type continental mass. the CMC Group II metasediments are a succession
A single grain with a Silurian (∼428-Ma) age may of partly pelitic, psammitic, and volcanosedimen-
indicate the maximum age of deposition of the tary rocks whose depositional ages range from pre-
metapelites. This is not, however, a reliable prov- Ordovician (Jocote Unit) to lower mid-Devonian
enance group (Gehrels et al. 2006), and it must be (CB64 and possibly the Custepec Unit) and which
treated with caution. Furthermore, since the meta- originally underlay CMC Group I metasediments.
pelites from the Sepultura Unit are in the vicinity of The metapelites from the Sepultura Unit that con-
the calcsilicates and metagreywackes of Group I, it tain only Grenville-age cores either are part of this
is possible that the pelitic protoliths have deposi- early Paleozoic succession of CMC Group II or
tional ages similar to those of Group I metasedi- represent stratigraphic intervals within CMC
ments with different provenances. Group I metasediments with different provenances.
In the Custepec Unit of the central CMC, high- Similar Early Ordovician S-type granites (Rabinal
grade metamorphism and anatexis at depths prob- Suite; Ortega-Obregón et al. 2008) that intrude the
ably greater than 25 km (Estrada-Carmona et al. San Gabriel Unit south of the Polochic Fault Zone
2007) are such that only a few detrital zircon cores in central Guatemala suggest a correlation of CMC
were large enough to retain information on their Group II metasediments with early Paleozoic strata
origin. These provenance ages are poorly con- from Guatemala (fig. 8f). Furthermore, the age dis-
strained but are consistent with SHRIMP data from tribution of zircon cores from Group II samples
a similar sample from the same area (Weber et al. is also similar to that from the pre-Devonian
2007) that contains two main provenance groups, sediments of Belize (fig. 8c; Martens et al. 2006;
one with Grenville-age zircons (∼1:0–1:2 Ga) and U. Martens, B. Weber, and V. Valencia, unpub.
the other with ∼1:5–1:6-Ga zircons (fig. 8c). manuscript).
In the southern part of the CMC, metamorphic Early Paleozoic metasedimentary units that are
conditions reached only the lower amphibolite intruded by Ordovician rift-related leucogranites
facies, and consequently, major lead loss in the are common in the Acatlán Complex (Mixteca
zircon cores did not occur. Detrital zircons from Terrane; fig. 1; Keppie et al. 2008a and references
paragneiss CB64 from south of the Polochic Fault therein). Recently published detrital zircon ages
have provenance groups similar to those of the from most of the pre-Silurian protoliths from the
Custepec Unit, with an additional cluster of small Acatlán Complex have age distributions with
Silurian to lower Devonian zircons that may mainly 1.0–1.2-Ga zircons but also with appreciable
indicate the maximum age of deposition (fig. 8c). ∼1:5–1:6-Ga peaks (fig. 8d; Sanchez-Zavala et al.
The Silurian to lower Devonian zircons have both 2004; Talavera-Mendoza et al. 2005; Vega-Granillo
igneous and metamorphic U=Th signatures. There- et al. 2007; Morales-Gámez et al. 2008). Some post-
fore, the possibility of a Silurian to lower Devonian Ordovician metasedimentary rocks from the Acat-
metamorphic overprint can be considered, and lán Complex that contain detrital zircons from
hence, lower Devonian also might be a mini- Ordovician plutonic rocks also have age patterns
mum age of deposition. This interpretation is that indicate Grenville-type (1.0–1.2-Ga) and addi-
further supported by Rb-Sr and Ar-Ar ages of tional ∼1:5–1:6-Ga sources (fig. 8d; Sanchez-Zavala
Journal of Geology C H I A PA S MA S S I F AG E S 633

et al. 2004; Grodzicki et al. 2008). These age dis- CMC group I metasediments have provenance
tributions are similar to those observed from the ages that are dominated by 500–650-Ma sources that
CMC Group II metasediments, indicating similar- are typical for Gondwanan protoliths, and they cor-
ities between the Maya Block and at least part of the relate with the SRF (fig. 8). CMC Group II metasedi-
early Paleozoic precursors that constitute the poly- ments are dominated by ∼1:0–1:2-Ga sources typical
metamorphic Acatlán Complex. However, none of for Grenville-type orogens, and they show prove-
the reported age spectra from metasedimentary nances and depositional ages similar to those of early
rocks of the Acatlán Complex or any other Paleo- Paleozoic strata from the Maya Mountains of Belize
zoic rocks from southern Mexico show a single (Martens et al. 2006; U. Martens, B. Weber, and
∼1:5–1:6-Ga probability peak like that in our data V. Valencia, unpub. manuscript) and from central
from the Jocote Unit. Guatemala (San Gabriel Unit; fig. 8f; Ortega-Obregón
Paleozoic sedimentary rocks covering the et al. 2008). Ordovician to Devonian plutons
1.0-Ga-old Oaxaquia Terrane (fig. 8e, 8f; Gillis et al. that intrude older sedimentary rocks have also
2005) do not contain any appreciable amount of been reported from the Cuchumatanes area in
∼1:5–1:6-Ga zircons. This indicates either (1) that Guatemala (Solari et al., forthcoming) and from the
