You are on page 1of 23

Lithos 300–301 (2018) 177–199

Contents lists available at ScienceDirect

Lithos

journal homepage: www.elsevier.com/locate/lithos

Late Mesoproterozoic to Early Paleozoic history of metamorphic


basement from the southeastern Chiapas Massif Complex, Mexico,
and implications for the evolution of NW Gondwana
Bodo Weber a,⁎, Reneé González-Guzmán a, Román Manjarrez-Juárez a, Alejandro Cisneros de León b,c,
Uwe Martens d, Luigi Solari d, Lutz Hecht e, Victor Valencia f
a
Departamento de Geología, Centro de Investigación Científica y de Educación Superior de Ensenada (CICESE), Carr. Ensenada-Tijuana 3918, 22860 Ensenada, B.C., Mexico
b
Instituto de Geología, Universidad Nacional Autónoma de México (UNAM), Ciudad Universitaria, 04510 Coyoacán, DF, Mexico
c
Institut für Geowissenschaften, Ruprecht-Karls Universität Heidelberg, Im Neuenheimer Feld 236, 69120 Heidelberg, Germany
d
Centro de Geociencias, Campus UNAM 3001, 76230 Juriquilla, Querétaro, Mexico
e
Museum für Naturkunde, Leibniz Institut für Evolutions- und Biodiversitätsforschung, Invalidenstrasse 43, 10115 Berlin, Germany
f
Department of the Environment, Washington State University, Pullman, WA 99164, USA

a r t i c l e i n f o a b s t r a c t

Article history: In this paper, U-Pb zircon geochronology, Lu-Hf and Sm-Nd isotope systematics, geochemistry and
Received 4 September 2017 geothermobarometry of metaigneous basement rocks exposed in the southeastern Chiapas Massif Complex are pre-
Accepted 7 December 2017 sented. Geologic mapping of the newly defined “El Triunfo Complex” located at the southeastern edge of the Chiapas
Available online 13 December 2017
Massif reveals (1) partial melting of a metamorphic basement mainly constituted by mafic metaigneous rocks
(Candelaria unit), (2) an Ediacaran metasedimentary sequence (Jocote unit), and (3) occurrence of massif-type an-
Keywords:
Geochronology
orthosite. All these units are intruded by undeformed Ordovician plutonic rocks of the Motozintla suite. Pressure and
Isotope systematics temperature estimates using Ca-amphiboles, plagioclase and phengite revealed prograde metamorphism that
Geothermobarometry reached peak conditions at ~650 °C and ~6 kbar, sufficient for partial melting under water saturated conditions. Rel-
Rodinia breakup ict rutile in titanite and clinopyroxene in amphibolite further indicate a previous metamorphic event at higher P-T
Famatinian orogeny conditions. U-Pb zircon ages from felsic orthogneiss boudins hosted in deformed amphibolite and migmatite yield
Chiapas Massif Complex crystallization ages of ~1.0 Ga, indicating that dry granitic protoliths represent remnants of Rodinia-type basement.
Additionally, a mid-Tonian (~920 Ma) metamorphic overprint is suggested by recrystallized zircon from a banded
gneiss. Zircon from folded amphibolite samples yield mainly Ordovician ages ranging from ~457 to ~444 Ma that are
indistinguishable from the age of the undeformed Motozintla plutonic suite. Similar ages between igneous- and
metamorphic- zircon suggest a coeval formation during a high-grade metamorphic event, in which textural discrep-
ancies are explained in terms of differing zircon formation mechanisms such as sub-solidus recrystallization and
precipitation from anatectic melts. In addition, some amphibolite samples contain inherited zircon yielding
Stenian-Tonian ages around 1.0 Ga. Lu-Hf and Sm-Nd isotopes and geochemical data indicate that the protoliths
of the amphibolite have E-MORB characteristics and were derived from a depleted mantle source younger than
the Rodinia-type basement. Inasmuch as similar amphibolites also occur in the Ediacaran metasedimentary rocks
as dykes or lenses, Late Neoproterozoic magmatism in a rift setting is suggested. Hence, the geologic record of the
El Triunfo Complex includes evidences for Rodinia assemblage, Tonian circum-Rodinia subduction, and breakup
during the Late Neoproterozoic. Metamorphism, and partial melting are interpreted in terms of a convergent margin
setting during the Ordovician. The results place the southern Chiapas Massif along with Oaxaquia and similar North-
ern Andes terranes on the NW margin of Gondwana interpreted as the extension of the Famatinian orogen that
evolved during the closure of the Iapetus Ocean.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction 1983). Whereas some Mexican terranes are the product of Mesozoic
Cordilleran tectonics, there is a belt spanning the basement of Tamauli-
The geology of Mexico and Central America is characterized by a pas, southeastern Mexico, and Central America north of the Nicaragua-
series of distinct tectonostratigraphic terranes (e.g., Campa and Coney, Honduras border that is composed of pre-Mesozoic terranes whose
Gondwanan affinity is well established by paleontologic,
⁎ Corresponding author. geochronologic, and provenance constraints (e.g., Centeno-García and
E-mail address: bweber@cicese.mx (B. Weber). Keppie, 1999; Keppie, 2004; Keppie and Ortega-Gutierrez, 1999; Keppie

https://doi.org/10.1016/j.lithos.2017.12.009
0024-4937/© 2017 Elsevier B.V. All rights reserved.
178 B. Weber et al. / Lithos 300–301 (2018) 177–199

and Ramos, 1999; Landing et al., 2007). In the Late Neoproterozoic In the southern part of the Chiapas Massif, the Permian plutons are
and Early Paleozoic, these terranes occupied positions proximal to minor relative to Ordovician igneous rocks (Estrada-Carmona et al.,
Amazonia or West Africa, along with other peri-Gondwanan terranes 2012; Weber et al., 2008) that intruded metasedimentary and
now located in eastern North America and the Variscan belt of Europe metaigneous host rocks henceforth referred to as the “El Triunfo
(e.g., Murphy et al., 2004; Nance et al., 2008). Complex”. A correlative rock association in Central Guatemala was
The backbone of central, eastern, and southern Mexico is the possibly displaced from southern Chiapas by the Cenozoic, left lateral
Oaxaquia microcontinent (Ortega-Gutierrez et al., 1995), characterized Polochic Fault (Fig. 1; Ortega-Obregón et al., 2008; Solari et al., 2009,
by ~ 0.99 Ga granulites with Late Mesoproterozoic (~ 1.25–1.0 Ga) 2010a, 2013). Similar pluton-host rock relations are also present in
protoliths, including ca. 1.01 Ga massif-type anorthosite and the Maya Mountains of Belize, where Silurian to earliest Devonian gran-
anorthosite-mangerite-charnockite (AMC) series rocks (Keppie et al., ites intrude metasedimentary units (Martens et al., 2010; Steiner and
2003; Solari et al., 2004; Keppie and Ortega-Gutiérrez, 2010; Weber Walker, 1996; Weber et al., 2012).
et al., 2010). In the Phanerozoic, crustal reworking substantially modi- Host rocks such as those described above have been the focus of recent
fied vast areas of Oaxaquia, obscuring any evidence of its Precambrian geochronologic and geochemical work because they offer a unique
geologic evolution. This feature is well exemplified by the Maya block glimpse into the pre-Early Paleozoic geologic evolution of the
(Dengo, 1969; Fig. 1), where Precambrian granulites are only exposed Maya block and allow assessing its relations to Amazonia, Oaxaquia and
along the terrane's westernmost margin (Guichicovi Complex, Fig. 1, other peri-Gondwanan terranes. For instance, detrital zircon geochronol-
Ruiz et al., 1999; Weber and Hecht, 2003; Weber and Köhler, 1999). In ogy and isotopic signatures of metasedimentary units in Belize allowed
contrast, the largest basement exposure of the Maya block, the Chiapas constraining the time of deposition as latest Neoproterozoic or Cambrian,
Massif (Fig. 1), is dominated by Late Permian batholithic plutons and and imply provenance from Late Mesoproterozoic to early Neoproterozoic
local Paleozoic metamorphic host rocks (Damon et al., 1981; Schaaf (1.2–0.9 Ga) and minor Early Mesoproterozoic (1.6–1.4 Ga) sources
et al., 2002). So far, evidence for reworked Stenian-Tonian basement (Martens et al., 2010; Solari et al., 2009; Weber et al., 2008, 2012). Fur-
in the Chiapas Massif has been indirect, such as the abundant ~1.0 Ga thermore, a depositional age of ca. 600–580 Ma was obtained by
zircon cores in the Permian granites and gneisses (Weber et al., 2005) chemostratigraphy for metasedimentary host rocks of the El Triunfo Com-
and amphibolites yielding depleted mantle Nd model ages of ~1.4 Ga plex, suggesting an origin in a basin created by Rodinia breakup and the
(Schaaf et al., 2002). opening of the Iapetus Ocean (González-Guzmán et al., 2016a).

(a)
(b)

Fig. 1. (a) Plate tectonic overview of Central America, showing Oaxaquia, Maya, and Chortis blocks as well as the Acatlán Complex and major Proterozoic basement exposures of the Northern
Andes. (b) Geologic map showing pre-Mesozoic rocks exposed in southern Mexico and Central America (modified after Martens et al., 2012; Ortega-Gutierrez et al., 1992, 2007; French
and Schenk, 1997). Abbreviations: ACu = Altos Cuchumatanes, BVFZ = Baja Verapaz Fault Zone, CMC = Chiapas Massif Complex, CUI = Cuicateco Terrane, CHA = Chatino Terrane,
JCFZ = Jocotlán-Chamaleón Fault Zone, M = Mixteca Terrane, MFZ = Motagua Fault Zone, MM = Maya Mountains, PFZ = Polochic Fault zone, Z = Zapoteco Terrane.
B. Weber et al. / Lithos 300–301 (2018) 177–199 179

In contrast, the geologic evolution of the metaigneous host rocks of 1975) turning into the NW-SE trending Tonalá shear zone that runs
the El Triunfo Complex is mostly unconstrained. González-Guzmán parallel to the Pacific coast along which syntectonic granitoids intruded
et al. (2016a) hypothesized that these metaigneous rocks may have in the middle Miocene (Fig. 2, Molina-Garza et al., 2015). These left
represented the crystalline basement flooring the basin where lateral fault systems reflect the relative eastward movement of the
Neoproterozoic sedimentary protoliths of the El Triunfo Complex were Chortís block of Central America (e.g., Ratschbacher et al., 2009). Recent
deposited. The evaluation of this idea is critical because it suggests pre- seismo-tectonic studies at the Polochic-fault revealed that present-day
viously unknown pre-Late Neoproterozoic components in the southern left lateral movement is about 4.8 mm/yr (Authemayou et al., 2011).
Chiapas Massif. Hence, in this study, which presents U-Pb zircon geo- By that rate and by assuming a constant velocity, a time span of
chronology, geochemistry, and isotopic (Lu-Hf, Sm-Nd) characteriza- ~ 50 Ma is necessary to accommodate ~ 135 km offset as proposed by
tion of the metaigneous rocks of the El Triunfo Complex, we show that Burkart (1983). This calculation implies that since the intrusion of
the southern Chiapas Massif indeed contains relict Stenian-Tonian, plutons during the middle Miocene, left lateral movement along the
Rodinian-type basement that was intruded by mafic dike swarms in Polochic Fault probably did not exceed ~25 km.
the Neoproterozoic. Furthermore we constrain P, T conditions of the The Guatemala Suture Complex contains high-pressure (HP) meta-
intense Ordovician metamorphism and anatexis that reworked the El morphic belts and serpentinite mélange complexes, which are
Triunfo Complex, and present plausible paleotectonic scenarios for the interpreted to be the result of the Late Cretaceous collision of the
Maya block in relation to Rodinia and Gondwana. North American Plate with Greater Antilles Arc (Martens et al., 2012;
Pindell and Kennan, 2009) predating the Tertiary lateral movements.
2. Geologic setting The Chuacús complex, between the Baja Verapáz shear zone and the
Motagua fault system (Fig. 1) that contains also HP rocks, is a piece of
2.1. The Maya block continental crust from the southern Maya block that was first subducted
beneath the Greater Antilles Arc and then back-thrusted over the
The Maya block (Dengo, 1969) is the southeasternmost of the southern Maya block (Martens et al., 2012; Ortega-Gutierrez et al.,
Mexican terranes and originally includes the Yucatan peninsula and 2007; Ortega-Obregón et al., 2008). North of the Baja Verapáz shear
southeastern Mexico from the Tehuantepec isthmus to Guatemala zone the Paleozoic basement of the Maya block is also affected by the
(Fig. 1). Pre-Mesozoic rocks are exposed only in the southern part of Late Cretaceous HP metamorphic event to certain extent (Solari et al.,
the Maya block, namely in the La Mixtequita (Guichicovi Complex), 2013).
the Chiapas Massif, Central Guatemala, and the Maya Mountains of Ordovician to Early Devonian igneous rocks occur along an ENE-
Belize (Fig. 1). WSW trending belt along the southern border of the Maya block
The southern boundary of the Maya block is defined by the roughly (Fig. 1, Estrada-Carmona et al., 2012). In the Maya Mountains of Belize
E-W trending Polochic-, Motagua, and Jocotlán-Chamaleón fault Late Silurian to Early Devonian (~420–400 Ma) granites and granodio-
systems, which separate the North American from the Caribbean plate rites intrude sedimentary rocks of the Baldy unit, covered by a contem-
(Chortís block; Fig. 1). These Fault zones are complex composite porary conglomerate with rhyolitic tuff of the Bladen Formation
structural boundaries on which left lateral displacement is accommo- (Martens et al., 2010; Steiner and Walker, 1996; Weber et al., 2012).
dated along several subordinate faults (Authemayou et al., 2011; In central Guatemala north to the Baja Verapáz shear zone the Early Or-
Guzmán-Speziale and Meneses-Rocha, 2000; Ratschbacher et al., dovician (471+3 / −5 Ma, Solari et al., 2013) Rabinal granite intruded
2009). To the west, in Chiapas, these individual faults splay into diffuse metasedimentary rocks of the San Gabriel sequence and the Barillas
structural lineaments (Burkart and Self, 1985; Muehlberger and Ritchie, complex (Ortega-Gutierrez et al., 2007; Ortega-Obregón et al., 2008).

Cenozoic cover
16.5°N

Cretaceous Limestone
(Sierra Madre Fm.)
0 20 40Km Jurassic-Triassic red beds
(Todos Santos Fm.)
La Permian Lime- and Silstone
An (Paso Hondo Fm.)
go Carboniferous
Arriaga Villa Flores stu
ra (Santa Rosa Fm.)
Da Paleozoic (meta)sedimentary
m
16.0°

Tonala rocks (not differentiated)


CHIAPAS MASSIF (CRYSTALLINE) COMPLEX

Miocene igneous rock


Permo-Triassic igneous and medium-
to high-grade metamorphic rocks
Igneous and metamorphic
rocks (not differentiated)
Pijijiapan
Pa Sepultura Unit - meta sedimentary,
Carboniferous protoliths
cif
15.5°

ic Custepec Unit - garnet amphibolite


O gneiss and marble
ce Mapastepec Motozintla Metamorphic rocks (not differentiated)
an
El Triunfo
EL TRIUNFO COMPLEX
N
Gu Mé

Late Ordovician metamorphic rocks


including Ordovician granitoids
at xic
em o

area of
Fig. 3 Jocote unit metasedimentary
ala

Ediacarian protolith
15.0°

Candelaria unit and anorthosite suite


Meso- to Neoproterozoic protoliths
-93.5° 93.0° -92.5° W

Fig. 2. Simplified geologic map of southern Chiapas, modified from Servicio Geológico Mexicano geologic maps 1:250,000, Tuxtla Gutiérrez (Martínez-Amador et al., 2005) and Huixtla
(Jiménez-Hernández et al., 2005).
180 B. Weber et al. / Lithos 300–301 (2018) 177–199