the lower Paleozoic sediments of the Maya Block Mexican-Guatemalan border near Motozintla
are derived not from Oaxaquia but from another (Salazar-Juárez et al. 2007). This suggests that pre-
Grenville-type orogen that also included ∼1:5– Ordovician to lower Devonian strata such as the
1:6-Ga-old precursors or (2) that the supply Jocote Unit (this article), the San Gabriel Unit
of ∼1:5–1:6-Ga zircons did not play a role in (Ortega-Obregón et al. 2008), and strata of similar
the Paleozoic sediments overlying the Oaxacan age from the Maya Mountains of Belize (Martens
Complex. et al. 2006; U. Martens, B. Weber, and V. Valencia,
unpub. manuscript), all of which are discordantly
overlain by Late Carboniferous to Permian “flysch-
Conclusions and Paleogeographic Implications
type” strata, are typical of the entire southern Maya
Although medium- to high-grade metamorphic Block and can be stratigraphically correlated into the
conditions during the Late Permian caused partial CMC (fig. 8f). Therefore, we suggest an autochtho-
or total lead loss in many analyzed zircon cores, the nous position for the CMC with respect to the un-
data provide useful information that enables us to metamorphosed Maya Block in the Paleozoic. In such
correlate the precursors of the CMC with Paleozoic a scenario, the sedimentary precursors of the CMC,
sedimentary rocks from the Maya Block and other including the Late Carboniferous to Early Permian
terranes, particularly the Mixteca Terrane. Impor- strata similar to the SRF, must have been buried to
tantly, the data reveal that two different types of depths of about 15 km (5–6 kbar) in ∼20 m:yr: where
sedimentary protoliths can be distinguished. they were metamorphosed at ∼260–248 Ma (Weber

Figure 8. a–e, Age probability and histogram plots of detrital zircon ages (all histograms show number of analyses
within each 100-Ma interval) from Chiapas Massif Complex (CMC) Group I metasediments (a, top; samples CB44 and
CB54); the Santa Rosa Formation in Chiapas (a, bottom; data from Weber et al. 2006 and forthcoming); metasedimentary
rocks from the Acatlán Complex with mainly Pan-African–Brasiliano zircons (b): Carboniferous units (top; Cosoltepec
Formation, Salada and Coatlaco units; data from Talavera-Mendoza et al. 2005; Keppie et al. 2006; Grodzicki et al. 2008;
Morales-Gámez et al. 2008) and pre-Silurian units (bottom; Las Minas and Mal Paso units; data from Keppie et al.
2008b); CMC Group II metasediments (c, top three plots; data from Weber et al. 2007 and this work); the Maya series
from the Maya Mountains in Belize (c, bottom; data from U. Martens, pers. comm.); metasedimentary rocks from the
Acatlán Complex with mainly mid-Proterozoic ages (d): post-Ordovician units (top; Tecomate Formation and Canoas
Unit; data from Sanchez-Zavala et al. 2004; Grodzicki et al. 2008) and metasedimentary units intruded by Ordovician
granites (bottom; Tecomate and Xayacatlán formations, Esperanza Suite, and Amate Unit; data from Sánchez-Zavala
et al. 2004; Talevera-Mendoza et al. 2005; Vega-Granillo et al. 2007; Morales-Gámez et al. 2008); and Paleozoic
sedimentary rocks, overlaying the mid-Proterozoic Oaxacan complex (Oaxaquia Terrane) (e; data from Gillis et al. 2005).
f, Simplified revised pre-Mesozoic stratigraphy of the Maya Block, comparing Chiapas (this work, revised from
Hernandez-García 1973 and considering information from Salazar-Juárez et al. 2007), Guatemala (revised from Clemons
and Burkart 1971 and considering information from Ortega-Gutiérrez et al. 2007; Ortega-Obregón et al. 2008; and Solari
et al., forthcoming), and Belize (revised from Dixon 1956 but including information from Martens et al. 2006;
U. Martens, B. Weber, and V. Valencia, unpub. manuscript). Additional columns of comparable units are from the
Mixteca Terrane (e.g., Keppie et al. 2008a) and from Oaxaquia (Gillis et al. 2005).
634 B. WEBER ET AL.

et al. 2007 and this article). The marked difference in the Maya Mountains (Martens el at. 2006) indicate
the age spectra of the Group I and Group II detrital ongoing extensional tectonics in the Maya Block.
zircons from the Maya Block suggests a major change Parts of the CMC Group II metasediments that
in provenance due to a different paleogeographic contain Silurian zircons indicate synchronous
situation, river drainage, or marine currents. sedimentation from combined Grenville-type and
During the early Paleozoic, terranes with Rio Negro–Juruena sources. Silurian rocks are sup-
Grenville affinity, such as Oaxaquia and Chortis, posed to be absent in the Mixteca Terrane (Keppie
were located on or outboard of the western margin et al. 2008a), but Silurian sedimentary rocks with a
of Gondwana and so could have provided detritus Gondwanan fauna rest unconformably on Oaxaquia
into a basin or to the continental rise of northern in northeastern Mexico (Stewart et al. 1999), indi-
Gondwana (fig. 9a, 9b). Since ∼1:5–1:6-Ga precur- cating a possible connection between the Maya
sors are not present in these terranes, we suggest Block and northeastern Mexico (Sierra Madre
that the detrital zircons of this age were shed from Terrane; fig. 9b).
the ∼1:5–1:6-Ga-old Rio Negro–Juruena Province of During the Late Carboniferous and Early
western Amazonia (Cordani et al. 2000 and refer- Permian, sediments were probably shed southwest-
ences therein). Occasional 1.4–1.5-Ga grains may ward from West Africa as a result of the erosion of the
have come from the Rondônia-San Ignacio Province newly formed mountain chains that resulted from
adjacent to the Rio Negro–Juruena Province (fig. 9a, the collision of Gondwana with Laurentia (fig. 9c,
9b; Geraldes et al. 2001). 9d). This orogenic belt contains components of
During the Cambrian to Ordovician, peri- Pan-African–Brasiliano crust that was probably
Gondwanan terranes (e.g., Avalonia and Carolina; not exposed to erosion before the Carboniferous.
fig. 9a) were transported from Gondwana to The outstanding accordance of detrital zircon age
Laurentia (e.g., Murphy et al. 2006) as the early patterns from Carboniferous rocks of the Maya
Paleozoic Iapetus Ocean closed and the Rheic Block with sedimentary precursors of similar age
Ocean opened (fig. 9a, 9b). This drift was accom- from the Acatlán Complex (fig. 8b) requires a
panied by rifting along the Gondwanan margin that reexamination of the paleogeographic models dis-
led to bimodal magmatism in southern Mexico cussed for the Acatlán Complex. A direct correlation
during Ordovician times and to the synchronous between the Mixteca Terrane and the Maya Block is
deposition of sediments, especially in the Mixteca unlikely, however, because the Acatlán Complex is
Terrane (e.g., Keppie et al. 2008a). Interestingly, the thought to have been located west of Oaxaquia,
observed similarities between early Paleozoic rocks facing the Rheic Ocean, during the Carboniferous
from the Acatlán Complex and CMC Group II and not east of Oaxaquia, as is thought for the Maya
metasediments, both of which contain Ordovician Block (e.g., Keppie 2004). However, if similar detrital
leucogranites, indicate comparable geological envi- age patterns indicate similar sources and trans-
ronments for the Mixteca Terrane and the Maya portation currents, then it is likely that the
Block. However, according to paleogeographic mod- Carboniferous units of the Acatlán Complex, with
els proposed by Keppie et al. (2008a and references mainly Neoproterozoic Gondwanan sources (fig. 8b),
therein), the Mixteca Terrane and the Maya Block were deposited in close proximity to and within the
were located in quite different paleogeographic same geological realm as synchronous sediments
positions during the Paleozoic, but both were from the Maya Block (CMC Group I and the SRF).
located along the Gondwana margin. As illustrated This favors a position for the Carboniferous units of
in figure 9a, several individual (pull-apart) basins the Acatlán Complex north of the Maya Block (X in
may have formed during early Paleozoic drift-rift fig. 9c), similar to that proposed by Vega-Granillo
tectonics. In such a scenario, sediments that con- et al. (2007) for the entire Mixteca Terrane. This
stitute the Jocote Unit in the Maya Block were hypothesis requires that these units came into con-
probably shed through a draining system that was tact with the rest of the Acatlán Complex until
provided by detritus mainly from the Rio Negro– Permian dextral faulting and thrusting occurred
Juruena Province (fig. 9a). In the Mixteca Terrane, during the final amalgamation of Pangea.
instead, the continental rise or rift basins were filled In our model, progressive movement of the
preferentially by detritus from Oaxaquia or similar continent-continent collision from northwest to
Grenville-type sources, and detritus from the Rio southeast (zipper tectonics; Hatcher 2002) caused
Negro–Juruena Province was less abundant. the westward tectonic escape of the Maya Block
Rift granites of Silurian age (Steiner and Walker (fig. 9c, 9d). During the Early Permian, arc magma-
1996) and a Late Silurian to Early Devonian tuff tism occurred in Oaxaquia (∼275 Ma; Solari et al.
intercalated with sandstones and conglomerates of 2001), the Mixteca Terrane (∼288 Ma; Yañez et al.
Journal of Geology C H I A PA S MA S S I F AG E S 635

Figure 9. a–c, Geotectonic reconstructions of northwest (after present coordinates) Gondwana, showing the proposed
positions of Proterozoic and Paleozoic crustal blocks of the Mexico, Central America (Chortis), Avalonia, and Carolina
terranes (modified after Keppie et al. 2008a). The present-day outlines of the Amazonian craton are shown, together with
the Rio Negro–Juruena Province (RNJ ), the Rondônoa–San Ignacio Province (RO), and the Pan-African–Brasiliano belts
(after Cordani et al. 2000). Arrows indicate probable sediment transport directions. a, Cambrian to Ordovician
separation of Avalonia from Gondwana, drift-rift producing independent (pull-apart) basins and rift magmatism in
the Mixteca Terrane and the Maya Block. b, Silurian to Early Devonian opening of the Rheic Ocean and further rifting in
the Maya Block. c, Late Carboniferous reconstruction showing the Acatlán Complex on a proposed active margin (after
Keppie et al. 2008a), sediment transport mainly from West Africa, and the Maya Block moving westward because of
tectonic escape induced by Gondwana-Laurentia collision. X indicates our hypothetical position of Carboniferous
sediments with Gondwana sources in the Acatlán Complex; open triangles west of the Maya Block illustrate
the proposed westerly-dipping subduction in the early Permian. d, Cartoon illustrating the formation of the Chiapas
Massif Complex (CMC) within the Maya Block. Top, Early Permian deposition of sediments of the uppermost Santa
Rosa Group from western Gondwana, overlying discordantly pre-Ordovician to Early Devonian sedimentary rocks;
approach of Oaxaquia toward northeast and westward motion of Gondwana. Bottom, stacking of the southern
Maya Block against Oaxaquia; flat, eastward subduction causes high-T metamorphism and batholith intrusion in
the CMC. Ac ¼ Acatlán Complex (Mixteca Terrane); BR ¼ Brasiliano belts; CA ¼ Colombian Andes; Car ¼ Carolina;
Cho ¼ Chortis; F ¼ Florida; M ¼ Mérida Terrane; Oax ¼ Oaxaquia; SM ¼ Sierra Madre Terrane.