North of the Polochic Fault zone another Ordovician granite (461+ 6 southern CMC and from Belize (González-Guzmán et al., 2016a; Weber
/ −3 Ma, Solari et al., 2010a) intruded medium- to high-grade gneiss et al., 2008).
in the Altos Cuchumatanes of Guatemala (Fig. 1). The same area is Unlike the northwestern and central CMC, in the southern CMC the
also intruded by Pennsylvanian granite (317–312 Ma, Solari et al., Permian metamorphic overprint reached greenschist facies conditions
2010a). West of the Mexican-Guatemalan border, in southern only. This interpretation is based on 40Ar-39Ar and Rb-Sr dates of
Chiapas, metasedimentary rocks are intruded by early Ordovician white mica and amphibole yielding Ordovician-Devonian ages
(~ 482–470 Ma) S-type granite dikes and a Late Ordovician (449 ± 7 (Estrada-Carmona et al., 2012; Pompa-Mera, 2009).This implies that
to 446 ± 6 Ma) granitic to granodioritic pluton (Estrada-Carmona the isotope systems of these minerals did not re-equilibrate during the
et al., 2012). Permian event, suggesting peak temperatures below 500 °C. The general
configuration of the basement, however, is further complicated by
Neogene deformation and syntectonic magmatism, which led to local
2.2. The Chiapas Massif Complex (CMC) hydrothermal alteration and mineralization, namely in the Tolimán
Cu-Au mining district.
The CMC, which covers an area of N 20,000 km2 parallel to the Pacific
coast in southeastern Mexico, is the most voluminous Permian batho-
3. The El Triunfo Complex
lithic complex in Mexico. It forms most of the crystalline basement of
the southwestern Maya block (Fig. 1) and is dominated by weakly to
The studied region extends from the Mexican-Guatemalan border
strongly deformed granitoids ranging in composition from granite to
about 40 km westward to the village of El Triunfo (Fig. 3). Due to strong
minor gabbro intruding orthogneiss, anatectite, amphibolite, and
topography ranging from a few hundred meters to over 3000 m above
metasedimentary rocks (Schaaf et al., 2002; Weber et al., 2007). The
sea level and dense tropical vegetation, fieldwork is mostly restricted
main tectonothermal event that led to the formation of the igneous
to riverbeds and valleys that are somehow related to the splaying
Massif started with the intrusion of orthogneiss protoliths in the late
Polochic Fault zone. In the studied area, however, no major displace-
Early Permian (~ 272 Ma) and culminated with medium- to high-
ment is observed on each side of these east-west trending valleys. To-
grade metamorphism, anatexis, and intrusion of the major magmatic
wards the south, the El Triunfo Complex is intruded by a Miocene
masses in the Late Permian (~ 254–250 Ma, Weber et al., 2007).
granodiorite that is deformed and mylonitized along the ESE-WNW
Batholithic rocks that intruded metasedimentary sequences of the
oriented Tonalá shear zone (Fig. 3, Molina-Garza et al., 2015). Towards
Sepultura Unit dominate the northwestern part of the Chiapas Massif
the north, the crystalline basement is covered by Jurassic redbeds of
(Fig. 2). The Sepultura unit, a likely metamorphic equivalent of the
the Todos Santos Formation or it is overthrusted by Cretaceous
Carboniferous Santa Rosa Formation (Fig. 2), is composed of meta-
limestones. Weakly metamorphosed sedimentary rocks (phyllite and
greywacke, quartzite, metapelite, and calcsilicate rocks that reached
calcsilicate rocks) probably correspond to the Carboniferous Santa
peak metamorphic conditions during the Late Permian of 730–780 °C
Rosa Formation. Permian granitoids of the Chiapas Batholith (mainly
and ~5.8 kbar (Hiller et al., 2004; Weber et al., 2007). Further southeast
tonalite, granodiorite, and diorite) form the high mountain ranges to
in the central CMC peak metamorphic conditions during the Late
the NW and they occur also in the SE, south of the Polochic Fault.
Permian reached N 800 °C and 9 kbar in the Custepec unit that is mainly
composed of garnet-amphibolite with minor calcsilicate and metapelite
(Fig. 2, Estrada-Carmona et al., 2009). The U-Pb ages obtained from 3.1. The Candelaria unit
zircon cores of garnet-amphibolite from the Custepec unit suggest
that the precursors are of sedimentary origin, having 1.2–1.0 Ga, and Estrada-Carmona et al. (2012) defined the Candelaria unit as
~ 1.5 Ga detrital sources, similar to metasedimentary rocks from the comprised by banded and folded amphibolite interbedded with calc-

Longitude
552000 558000 564000 570000 576000 582000 588000 594000
1702000

1205-2
R0302 ault
hic F
CAL-03
R0202 Poloc 1202-3
1205-1
R0103 RE1101
Motozintla
R0206 1203-2
R0301
CH-02
Latitude

TR02 CH-03
1696000

El Triunfo
PAN-14
Tona
R0401 la Fau
?
la

lt
Gu exico

9511-9
ma
ate

N
M

Belisario
1690000

0 5Km Dominguez

Alluvion The El Triunfo Complex: Foliation


Miocene igneous rocks
Limestone Ordovician Motozintla plutonic suite Strike-slip fault
Jurassic redbeds (Todos Santos Fm.) Jocote unit (metasedimentary) Thrust fault
Permian granitoid Candelaria unit (metaigneous)
Anorthosite and associated rocks R0206 Sampling site
Carboniferous(?) phyllite (Santa Rosa Fm.)

Fig. 3. Geologic map of the study area from the Mexican-Guatemalan border to El Triunfo. (Note: Background elevation model is based on NASA Shuttle Radar Topography Mission - SRTM).
B. Weber et al. / Lithos 300–301 (2018) 177–199 181

silicate rocks, impure marble, quartzite, and minor biotite gneiss. foliation of the Candelaria unit dips moderately or steeply into SW to
Indeed, most of the Candelaria unit is formed by folded hornblende SE directions, preferentially to the south (Fig. 3). The variations in the
gneiss but there are some important new findings: (1) some outcrops directions of the foliation planes throughout the western study area fur-
show felsic orthogneisses contained or interbedded with the folded ther suggest a second (D2) open folding of the sequence with a calculat-
hornblende and biotite gneiss, often as meter-scale boudins or pinch- ed fold axis (π) plunging to the SE (130°/30°). In the eastern part, SE of
and-swell structures (Fig. 4a, c) (2) Partial melting of the Candelaria Motozintla (Fig. 3) the orthogneiss is strongly deformed with sheath-
unit is observed in most outcrops at different scale leading to leucosome and chevron-folds having interlimb angles b 30°, ptygmatic folds and
veins or layers that are deformed and folded (Fig. 4). Some areas are subvertical foliation planes trending E-W to NE-SW (Fig. 4d). Due to
dominated by migmatite with neosome forming several meter thick its deformation style and field relations with the nearby anorthosite
felsic dikes. (3) A second generation of mafic rocks (hornblende gabbro) this gneiss differs from the typical Candelaria unit gneiss and is hence-
intruded the folded basement and (4) marbles intercalated with am- forth referred to as the “Chipilín” gneiss (Cisneros de León et al.,
phibolites are now defined to be part of the metasedimentary Jocote 2017). However, in the geologic map presented here, the “Chipilín”
unit (González-Guzmán et al., 2016a). gneiss is included into the Candelaria unit, because no mappable limit
Therefore, the Candelaria unit is here redefined as a sequence of defines it as a distinguishable unit.
banded orthogneiss mainly constituted of amphibolite (hornblende
gneiss) and hornblende-biotite gneiss with variable grade of 3.2. The Jocote unit
migmatization, as well as quartz-feldspar orthogneiss. Leucocratic
layers in amphibolite are often folded into tight to isoclinal folds Metasedimentary rocks of the Jocote unit (Estrada-Carmona et al.,
(Fig. 4b), mostly subparallel to the foliation planes. Major leucosome 2012; González-Guzmán et al., 2016a; Weber et al., 2008) include
accumulations in strongly migmatized areas of the Candelaria unit are pelitic and psammitic sillimanite-rich schist, minor quartzite, calc-
parallel to the penetrative foliation (Fig. 4c) or the leucosome forms silicate rocks and marbles, locally intercalated with metabasite. Al-
crosscutting dikes. In the western part of the study area, penetrative though the Jocote unit seems to be structurally above the Candelaria

(a) leucosome (b)

amphibolite
orthogneiss
boudin

(c)

leucosome

orthogneiss
boudins (e)

10 cm

(d)

Fig. 4. Field photographs from outcrops of the Candelaria unit: (a, c) felsic orthogneiss boudins in mafic hornblende-biotite migmatic gneiss (outcrop 1205-1); (b) isoclinal folds of
leucosome in amphibolite (outcrop 1203-2); (d) banded and isoclonally folded “Chipilín” gneiss (outcrop of samples CH-02, mafic, CH-03 felsic bands; (e) leucosome veinlets and
patches in amphibolite (outcrop 1203-2).
182 B. Weber et al. / Lithos 300–301 (2018) 177–199

unit, it shares similar metamorphic and deformation histories, and sim- by Gehrels et al. (2008) and Johnston et al. (2009). Ablation was
ilar mafic magmas intruded both Candelaria and Jocote units prior to de- performed with a Photon Machines Analyte G2 Excimer laser using
formation and metamorphism. Peak metamorphic conditions of the spot diameters of 30, 25, or 10 μm, a fluence of ~4 J/cm2 and a repetition
Jocote unit have been estimated at ~6 kbar and ~650 °C and the timing rate of 8 Hz (Gehrels et al., 2008). All measurements were made in
of metamorphism at 438+23−12 (González-Guzmán et al., 2016a). The static mode. For 30 and 25 μm spot analyses, Faraday detectors
depositional age of the Jocote unit was estimated in calcite marble with 3 × 1011 ohm resistors were used for 238U, 232Th, 208Pb-206Pb,
bands by chemostratigraphy (Sr and C isotopes) between 600 and and discrete-dynode ion counters for 204Pb and 202Hg. Using a spot
580 Ma. Chemical and isotopic evidence further indicate a pelagic diameter of 10 μm Faraday detectors were used for 238U, 232Th, and
zircon-depleted environment for sedimentation (González-Guzmán discrete-dynode ion counters for 208-204Pb. 204Hg interference with
204
et al., 2016a). The results were interpreted in terms of deposition during Pb is accounted for by measurement of 202Hg during laser ablation
the opening of the Eastern Iapetus Ocean in the Ediacaran Period and subtraction of 204Hg according to the natural 202 Hg/ 204 Hg of
evidencing for the first time Rodinia breakup in southern Mexico. 4.35. Common Pb correction is accomplished by using the measured
204
Pb and assuming an initial Pb composition from Stacey and
3.3. Massif-type anorthosite Kramers (1975) with uncertainties of 1.0 for 206 Pb/204Pb and 0.3
for 207Pb/204Pb. As a primary standard an in-house Sri Lanka zircon
Massif-type anorthosites, as they are constituted by almost pure pla- was used with an ID-TIMS age of 563.5 ± 3.2 Ma (2σ, Gehrels et al.,
gioclase, suffer extreme weathering in tropical regions. Occasionally, 2008). During the analytical sessions the secondary reference zircon
weathered outcrops contain relics of fresh anorthosite that can be con- R33 (419.3 ± 0.4 Ma, Black et al., 2004) was measured, yielding a
fused with xenoliths. The Fe-Ti oxide content of the anorthosite that also Concordia age of 421.4 ± 4.3 Ma (2σ, n = 10, MSWD = 0.63).
forms centimeter-decimeter thick ilmenitite lenses is a good feature Zircons from sample 9511-9 were analyzed at Centro de
that helps identifying it in the field. In the study area anorthosite oc- Geociencias, UNAM, Juriquilla, Mexico, using a Resolution M-050
curs south of the Polochic Fault from SW of Motozintla eastward 193 nm excimer laser coupled to a Thermo X-ii quadrupole ICPMS. Pri-
across the Guatemalan border and north of the Tonalá shear zone mary standard was reference zircon 91500 (1065 ± 5 Ma, Wiedenbeck
in the west of the study area (Cisneros de León et al., 2017). The gen- et al., 1995) and Pleŝovice zircon (336.7 ± 0.65 Ma, Sláma et al., 2008)
eral characteristics of the massif-type anorthosites in the El Triunfo was used as a secondary standard yielding a Concordia age of
Complex are: (1) they are associated with meter to tens of meters 334.8 ± 1.4 Ma (2σ, n = 9, MSWD = 0.12) during the analytical ses-
thick rutile-bearing ilmenitites that are mined for titanium; sion. For details on the analytical procedure and data reduction see
(2) near the villages of El Triunfo and Motozintla, anorthosite is in Solari et al. (2010b) and Solari and Tanner (2011).
contact with migmatized orthogneiss of the Candelaria unit suggest- Zircons from samples CH-02 and CH-03 were analyzed using a New
ing a genetic relationship with this basement unit; (3) intrusive con- Wave Nd-YAG UV 213 nm laser coupled to a Thermo Element 2 single
tacts with Permian plutonic rocks are typically observed along its collector, double-focusing, magnetic sector ICPMS at Radiogenic Isotope
horizontal extension and (4) penetrative alteration of the rheologi- and Geochronology Lab (RIGL) at Washington State University, Pullman
cally weak anorthosite is mainly controlled by Miocene to Recent WA, following the procedures and parameters similar to Chang et al.
shear zones. (2006). At this facility Pleŝovice zircon was used as primary standard
and FC-1 (1099 ± 2 Ma, Paces and Miller, 1993) as secondary standard
4. Analytical methods yielding a Concordia age of 1096.9 ± 6.4 (2σ, n = 45, MSWD = 0.77).
The results are given in Figs. 9 and 10, and the full dataset can be
4.1. Mineral chemistry found in Appendix A (supplementary data). The data were plotted
using ISOPLOT 3.75 (Ludwig, 2012), using error ellipses in Concordia
Mineral analyses were performed on three typical amphibolite diagrams of 1σ.
(R0301, R0202 and R0206) and one othogneiss boudin (1203-2a)
samples using a Jeol JXA-8500F electron microprobe, at Museum 4.3. Sm-Nd and Lu-Hf isotopes by isotope dilution analyses
für Naturkunde, Leibniz Institut für Evolutions- und
Biodiversitätsforschung, Berlin, Germany. Operating conditions were Chemical preparation and element separation for Sm-Nd and Lu-Hf
15 kV accelerating voltage, 15 nA beam current, and beam diameter of isotope analysis were performed in PicoTrace® cleanlab facilities at
1–2 μm for most minerals. Counting times were 20 s on peak and 10 s Departamento de Geología, CICESE, following the procedure based on
on background. The reproducibility of test measurements on a standard Sprung et al. (2010) with some modifications (González-Guzmán
was generally much better than 3% (1σ), except for values close to the et al., 2016b). Between 80 and 150 mg of whole rock powder was
detection limit, where the reproducibility may be reduced to about weighed into a digestion vessel and spiked with mixed 149Sm-145Nd
10%. Analyses were calibrated using Smithsonian international mineral and 180Hf-176Lu spikes. After dissolution with conc. HF, HNO3, and
standards. Representative analyses data are given in Table 1. HClO4 either in a PicoTrace® pressure digestion system (DAS®, mafic
rocks and anorthosites) or in a Parr® acid digestion vessel (felsic gneiss
4.2. U-Pb geochronology samples), evaporation, and sample-spike equilibration, the sample was
loaded with 3 N HCl on Teflon® columns filled with ~2 mL of Eichrom
Sample preparation and zircon separation was performed at Ln-Spec® resin. Most of the matrix elements including Light Rare
Departamento de Geología, CICESE, Ensenada, Mexico. Zircon was con- Earth Elements (LREE) for further element purification for Sm-Nd
centrated first by using a Wilfley® table, Frantz® isodynamic magnetic were eluted sequentially with 2–3 M and 3 M HCl. The HREE fraction
separator, followed by density separation using Diiodmethane. Zircon was eluted with 6 M HCl and evaporated to dryness. After the HREE
grains were mounted in epoxy resin and polished to half of the smallest cut the column was rinsed with 6 M HCl. Subsequently, Ti was eluted
crystal thickness. Cathodoluminescence (CL) imaging of the zircon using a mixture of 0.45 M HNO3, 0.09 M citric acid (HCit), and 1 vol%
mounts was performed with a JEOL JSM-35c electron microscope with of H2O2. Prior to Hf collection Zr was extracted from the column with
an attached CL detector at CICESE. a mixture of 6 M HCl and 0.06 M HF. Finally, Hf was collected with
For most samples the U-Pb geochronology of zircon by laser ablation 12 mL of 2 M HF and gently evaporated to dryness. A second cleanup
multicollector inductively coupled plasma mass spectrometry (LA-MC- was performed for both the HREE- and the Hf-cuts to improve Lu/Yb
ICPMS) was conducted on a Nu-Plasma instrument at the Arizona on the HREE- and to eliminate the Lu-tail on the Hf-cut. Separation of
LaserChron Center at Tucson, Arizona, following the procedures given Sm and Nd was performed using standard procedure (e.g., Weber
Table 1
Representative mineral chemistry composition (oxide [wt%]; cation [p.f.u.]) of amphibolites (R0301, R0206, R0202) and gneiss (1203-2a).

Sample R0301 (amphibolite) R0206 (amphibolite) R0202 (amphibolite) 1203-2a (gneiss)


ID
Amp Pl Pair Amp Pl Pair Amp Pl Fsp

Bulk Rim An(43–77) An(32–42) Amp Pl Inner Outer Rim An(19–27) An(46–54) Amp Pl Inner Outer Rim An(19–27) An(46–54) Bt Bt Ms Pl K-rich Na-rich
core core core core

(n = (n = (n = 8) (n = 8) (46–50) (53–55) (n = (n = (n = (n = 6) (n = 10) (28–31) (32−33) (n = (n = (n = (n = 10) (n = 14) (n = 3) (n = (n = (n = (n = (n =


48) 4) 8) 22) 8) 14) 20) 4) 12) 14) 4) 10) 12)

SiO2 (%) 47.33 43.08 53.47 59.88 46.45 57.09 54.38 44.08 41.75 55.83 62.97 49.53 59.54 50.80 47.27 43.70 57.16 63.57 37.38 36.56 47.66 61.79 64.33 68.49
TiO2 1.09 0.52 0.01 0.01 0.78 0.01 0.12 0.84 0.46 0.04 0.04 0.37 0.01 0.30 0.67 0.76 0.01 0.01 1.83 2.18 0.30 0.01 0.01 0.01
Al2O3 8.73 13.74 29.42 25.15 11.74 27.01 2.12 10.33 13.33 27.60 22.96 6.56 25.32 6.05 9.25 13.17 26.85 22.69 16.57 15.58 30.75 23.89 18.55 20.91