1991), and the CMC (∼272 Ma; Weber et al. 2007) ued westward movement of the Maya Block and,
that is explained by eastward oblique subduction consequently, convergence and accretionary tecton-
(e.g., Keppie et al. 2008a). In our hypothetical model, ics started by the Early Permian (∼280 Ma; fig. 9d).
coeval northeastward movement of Oaxaquia and We consider the development of a short-lived
part of the Acatlán Complex (Elías-Herrera and westerly-dipping subduction beneath Oaxaquia a
Ortega-Gutiérrez 2002) probably prevented contin- possible explanation for arc magmatism in the
636 B. WEBER ET AL.

CMC during the Early Permian (fig. 9d), before the detailed work in the CMC and future reexamina-
development of the easterly-dipping subduction tions of the existing tectonic models.
zone in the Late Permian. We have two arguments
in favor of polarity reversal. (1) Paleogeographic
reconstructions (e.g., Dickinson and Lawton 2001)
ACKNOWLEDGMENTS
place the Maya Block east of Oaxaquia, but in the
Early Permian, it was probably too far east to explain This work was supported by Consejo Nacional de
arc magmatism in the CMC by an easterly-dipping Ciencia y Tecnología project D41083-F. Many
subduction zone in the proto–Pacific Ocean. (2) thanks to S. Rosas-Montoya, V. Pérez-Arroyoz,
Ultramafic rocks that occur in the southern CMC and G. Rendón-Márquez (CICESE) for their help
may be relics of oceanic crust; however, nothing is with sample preparation. We appreciate discussions
known about the age and emplacement mechanisms with U. Martens (Stanford University), F. Ortega-
of these ultramafic rocks. Gutiérrez (UNAM), and M. Elías-Herrera (UNAM).
We suggest that convergence culminated in We are grateful to J. D. Keppie (UNAM) and
the collision of the western Maya Block with R. D. Nance (Ohio University) for their helpful
eastern Oaxaquia in the Late Permian and produced and comprehensive reviews. We wish to thank
stacking, deformation, and metamorphism of the J. Morales-Contreras, G. Solís-Pichardo, and M. S.
Paleozoic sedimentary rocks in the CMC (fig. 9d). Hernández-Bernal of UNAM for assistance with
The high-grade metamorphic event was followed in Sm-Nd isotope data acquisition. Many thanks
the Late Permian to Early Triassic by the intrusion for their logistical support during field work in
of voluminous calc-alkaline plutons (Damon et al. the Sepultura area go to J. C. Pisaño-Soto and
1981; Schaaf et al. 2002) that constitute most of the P. Hernández-Martínez (Comisión Nacional de
CMC as a result of eastward subduction (fig. 9d) and Áreas Naturales Protegidas, Tuxtla Gutiérrez) and
that obscure the prebatholithic units for most parts to A. Pohlenz and M. Pohlenz for their hospitality
of the CMC. In conclusion, strong parallels between and their logistical support at the Finca Custepec.
the geologic history and provenance data of the The Arizona LaserChron Center is partially sup-
Maya Block and the reported data from the Acatlán ported by a grant from the Instrumentation and
Complex suggest a correlation between the two Facilities Program, Division of Earth Sciences,
Paleozoic crustal blocks that requires both more NSF-EAR 0443387.

REFERENCES CITED

Alvarez, D. J.; Alvarez, W.; Asaro, F.; and Michel, H. V. Clemons, R. E., and Burkart, B. 1971. Stratigraphy of
1980. Extraterrestrial cause for the Cretaceous- northwestern Guatemala. Bol. Soc. Geol. Mex. 32:
Tertiary extinction. Science 208:1095–1108. 143–158.
Bateson, J. H., and Hall, I. H. S. 1977. The geology of the Cordani, U. G.; Sato, K.; Teixeira, W.; Tassinari, C. C. G.;
Maya Mountains, Belize. Inst. Geol. Sci. Overseas and Basei, M. A. S. 2000. Crustal evolution of the
Mem. 3. London, Her Majesty’s Stationery Office, 44 p. South American Platform. In Cordani, U. G.; Milani,
Blum, J. D.; Chamberlain, C. P.; Hingston, M. P.; Koeberl, C.; E. J.; Thomas-Filho, A.; and Campos, D. A., eds. Tec-
Marin, L. E.; Schuraytz, B. C.; and Sharpton, V. L. 1993. tonic evolution of South America, Int. Geol. Cong.,
Isotopic comparison of K/T boundary impact glass 31st (Rio de Janeiro, 2000), Proc. p. 19–40.
with melt rock from the Chicxulub and Manson im- Damon, P. E., and Montesinos, E. 1978. Late Cenozoic
pact structures. Nature 364:325–327. volcanism and metallogenesis over an active Benioff
Bohnenberger, O. H. 1966. Nomenclatura de las capas zone in Chiapas, Mexico. Ariz. Geol. Soc. Dig. 11:
Santa Rosa en Guatemala. Publ. Geol. Inst. Centroam. 155–168.