B. Weber et al. / Lithos 300–301 (2018) 177–199


Cr2O3 0.12 0.06 0.00 0.00 0.01 0.00 0.01 0.02 0.01 0.01 0.01 0.01 0.01 0.00 0.01 0.01 0.01 0.01 0.00 0.01 0.01 0.01 0.01 0.01
Fe2O3 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00
FeO 16.55 17.87 0.07 0.17 17.16 0.16 14.32 19.97 21.46 0.34 0.33 16.66 0.42 14.67 16.28 18.49 0.07 0.26 19.31 22.78 5.91 0.07 0.03 0.09
MnO 0.27 0.23 0.00 0.02 0.26 0.00 0.28 0.28 0.28 0.03 0.02 0.29 0.02 0.30 0.31 0.27 0.00 0.04 0.13 0.25 0.03 0.01 0.01 0.01
MgO 11.30 9.41 0.00 0.02 9.86 0.01 14.82 8.85 6.99 0.00 0.01 11.89 0.00 13.59 11.43 8.17 0.01 0.01 10.39 8.94 1.69 0.00 0.00 0.01
CaO 11.60 11.01 12.38 7.26 11.69 9.51 12.23 11.58 11.52 10.31 4.87 12.00 7.46 11.45 11.15 11.46 7.47 5.31 0.17 0.03 0.04 5.62 0.03 0.75
Na2O 1.07 1.24 4.55 7.66 1.24 6.33 0.33 1.46 1.79 5.73 9.08 0.92 7.61 0.66 1.11 1.24 6.40 7.97 0.26 0.09 0.29 8.46 0.60 11.08
K2O 0.28 0.39 0.04 0.05 0.37 0.06 0.04 0.45 0.52 0.04 0.06 0.25 0.05 0.15 0.33 0.53 0.05 0.68 8.64 9.22 10.18 0.25 14.76 0.32
Total 98.34 97.55 99.95 100.21 98.28 100.17 98.67 98.56 98.44 99.91 100.34 98.48 100.43 97.98 97.81 97.79 98.85 99.79 94.68 95.64 96.84 100.10 98.33 101.67
Si (p.f.u.) 6.879 6.321 6.205 5.304 6.836 5.697 7.350 6.966 6.238 5.820 4.842 7.184 5.341 7.345 6.796 6.479 5.481 5.075 5.691 5.643 6.377 10.970 12.004 11.800
Ti 0.119 0.058 – – 0.114 – 0.038 0.034 0.062 – – 0.041 – 0.025 0.075 0.085 – – 0.209 0.253 0.031 – – –
Alt 1.495 2.377 – – 1.563 – 0.859 1.429 2.450 – – 1.143 – 0.907 1.680 2.302 – – – 2.357 4.847 – – –
Cr 0.014 0.006 – – 0.013 – 0.001 0.000 0.001 – – 0.001 – 0.001 0.001 0.002 – – – 0.001 0.001 – – –
Fe3+ 0.531 0.891 – – 0.558 – 0.415 0.475 0.607 – – 0.358 – 0.637 0.714 0.472 – – 2.458 2.940 – – – –
Fe2+ 1.480 1.301 – – 1.466 – 1.582 1.787 2.117 – – 1.675 – 1.065 1.298 1.823 – – – – 0.661 – – –
Mn 0.033 0.029 – – 0.033 – 0.034 0.034 0.035 – – 0.036 – 0.036 0.037 0.033 – – 0.017 0.032 0.003 – – –
Mg 2.450 2.056 – – 2.419 – 2.720 2.274 1.490 – – 2.561 – 2.985 2.398 1.805 – – 2.358 2.057 0.337 – – –
Ca 1.806 1.732 2.374 1.391 1.801 1.822 1.853 1.858 1.848 1.976 0.933 1.869 1.430 1.767 1.735 1.820 1.573 1.017 0.028 0.005 0.005 1.069 0.006 0.138
Na 0.302 0.354 1.579 2.657 0.306 2.194 0.208 0.316 0.531 1.988 3.152 0.264 2.640 0.161 0.328 0.356 2.441 2.766 0.076 0.028 0.075 2.912 0.216 3.701
K 0.051 0.073 0.067 0.051 0.053 0.060 0.033 0.063 0.115 0.054 0.026 0.047 0.055 0.021 0.065 0.100 0.028 0.134 1.679 1.816 1.737 0.056 3.513 0.071
IV
Al 1.121 1.679 – – 1.164 – 0.650 1.034 1.762 – – 0.816 – 0.655 1.204 1.521 – – 2.309 2.357 1.623 – – –
VI
Al 0.374 0.698 – – 0.399 – 0.209 0.395 0.688 – – 0.327 – 0.252 0.476 0.782 – – 0.665 0.477 3.225 – – –
Fe* 0.585 0.491 – – 0.578 – 0.709 0.616 0.468 – – 0.607 – 0.714 0.593 0.368 – – – – – – – –
Mg# 0.623 0.515 – – 0.617 – 0.626 0.542 0.413 – – 0.594 – 0.737 0.650 0.497 – – – – – – – –
T 542 617 – – 548 – 464 523 642 – – 491 – 460 551 600 – – – – – – – –
P1 3.84 5.50 – – 3.97 – 2.61 3.71 5.71 – – 3.12 – 2.74 4.24 5.29 – – – – – – – –
P2 4.54 6.73 – – 4.71 – 2.97 4.42 7.00 – – 3.64 – 3.23 5.15 6.53 – – – – – – – –

Fe* = Fe3+ / (Fe3+ + VIAl). Mg# = Mg + Fe2+). O basis: amphiboles (Amp) = 23 oxygens (Leake et al., 1997); plagioclase (Pl) = 8 oxygens; micas, biotite (Bt) and muscovite (Ms) = 22 oxygens. Pressure (P, kbar) and Temperature (T, °C) estimates
are based on the empirical Ca-amphibole geothermobarometry after (1) Gerya et al. (1997) and (2) Zenk and Schulz (2004). Error ranges: T (±50 °C); P (±1.2 kb).

183
184 B. Weber et al. / Lithos 300–301 (2018) 177–199

et al., 2012) first with quartz-glass columns filled with DOWEX AG 5. Petrology of the Candelaria unit
50W-X8 resin to separate REE and then Ln-Spec® resin to separate
Sm and Nd. 5.1. Petrographic features
Individual zircon grains, previously measured with LA-ICPMS, were
picked from the mount and dissolved in small Teflon capsules at 190 The typical Candelaria unit (amphibolite and gneiss) is banded and
°C for one week in Parr® bombs. Lu and Hf separation was performed contains hornblende-rich and biotite-rich layers both defining the pen-
following the procedure described by Weber et al. (2012) that is etrative foliation. The hornblende-rich layers are mainly composed of
based on the method by Nebel-Jacobsen et al. (2005). brown-green hornblende that is zoned from brownish-green or
Lutetium and hafnium isotope ratios of solutions were measured on greenish-brown in the core to dark green at the rim, often containing
a Thermo Neptune Plus® MC-ICP-MS installed at LEI, UNAM, in static exsolutions of opaque (ilmenite or rutile) needles in the core or along
mode on Faraday cups. The sample solutes were introduced to the grain boundaries (Hbl1, Fig. 5a, b). Subhedral plagioclase fill spaces be-
plasma via an Aridus® desolvating sample introduction system using tween hornblende crystals together with quartz, mostly anhedral
an Ar carrier gas and a blended Ar + N2 sweep gas. Samarium and brown biotite, ilmenite, and apatite. Biotite-rich layers often contain
neodymium isotope ratios were measured on a Thermo-Scientific® abundant plagioclase and quartz and only minor brownish-green horn-
TRITON® TIMS at LUGIS, UNAM. blende. These layers typically contain relatively large anhedral poikilitic
Isotope dilution analysis was performed on the same solution as for amphiboles with colorless to pale green cores of actinolite or actinolitic
isotope composition, following the approach of Boelrijk (1968) and Chu hornblende having numerous quartz inclusion, sometimes symplectitic
et al. (2011). Besides that, the mass bias for spiked Lu was corrected by textures, and green to bluish-green rims (Hbl2, Fig. 5b). Brownish-green
adding Re to every sample and by normalizing to external Lu + Re stan- hornblende (Hbl1) was probably a Ti-rich high-temperature
dard solutions with known isotopic composition (González-Guzmán hornblende present in the protolith that lost part of its Ti by exsolution
et al., 2016b; Scherer et al., 1999). Correction for mass bias for Sm, Nd, of ilmenite or rutile during metamorphism. The pale green poikilitic
and Hf was achieved by normalizing to 152Sm/147Sm = 1.7845352, actinolite, instead (Hbl2), appears to form by pseudomorphic replace-
Nd: 146Nd/144Nd = 0.7219, and Hf: 179Hf/177Hf = 0.7325, respectively. ment of pyroxene followed by continued prograde reaction forming
The data reduction was performed offline using IsotopeHf® and bluish-green hornblende (Thomson et al., 1985). This hypothesis is
IsotopeNd® programs (González-Guzmán et al., 2016b) written in R, a further supported by the presence of (relict?) clinopyroxene rarely ob-
free software environment for statistical computing and graphics. The served in hornblende gneiss samples (Fig. 5f). Abundant opaque min-
results are listed in Table 2. Over the course of this work the JMC-475 erals are mostly ilmenite typically surrounded by anhedral titanite
Hf standard gave a mean 176Hf/177Hf of 0.282150 ± 0.000007 (n = (Fig. 5a). Occasionally, and mostly restricted to particular layers in the
41) and the La Jolla Nd isotope standard yielded a mean 143Nd/144Nd rock, hornblende is partially replaced by epidote (and/or clinozoisite)
of 0.511846 ± 0.000003 (n = 3). The reported 176Hf/177Hf and forming symplectitic intergrowth with quartz (Fig. 5d). Epidote is a
143
Nd/144Nd values in Table 2 were normalized to the well-accepted common secondary phase, whereas green chlorite replacing biotite is
176
Hf/177Hf for the JMC-475 of 0.282160 and 143Nd/144Nd for the La less abundant.
Jolla standard of 0.511860, respectively. On the basis of duplicate sam- Partial melting is observed at most outcrops as neosome veinlets and
ple dissolution and repeated analyses of reference rock standards, dikes from millimeter to meter scale. The neosome is composed mainly
total uncertainties are estimated around 0.003% for both 176Hf/177Hf of plagioclase and quartz, minor K-feldspar, sometimes with abundant
and 143Nd/144Nd, 0.5% for parent/daughter ratios (176Lu/177Hf and either reddish (Fig. 5a) or brownish biotite, green hornblende, euhedral
147
Sm/144Nd), and 2% for element concentrations by ID analyses. titanite, accessory apatite, and zircon as well as secondary actinolite
needles and epidote.
The felsic gneiss boudins consist of quartz (with dynamic recrystal-
4.4. Lu-Hf analyses of zircon by laser ablation MC-ICPMS lization), plagioclase, and perthitic K-feldspar, the latter partially
replaced by muscovite (Fig. 5c). Large (up to 0.5 mm) apatite, brown
In-situ Lu-Hf isotope composition of zircon crystals was determined biotite, and zircon are primary accessory phases. Quartz contains small
at the Radiogenic Isotope and Geochronology Lab (RIGL) at Washington rutile needles suggesting exsolution of TiO2 from high temperature
State University. Analyses were conducted using ThermoFinnigan quartz and relict rutile is present as inclusions in titanite (Fig. 5e).
Neptune® mass spectrometer coupled to a New Wave 213 nm
Nd:YAG laser, using a spot size of 40 μm, a laser fluence of ~ 7 J/cm2, 5.2. Mineral chemistry and geothermobarometry
and repetition rate of 10 Hz. This study used the same instrument con-
figuration, operating parameters, and data reduction methods outlined According to the classification after Leake et al. (1997), the amphib-
by Fisher et al. (2014a), with the exception that U-Pb ages were not si- olites contain calcic amphiboles with actinolithic cores or patches
multaneously determined. The output from the ablation cell was mixed (Figs. 6a, d and 7a, b) that display high Si and Mg# (Mg / (Mg
with N2 gas and delivered directly to the Neptune MC-ICPMS. To reduce + Fe2+)) and low AlTOT contents. The main amphibole crystals are ei-
inter-laboratory bias, the Plešovice zircon standard (176Hf/177Hf = ther magnesio-hornblende (R0301 and R0202) or ferro-edenite
0.282482 ± 13, Sláma et al., 2008) was regularly analyzed between (R0206). Towards the rims they display significantly higher Al (2.0–
sample blocks and used to correct the measured 176Hf/177Hf of un- 2.6 AlTOT p.f.u. out of 13 cations) with lower Si and Mg#, respectively
knowns. Given the potentially large range of (Lu + Yb)/Hf in zircon (Table 1), reaching ferro-tschermakitic (R0301, R0202) and ferro-
samples, accurate correction for the isobaric interference of 176Yb and pargasitic compositions (R0206). The compositional zoning and their
176
Lu on 176Hf is imperative, and should be assessed using quality con- chemical features suggest that amphibole crystals are metamorphic
trol zircons interspersed with samples (Fisher et al., 2014b). The stan- (not igneous) and they reflect changing metamorphic conditions
dard zircon analyses gave 176Hf/177Hf of 0.282440 ± 0.000037 (1σ, n along a prograde metamorphic path (e.g., Ernst and Liu, 1998; Raase
= 21) for Pleŝovice, 0.282156 ± 0.000022 (1σ, n = 8) for FC-1 and et al., 1986; Spear, 1981; Zenk and Schulz, 2004).
0.282105 ± 0.000022 (1σ, n = 11) for synthetic zircon, respectively. Plagioclase compositions are variable, in particular in sample R0301
Analyses of these quality control zircons agree well with published where core plagioclase goes from bytownite over labradorite to
S-MC-ICPMS isotope compositions of purified Hf from these zircons, andesine (Fig. 7c, An77 to An42). The zoned An-rich plagioclase is
attesting to the accuracy of the interference correction methods probably magmatic, whereas a rim with An34 is supposed to be
employed. Internal 2σ precision was typically ~ 1.2 εHf. The results metamorphic. Core plagioclase from other two samples is mainly
are included in Table 2. andesine (~ An45, sample R0202) or andesine to labradorite (sample
Table 2
Lu-Hf and Sm-Nd data for whole rocks and Lu-Hf data on zircon.

Sample ID Rock type Lu (ppm) Hf (ppm) Lu-Hf isotope ratios Age correcteda Model age Sm (ppm) Nd (ppm) Sm-Nd isotope ratios Age correcteda Model age
176 176 176 b 147 143 143
Lu Hf ±s.e. ε Hf t (Ma) Hf ε Hf (t) (Ma) Sm Nd ±s.e. ε Nd Nd ε Nd(t) (Ma)c
177 177 177 144 144 144
Hf Hf 0 Hf(t) Nd Nd 0 Nd(t)

Isotope dilution analyses


1205-1b Orthogneiss (boudin) 0.59 28.72 0.00293 0.282251 4 −18.9 1005 0.282196 1.7 1451 13.9 75.1 0.1116 0.512096 4 −10.6 0.511352 0.5 1402
1203-2a Orthogneiss (boudin) 0.59 25.64 0.00328 0.282230 4 −19.6 1016 0.282167 0.9 1497 21.9 138.3 0.0958 0.512073 4 −11.0 0.511441 2.0 1246
1203-2a-b Zircon (orthogneiss) 0.00069 0.282213 3 −20.2 1016 0.282200 2.1 1560
1203-2a-a Zircon (orthogneiss) 0.00063 0.282154 2 −22.3 1016 0.282142 0.0 1669
CH-02 Gneiss (Chipilin) (mafic) 0.29 2.32 0.01804 0.282691 4 −3.3 919 0.282373 6.2 1387 6.1 33.5 0.1105 0.512286 35 −6.9 0.511608 3.5 1113
CH-03 Gneiss (Chipilin) (felsic) 0.37 3.99 0.01331 0.282468 4 −11.2 919 0.282234 1.2 1593 12.4 70.6 0.1066 0.511921 29 −14.0 0.511267 −3.2 1584
TR-02 Anorthosite 0.10 0.94 0.01508 0.282592 3 −6.8 1000 0.282308 5.5 1435 0.4 2.3 0.1101 0.512240 3 −7.8 0.511518 3.3 1174
PAN-14 Anorthosite 0.3 1.8 0.0951 0.512108 31 −10.3 0.511484 2.7 1195
1202-3 Anorthosite 0.03 0.51 0.00733 0.282328 6 −16.2 1000 0.282190 1.4 1525 1.1 6.9 0.1001 0.512075 6 −11.0 0.511418 1.4 1291
NF-01 Anorthosite 0.16 8.87 0.00257 0.282281 4 −17.8 1000 0.282232 2.9 1393 1.2 7.1 0.0984 0.512124 4 −10.0 0.511479 2.6 1208
CH-01 Anorthosite 0.03 0.32 0.01325 0.282511 4 −9.7 1000 0.282261 3.9 1499 1.1 6.0 0.1118 0.512156 4 −9.4 0.511423 1.5 1317
1203-2b Amphibolite (E-MORB) 0.34 1.97 0.02447 0.282899 4 4.0 600 0.282623 7.7 1239 3.2 10.9 0.1794 0.512853 6 4.2 0.512148 5.5 −
1203-2c Amphibolite (E-MORB) 0.36 1.92 0.02681 0.282895 4 3.9 600 0.282593 6.6 1504 3.8 11.8 0.1952 0.512855 4 4.2 0.512088 4.4 −
1203-2c-2 Zircon (amphibolite) 0.00037 0.282702 6 −2.9 445 0.282699 6.9 831

B. Weber et al. / Lithos 300–301 (2018) 177–199


1203-2c-5 Zircon (amphibolite) 0.00108 0.282251 7 −18.9 996 0.282230 2.7 1510
1203-2c-6 Zircon (amphibolite) 0.00028 0.282245 3 −19.1 464 0.282243 −8.6 1674
1205-1a Amphibolite (E-MORB) 0.57 3.35 0.02423 0.282889 6 3.7 600 0.282616 7.4 1255
1205-2a Amphibolite (E-MORB) 0.45 2.90 0.02193 0.282988 3 7.2 600 0.282741 11.8 764 3.9 16.0 0.1467 0.512850 3 4.1 0.512273 8.0 465
1205-3 Amphibolite (E-MORB) 0.46 2.91 0.02267 0.282915 3 4.6 600 0.282660 9.0 1044 5.1 17.9 0.1718 0.512754 4 2.3 0.512078 4.2 1044
R0103 Amphibolite (E-MORB) 0.25 1.67 0.02114 0.282865 3 2.8 600 0.282627 7.8 1105 4.7 15.8 0.1794 0.512774 30 2.7 0.512069 4.0 –
R0202 Amphibolite (E-MORB) 0.52 2.86 0.02593 0.282907 3 4.3 600 0.282615 7.4 1349 3.6 12.4 0.1748 0.512759 3 2.4 0.512072 4.0 1100
R0206 Amphibolite (E-MORB) 0.36 2.33 0.02226 0.282870 3 3.0 600 0.282619 7.5 1165 6.6 23.9 0.1674 0.512702 3 1.2 0.512044 3.5 1112
R0302 Amphibolite (E-MORB) 0.43 3.18 0.01918 0.282885 3 3.5 600 0.282669 9.3 938 5.6 19.8 0.1692 0.512772 4 2.6 0.512107 4.7 927
RE1101 Amphibolite (E-MORB) 0.41 2.72 0.02162 0.282884 3 3.5 600 0.282640 8.3 1077 4.3 13.7 0.1902 0.512852 6 4.2 0.512104 4.7 –
CAL-03 Amphibolite (E-MORB) 0.33 2.11 0.02234 0.282870 3 3.0 600 0.282618 7.5 1171 5.1 16.3 0.1878 0.512838 10 3.9 0.512100 4.6 –