Investig. Tecnol. Ind. 1:47–51. Damon, P. E.; Shafiqullah, M.; and Clark, K. 1981. Age
Burkart, B.; Deaton, B. C.; Dengo, C.; and Moreno, G. trends of igneous activity in relation to metallogenesis
1987. Tectonic wedges and offset Laramide structures in the southern Cordillera. Ariz. Geol. Soc. Dig. 14:
along the Polochic Fault of Guatemala and Chiapas, 137–153.
Mexico: reaffirmation of large Neogene displace- Dengo, G. 1985. Mid America: tectonic setting for the
ments. Tectonics 6:411–422. Pacific margin from southern Mexico to northwestern
Burkart, B., and Self, S. 1985. Extension and rotation of Colombia. In Nairn, A. E. M., and Stehli, F. G., eds.
crustal blocks in northern Central America and effect The Pacific Ocean (Vol. 7a of The oceanic basins and
on the volcanic arc. Geology 13:22–26. margins). New York, Plenum, p. 123–180.
Journal of Geology C H I A PA S MA S S I F AG E S 637

Dickinson, W. R., and Lawton, T. F. 2001. Carboniferous eds. Variscan-Appalachian dynamics: the building of
to Cretaceous assembly and fragmentation of México. the late Paleozoic basement. Geol. Soc. Am. Spec. Pap.
Geol. Soc. Am. Bull. 113:1142–1160. 364:199–208.
Dixon, C. G. 1956. Geology of southern British Honduras Hernández-García, R. 1973. Paleogeografía del Paleozoico
with notes on adjacent areas. Belize, Government de Chiapas, México. Bol. Asoc. Mex. Geol. Pet. 25:
Printer, 92 p. 79–113.
Elías-Herrera, M., and Ortega-Gutiérrez, F. 2002. Cal- Hiller, R.; Weber, B.; Hecht, L.; Ortega-Gutiérrez, F.;
tepec Fault zone: an Early Permian dextral transpres- Schaaf, P.; and López-Martínez, M. 2004. The Sepul-
sional boundary between the Proterozoic Oaxacan and tura Unit: a medium to high grade metasedimentary
Paleozoic Acatlán complexes, southern Mexico, and sequence in the Chiapas Massif, SE México. Reunión
regional tectonic implications. Tectonics 21:1–19. Nacional de Ciencias de la Tierra, Querétaro, México,
Estrada-Carmona, J.; Weber, B.; Hecht, L.; and Martens, Libro de Resumenes. México, Sociedad Geológica
U. 2007. Petrogenesis and P-T conditions of metamor- Mexicana, p. 200.
phic rocks from the Chiapas Massif Complex in Jiménez-Hernández, A.; Jaimez-Fuentes, A.; Motolinía-
the Custepec area, Chiapas, Mexico. Abstract García, O.; Pinzón-Salazar, T.; and Membrillo-Ortega,
U51A-04. EOS: Trans. Am. Geophys. Union 88 (Jt. H. 2005. Carta Geológica: Minera Huixtla D15-2
Assembly Suppl.). Chiapas. Pachuca, Servicio Geológico Mexicano, scale
Fedo, C. M.; Sircombe, K. N.; and Rainbird, R. H. 2003. 1∶250,000.
Detrital zircon analysis of the sedimentary record. Keppie, J. D. 2004. Mexican terranes revisited: a 1.3
Rev. Mineral. Geochem. 53:277–303. billion year odyssey. Int. Geol. Rev. 46:765–794.
French, C. D., and Schenk, C. J. 1997. Map showing Keppie, J. D.; Dostal, J.; Murphy, J. B.; and Nance, R. D.
geology, oil and gas fields, and geologic provinces of 2008a. Synthesis and tectonic interpretation of the
the Caribbean Region. U.S. Geol. Surv. Open-File Rep. westernmost Paleozoic Variscan orogen in southern
97-470-K. México: from rifted Rheic margin to active Pacific
Gehrels, G.; Valencia, V.; and Pullen, A. 2006. Detrital margin. Tectonophysics, forthcoming, doi:10.1016/
zircon geochronology by laser ablation multi- j.tecto.2008.01.012.
collector ICPMS at the Arizona LaserChron Center. Keppie, J. D.; Dostal, J.; Ramos-Arias, M. A.; Morales-
In Olszewski, T., ed. Geochronology: emerging oppor- Gámez, M.; Miller, R. D., Nance, R. D.; Murphy,
tunities. Paleontol. Soc. Pap. 12:67–76. J. B.; Ortega-Rivera, A.; and Lee, J. W. K. 2008b.