Laser ablation MC-ICPMS


CH02_1 Zircon (Chipilin gneiss) 0.00085 0.281978 29 −28.6 990 0.281962 −7.0 2012
CH02_2 Zircon (Chipilin gneiss) 0.00068 0.282020 33 −27.1 845 0.282009 −8.6 1976
CH02_3 Zircon (Chipilin gneiss) 0.00065 0.282043 31 −26.3 913 0.282031 −6.2 1910
CH02_4 Zircon (Chipilin gneiss) 0.00093 0.282015 27 −27.2 926 0.281998 −7.1 1967
CH02_5 Zircon (Chipilin gneiss) 0.00075 0.282025 33 −26.9 814 0.282013 −9.1 1979
CH02_6 Zircon (Chipilin gneiss) 0.00065 0.281964 36 −29.0 911 0.281952 −9.1 2058
CH02_7 Zircon (Chipilin gneiss) 0.00091 0.282078 32 −25.0 868 0.282063 −6.1 1868
CH02_8 Zircon (Chipilin gneiss) 0.00010 0.281892 25 −31.6 901 0.281890 −11.5 2177
CH02_9 Zircon (Chipilin gneiss) 0.00059 0.282078 43 −25.0 990 0.282067 −3.2 1818
CH02_10 Zircon (Chipilin gneiss) 0.00072 0.281964 31 −29.0 845 0.281952 −10.6 2081
CH03_1 Zircon (Chipilin gneiss) 0.00062 0.282051 37 −26.0 885 0.282040 −6.6 1904
CH03_2 Zircon (Chipilin gneiss) 0.00067 0.282065 26 −25.5 943 0.282053 −4.8 1861
CH03_3 Zircon (Chipilin gneiss) 0.00072 0.282059 36 −25.7 902 0.282046 −6.0 1887
CH03_4 Zircon (Chipilin gneiss) 0.00053 0.282037 21 −26.5 918 0.282028 −6.3 1916
CH03_5 Zircon (Chipilin gneiss) 0.00046 0.282031 25 −26.7 952 0.282022 −5.7 1914
CH03_6 Zircon (Chipilin gneiss) 0.00063 0.282000 24 −27.8 938 0.281988 −7.2 1981
CH03_7 Zircon (Chipilin gneiss) 0.00056 0.281933 34 −30.1 902 0.281923 −10.3 2115
a
Initial (t) isotope ratios and ε-values are recalculated for measured 207Pb/206Pb age from laser ablation ICPMS for zircon (see Supplementary data Appendix A), the Concordia intercept age (U-Pb zircon) if available for whole rock data (and zircon
from sample 1203-2a) and to either 1000 (gneiss and anorthosite) or 600 Ma (E-MORB) estimated ages.
b
Lu-Hf depleted mantle model ages are calculated using present day depleted mantle model 176Hf/177Hf = 0.283224 (Vervoort et al., 2000) and 176Lu/177Hf = 0.03836 (calculated for εHf = 0 at 4500 Ma by using λ = 1.867 × 10-11a-1, Scherer
et al., 2001; Söderlund et al., 2004); for zircon the model ages are calculated in two-steps: 1st step 176Lu/177Hf of sample to 207Pb/206Pb age, 2nd step before zircon crystallization assuming 176Lu/177Hf of 0.01 for an average felsic crust.
c
Sm-Nd depleted mantle model ages are calculated after DePaolo (1981). Average measured 176Hf/177Hf of the JMC-745 standard solution at Juriquilla (UNAM) was 0.282149 ± 0.000007 (n = 41). Average measured 143Nd/144Nd of the La Jolla
standard at LUGIS (UNAM) was 0.511846 ± 0.000003 (n = 3); Standard zircon analyses with LA-MC-ICPMS at WSU, Pullman, gave 176Hf/177Hf of 0.282440 ± 0.000037 (n = 21) for Pleŝovice, 0.282156 ± 0.000022 (n = 8) for FC-1 and 0.282105 ±
0.000022 (n = 11) for synthetic zircon, respectively.

185
186 B. Weber et al. / Lithos 300–301 (2018) 177–199

(a) (b)

(c) (d)

(e) (f)

Fig. 5. Microphotographs of (a) biotite-bearing amphibolite with zoned brownish-green hornblende (Hbl1), biotite (Bt1), and ilmenite surrounded by titanite; neosome veinlet composed
of quartz, plagioclase, and reddish-brown biotite (Bt2); (b) hornblende-biotite gneiss with small dark-green hornblende (Hbl1) and large poikilitic hornblende (Hbl2) with pale actinolitic
cores and green rims or patches suggesting metamorphic growth after pyroxene; (c) quartz-feldspar gneiss with K-feldspar partially reacted to muscovite (crossed Nicols);
(d) amphibolite with symplectitic textures of epidote + quartz (crossed Nicols); (e) hornblende gneiss from banded “Chipilín” gneiss showing titanite accumulation with rutile
inclusion (f) hornblende gneiss with Hbl1 and clinopyroxene.

R0206), whereas these plagioclase crystals have well-defined metamor- temperature estimates (see error bar in Fig. 7b) this is considered as a
phic rims of oligoclase (inset Figs. 6e and 7c) that is typical for metamor- qualitative approach only, used to show the prograde metamorphic
phic plagioclase at amphibolite facies conditions. Besides that, sample evolution of amphibole.
R0202 has additional albite rims, probably grown during greenschist fa- To further constrain the temperature estimates, the hornblende-
cies overprint together with chlorite (replacing biotite) and epidote plagioclase geothermometer after Holland and Blundy (1994) was
+ quartz symplectites (Figs. 6b, c and 7c). applied, using the calibration based on the reaction edenite + 4
Inasmuch as Al and Si contents with respect to Mg# for any Ca- quartz = tremolite + albite that is recommended for quartz saturated
amphibole depend on metamorphic pressure and temperature, samples. The calculations were conducted in amphibole-plagioclase
this chemical relationship can be used to estimate P-T conditions for pairs that are in contact for samples R0301 and R0206. Unfortunately,
any Ca-amphibole-bearing metamorphic assemblage that includes pla- no direct contact pair was available from sample R0202, therefore,
gioclase and quartz (Gerya et al., 1997; Zenk and Schulz, 2004). Fig. 7d average compositions were used. The resulting peak temperatures
illustrates the results of P-T estimates for all individual points measured are within errors consistent for all three samples between 650 °C and
on the Ca-amphiboles, indicating increasing P-T conditions from core to 662 °C (±40 °C) at an estimated peak pressure of 6 kbar (Fig. 7d).
rim reaching peak conditions at ca. 6 kbar and N 600 °C. Note that the To further constrain the pressure estimates, the Si-in-phengite
resulting pressure is generally 0.75 kbar higher using the approach barometer after Massonne and Schreyer (1987), which is based on the
from Zenk and Schulz (2004) compared to those by Gerya et al. pressure dependence of the Si excess from ideal muscovite in the
(1997), however, the authors suggest an even larger error of tetrahedral occupancy, was applied to white mica from an orthogneiss
±1.2 kbar for the pressure estimates, thus the average of both values boudin (sample 1203-2a). The Si-contents of measured white mica
is used and illustrated in Fig. 7d. Due to the large uncertainties in the crystals range from 3.12 to 3.24 p.f.u., suggesting pressures in the
pressure calculation and another ca. ± 50 °C uncertainly in the range from 5.1 to 8.0 kbar at a supposed temperature of 650 °C
B. Weber et al. / Lithos 300–301 (2018) 177–199 187

(a) (b)

(c) (d)

(e) (f)

Fig. 6. Backscattered electrons images of analyzed amphibolites (a–e) Notes: relict patchy actinolite cores in zoned hornblende (a, d) having Al-poor cores and Al-rich rims (magenta
colored numbers are measured AlTOT compositions); epidote + quartz symplectites (b, c); white numbers on zoned plagioclase are measured An contents (b, c, inset e);
(f) orthogneiss 1203-2a: perthitic K-feldspar partially replaced by fibrous phengite (Ms).

(Fig. 7d). Inasmuch as two generations of white mica are observed, the age distributions with concordant ages at 1016.9 ± 4.2 Ma (sample
phengite with higher Si-contents (N 3.2 p.f.u.) may be relict from a pre- 1203-2a) and 1005.1 ± 7.5 Ma (sample 1205-1b) on one hand and
vious event at higher pressures. 965 ± 7.5 Ma (sample 1203-2a) and 950 ± 8 Ma (sample 1205-1b)
on the other (Fig. 9a, c). As illustrated in age probability density and
6. U-Pb geochronology histogram plots (Fig. 9b, d), the deconvolved ages applying the
“unmix” algorithm of Isoplot (Ludwig, 2012) suggest two populations
6.1. Felsic and banded gneiss with relative maxima at ~1017 Ma and ~1005 Ma, reflecting the igneous
age of the protoliths, and a second one at ~970 and ~950 Ma, reflecting
Sample 1203-2a is a quartz-feldspar gneiss boudin from the ancient lead loss.
Candelaria unit type locality at the Candelaria River and 1205-1b is a Another granitic gneiss sample is from NE of Belisario Domínguez
similar sample from another quartz-feldspar gneiss boudin from north (9511-9, Fig. 3). Similar to the other felsic gneisses, most zircons display
of the village of El Triunfo (Figs. 3, 4a, c). Cathodoluminescence (CL) low luminescent U-rich cores, partly with oscillatory zoning and high
images show oscillatory zoning in the zircon cores, with resorbed em- luminescent U-poor rims (not shown). Forty-three laser spot analyses
bayments filled with low luminescent zircon, and thin high luminescent were performed, 16 on cores, the rest on high-luminescent rims or
metamorphic overgrowth (Fig. 8a, b). Forty-two and 34 laser spots were unzoned zircon. The zircon cores reflect mainly igneous crystallization
measured on zircons from these two samples. The resulting 207Pb/206Pb whereas the rims metamorphic growth. However, some cores were par-
apparent ages range from ~1096 to ~933 Ma. Both samples have similar tially reset during metamorphism or the zircons suffered recent lead
188 B. Weber et al. / Lithos 300–301 (2018) 177–199

Calcic amphiboles: CaB > 1.50; (Na+K)A < 0.50 Or


(a) 1.0
Tremolite
(c) 100

Magnesio-
0.8 Actinolite

20
hornblende
Tshermakite 80
Mg/(Mg+Fe)
0.6

40
60
0.4

Ferro- Ferro- Ferro-

60
0.2 tshermakite hornblende actinolite 40

0.0

80
5.5 6.0 6.5 Si 7.0 7.5 8.0 20
Calcic amphiboles: CaB > 1.50; (Na+K)A > 0.50
(b) 1.0
Oli And Lab By

0
10
Ab An

80

60

40

20
10
0.8

80
0
20
Pargasite Edenite
Mg/(Mg+Fe)

0.6
Oli And Lab By

0
10
Ab An

80

60

40

20
10
0.4

80
20
Ferro- Ferro-
0.2 pargasite edenite
Oli And Lab By

0
10
0.0
Ab An
80

60

40

20
10

5.5 6.0 6.5 7.0 7.5 8.0


Si
0

10
(d) R0206 R0301 Ca-amp core main rim
R0301
Ky
Sil R0202
3 .4 R0206
8
.) =
f.u
e (p. 650° 6 kb
in Ph 1203-2a
Si Si-in-Phe
6
P (kbar)

Jocote unit
3.3 G.G. 2016

ut
“o
4 Ep R0202
3.2 tz
Q

Hbl+Pl+
“out

+
p Ep Qtz+F
+E
Chl

hl z
3.1 C
t+ Qt
2 Ac Ab+ And
+ Granite
wet solidus

HB94a
0
300 400 500 600 700 800
T(°C)

Fig. 7. Mineral chemistry. (a, b) Classification diagrams for calcic-amphiboles measured in amphibolite samples; (c) ternary feldspar classification diagrams for measured plagioclase in
amphibolite samples (Ab = albite, An = anorthite, Or = orthoclase); (d) pressure-temperature diagram showing the results from geothermobarometric calculations of individual
measurements on zoned calcic amphiboles (average of calculations after Gerya et al. (1997), and Zenk and Schulz (2004), note: the error bar corresponds to these
calculations), hornblende-plagioclase geothermometer after Holland and Blundy (1994) (dashed and point-dashed black lines, HB94a: using calibration a, based on the
reaction edenite + 4 quartz = tremolite + albite), and phengite geobarometer after Massonne and Schreyer (1987) (grey dashed lines show isopleths for Si-contents in
phengite per formula unit; pressure estimates using average peak temperature of 650 °C.

loss. An upper Concordia intercept of 21 either concordant or normally Zircon data from two samples taken at the same outcrop of the
discordant analyses yielded an age of 1003 +23/−18 Ma (Fig. 9e), which banded “Chipilin” gneiss (one felsic and one mafic portion, CH-03 and
is interpreted to be the best estimate for igneous crystallization of the CH-02) in the eastern part of the study area (Fig. 3) are presented to-
rock. Another group of twelve analyses, mostly from high luminescent gether in Fig. 9g and h. Most of the zircon grains are rounded or short
rims, yield a Concordia age at 896 ± 7 Ma. On age probability density prismatic bi-pyramidal with diffuse grayish luminescence or patchy
and histogram plots (Fig. 9f), the “unmix” algorithm clearly separates fields on CL images (Fig. 8c, d). They rarely display relics from oscillatory
two populations one at 1004 ± 13 Ma and another at 907 ± 20 Ma. zoning but they do contain inherited cores, some of which appear
B. Weber et al. / Lithos 300–301 (2018) 177–199 189

Fig. 8. Cathodoluminescence (CL) images of typical zircon grains from (a, b) orthogneiss boudins, (c, d) the “Chipilín” gneiss, and (e) an amphibolite. Numbers are 207Pb/206Pb ages of
(small) individual laser spots in Ma or εHf(i) values (large spots in CH-02 and CH-03).

partially erased on CL images probably due to later stage recrystalliza- (paleo-, melano-, or mesosome) samples were analyzed. It needs to be
tion processes. Consequently, their U-Pb isotope ratios yield strongly noted that some of the mafic samples contained few zircons, which
discordant ages, but some of the cores have apparent 207Pb/206Pb ages may have been derived from mm-sized neosome veinlets that could
N1.4 Ga, indicating inheritance from an Early Mesoproterozoic not be separated from the rock. Such special circumstance was consid-
continental component similar to some of the detritus in the overlying ered in the interpretation of the data.
Jocote unit (González-Guzmán et al., 2016a; Weber et al., 2008) or Sample 1203-1b is banded hornblende- and biotite-bearing quartz-
other metasedimentary sequences in the region (e.g., Martens et al., feldspar gneiss from the Candelaria unit. Two types of zircon are
2010; Weber et al., 2012). Forty-two out of 47 analyzed laser spots distinguished on CL images: (1) short- to long prismatic well-developed
yield a Concordia age at 911.2 ± 9.2 Ma or an upper Concordia intercept bi-pyramidal zircon with oscillatory zoning and (2) zircon with diffuse
age (forced to a lower intercept at 250 ± 10 Ma that corresponds to the internal structure (recrystallization) or cores with bi-pyramidal sub-
age of granitic intrusions close by) at 919 ± 13 Ma (Fig. 9g). Noteworthy, rounded to elliptic forms (Fig. 8e). The former yield Ordovician whereas
(1) no significant number of laser spots yielded ages around 1000 Ma the latter Stenian-Tonian ages (~1 Ga). Twenty-seven out of 40 analyzed
(Fig. 9h), as observed in the abovementioned samples, and (2) the laser spots yield a Concordia age of 454.9 ± 3.1 Ma. The remaining data
Concordia age corresponds to the second age population mainly of meta- are from inherited zircon or mixed ages of which an upper intercept age
morphic rims of sample 9511-9 and it seems the data suggest an increas- of 971 ± 90 Ma is calculated (Fig. 10a).
ing influence of the ca. 0.9 Ga event from west to east of the study area. Sample 1203-2c is a folded amphibolite adjacent to the dated felsic
The Th/U values of the ~0.9 Ga zircons range from 0.16 to 1.6 with an av- gneiss boudin (1203-2a, see above). Only nine zircon grains were
erage at 0.7. These values are not very helpful to decide whether zircon is obtained of which five yield a concordant age at 457 ± 20 Ma and
either of igneous or metamorphic origin, in particular when the U con- four laser spots date inherited components resulting in an upper
tents are low (as low as 21 ppm; 150 ppm by average). However, consid- intercept age at 1014 ± 140 Ma (Fig. 10d).
ering the CL images a metamorphic origin is preferred for interpretation. Hosting the other felsic gneiss boudin analyzed in this study (1205-
1b, see Fig. 4a), the biotite-hornblende gneiss sample 1205-1c contains
6.2. Deformed amphibolite, mafic gneiss, and migmatite large (N200 μm) zircon grains either with well-developed oscillatory
zoning or zircon with large cores and overgrowth rims, also with oscil-
In order to constrain the age of the deformed mafic rocks, zircons latory zoning (Fig. 11f). Ten out of 27 analyzed laser spots yield a
from nine amphibolite, hornblende-biotite gneiss, and migmatite Concordia age of 451 ± 9 Ma but with a low probability of 0.14
190 B. Weber et al. / Lithos 300–301 (2018) 177–199

(a) 1203-2a, orthogneiss boudin (b) n


0.19 Candelaria river 1120

12
1080
0.18
10
Pb/ U

1040
238

0.17 Conc. Age = 8


1000 Age ±2σ fraction
1016.9 ± 4.2 Ma 970.4 ± 14 0.32
206

960 MSWD = 0.59, p.= 0.98 1008 ± 7.4 0.68 6


0.16
n = 21 relative misfit = 0.97
920 4
Conc. Age =
0.15
880 965 ± 7.5 Ma 2
MSWD = 0.59, p.= 0.93, n = 11
0.14 0
1.3 1.5 1.7 1.9 2.1 800 900 1000 1100 1200