Gehrels, G.; Valencia, V.; and Ruiz, J. 2008. Enhanced Ordovician–earliest Silurian rift tholeiites in the Acat-
precision, accuracy, efficiency, and spatial resolution lán Complex, southern Mexico: evidence of rifting on
of U-Pb ages by laser ablation–multicollector– the southern margin of the Rheic Ocean. Tectonophys-
inductively coupled plasma–mass spectrometry. ics, forthcoming, doi:10.1016/j.tecto.2008.01.010.
Geochem. Geophys. Geosyst. 9:Q03017, doi: Keppie, J. D., and Morán-Zenteno, D. J. 2005. Tectonic
10.1029/2007GC001805. implications of alternative Cenozoic reconstructions
Geraldes, M. C.; Van Schmus, W. R.; Condie, K. C.; for southern México and the Chortis Block. Int. Geol.
Bell, S.; Teixeira, W.; and Babinski, M. 2001. Protero- Rev. 47:473–491.
zoic geologic evolution of the SW part of the Amazo- Keppie, J. D.; Nance, R. D.; Fernández-Suárez, J., Storey,
nian Craton in Mato Grosso state, Brazil. Precambrian C. D.; Jeffries, T. E.; and Murphy, J. B. 2006. Detrital
Res. 111:91–128. zircon data from the eastern Mixteca Terrane, southern
Gillis, R. J.; Gehrels, G. E.; Ruiz, J.; and Flores de Dios Mexico: evidence for an Ordovician–Mississippian
González, L. A. 2005. Detrital zircons provenance of continental rise and a Permo-Triassic clastic wedge
Cambrian-Ordovician and Carboniferous strata of the adjacent to Oaxaquia. Int. Geol. Rev. 48:97–111.
Oaxaca terrane, southern Mexico. Sediment. Geol. Keppie, J. D.; and Ortega-Gutiérrez, F. 1999. Middle
182:87–100. American Precambrian basement: a missing part of
Grodzicki, K. R.; Nance, R. D.; Keppie, J. D.; Dostal, J.; the reconstructed 1 Ga orogen. In Ramos, V. S., and
and Murphy, J. B. 2008. Structural, geochemical and Keppie, J. D., eds. Laurentia-Gondwana connections
geochronological analysis of metasedimentary and before Pangea. Geol. Soc. Am. Spec. Pap. 336:199–210.
metavolcanic rocks of the Coatlaco area, Acatlán Keppie, J. D., and Ramos, V. A. 1999. Odyssey of terranes
Complex, southern Mexico. Tectonophysics, forth- in the Iapetus and Rheic oceans during the Paleozoic.
coming, doi:10.1016/j.tecto.2008.01.016. In Ramos, V. A., and Keppie, J. D., eds. Laurentia-
Guzmán-Speziale, M. 2000. The North America- Gondwana connections before Pangea. Geol. Soc.
Caribbean plate boundary west of the Motagua- Am. Spec. Pap. 336:267–276.
Polochic fault system: a fault jog in southeastern Krogh, T. E.; Kamo, S. L.; Sharpton, V. L.; Martin, L. E.;
México. J. S. Am. Earth Sci. 13:459–468. and Hildebrand, A. R. 1993. U-Pb ages of single
Hatcher, R. D. 2002. Alleghanian (Appalachian) orogeny, shocked zircons linking distal K/T ejecta to the
a product of zipper tectonics: rotational transpressive Chicxulub crater. Nature 366:731–734.
continent-continent collision and closing of ancient López-Ramos, E. 1979. Geología de México. Tomo III.
oceans along irregular margins. In Martínez-Catalán, México, s. n., 446 p.
J. R.; Hatcher, R. D.; Arenas, R.; and Díaz-García, F., Ludwig, K. R. 2003. User’s manual for Isoplot 3.00: a
638 B. WEBER ET AL.

geochronological toolkit for Microsoft Excel. Berkeley 2004. Tectonostratigraphic analysis at the southern
Geochronol. Cent. Spec. Publ. 4. Berkeley, CA, margin of the Maya Block: where is the limit? Reunión
Berkeley Geochronology Center, 70 p. Nacional de Ciencias de la Tierra, Querétaro, México,
Martens, U.; Weber, B.; and Valencia, V. 2006. Zircon Libro de Resumenes. México, Sociedad Geológica
geochronology implies the existence of pre-Devonian Mexicana, p. 204.
sedimentary rocks with Grenvillian provenance in the Ortega-Gutiérrez, F.; Solari, L. A.; Ortega-Obregón, C.;
Maya Mountains of Belize. Geol. Soc. Am. Abstr. Elías-Herrera, M.; Martens, U.; Morán-Icál, S.;
Program 38:504. Chiquín, M.; Keppie, J. D.; Torres de León, R.; and
Martínez-Amador, H.; Rosendo-Brito, B.; Fitz-Bravo, C.; Schaaf, P. 2007. The Maya-Chortis boundary: a tec-
Tinajera-Fuentes, E.; and Beltrán-Castillo, H. D. 2005. tonostratigraphic approach. Int. Geol. Rev. 49:996–
Carta Geológica: Tuxtla Gutiérrez, Chiapas E5-11. 1024.
Pachuca, Servicio Geológico Mexicano, scale Ortega-Obregón, C. 2005. Caracterización estructural
1∶250,000. petrológica y geoquímica de la zona de cizalla “Baja
Morales-Gámez, M.; Keppie, J. D.; and Norman, M. 2008. Verapaz,” Guatemala. MS thesis, UNAM, 99 p.
Ordovician-Silurian rift-passive margin on the Mexi- Ortega-Obregón, C.; Solari, L. A.; Keppie, J. D.;
can margin of the Rheic Ocean overlain by Ortega-Gutiérrez, F.; Solé, J.; and Morán-Icál, S.
Carboniferous-Permian periarc rocks: evidence from 2008. Middle-Late Ordovician magmatism and Late
the eastern Acatlán Complex, southern Mexico. Cretaceous collision in the southern Maya Block,
Tectonophysics, forthcoming, doi:10.1016/j.tecto Rabinal-Salamá area, central Guatemala: implications
.2008.01.014. for North America–Caribbean plate boundary. Geol.
Morán-Zenteno, D. 1984. Geología de la República Soc. Am. Bull. 120:556–570.
Mexicana. México, Universidad Nacional Autónoma Pantoja-Alor, J., and Robinson, R. 1967. Paleozoic sedi-
de México (UNAM), Instituto Nacional de Estadística, mentary rocks in Oaxaca, Mexico. Science 157:
Geografía e Informática, 88 p. 1033–1035.
Muehlberger, W. R., and Ritchie, A. W. 1975. Caribbean- Pompa-Mera, V.; Solís-Pichardo, G.; Schaaf, P.; Weber, B.;
Americas plate boundary in Guatemala and southern and Hernández-Treviño, T. 2007. Geoquímica y geo-
México as seen on Skylab IV orbital photography. cronología de algunas rocas plutónicas del sector
Geology 3:232–235. oriental del Macizo de Chiapas, México. Geos 27:88.
Murillo-Muñeton, G. 1994. Petrologic and geochrono- Ruiz, J.; Tosdal, R. M.; Restrepo, P. A.; and Murillo-
logic study of Grenville-age granulites and post- Muñeton, G. 1999. Pb isotope evidence for Colombia–
granulite plutons from the La Mixtequita area, state southern México connections in the Proterozoic. In
of Oaxaca in southern Mexico, and their tectonic Ramos, V. A., and Keppie, J. D., eds. Laurentia-
significance. MS thesis, University of Southern Gondwana connections before Pangea. Geol. Soc.
California, Los Angeles, 163 p. Am. Spec. Pap. 336:183–198.
Murphy, J. B.; Gutiérrez-Alonso, G.; Nance, D. R.; Salazar-Juárez, J.; Schaaf, P.; Solís-Pichardo, G.; Ortega-
Fernández-Suárez, J.; Keppie, J. D.; Quesada, C.; Gutiérrez, F.; Elías-Herrera, M.; and Weber, B. 2007.
Strachan, R. A.; and Dostal, J. 2006. Origin of the Geología, petrología y geoquímica del Macizo de Chia-
Rheic Ocean: rifting along a Neoproterozoic suture? pas, área de Motozintla de Mendoza. Geos 27:87–88.
Geology 34:325–328. Sanchez-Zavala, J. L.; Ortega-Gutiérrez, F.; Keppie, J. D.;
Nance, R. D.; Miller, B. V.; Keppie, J. D.; Murphy, J. B.; and Jenner, G. A.; Belousova, E.; and Macías-Romo, C.
Dostal, J. 2006. Acatlán Complex, southern Mexico: 2004. Ordovician and Mesoproterozoic zircons from
record spanning the assembly and breakup of Pangea. the Tecomate Formation and Esperanza granitoids,
Geology 34:857–860. Acatlán Complex, southern Mexico: local provenance
———. 2007. Vestige of the Rheic Ocean in North in the Acatlán and Oaxacan complexes. Int. Geol. Rev.
America: the Acatlán Complex of southern México. 46:1005–1021.
In Linnemann, U.; Nance, R. D.; Kraft, P.; and Zulauf, Schaaf, P.; Weber, B.; Weis, P.; Groß, A.; Ortega-Gutiérrez,
G., eds. The evolution of the Rheic Ocean: from F.; and Köhler, H. 2002. The Chiapas Massif (México)
Avalonian-Cadomian active margin to Alleghenian- revised: new geologic and isotopic data for basement
Variscan collision. Geol. Soc. Am. Spec. Pap. 423: characteristics. Neues Jahrb. Geol. Palaeont. Abh. 225:
437–452. 1–23.
Ortega-Gutiérrez, F.; Mitre-Salazar, L. M.; Roldan- Sedlock, R. L.; Ortega-Gutiérrez, F.; and Speed, R. C. 1993.
Quintana, J.; Aranda-Gómez, J. J.; Morán-Zenteno, D.; Tectonostratigraphic terranes and tectonic evolution
Alaniz-Álvarez, S. A.; and Nieto-Samaniego, A. N. 1992. of Mexico. Geol. Soc. Am. Spec. Pap. 278, 153 p.
Carta geológica de la República Mexicana. México, Solari, L. A.; Dostal, J.; Ortega-Gutiérrez, F.; and Keppie,
UNAM, Instituto de Geología, scale 1∶2,000,000. J. D. 2001. The 275 Ma arc-related La Carbonara stock
Ortega-Gutiérrez, F.; Ruiz, J.; and Centeno-García, E. in the northern Oaxacan Complex of southern
1995. Oaxaquia, a Proterozoic microcontinent ac- Mexico: U-Pb geochronology and geochemistry. Rev.
creted to North America during the late Paleozoic. Mex. Cienc. Geol. 18:149–161.