(c) 1205-1b, orthogneiss boudin (d) 10


north of El Triunfo 1080
0.18
1040 8

Conc. Age =
Pb/238U

0.17
950 ± 8 Ma 1000 Age ±2σ fraction 6
MSWD = 1.3 953.1 ± 16 0.25
p.= 0.12 960 998.3 ± 6.4 0.75
206

0.16 Conc. Age = relative misfit = 0.94


n = 14 4

920 1005.1 ± 7.5 Ma


MSWD = 0.27, p.= 0.98
0.15 2
880
n=8

0.14
1.3 1.5 1.7 1.9 800 900 1000 1100 1200
7
(e) 9511-9, felsic gneiss (f)
0.18
Belisario Dominguez 1050 6

Conc. Age = 5
0.16
896 ± 7 Ma 950
Age ±2σ fraction
Pb/ U

MSWD = 0.9 4
238

907.1 ± 20 0.27
p.= 0.61 1004 ± 13 0.73
n = 12 850
0.14 Upper Intercept at relative misfit = 0.88 3
206

+23
1003 /-18 Ma
750 MSWD = 0.78, n = 21 2
0.12

0.10
0.9 1.1 1.3 1.5 1.7 1.9 2.1 800 900 1000 1100 1200

(g) CH-2/CH-3 (h) 926 ± 11 Ma


1300 18
0.22 “Chipilin” gneiss
16
east of Motozintla
14
Pb/ U

1100
12
238

0.18
10
Conc. Age =
206

8
900 911.2 ± 9.2 Ma
0.14 MSWD = 0.7, p. = 0.98, n = 42 6

4
Upper Intercept at 919 ± 13 Ma ~1.17 Ga ~1.43 Ga
2
MSWD = 0.52, (forced to 250 ± 10 Ma)
0.10 0
1.0 1.4 1.8 2.2 2.6 3.0 800 900 1000 1100 1200 1300 1400
207
Pb/235U 207
Pb/206Pb age (Ma)

Fig. 9. U-Pb isotope data of felsic orthogneiss samples from the Candelaria unit (a–f) and the banded “Chipilín” gneiss (g, h): (a, c, e, f) Wetherill Concordia diagrams of U-Pb isotope data
and (b, d, f, h) corresponding relative probability and histogram plots of 207Pb/206Pb ages applying the unmix algorithm of Isoplot 3.75 (Ludwig, 2012). (Stippled error ellipses are data not
used for age calculation; errors on age calculations are 2σ, error ellipses are plotted at the 1σ level; MSWD = mean standard weighted deviates of Concordia ages correspond to MSWD of
concordance and equivalence; p. = probability of concordance and equivalence).
650
(a) 1203-1b Conc. Age = 0.20 1203-2c 1100
0.10 1205-1c
1100 454.9 ± 3.1 Ma 550
0.18 0.09
MSWD = 0.83, p.= 0.998
550
0.16
(d) 0.08
(g)
900
n = 27 900
450
Pb/238U

Pb/238U
0.14 0.08 Intercepts at Intercepts at
Intercepts at 0.12 700 0.06
700 450 446 ± 63 & 350 429 ± 67 &
462 ± 13 & 1019 ± 140 Ma 1037 ± 93 Ma
0.10 0.07
206

971± 90 Ma

206
0.08 500 0.04 250
MSWD = 0.15 MSWD = 2.6
500
MSWD = 1.05
0.06 0.06 300 Conc. Age = 150
0.04 0.02
300 350 457 ± 20 Ma Conc. Age = 451 ± 9 Ma
MSWD = 0.16, p.= 0.997, n = 5 50
MSWD = 1.5, p.= 0.14, n = 10
0.02 0.05 0.00 0.00
0.0 0.4 0.8 1.2 1.6 2.0 2.4 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.4 0.8 1.2 1.6 2.0 0.0 0.2 0.4 0.6 0.8

(b) R0401 Conc. Age = R0103 1205-2

B. Weber et al. / Lithos 300–301 (2018) 177–199


1100 540 0.09
0.18 446.5 ± 3.8 Ma 540
0.085 MSWD = 0.91, p.= 0.66 0.085 (e) (h) 500
900 n = 26

Pb/238U
0.14
Pb/ U

0.07
238

Intercepts at 0.075 460 400


700 0.075 460
429 ± 11 &
0.10
924 ± 85 Ma

206
206

500
MSWD = 0.8 0.05 Conc. Age =
0.065 0.065 Conc. Age = 300
0.06 380 380 449.5 ± 3.4 Ma 443.9 ± 2.9 Ma
300
MSWD = 0.86, p.= 0.77
MSWD = 1.2, p.= 0.18 n = 31
0.02 0.055 n = 19
0.055 0.03
0.0 0.4 0.8 1.2 1.6 2.0 0.35 0.45 0.55 0.65 0.2 0.4 0.6 0.8 0.2 0.3 0.4 0.5 0.6 0.7

(c) R0302 0.082 Conc. Age = R0202 R0206 600


1100
0.18 449.9 ± 4.3 Ma 0.095
0.09 550
0.078 MSWD = 0.56, p.= 0.95 480 (f) (i)
900 n = 11 0.085

Pb/238U
0.14
Pb/ U

500
0.08
0.074
238

Intercepts at
700
450
0.075
0.10
441 ± 19 & 440
0.07
206

959 ± 66 Ma 0.070
206

500 0.065 400


MSWD = 0.8
Conc. Age = Conc. Age =
0.06 0.066 0.06
300
400 350
452.6 ± 5.8 Ma 0.055 452.5 ± 2.7 Ma
MSWD = 0.23, p. = 1.0 300 MSWD = 0.81, p.= 0.82
0.02 0.062 n = 25 n = 26
0.05 0.045
0.0 0.4 0.8 1.2 1.6 2.0 0.42 0.46 0.50 0.54 0.58 0.62 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2
207
Pb/235U 207
Pb/235U 207
Pb/235U 207
Pb/235U

Fig. 10. Wetherill Concordia diagrams of U-Pb isotope data of deformed amphibolites and other magmatic mafic gneisses of the Candelaria unit (errors on age calculations are 2σ, error ellipses are plotted at the 1σ level; MSWD = mean standard
weighted deviates of Concordia ages correspond to MSWD of concordance and equivalence; p. = probability of concordance and equivalence).

191
192 B. Weber et al. / Lithos 300–301 (2018) 177–199

(Fig. 10g), eight spots have younger, discordant ages suggesting meta- Sample R0401 is a hornblende-biotite gneiss from the type locality
morphic overprint and lead loss during a subsequent thermal event. of the Candelaria migmatite. In order to detect the protolith age, the
The remaining data are from inherited zircon cores or have mixed sample was taken from the paleosome, however, small neosome
ages. An upper intercept age is calculated at 1037 ± 93 Ma. veinlets cannot be excluded. The zircon grains vary from short to long
Sample R0103 is a banded biotite-hornblende gneiss (migmatite) prismatic and on CL images they display either zoned internal structure,
from the same outcrop. Zircon grains have different sizes and shapes, patchy recrystallization, or clearly defined and partly resorbed cores
different kinds of cores, sometimes with visible oscillatory zoning, with overgrowths (Fig. 11b). The high luminescent cores yield an
sometimes with patchy recrystallization, and in addition with upper Concordia intercept at 924 ± 85 Ma (Fig. 10b). Overgrowth
overgrowth rims (Fig. 11a). Nineteen out of 38 analyzed laser spots rims, long prismatic zoned zircon grains, and recrystallized zircon
yield a Concordia age of 449.5 ± 3.4 Ma (Fig. 10e). The zircon cores yield Ordovician ages with a Concordia age at 446.5 ± 3.8 Ma
have essentially Stenian-Tonian ages but some recrystallized cores (Fig. 10b). The former reflects the protolith age, the latter the time of
yield Ordovician or mixed ages. Besides, some zircon grains show anatexis.
additional lead loss or have rims yielding Permian ages (Fig. 11a) that Twenty-five zircon grains from amphibolite sample R0202 that was
reflect the time of intrusion of nearby Permian granitoids and later also chosen for metamorphic P-T estimates (see Section 4.2) were
stage metamorphic overprint. measured and they all together yield a Concordia age at 452.6 ±
Sample 1205-2 is a typical amphibolite from the deformed 5.8 Ma. Sector zoning and patchy luminescence of many irregular zircon
Candelaria Unit but with abundant zircon grains, which are mostly grains (Fig. 11b) suggests metamorphic zircon growth (Corfu et al.,
elliptic and sub-rounded, showing weak concentric zoning on CL images 2003). Other zircon grains are subrounded and most of the former
(Fig. 11d). All analyzed laser spots have Ordovician ages yielding a oscillatory zoning is erased. Thus, the age is most likely related to upper
Concordia age of 443.9 ± 2.9 Ma (Fig. 10h). amphibolite facies metamorphism and not to magmatic crystallization.

Fig. 11. CL images of typical zircon grains from amphibolites and other magmatic mafic gneisses. Numbers are 207Pb/206Pb ages of individual laser spots.
B. Weber et al. / Lithos 300–301 (2018) 177–199 193

Amphibolite R0206 is another sample used for geothermobarometry.


Twenty-six out of 35 analyzed laser spots yield a Concordia age at
(a)
452.5 ± 2.7 Ma (Fig. 10i). The Ordovician zircon grains display either 100
oscillatory zoning indicating igneous origin or zoning typical for meta-

Rock/Chondrite
morphic zircon (Fig. 11d). Inasmuch as there is no age difference, coex-
istence of both types of zircon (igneous and metamorphic) in the same
sample suggests the presence of anatectic veinlets of neosome in the
metamorphic rock. Only one grain is inherited from a ~ 1.0 Ga host 10
rock (Fig. 11d).
Finally, sample R0302 is another migmatic amphibolite of which mi-
croprobe data are presented in Section 4.2. The zircon grains are either this study E-MORB
small (b 50 μm) euhedral, short prismatic, bi-pyramidal with weak os- (1) N-MORB
cillatory zoning, or larger, long prismatic to sub-rounded and elliptic 1
grains, having high luminescent sectors (Fig. 11e). The small euhedral La Ce Pr Nd Pm Sm Eu Gd Tb Dy Ho Er Yb Lu
zircon grains yield a Concordia age at 449.9 ± 4.3 Ma, the large ones
have Stenian-Tonian ages and give an upper Concordia intercept age
at 959 ± 66 Ma (Fig. 10c). The small size and form of the Ordovician zir- 50 (b)
con grains suggest that they grew in the anatectic liquid of the quartz-
feldspar veinlets, whereas the larger Stenian-Tonian zircons are from

Rock/P Mantle
inherited components.
10
7. Geochemistry of amphibolite

To characterize the magmatic protoliths of the deformed mafic rocks


(mainly amphibolite) from the Candelaria unit, major and trace element
data of whole rock samples are presented together with data of similar
rocks reported by Estrada-Carmona et al. (2012, Appendix B, supple-
mentary file). In chondrite-normalized REE plots and multielement 1
plots (spider diagrams) normalized to primitive mantle (after Sun and Th Nb Ta La Ce Nd Zr Hf Sm Eu Ti Gd Y Yb
McDonough, 1989, Fig. 12) the chemical compositions of the amphibo-
lites are comparable to those from typical E-MORB's (evolved mid- Fig. 12. (a) Chrondrite normalized REE and (b) primitive mantle normalized trace element
ocean ridge basalt). The analytical procedures as well as considerations patterns of amphibolites from the Candelaria unit (normalization values and E-MORB and
on element mobility and additional classification and tectonic discrimi- MORB references after Sun and McDonough, 1989). (1) Data from Estrada-Carmona et al.
(2012).
nation diagrams are included in Appendix B (supplementary file).

8. Lu-Hf and Sm-Nd isotope systematics +7.2. Similarly the Nd isotope ratios yield εNd0 values between +1.2
and + 4.2 (Table 2). Recalculated to an assumed magmatic age of
Fig. 13a shows the Hf isotope ratios of all analyzed samples in a about 600 Ma, the resulting εHf(600 Ma) values are between + 6.6 and
176
Hf/177Hf vs. time isotope evolution diagram. Felsic orthogneiss and +11.8 and the εNd(600 Ma) values lie between +4.2 and +8.0, suggest-
anorthosite samples together with two zircon grains from the ing together with relatively high 147Sm/144Nd values (0.147–0.195,
orthogneiss analyzed by isotope dilution analyses have initial Hf isotope Fig. 13b) that all amphibolites originated from strongly to moderately
ratios between 0.282142 and 0.282308, corresponding to εHf(i) depleted sources. The most radiogenic values (sample 1205-2a)
between 0.0 and +5.5, mainly in the range of Hf-isotope ratios of typical plot near the depleted mantle evolution line at 600 Ma (Fig. 13a) and
anorthosite-mangerite-charnockite suite rocks from Oaxaquia (Fig. 13a, will plot far above it if recalculated to 1.0 Ga, which implies that
Weber et al., 2010). magmas were separated from a depleted mantle source during the
Of particular interest are the two samples from the banded Chipilín Neoproterozoic probably after Tonian times. The lower values of most
gneiss (see. Fig. 4d): whereas the initial isotope ratio of the whole amphibolite samples indicate also a significant amount of crustal
rock sample from the felsic gneiss band is indistinguishable from the contamination, which is consistent with an interpretation in terms of
Oaxaquia-type rocks, the whole rock sample from the mafic band is E-MORB magmas as suggested from the chemical data.
more radiogenic yielding εHf(919Ma) of + 6.2 and εNd(919Ma) of + 3.2. Using the isotopically most radiogenic amphibolite (1205-2a:
On the other hand, zircon grains from both samples (CH-02 and CH- εHf(600 Ma) = 11.8 and εNd(600Ma) = 8.0) as an endmember for the
03) that were analyzed by laser ablation MC-ICPMS have negative E-MORB magmatism and the average of felsic gneiss boudins as an
εHf(i) between − 3.2 and − 11.5 and depleted mantle model ages endmember for crustal contamination (εHf (600 Ma) = − 7.6 and
(two step) between 1.87 Ga and 2.18 Ga that are significantly older εNd(600 Ma) = − 4.0), the other analyzed amphibolites plot close to
than those from the whole rock samples (1.39 Ga and 1.59 Ga) or zircon a hypothetical mixing line on a εHf(600 Ma) vs. εNd(600 Ma) diagram
from typical Oaxaquia rocks. It is widely accepted that most of the Hf (Fig. 13c), suggesting mixing of not more than 6% of continental
from a rock is concentrated in zircon, for which reason this huge differ- crust to the magmas coming from the depleted source used for the
ences between whole rock and zircon is an uncommon feature. Howev- model.
er, both samples only contain little amount of small zircon crystals, The crustal contamination of a basaltic parental magma is further
suggesting that most Hf of the rock was originally hosted in other documented by inherited zircon grains that have significantly lower
176
phases like rutile, which is now mostly converted to titanite (see Hf/177Hf values compared to the whole rock: three dated zircon
Fig. 5e). The zircon grains, instead, are inherited from older continental grains from sample 1203-2c (numbered triangles in Fig. 13a) were ana-
rocks and, whereas the U-Pb system was reset by recrystallization lyzed for Lu-Hf isotopes with isotope dilution analyses. One ca. 1.0 Ga
around 920 Ma, the Hf-isotopic compositions remained unchanged. old zircon grain (#5) has a 176Hf/177Hf(996 Ma) of 0.282230, which is
The pre-Ordovician amphibolites have present-day 176Hf/177Hf within errors indistinguishable from an Ordovician zircon grain (#6)
values that correspond to positive εHf0 values between + 3.0 and yielding a 176Hf/177Hf(464 Ma) of 0.282243, indicating that this zircon
194 B. Weber et al. / Lithos 300–301 (2018) 177–199

0.28300 0.5130
(a) amphibolite (b)
Chipilin gneiss 0.5128
1205-2a anorthosite
0.28280 CH 0.5126
UR orthogneiss (boudin)

Nd/144Nd
(1)
2 Oaxaquia (zircon)
0.5124
Belize (det. zircon)(2)
0.28260 1203-2c

143
0.5122
amphibolite (3)
0.5120 amphibolite (4)
0.28240 De anorthosite (4)
Hf/177Hf

ple 0.5118
te 0.08 0.10 0.12 0.14 0.16 0.18 0.20
d
m 147
Sm/144Nd
0.28220
6 5 an
176

tle 14

10 (c)
0.28200
6

εHf(0.6 Ga)
2
0.28180
-2

y
rra
la
ria
-6

st
0.28160

re
r
Te
0 400 800 1200 1600 -10
-10 -6 -2 2 6 10
Time (Ma) εNd(0.6 Ga)

Fig. 13. (a) 176Hf/177Hf vs. time diagram, illustrating isotopic evolution of whole rock and zircon (triangles) samples from the Candelaria unit. Stippled red and blue lines correspond to
zircons from orthogneiss and amphibolite, respectively. (1) Zircon and whole rock isotope dilution data from Weber et al. (2010) and (2) detrital zircon isotope dilution data from
Weber et al. (2012). CHUR = the evolution lines of the chondritic uniform reservoir after Bouvier et al. (2008); DMM = depleted model mantle after Vervoort et al. (2000). – Filled
circles and squares are whole rock data, filled triangles are zircon data. (b) 147Sm/144Nd vs. 143Nd/144Nd diagram showing data from this work (symbols same as in a) and from
(3) Estrada-Carmona et al. (2012) and (4) Cisneros de León et al. (2017). (c) εNd vs. εHf diagram recalculated to an assumed age of 600 Ma for amphibolite protoliths. The terrestrial
array is as follows: εHf = 1.50 × εNd + 1.57 (Vervoort et al., 2011). A theoretical mixing line is included for a two-component mixture with the following parameters: endmember A
(depleted amphibolite 1205-2a) εNd(600 Ma) = +8.0; εHf(600 Ma) = +11.8; assuming 10 ppm Nd and 2 ppm Hf; endmember B (enriched basement boudins) εNd(600 Ma) = −4.1;
εHf(600 Ma) = −7.6; assuming 75 ppm Nd and 10 ppm Hf). For further explanations see text.