Geology 23:1127–1130. Solari, L. A.; Ortega-Gutiérrez, F.; Elías-Herrera, M.;
Ortega-Gutiérrez, F.; Solari, L.; and Ortega-Obregón, C. Schaaf, P.; Norman, M.; Torres de León, R.; Ortega-
Journal of Geology C H I A PA S MA S S I F AG E S 639

Obregón, C.; Chiquín, M.; and Morán-Ical, S. Forth- from the Chiapas Massif, southeastern Mexico. Int.
coming. U-Pb zircon geochronology of Paleozoic units Geol. Rev. 47:509–529.
in western and central Guatemala: insights into the Weber, B.; Gruner, B.; Hecht, L.; Molina-Garza, R. S.; and
tectonic evolution of Middle America. In Pindell, J., Köhler, H. 2002. El descubrimiento de basamento
and James, K. H., eds. Origin of the Caribbean Plate. metasedimentario en el macizo de Chiapas: la
J. Geol. Soc. Lond. Spec. Publ. “Unidad La Sepultura.” Geos 22:2–11.
Stacey, J. S. K., and Kramers, J. D. 1975. Approximation of Weber, B.; and Hecht, L. 2003. Petrology and geochemistry
terrestrial lead isotope evolution by a two-stage model. of metaigneous rocks from a Grenvillian basement
Earth Planet. Sci. Lett. 26:207–221. fragment in the Maya Block: the Guichicovi
Steiner, M. B., and Walker, J. D. 1996. Late Silurian Complex, Oaxaca, southern Mexico. Precambrian
plutons in Yucatan. J. Geophys. Res. 101:17,727– Res. 124:41–67.
17,735. Weber, B.; Iriondo, A.; Premo, W. R.; Hecht, L.; and Schaaf,
Stewart, J. H.; Blodgett, R. B.; Boucot, A. J.; Carter, J. L.; P. 2007. New insights into the history and origin of
and Lopez, R. 1999. Exotic Paleozoic strata of the southern Maya Block, SE México: U-Pb-SHRIMP
Gondwanan provenance near Ciudad Victoria, zircon geochronology from metamorphic rocks of the
Tamaulipas, Mexico. In Ramos, V. A., and Keppie, J. D., Chiapas Massif. Int. J. Earth Sci. 96:253–269.
eds. Laurentia-Gondwana connections before Pangea. Weber, B., and Köhler, H. 1999. Sm-Nd, Rb-Sr and U-Pb
Geol. Soc. Am. Spec. Pap. 336:227–252. isotope geochronology of a Grenville terrane in
Talavera-Mendoza, O.; Ruiz, J.; Gehrels, G. E.; southern Mexico: origin and geologic history of the
Meza-Figueroa, D. M.; Vega-Granillo, R.; and Campa- Guichicovi Complex. Precambrian Res. 96:245–262.
Uranga, M. F. 2005. U-Pb geochronology of the Acatlán Weber, B.; Schaaf, P.; Valencia, V.; Iriondo, A.; and Ortega-
Complex and implications for the Paleozoic paleoge- Gutiérrez, F. 2006. Provenance ages of late Paleozoic
ography and tectonic evolution of southern Mexico. sandstones (Santa Rosa Formation) from the Maya
Earth Planet. Sci. Lett. 235:682–699. Block, SE México: implications on the tectonic
Vega-Granillo, R.; Talavera-Mendoza, O.; Meza-Figueroa, evolution of western Pangea. Rev. Mex. Cienc. Geol.
D.; Ruiz, J.; Gehrels, G. E.; López-Martínez, M.; and de 23:262–276.
la Cruz-Vargas, J. C. 2007. Pressure-temperature-time Weber, B.; Valencia, V. A.; Schaaf, P.; and Ortega-
evolution of Paleozoic high-pressure rocks of the Gutiérrez, F. Forthcoming. Detrital zircon ages from
Acatlán Complex (southern Mexico): implications for the lower Santa Rosa Formation, Chiapas: implica-
the evolution of the Iapetus and Rheic oceans. Geol. tions on regional Paleozoic stratigraphy. Rev. Mex.
Soc. Am. Bull. 119:1249–1264. Cienc. Geol.
Weber, B. 1998. Die magmatische und metamorphe Whitaker, A.; Cope, J. C. W.; Cowie, J. W.; Gibbons, W.;
Entwicklung eines kontinentalen Krustensegments: Hailwood, E. A.; House, M. R.; Jenkins, D. G.; et al.
isotopengeochemische und geochronologische Unter- 1991. A guide to stratigraphical procedure. J. Geol. Soc.
suchungen am Mixtequita-Komplex, Südostmexiko. Lond. 148:813–824.
Muench. Geol. Hefte A24:1–176. Yañez, P.; Ruiz, J.; Patchett, P. J.; Ortega-Gutiérrez, F.; and
Weber, B.; Cameron, K. L.; Osorio, M.; and Schaaf, P. 2005. Gehrels, G. E. 1991. Isotopic studies of the Acatlán
A late Permian tectonothermal event in Grenville complex, southern Mexico: implications for North
crust of the southern Maya Terrane: U-Pb zircon ages American tectonics. Geol. Soc. Am. Bull. 103:817–827.

You might also like