grain lost all its radiogenic lead during Ordovician metamorphism The results of U-Pb zircon dating from felsic orthogneiss samples
whereas the Hf-isotopes remained unaffected. The third analyzed zircon imply that at least part of the Candelaria unit represents ~1.0 Ga base-
grain (#2) is Ordovician, too, but its initial Hf-isotope ratio is higher ment. Besides that, Lu-Hf and Sm-Nd isotope data indicate that
(176Hf/177Hf(445 Ma) = 0.282699) and it plots close to the Hf-isotope orthogneiss from the Candelaria unit and massif-type anorthosites are
evolution line of the whole rock sample, suggesting that this zircon crys- isotopically similar to typical Oaxaquia rocks of Late Mesoproterozoic
tallized from the host rock probably during anatexis. age. The occurrence of massif-type anorthosite with economically im-
portant rutile-bearing ilmenitite lenses (Cisneros de León et al., 2017)
9. Discussion – although they could not be convincingly dated so far – leads us to hy-
pothesize that the gneisses of the Candelaria unit together with the
9.1. Evidence for latest Mesoproterozoic and Neoproterozoic basement massif-type anorthosite and associated rocks were formerly a Rodinia-
evolution type basement inlier similar to those defining Oaxaquia (e.g., Cameron
et al., 2004; Keppie et al., 2003; Schulze, 2011; Weber and Hecht, 2003).
Since Late Mesoproterozoic granulite facies basement was first de- Obviously, later metamorphic overprints destroyed possible former
scribed in the Guichicovi Complex west of the Tehuantepec Isthmus granulite facies mineral assemblages. However, rutile inclusions in titanite
(Murillo-Muñetón, 1994; Ruiz et al., 1999; Weber and Köhler, 1999), from the “Chipilín” gneiss together with clinopyroxene relicts observed in
its continuation towards the southeast into the Chiapas Massif was in- some samples from the Candelaria unit may suggest former granulite fa-
ferred mainly from inherited zircon in orthogneisses of the CMC and cies conditions. The same goes for the anorthosite that contains abundant
Sm-Nd depleted mantle model ages similar to those from the rutile that is mostly retrogressed to ilmenite and titanite. Exsolution of
Mesoproterozoic basement (Schaaf et al., 2002; Weber et al., 2005). baddeleyite and zircon from rutile also indicates high Zr-content and
However, medium- to high-grade Permo-Triassic metamorphism ham- therefore relatively high temperatures (Cisneros de León et al., 2017).
pered further interpretation of the basement rocks, for example, those Interestingly, some zircons of the orthogneiss boudins from the
from the Custepec unit in the central CMC (Estrada-Carmona et al., western study area display weak disturbance in the U-Pb isotope sys-
2009; Weber et al., 2007). Some Ordovician-Devonian cooling ages tem that is well-marked in sample 9511-9 from northeast of Belisario
(Ar-Ar hornblende and white mica) reported from the El Triunfo Domínguez, giving a second probability density maximum for
207
Complex, instead (Estrada-Carmona et al., 2012; Pompa-Mera, 2009), Pb/206Pb ages at 907 ± 20 Ma (see Fig. 9f). Most zircon grains from
indicate that the Permo-Triassic metamorphism did not exceed the banded “Chipilín” gneiss in the eastern study area yield a Concordia
greenschist facies and that the El Triunfo Complex was at higher crustal age at 911 ± 9 Ma, or an upper intercept age at 919 ± 13 Ma, and
level at this time. Therefore, the El Triunfo Complex is suitable for ~1.0 Ga zircon is absent. As CL images from these zircon grains indicate
looking further back into the geologic history. However, nothing is metamorphic origin, we interpret that the “Chipilín” gneiss displays a
known so far whether a major tectonic suture separates the El Triunfo mid-Tonian metamorphic overprint that affected other gneisses in the
Complex from the rest of the CMC or if the crustal level gets continuous- region to a minor extent. This leads us to suspect that this overprint is
ly deeper towards the NW. structurally controlled i.e., reactivation of former tectonic sutures.
B. Weber et al. / Lithos 300–301 (2018) 177–199 195

A Tonian metamorphic event around 920 Ma was yet not reported the gabbroic to granitic Motozintla suite intruded the El Triunfo
from Oaxaquia, although there are quite some lines of evidence like Complex between ~452 and ~446 Ma. Zircon data from a total of nine
(1) titanite from the Novillo gneiss yielding a 928 ± 2 Ma U-Pb age amphibolite, biotite-amphibole gneiss, and migmatite samples from
(Keppie et al., 2006) and (2) six garnet – whole rock ages between the El Triunfo Complex yield ages ranging from ~ 457 Ma to ~ 444 Ma
933 ± 6 and 911 ± 12 Ma from the Guichicovi Complex (Weber and (Fig. 10) indicating regional medium- to high-grade metamorphism
Köhler, 1999). It is important to note that widespread hydration of gran- and anatexis contemporaneous with crystallization of the
ulite to amphibolite in rocks from the northern Guichicovi Complex, abovementioned plutonic suite. These ages are obtained from zircon
was explained in terms of the lower crust position above a subduction overgrowth around inherited Stenian-Toninan cores or recrystallized
zone (Weber and Hecht, 2003). Although there is no clear evidence (metamorphic) zircon, and from oscillatory-zoned zircons, often both
for an active continental margin of Tonian age like arc-related igneous occurring within the same rock (Figs. 9e, 11). Whereas the former
rocks, in a global geodynamic context, the ca. 920 Ma metamorphic form either under subsolidus conditions (metamorphic overgrowth
overprint is best explained in terms of circum-Rodinia subduction that and recrystallization) or under supersolidus conditions (magmatic
started after the supercontinent Rodinia was assembled (e.g., Evans overgrowth) during anatexis, the latter are most likely from neosome
et al., 2016; Li et al., 2008, and references therein). Obviously, such a hy- veinlets crystallized from the anatectic melt that penetrated the
pothesis needs further proof but this is beyond the scope of this amphibolite.
contribution. In regional metamorphic terranes, partial melting can be triggered
Most of the amphibolite samples analyzed from the Candelaria unit by an externally derived H2O-rich fluid leading to fluid-induced melt-
also contain an inherited Stenian-Tonian zircon component. It remains ing, or by hydrate-breakdown melting involving dehydration reactions
unclear, however, to what extent the mafic rocks of the Candelaria and decomposition of a hydrous phase like muscovite, biotite or amphi-
unit belong to the Stenian-Tonian basement or to what extent they bole (e.g., Berger et al., 2008; Brown, 2013; Rubatto et al., 2009). Both
are E-MORB protoliths that intruded the crust subsequently by under- amphibole and biotite are present in most mafic rocks from the
plating, assimilating part of the Proterozoic lower crust. From the isoto- Candelaria unit. However, fluid-absent hydrate-breakdown melting is
pic point of view, like strongly positive initial εHf and εNd values, there not indicated because the neosome does not carry a peritectic anhy-
is evidence to suggest that the E-MORB amphibolites are younger than drous phase such as garnet or pyroxene that appears as a side product
the Rodinia-type basement. Indirect evidence for the age of this proba- of incongruent melting reactions (e.g., Vielzeuf et al., 1990). Instead,
bly rift-related magmatism comes from secondary zircon coronas biotite- and hornblende-bearing leucosome without anhydrous
around rutile and ilmenite found in anorthosite and related rocks that minerals in a biotite- and hornblende-bearing host suggest fluid-
yielded U-Pb dates around 600 Ma (Cisneros de León et al., 2017) and present melting (e.g., Brown, 2013). Additionally, the P-T estimates
a similar Ediacaran depositional age of the Jocote unit, into which from calcic amphiboles and the results from plagioclase-hornblende
these magmas probably intruded syn-sedimentarily (González- geothermometric calculations indicate peak temperatures not
Guzmán et al., 2016a). On the other hand, due to strong deformation above ~ 660 °C, which is insufficient for fluid absent melting.
and folding of these amphibolites, an Ordovician age can be discarded. Due to the fact that crystalline rocks of the middle and lower crust
have very low porosity, the amount of fluid necessary to produce signif-
9.2. Deformation and competence contrast icant amount of melt can only be achieved by an external H2O-rich fluid
source. Such fluids can be extracted from subsolidus dehydration reac-
At a first glance isoclinal D1 folding and prograde metamorphism tions of (1) pelitic sediments or (2) hydrated basaltic igneous rocks, or
suggest regional shortening in a convergent tectonic setting. However, (3) from crystallization fluids of nearby plutonic rocks. The Motozintla
folding in amphibolite and granulite facies rocks during rifting or oro- plutonic suite, which has arc-like chemical characteristics (González-
genic collapse, or accompanying pluton emplacement, is a common Guzmán, 2016), crystallized contemporaneously with the neosome
phenomenon at all scales (Harris et al., 2002). Boudins and pinch-and- and metamorphic zircon of the Candelaria unit. Therefore, it is a poten-
swell structures (mostly asymmetric) observed elsewhere in the tial source for crystallization fluids, whereas fluids from dehydration re-
Candelaria unit suggest extension during ductile and probably also actions from the Jocote metasedimentary rocks would have been
non-coaxial deformation. An extensional tectonic setting should there- consumed in situ by anatexis of these rocks themselves. Dehydration
fore be considered also for the formation of tight and isoclinal folds of hydrated basaltic and metasedimentary rocks from the subducting
(Harris et al., 2002). plate are also possible sources for external fluids. Thus, it is plausible
Interestingly, the observed boudin structures in the Candelaria unit to suggest that regional metamorphism that reached partial melting of
indicate that the quartz-feldspar gneiss is the competent layer and the the Candelaria unit in the late Ordovician is the result of subduction
amphibolite the incompetent matrix that deformed faster. Mafic rocks along a convergent continental margin and not rifting along a passive
are known to be more competent than felsic rocks in the upper crust. margin as proposed earlier (Estrada-Carmona et al., 2012).
At higher temperature of the lower crust, however, the viscosity of a
dry metagranite (charnockite) is getting higher than the viscosity of a 9.4. Considerations on paleogeography
metagabbro (Talbot, 1999; Talbot and Sokoutis, 1992).
These observations lead us to suspect that the Candelaria unit was The discovery of ~1.0 Ga outcrops together with massif-type anor-
affected by extension at middle to lower crustal level, followed by basal- thosite in the southern Chiapas Massif (Cisneros de León et al., 2017)
tic underplating, intrusion of magmas from the depleted mantle, uplift, imply Rodinia-type basement in the El Triunfo Complex. It remains un-
further extension and deposition of the Ediacaran Jocote unit known, however, if this basement continues into the central CMC that is
(González-Guzmán et al., 2016a) above this thinned Proterozoic crust strongly affected by Late Permian high-grade metamorphism or if a
prior to the Ordovician tectonothermal event. major tectonic suture separates the El Triunfo Complex from the rest
of the CMC. It is also likely that left-lateral displacement related to the
9.3. The Ordovician tectonothermal event branched Polochic-and Tonalá fault systems disrupted the El Triunfo
Complex into different discrete blocks that may occur elsewhere along
Besides Permian arc magmatism, the late Ordovician tectonothermal the southern border of the North American plate. However, the discov-
event is by far the most important in the El Triunfo Complex. González- ery of the 1.0 Ga basement of the El Triunfo Complex is an important
Guzmán (2016) defined this event as “Mochonian Orogeny” according second pinpoint besides the Guichicovi Complex (e.g., Weber and
to the local Mochó tribe native in the Motozintla to El Triunfo area. As re- Köhler, 1999) to favor Rodina-type basement probably for the entire
ported by Estrada-Carmona et al. (2012) and González-Guzmán (2016) southern Maya block and to consider it as part of Oaxaquia.
196 B. Weber et al. / Lithos 300–301 (2018) 177–199

Similar basement inliers occur in the Northern Andes (Cardona et al., Little is known about the breakup of Rodinia from the Mexican base-
2010; Cordani et al., 2005; Ramos, 2010) and along almost the ment except some alkaline basaltic dikes intruding the Stenian-Tonian
entire central and southern Andes (Ramos, 2010). Although some of Novillo gneiss in Tamaulipas (northeastern Mexico) at ~ 546 Ma that
the ~ 1.0 Ga basement inliers of Northern Andes are allochthonous were interpreted in terms of rift basalts comparable to their Late
and display significant Mesozoic right-lateral displacement, the Neoproterozoic counterpart in Baltica (Keppie et al., 2006; Pisarevsky
Mesoproterozoic Putumayo orogen, mostly buried beneath the Andean et al., 2008). The Jocote unit, recently proven to be of Ediacaran age
foreland is considered autochthonous with respect to northeastern (González-Guzmán et al., 2016a), is so far the only documented Late
Amazonia (Ibanez-Mejia et al., 2011). Recently ~1.0 Ga granulite facies Neoproterozoic rift basin deposit in southern Mexico. Together with
basement was also discovered from wells offshore NW Venezuela (1) the intrusion of E-MORB magmas into both Candelaria and Jocote
(Baquero et al., 2015), suggesting that similar basement inliers are not units, (2) structural features that suggest crustal extension prior to
local occurrences with exotic character only. They rather represent an Ordovician magmatism, as well as (3) indirect age evidence from
orogenic belt extending along NW South America from Venezuela to ~ 600 Ma old exsolved zircon detected in anorthosite (Cisneros de
Colombia, Ecuador, and Peru including Oaxaquia, which formed by in- León et al., 2017), the geologic record of the El Triunfo Complex can
teraction of NW Amazonia with Baltica in the Late Mesoproterozic, dur- now be considered to include Rodinia breakup during the Late
ing assemblage of the Rodinia supercontinent (Fig. 14a, Cardona et al., Neoproterozoic (Fig. 14b).
2010; Ibanez-Mejia et al., 2011; Li et al., 2008; Weber et al., 2010; High-grade metamorphism and anatexis triggered by an external
Weber and Schulze, 2014). It is widely accepted that supercontinent as- fluid, implies a convergent margin setting for the El Triunfo Complex
semblage is followed by circum-supercontinent subduction (e.g., Evans during the Late Ordovician. The new evidence presented in this
et al., 2016, and references therein), of which the ca. 920 Ma metamor- paper contrasts with interpretations where Oaxaquia including the
phic zircon-growth or recrystallization observed in zircons from the Maya block and the Acatlán Complex are placed on the southern
Chipilín gneiss possibly is an evidence. Such a scenario requires a paleo- margin of the Rheic Ocean in Ordovician times (Estrada-Carmona
geographic position of this portion of Oaxaquia close to the outer limit of et al., 2012; Keppie, 2004; Keppie et al., 2008; Morales-Gamez
Rodinia (Fig. 14a) as suggested earlier also from similar garnet ages as et al., 2008).
well as hydration and regression observed in the Guichicovi Complex The Maya block has for long been correlated with the Mérida Andes
(Weber and Hecht, 2003). of Venezuela (Bellizzia and Pimentel, 1994). Van der Lelij et al. (2016)

(a) SP (b) SP

Oax WA
BA BA WA

AM SF
Oax

G AM
C
R
LAU
ca. K LAU
ca. 600 Ma
920 Ma

(c) AMCG suites


Mesoproterozoic arc terranes
LAU Rheic Rodinia Orogens (Grenville Belts)
SP
Ocean Rodinia-type peri-Gondwanan terranes
WA Paleozoic Gondwana margin
F
n Y Avalonian terranes
ea MA
s Oc Ac S Southern Chiapas Massif Complex
tu
Oax

Pu
e (El Triunfo Complex)
Famatinia

Iap G AM
Plate movement
At
na

ca.
rc

450 Ma Ar Sunsas

Fig. 14. Neoproterozoic to Ordovician geotectonic reconstruction of western Rodinia and Gondwana (modified and reinterpreted considering previous reconstructions after Murphy et al.,
2004; Li et al., 2008; Pisarevsky et al., 2008; Keppie et al., 2012; Torsvik et al., 2012; Van der Lelij et al., 2016; Weber and Schulze, 2014; Weber et al., 2012). Note, the Amazon craton is
shown in its present-day orientation within South America with the South Pole (SP) towards the top of the image. (a) Tonian: Rodinia established, Oaxaquia sitting between Amazonia and
Baltica facing the ocean and possible subduction towards the south. Abbreviations: AM = Amazonia, BA = Baltica-Sveconorwegian, C = Congo, G = Greenland, LAU = Laurentia, Oax =
Oaxaquia, R = Rio de la Plata, SF = São Francisco, WA = West Africa. (b) Ediacaran. Incipient rifting between Laurentia, Baltica, and Amazonia/Oaxaquia with a superplume beneath the
triple junction. (c) Late Ordovician. Famatinian active continental margin facing the Iapetus Ocean extents along NW Amazonia to Central Venezuela. Unlike previous models, opening of
the Rheic Ocean and separation of Avalonian type terranes from the Gondwana margin occurred in the Cambrian to Early Ordovician further northeast not involving Oaxaquia and other
peri-Gondwanan terranes of the NW Amazonia realm. Abbreviations: Ac = Acatlán Complex, Ar = Arequipa, At = Antofalla, CA = Colombian Andes, Ch = Chortís Block, F = Florida, G =
Garzón Massif, MA = Merida Andes, Oax = Oaxaquia, Pu = Putumayo, Y = Yucatán.
B. Weber et al. / Lithos 300–301 (2018) 177–199 197

reported new evidence from the Mérida Andes and the Santander Mas- References
sif of Colombia indicating that they represent an Early Paleozoic active
Authemayou, C., Brocard, G., Teyssier, C., Simon-Labric, T., Guttiérrez, A., Chiquín, E.,
margin, which lasted from ~ 500 Ma to ~ 415 Ma coeval with the Morán, S., 2011. The Caribbean–North America–Cocos Triple Junction and the
Famatinian arc and related intrusions in Peru (Bahlburg et al., 2009). Al- dynamics of the Polochic–Motagua fault systems: pull-up and zipper models. Tectonics
though there is some difference in the interpretation of timing, over- 30. https://doi.org/10.1029/2010TC002814.
Bahlburg, H., Vervoort, J.D., Du Frane, S.A., Bock, B., Augustsson, C., Reimann, C., 2009.
whelming accordance is found between Mérida Andes and Santader Timing of crust formation and recycling in accretionary orogens: insights learned
Massif of NW South America (Van der Lelij et al., 2016) and the southern from the western margin of South America. Earth Science Reviews 97:215–241.
Maya block (Chiapas, Guatemala, Belize) like (1) Rodinia-type base- https://doi.org/10.1016/j.earscirev.2009.10.006.
Baquero, M., Grande, S., Urbani, F., Hall, C., Armstrong, R., 2015. New evidence for Putumayo
ment inliers, (2) early Ordovician to late Silurian igneous rocks, crust in the basement of the Falcón Basin and Guajira Peninsula, northwestern
(3) Barrovian medium to high-grade metamorphism, (4) anatexis of Venezuela. In: Bertolini, C., Mann, P. (Eds.), Petroleum Geology and Potential
Grenvillian orthogneiss at ~ 450 Ma (comparable with Micarache Hydrocarbon of the Colombia-Venezuela Caribbean Margin. AAPG Memoir 108.
Bellizzia, A., Pimentel, N., 1994. Terreno Mérida: un cinturón alóctono Herciniano en la
orthogneiss of the Mérida Andes protolith age at 1009 ± 9 Ma and
Cordillera de Los Andes de Venezuela. V Simp. Boliv. Explor. Pet. En Las Cuencas
anatexis at 454 ± 10 Ma, Van der Lelij et al., 2016), and (5) Permo- Subandinas Mem., pp. 271–299
Triassic arc magmatism. Berger, A., Burri, T., Alt-Epping, P., Engi, M., 2008. Tectonically controlled fluid flow and
That is why we place the southern Maya block, along with Oaxaquia, water-assisted melting in the middle crust: an example from the Central Alps. Lithos,
Granites and Migmatites: Their Temporal, Spatial and Causal Relationships Granites
Chortís, and the Acatlán Complex at the NW Gondwana margin that is and Migmatites. European Geosciences Union General Assembly 102:pp. 598–615.
interpreted to be the extension of the Famatinian arc (Fig. 14c) in agree- https://doi.org/10.1016/j.lithos.2007.07.027.
ment with the model of Van der Lelij et al. (2016). Similar to NW South Black, L., Kamo, S., Allen, C., Davis, D., Aleinikoff, J., Valley, J., Mundil, R., Campbell, I., Korsch,
R., Williams, I., Foudoulis, C., 2004. Improved 206Pb/238U microprobe geochronology
America, magmatism in the southern Maya block ceased in the late Silu- by the monitoring of a trace-element-related matrix effect; SHRIMP, 8 ID–TIMS,
rian to earliest Devonian (Martens et al., 2010; Steiner and Walker, ELA–ICP–MS and oxygen isotope documentation for a series of zircon standards.
1996; Weber et al., 2012). Chemical Geology 205, 115–140.
Boelrijk, N., 1968. A general formula for “double” isotope dilution analysis. Chemical
Geology 3, 323–325.
10. Conclusions Bouvier, A., Vervoort, J.D., Patchett, P.J., 2008. The Lu–Hf and Sm–Nd isotopic composition
of CHUR: constraints from unequilibrated chondrites and implications for the bulk
composition of terrestrial planets. Earth and Planetary Science Letters 273, 48–57.
1. Metaigneous rocks of the El Triunfo Complex, namely the
Brown, M., 2013. Granite: from genesis to emplacement. Geological Society of America
Candelaria unit, contain orthogneisses of late Stenian to early Bulletin 125:1079–1113. https://doi.org/10.1130/B30877.1.
Tonian age (~ 1.0 Ga) indicating, together with Nd and Hf isotope Burkart, B., 1983. Neogene North American–Caribbean Plate boundary across northern
central America: offset along the Polochic Fault. Tectonophysics 99, 251–270.
evidence, that the basement of the southern Chiapas Massif is re-
Burkart, B., Self, S., 1985. Extension and rotation of crustal blocks in northern Central
lated to Rodinia assemblage similar to Oaxaquia or other Precam- America and effect on the volcanic arc. Geology 13, 22–26.
brian inliers along the northwestern and western margin of Cameron, K.L., Lopez, R., Ortega-Gutiérrez, F., Solari, L.A., Keppie, J.D., Schulze, C.,
Amazonia. 2004. U-Pb geochronology and Pb isotopic compositions of leached feldspars:
constraints on the origin and evolution of Grenville rocks from eastern and southern
2. A locally observed ~ 920 Ma metamorphic overprint is possibly Mexico. Geological Society of America Memoirs 197:755–769. https://doi.org/
related to subsequent circum-Rodinia subduction. 10.1130/0-8137-1197-5.755.
3. The El Triunfo Complex was intruded by E-MORB magmas that Campa, M.F., Coney, P.J., 1983. Tectono-stratigraphic terranes and mineral resource
distributions in Mexico. Canadian Journal of Earth Sciences 20, 1040–1051.
differ by their Nd and Hf isotopic composition from the gneisses, Cardona, A., Chew, D., Valencia, V.A., Bayona, G., Mišković, A., Ibañez-Mejía, M., 2010.
suggesting a rift-related event during Rodinia breakup and open- Grenvillian remnants in the Northern Andes: Rodinian and Phanerozoic paleogeo-
ing of the Iapetus Ocean probably contemporaneous with the de- graphic perspectives. Journal of South American Earth Sciences 29:92–104. https://
doi.org/10.1016/j.jsames.2009.07.011.
position of the Jocote unit above thinned continental crust during Centeno-García, E., Keppie, J.D., 1999. Latest Paleozoic-early Mesozoic structures in the
the Ediacaran. central Oaxaca Terrane of southern Mexico: deformation near a triple junction.
4. Medium- to high-grade metamorphism and anatexis with peak con- Tectonophysics 301, 231–242.
Chang, Z., Vervoort, J.D., McClelland, W.C., Knaack, C., 2006. U-Pb dating of zircon by
ditions at ~ 650 °C and ~ 6 kbar affected the El Triunfo Complex at LA-ICP-MS. Geochemistry, Geophysics, Geosystems 7. https://doi.org/10.1029/
~450 Ma contemporaneous with the intrusion of the granitic to gab- 2005GC001100.
broic Motozintla plutonic suite. This major regional tectonothermal Chu, Z.Y., Yang, Y.H., Jinghui, G., Qiao, G.S., 2011. Calculation methods for direct internal mass
fractionation correction of spiked isotopic ratios from multi-collector mass spectromet-
event is interpreted in terms of an active continental margin such
ric measurements. International Journal of Mass Spectrometry 299, 87–93.
as the Famatinian arc. Cisneros de León, A., Weber, B., Ortega-Gutiérrez, F., González, R., Maldonado, R., Solari, L.,
Schaaf, P., Manjarrez-Juárez, R., 2017. Grenvillian massif-type anorthosite suite in
Supplementary data to this article can be found online at https://doi. Chiapas, Mexico: magmatic to polymetamorphic evolution of anorthosites and their
Ti-Fe ores. Precambrian Research 295, 203–226.
org/10.1016/j.lithos.2017.12.009. Cordani, U.G., Cardona, A., Jimenez, D.M., Liu, D., Nutman, A.P., 2005. Geochronology of
Proterozoic basement inliers in the Colombian Andes: tectonic history of remnants
Acknowledgements of a fragmented Grenville belt. Geological Society of London, Special Publication
246:329–346. https://doi.org/10.1144/GSL.SP.2005.246.01.13.
Corfu, F., Hanchar, J.M., Hoskin, P.W.O., Kinny, P., 2003. Atlas of zircon textures. Reviews in
This contribution was supported by Consejo Nacional de Ciencia y Mineralogy and Geochemistry 53:469–500. https://doi.org/10.2113/0530469.
Tecnología (CONACYT, Convocatoria Ciencia Básica 2012, project Damon, P.E., Shafiqullah, M., Clark, K., 1981. Age trends of igneous activity in relation to
metallogenesis in the southern Cordillera. In: Dickinson, W. (Ed.), Relations of
#180588). We would like to thank Sergio Padilla-Ramírez Centro Tectonics to Ore Deposits in the Southern Cordillera. Arizona Geological Society
de Investigación Científica y de Educación Superior de Ensenada Digest 14, pp. 137–153.
B.C. (CICESE), México, for running the clean lab, Susana Rosas- Dengo, G., 1969. Problems of tectonic relations between Central America and the
Caribbean. Gulf Coast Association of Geological Societies Transactions 19, 311–320.
Montoya, Victor Pérez-Arroyo, and Gabriel Rendón-Márquez (all DePaolo, D.J., 1981. Neodymium isotopes in the Colorado Front Range and crust-mantle
CICESE) for their help with sample preparation and to Luis evolution in the Proterozoic. Nature 291, 193–196.
Gradilla-Martínez (CICESE) for performing CL images. Many thanks Ernst, W., Liu, J., 1998. Experimental phase-equilibrium study of Al- and Ti-contents of
calcic amphibole in MORB - a semiquantitative thermobarometer. American Mineral-
to Peter Schaaf and Gerardo Arrieta, Laboratorio Universitario de
ogist 83, 952–969.
Geología Isotópica (UNAM) for assistance running TIMS. Many Estrada-Carmona, J., Weber, B., Hecht, L., Martens, U., 2009. P-T-t trajectory of metamor-
thanks go to the reviewers Mauricio Ibanez-Mejia (Rochester, NY) phic rocks from the central Chiapas Massif Complex: the Custepec Unit, Chiapas,
and Roelant van der Lelij (Geological Survey of Norway) and to Lith- Mexico. Revista Mexicana de Ciencias Geológicas 26, 143–259.
Estrada-Carmona, J., Weber, B., Martens, U., López-Martínez, M., 2012. Petrogenesis of Or-
os editor Marco Scambelluri for their positive and constructive revi- dovician magmatic rocks in the southern Chiapas Massif Complex: relations with the
sions of the manuscript. This is a contribution to IGCP 648 early Palaeozoic magmatic belts of northwestern Gondwana. International Geology
“Supercontinent Cycles and Global Geodynamics”. Review 54, 1918–1943.
198 B. Weber et al. / Lithos 300–301 (2018) 177–199

Evans, D.A.D., Li, Z.X., Murphy, J.B., 2016. Four-dimensional context of Earth's history su- Leake, B.E., Woolley, A.R., Arps, C.E., Birch, W.D., Gilbert, M.C., Grice, J.D., Hawthorne, F.C.,
percontinents. In: Li, Z.X., Evans, D.A.D., Murphy, J.B. (Eds.), Supercontinent Cycles Kato, A., Kisch, H.J., Krivovichev, V.G., 1997. Nomenclature of Amphiboles: report of
through Erath History. Geological Society, London, Special Publ 424, pp. 1–14. the Subcommittee on Amphiboles of the International Mineralogical Association
Fisher, C.M., Vervoort, J.D., DuFrane, S.A., 2014a. Accurate Hf isotope determinations of Commission on New Minerals and Mineral Names. Mineral. Mag. 61, 295–321.
complex zircons using the “laser ablation split stream” method. Geochemistry, Li, Z.X., Bogdanova, S.V., Collins, A.S., Davidson, A., De Waele, B., Ernst, R.E., Fitzsimons,
Geophysics, Geosystems (G3) 15, 121–139. I.C.W., Fuck, R.A., Gladkochub, D.P., Jacobs, J., Karlstrom, K.E., Lu, S., Natapov, L.M.,
Fisher, C.M., Vervoort, J.D., Hanchar, J.M., 2014b. Guidelines for reporting zircon Hf isoto- Pease, V., Pisarevsky, S.A., Thrane, K., Vernikovsky, V., 2008. Assembly, configuration,
pic data by LA-MC-ICPMS and potential pitfalls in the interpretation of these data. and break-up history of Rodinia: a synthesis. Precambrian Research 160:179–210.
Chemical Geology 363, 125–133. https://doi.org/10.1016/j.precamres.2007.04.021.
French, C.D., Schenk, C.J., 1997. Map showing geology, oil and gas fields, and Geologic Ludwig, K.R., 2012. Isoplot 3.75. Berkeley Geochronol. Cent. Spec. Publ 5 p. 75.
Provinces of the Caribbean Region. US Geological Survey Open File Report 97-470-K. Martens, U., Weber, B., Valencia, V.A., 2010. U/Pb geochronology of Devonian and older
Gehrels, G.E., Valencia, V.A., Ruiz, J., 2008. Enhanced precision, accuracy, efficiency, and Paleozoic beds in the southeastern Maya block, Central America: its affinity with
spatial resolution of U-Pb ages by laser ablation-multicollector-inductively coupled peri-Gondwanan terranes. Geological Society of America Bulletin 122:815–829.
plasma-mass spectrometry. Geochemistry, Geophysics, Geosystems 9. https:// https://doi.org/10.1130/B26405.1.
doi.org/10.1029/2007gc001805. Martens, U.C., Brueckner, H.K., Mattinson, C.G., Liou, J.G., Wooden, J.L., 2012. Timing of
Gerya, T., Perchuk, L., Triboulet, C., Audren, C., Sez'ko, A., 1997. Petrology of the Tumanshet eclogite-facies metamorphism of the Chuacús complex, Central Guatemala: record
zonal metamorphic complex, eastern Sayan. Petrology 5, 503–533. of Late Cretaceous continental subduction of North America's sialic basement. Lithos
González-Guzmán, R., 2016. Estudio petrogenético del basamento cristalino de la porción 146–147:1–10. https://doi.org/10.1016/j.lithos.2012.04.021.
sureste del Macizo de Chiapas: Implicaciones tectónicas del Bloque Maya Sur. (Doc- Martínez-Amador, H., Rosendo-Brito, B., Fitz-Bravo, C., Tinajera-Fuentes, E., and Beltrán-
toral Thesis). CICESE, Ensenada, México, p. 278. Castillo, H.D. 2005. Carta Geológica-Minera Tuxtla Gutiérrez, Chiapas E5-11, 1:
González-Guzmán, R., Weber, B., Manjarrez-Juárez, R., Cisneros de León, A., Hecht, L., 250,000. Pachuca, Hgo., Servicio Geológico Mexicano.
Herguera-García, J.C., 2016a. Provenance, age constraints and metamorphism of Massonne, H.-J., Schreyer, W., 1987. Phengite geobarometry based on the limiting
Ediacaran metasedimentary rocks from the El Triunfo Complex (SE Chiapas, assemblage with K-feldspar, phlogopite, and quartz. Contributions to Mineralogy
Mexico): evidence for Rodinia breakup and Iapetus active margin. International and Petrology 96, 212–224.
Geology Review 58, 2065–2091. Molina-Garza, R.S., Geissman, J.W., Wawrzyniec, T.F., Alonso, T.A.P., Iriondo, A., Weber, B.,
González-Guzmán, R., Weber, B., Tazzo-Rangel, M.D., Solari, L., 2016b. Validation of diges- Aranda-Gómez, J., 2015. Geology of the coastal Chiapas (Mexico) Miocene plutons
tion and element separation methods and a new data reduction program and the Tonalá shear zone: Syntectonic emplacement and rapid exhumation during
(IsotopeHf®) for Lu-Hf isotope dilution analysis by MC-ICP-MS. Revista Mexicana sinistral transpression. Lithosphere L409:1. https://doi.org/10.1130/L409.1.
de Ciencias Geológicas 33, 254–269. Morales-Gamez, M., Keppie, J.D., Norman, M., 2008. Ordovician-Silurian rift-passive mar-
Guzmán-Speziale, M., Meneses-Rocha, J.J., 2000. The North America–Caribbean plate gin on the Mexican margin of the Rheic Ocean overlain by Carboniferous-Permian
boundary west of the Motagua–Polochic fault system: a fault jog in Southeastern periarc rocks: evidence from the eastern Acatlan Complex, southern Mexico.
Mexico. Journal of South American Earth Sciences 13:459–468. https://doi.org/ Tectonophysics 461:291–310. https://doi.org/10.1016/j.tecto.2008.01.014.
10.1016/S0895-9811(00)00036-5. Muehlberger, W.R., Ritchie, A.W., 1975. Caribbean-Americas plate boundary in Guatemala
Harris, L.B., Koyi, H.A., Fossen, H., 2002. Mechanisms for folding of high-grade rocks in ex- and southern Mexico as seen on Skylab IV orbital photography. Geology 3:232–235.
tensional tectonic settings. Earth Science Reviews 59:163–210. https://doi.org/ https://doi.org/10.1130/0091-7613(1975)3b232:cpbigaN2.0.co;2.
10.1016/S0012-8252(02)00074-0. Murillo-Muñetón, G., 1994. Petrologic and Geochronologic Study of Grenville-Age
Hiller, R., Weber, B., Hecht, L., Ortega-Gutierrez, F., Schaaf, P., López-Martínez, M., 2004. Granulites and Post-granulite Plutons From the la Mixtequita Area, State of Oaxaca
The Sepultura unit – a medium to high grade metasedimentary sequence in the in Southern Mexico, and Their Tectonic Significance. (Masters Thesis). University of
Chiapas Massif, SE Mexico. Reunión Nac. Cienc. Tierra Querétaro México Libro Southern California, Los Angeles.
Resumenes, p. 200. Murphy, J.B., Pisarevsky, S.A., Nance, R.D., Keppie, J.D., 2004. Neoproterozoic–Early
Holland, T., Blundy, J., 1994. Non-ideal interactions in calcic amphiboles and their bearing Paleozoic evolution of peri-Gondwanan terranes: implications for Laurentia-
on amphibole-plagioclase thermometry. Contributions to Mineralogy and Petrology Gondwana connections. International Journal of Earth Sciences 93:659–682. https://
116, 433–447. doi.org/10.1007/s00531-004-0412-9.
Ibanez-Mejia, M., Ruiz, J., Valencia, V.a., Cardona, A., Gehrels, G.E., Mora, A.R., 2011. The Nance, R.D., Murphy, J.B., Strachan, R.O.B.A., Keppie, J.D., Gutie, G., Rez, J.F.N., Quesada, C.,
Putumayo Orogen of Amazonia and its implications for Rodinia reconstructions: Linnemann, U.L.F., Lemos, R.D., Pisarevsky, S.A., Dresden, D., 2008. Neoproterozoic –
new U–Pb geochronological insights into the Proterozoic tectonic evolution of north- Early Palaeozoic Tectonostratigraphy and Palaeogeography of the Peri-Gondwanan
western South America. Precambrian Research 191:58–77. https://doi.org/10.1016/ Terranes: Amazonian Vs. West African Connections. pp. 345–383.
j.precamres.2011.09.005. Nebel-Jacobsen, Y., Scherer, E.E., Münker, C., Mezger, K., 2005. Separation of U, Pb, Lu, and
Jiménez-Hernández, A., Jaimez-Fuentes, A., Motolinía-García, O., Pinzón-Salazar, T., and Hf from single zircons for combined U–Pb dating and Hf isotope measurements by
Membrillo-Ortega, H., 2005. Carta Geológica-Minera Huixtla D15-2 Chiapas, 1: TIMS and MC-ICPMS. Chemical Geology 220, 105–120.
250,000. Servicio Geológico Mexicano, Pachuca, Hgo., México. Ortega-Gutierrez, F., Mitre-Salazar, L.M., Moran-Centeno, D.J., Alaniz-Álvarez, S.A., Nieto-
Johnston, S., Gehrels, G., Valencia, V., Ruiz, J., 2009. Small-volume U–Pb zircon geochronol- Samaniego, Á.F., 1992. Carta geológica de la República Mexicana.
ogy by laser ablation-multicollector-ICP-MS. Chemical Geology 259:218–229. https:// Ortega-Gutierrez, F., Ruiz, J., Centeno-Garcia, E., 1995. Oaxaquia, a Proterozoic
doi.org/10.1016/j.chemgeo.2008.11.004. microcontinent accreted to North America during the late Paleozoic. Geology 23:
Keppie, J.D., 2004. Terranes of Mexico revisited: a 1.3 billion year odyssey. International 1127–1130. https://doi.org/10.1130/0091-7613(1995)023b1127:OAPMATN2.3.CO;2.
Geology Review 46, 765–794. Ortega-Gutierrez, F., Solari, L.A., Ortega-Obregon, C., Elias-Herrera, M., Martens, U.,
Keppie, J.D., Ortega-Gutierrez, F., 1999. Middle American Precambrian basement: a Moran-Ical, S., Chiquin, M., Keppie, J.D., De Leon, R.T., Schaaf, P., 2007. The Maya-
missing piece of the reconstructed 1-Ga orogen. Geological Society of America Special Chortis boundary: a tectonostratigraphic approach. International Geology Review
Papers 336:199–210. https://doi.org/10.1130/0-8137-2336-1.199. 49, 996–1024.
Keppie, J.D., Ortega-Gutiérrez, F., 2010. 1.3–0.9 Ga Oaxaquia (Mexico): remnant of an arc/ Ortega-Obregón, C., Solari, L.A., Keppie, J.D., Ortega-Gutiérrez, F., Solé, J., Morán-Ical, S.,
backarc on the northern margin of Amazonia. Journal of South American Earth 2008. Middle–Late Ordovician magmatism and Late Cretaceous collision in the south-
Sciences 29:21–27. https://doi.org/10.1016/j.jsames.2009.07.001. ern Maya block, Rabinal-Salamá area, central Guatemala: implications for North
Keppie, J.D., Ramos, V.A., 1999. Odyssey of terranes in the Iapetus and Rheic oceans during America-Caribbean plate tectonics. Geological Society of America Bulletin 120:
the Paleozoic. Geological Society of America Special Papers 336:267–276. https:// 556–570. https://doi.org/10.1130/b26238.1.
doi.org/10.1130/0-8137-2336-1.267. Paces, J.B., Miller, J.D., 1993. Precise U-Pb ages of Duluth Complex and related mafic intru-
Keppie, J.D., Dostal, J., Cameron, K.L., Solari, L.A., Ortega-Gutierrez, F., Lopez, R., 2003. sions, northeastern Minnesota: Geochronological insights to physical, petrogenetic,
Geochronology and geochemistry of Grenvillian igneous suites in the northern paleomagnetic, and tectonomagmatic processes associated with the 1.1 Ga
Oaxacan Complex, southern Mexico: tectonic implications. Precambrian Research midcontinent rift system. Journal of Geophysical Research 98:13997–14013.
120, 365–389. https://doi.org/10.1029/93JB01159.
Keppie, J.D., Dostal, J., Nance, R.D., Miller, B.V., Ortega-Rivera, A., Lee, J.K.W., 2006. Circa Pindell, J.L., Kennan, L., 2009. Tectonic evolution of the Gulf of Mexico, Caribbean and
546 Ma plume-related dykes in the ∼1 Ga Novillo Gneiss (east-central Mexico): northern South America in the mantle reference frame: an update. Geological Society
evidence for the initial separation of Avalonia. Precambrian Research 147:342–353. of London, Special Publication 328:1–55. https://doi.org/10.1144/SP328.1.
https://doi.org/10.1016/j.precamres.2006.01.020. Pisarevsky, S.A., Murphy, J.B., Cawood, P.A., Collins, A.S., 2008. Late Neoproterozoic and
Keppie, J.D., Dostal, J., Murphy, J.B., Nance, R.D., 2008. Synthesis and tectonic interpreta- Early Cambrian palaeogeography: models and problems. Geological Society of
tion of the westernmost Paleozoic Variscan orogen in southern Mexico: from rifted London, Special Publication 294:9–31. https://doi.org/10.1144/sp294.2.
Rheic margin to active Pacific margin. Tectonophysics 461:277–290. https://doi.org/ Pompa-Mera, V., 2009. Geoquímica y Geocronología de los Complejos Intrusivos en el
10.1016/j.tecto.2008.01.012. sureste de Chiapas, México. (Masters Thesis). Universidad Nacional Autónoma de
Keppie, J.D., Dostal, J., Murphy, J.B., Galaz-Escanilla, G., Ramos-Arias, M.a., Nance, R.D., México (160 pp.).
2012. High pressure rocks of the Acatlán Complex, southern Mexico: large-scale Raase, P., Raith, M., Ackermand, D., Lal, R., 1986. Progressive metamorphism of mafic rocks
subducted Ordovician rifted passive margin extruded into the upper plate during from greenschist to granulite facies in the Dharwar Craton of South India. Journal of
the Devonian–carboniferous. Tectonophysics 560–561:1–21. https://doi.org/ Geology 94, 261–282.
10.1016/j.tecto.2012.06.015. Ramos, V.A., 2010. The Grenville-age basement of the Andes. Journal of South American
Landing, E., Westrop, S.R., Keppie, J.D., 2007. Terminal Cambrian and lowest Ordovician Earth Sciences 29:77–91. https://doi.org/10.1016/j.jsames.2009.09.004.
succession of Mexican West Gondwana: biotas and sequence stratigraphy of the Ratschbacher, L., Franz, L., Min, M., Bachmann, R., Martens, U., Stanek, K., Stubner, K.,
Tiñu Formation. Geological Magazine 144:909–936. https://doi.org/10.1017/ Nelson, B.K., Herrmann, U., Weber, B., Lopez-Martinez, M., Jonckheere, R., Sperner,
S0016756807003585. B., Tichomirowa, M., Mcwilliams, M.O., Gordon, M., Meschede, M., Bock, P., 2009.
B. Weber et al. / Lithos 300–301 (2018) 177–199 199

The North American-Caribbean Plate boundary in Mexico-Guatemala-Honduras. Sun, S., McDonough, W., 1989. Chemical and isotopic systematics of oceanic basalts:
Geological Society of London, Special Publication 328:219–293. https://doi.org/ implications for mantle composition and processes. Geological Society, London,
10.1144/SP328.11. Special Publications 42, 313–345.
Rubatto, D., Hermann, J., Berger, A., Engi, M., 2009. Protracted fluid-induced melting dur- Talbot, C.J., 1999. Can field data constrain rock viscosities? Journal of Structural Geology
ing Barrovian metamorphism in the Central Alps. Contributions to Mineralogy and 21, 949–957.
Petrology 158:703–722. https://doi.org/10.1007/s00410-009-0406-5. Talbot, C.J., Sokoutis, D., 1992. The importance of incompetence. Geology 20:951–953.
Ruiz, J., Tosdal, R.M., Restrepo, P.A., Murillo-Muñetón, G., 1999. Pb isotope evidence https://doi.org/10.1130/0091-7613(1992)020b0951:TIOIN2.3.CO;2.
for Colombia-southern Mexico connections in the Proterozoic. Geological Society Thomson, M.L., Bernett, R.L., Fleet, M.E., Kerrich, R., 1985. Metamorphic assemblages in
of America Special Papers 336:183–197. https://doi.org/10.1130/0-8137-2336- South-Range norite and footwall mafic rocks near Kirkwood mine, Sudbury, Ontario.
1.183. Canadian Mineralogist 23, 173–186.
Schaaf, P., Weber, B., Weis, P., Groβ, A., Ortega-Gutierrez, F., Kohler, H., 2002. The Chiapas Torsvik, T.H., Van der Voo, R., Preeden, U., Mac Niocaill, C., Steinberger, B., Doubrovine,
Massif (Mexico) revised: New geologic and isotopic data and basement characteris- P.V., van Hinsbergen, D.J.J., Domeier, M., Gaina, C., Tohver, E., Meert, J.G.,
tics. In: Miller, H. (Ed.), Contributions to Latin-American Geology. Neues Jahrbuch McCausland, P.J.A., Cocks, L.R.M., 2012. Phanerozoic polar wander, palaeogeography
Geologie Paläontologie Abhandlungen 225, pp. 1–23. and dynamics. Earth Science Reviews 114:325–368. https://doi.org/10.1016/
Scherer, E., Münker, C., Rehkämper, M., Mezger, K., 1999. Improved precision of Lu isotope j.earscirev.2012.06.007.
dilution measurements by MC-ICP-MS and application to Lu-Hf geochronology. Eos Van der Lelij, R., Spikings, R., Ulianov, A., Chiaradia, M., Mora, A., 2016. Palaeozoic to
(Suppl. 80), 1118. Early Jurassic history of the northwestern corner of Gondwana, and implications for
Scherer, E., Münker, C., Mezger, K., 2001. Calibration of the lutetium–hafnium clock. Sci- the evolution of the Iapetus, Rheic and Pacific Oceans. Gondwana Research 31,
ence 293, 683–687. 271–294.
Schulze, C., 2011. Petrología y geoquímica de las rocas del área de Pluma Hidalgo, Oaxaca Vervoort, J.D., Patchett, P.J., Albarede, F., Blichert-Toft, J., Rudnick, R., Downes, H., 2000. Hf-
e implicaciones tectónicas para el Proterozoico de Oaxaquia. (PhD Thesis). Nd isotopic evolution of the lower crust. Earth Planet. Sci. Lett. 181, 115–129.
Universidad Autónoma Nacional de México, México. Vervoort, J.D., Plank, T., Prytulak, J., 2011. The Hf-Nd isotopic composition of marine
Sláma, J., Košler, J., Condon, D.J., Crowley, J.L., Gerdes, A., Hanchar, J.M., Horstwood, M.S.A., sediments. Geochimica et Cosmochimica Acta 75, 5903–5926.
Morris, G.A., Nasdala, L., Norberg, N., Schaltegger, U., Schoene, B., Tubrett, M.N., Vielzeuf, D., Clemens, J.D., Pin, C., Moinet, E., 1990. Granites, granulites, and crustal differ-
Whitehouse, M.J., 2008. Plešovice zircon – a new natural reference material for entiation. In: Vielzeuf, D., Vidal, P. (Eds.), Granulites and Crustal Evolution. NATO ASI
U–Pb and Hf isotopic microanalysis. Chemical Geology 249, 1–35. Series. Springer, Netherlands, pp. 59–85.
Söderlund, U., Patchett, J.P., Vervoort, J.D., Isachsen, C., 2004. The 176Lu decay constant de- Weber, B., Hecht, L., 2003. Petrology and geochemistry of metaigneous rocks from a
termined by Lu–Hf and U–Pb isotope systematics of Precambrian mafic intrusions. Grenvillian basement fragment in the Maya block: the Guichicovi complex, Oaxaca,
Earth Planet. Sci. Lett. 219, 311–324. southern Mexico. Precambrian Research 124:41–67. https://doi.org/10.1016/S0301-
Solari, L.A., Tanner, M., 2011. U-Pb age, a fast data reduction script for LA-ICP-MS U-Pb 9268(03)00078-0.
geochronology. Revista Mexicana de Ciencias Geológicas 28, 83–91. Weber, B., Köhler, H., 1999. Sm–Nd, Rb–Sr and U–Pb geochronology of a Grenville
Solari, L.A., Keppie, J.D., Ortega-Gutierrez, F., Cameron, K.L., Lopez, R., 2004. Similar Terrane in Southern Mexico: origin and geologic history of the Guichicovi
to 990 Ma peak granulitic metamorphism and amalgamation of Oaxaquia, Mexico: Complex. Precambrian Research 96:245–262. https://doi.org/10.1016/S0301-
U-Pb zircon geochronological and common Pb isotopic data. Revista Mexicana de 9268(99)00012-1.
Ciencias Geológicas 21, 212–225. Weber, B., Schulze, C.H., 2014. Early Mesoproterozoic (N1.4 Ga) ages from granulite base-
Solari, L.A., Ortega-Gutiérrez, F., Elías-Herrera, M., Schaaf, P., Norman, M., de León, R.T., ment inliers of SE Mexico and their implications on the Oaxaquia concept – evidence
Ortega-Obregón, C., Chiquín, M., Ical, S.M., 2009. U-Pb zircon geochronology of from U-Pb and Lu-Hf isotopes on zircon. Revista Mexicana de Ciencias Geológicas 31,
Palaeozoic units in Western and Central Guatemala: insights into the tectonic evolu- 377–394.
tion of Middle America. Geological Society of London, Special Publication 328: Weber, B., Cameron, K.L., Osorio, M., Schaaf, P., 2005. A late Permian tectonothermal event
295–313. https://doi.org/10.1144/SP328.12. in Grenville crust of the Southern Maya terrane: U-Pb zircon ages from the Chiapas
Solari, L.A., Ortega-Gutiérrez, F., Elías-Herrera, M., Gómez-Tuena, A., Schaaf, P., 2010a. Re- Massif, Southeastern Mexico. International Geology Review 47, 509–529.
fining the age of magmatism in the Altos Cuchumatanes, western Guatemala, by LA– Weber, B., Iriondo, A., Premo, W.R., Hecht, L., Schaaf, P., 2007. New insights into the
ICPMS, and tectonic implications. International Geology Review 52:977–998. https:// history and origin of the southern Maya block, SE Mexico: U–Pb–SHRIMP zircon
doi.org/10.1080/00206810903216962. geochronology from metamorphic rocks of the Chiapas Massif. International Journal
Solari, L.A., Gómez-Tuena, A., Bernal, J.P., Pérez-Arvizu, O., Tanner, M., 2010b. U-Pb zircon of Earth Sciences 96:253–269. https://doi.org/10.1007/s00531-006-0093-7.
geochronology with an integrated LA-ICP-MS microanalytical workstation: Weber, B., Valencia, V.A., Schaaf, P., Pompa-Mera, V., Ruiz, J., 2008. Significance of prove-
achievements in precision and accuracy. Geostandards and Geoanalytical Research nance ages from the Chiapas Massif Complex (Southeastern Mexico): redefining the
34, 5–18. Paleozoic basement of the Maya Block and its evolution in a peri-Gondwanan realm.
Solari, L.A., Garcia-Casco, A., Martens, U., Lee, J.K.W., Ortega-Rivera, A., 2013. Late Journal of Geology 116:619–639. https://doi.org/10.1086/591994.
Cretaceous subduction of the continental basement of the Maya block (Rabinal Weber, B., Scherer, E.E., Schulze, C., Valencia, V.A., Montecinos, P., Mezger, K., Ruiz, J., 2010.
Granite, central Guatemala): tectonic implications for the geodynamic evolution of U–Pb and Lu–Hf isotope systematics of lower crust from central-southern Mexico –
Central America. Geological Society of America Bulletin 125:625–639. https:// geodynamic significance of Oaxaquia in a Rodinia Realm. Precambrian Research
doi.org/10.1130/B30743.1. 182:149–162. https://doi.org/10.1016/j.precamres.2010.07.007.
Spear, F.S., 1981. An experimental study of hornblende stability and compositional Weber, B., Scherer, E.E., Martens, U.K., Mezger, K., 2012. Where did the lower Paleozoic
variability in amphibolite. American Journal of Science 281, 697–734. rocks of Yucatan come from? A U–Pb, Lu–Hf, and Sm–Nd isotope study. Chemical
Sprung, P., Scherer, E.E., Upadhyay, D., Leya, I., Mezger, K., 2010. Non-nucleosynthetic Geology 312–313:1–17. https://doi.org/10.1016/j.chemgeo.2012.04.010.
heterogeneity in non-radiogenic stable Hf isotopes: implications for early solar Wiedenbeck, M., Alle, P., Corfu, F., Griffin, W.L., Meier, M., Oberli, F., von Quadt, A.,
system chronology. Earth and Planetary Science Letters 295, 1–11. Roddick, J.C., Spiegel, W., 1995. Three natural zircon standards for U-Th-Pb, Lu-Hf,
Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead isotope evolution by a trace element and REE analyses. Geostandards Newsletter 19, 1–23.
two-stage model. Earth and Planetary Science Letters 26, 207–221. Zenk, M., Schulz, B., 2004. Zoned Ca-amphiboles and related PT evolution in metabasites
Steiner, M., Walker, J., 1996. Late Silurian plutons in Yucatan. Journal of Geophysical from the classical Barrovian metamorphic zones in Scotland. Mineralogical Magazine
Research 101, 17727–17735. 68, 769–786.

You might also like