Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

The Geological Society of America

Special Paper 546

Petrogenesis of the crystalline basement


along the western Gulf of Mexico:
Postcollisional magmatism during the formation of Pangea

Henry E. Coombs
Andrew C. Kerr
School of Earth and Ocean Sciences, Cardiff University Main Building, Park Place, Cardiff CF10 3AT, UK

James Pindell
Tectonic Analysis Ltd., Chestnut House, Duncton, West Sussex GU28 0LH, UK

David Buchs
School of Earth and Ocean Sciences, Cardiff University Main Building, Park Place, Cardiff CF10 3AT, UK

Bodo Weber
Departamento de Geología, Centro de Investigación Científca y de Educación Superior de Ensenada (CICESE),
Carretera Ensenada-Tijuana 3918, Zona Playitas, 22860 Ensenada B.C., Mexico

Luigi Solari
Centro de Geociencias, Universidad Nacional Autónoma de México,
Campus Juriquilla, 76001 Querétaro, Querétaro, Mexico

ABSTRACT

The supercontinent of Pangea formed through the diachronous collision of Lau-


russia and Gondwana during the late Paleozoic. While magmatism associated with
its formation is well documented in the Variscan orogeny of Europe and Alleghanian
orogeny of the United States, little is known about the Sonora orogeny of northern
Mexico. This paper reports geochronology (U-Pb zircon), whole-rock geochemistry,
and Lu-Hf zircon isotope data on basement cores from the western Gulf of Mexico,
which were used to develop a tectonomagmatic model for pre- to post-Pangea amal-
gamation. Our results suggest the existence of three distinct phases of magmatism,
produced during different stages of continental assembly and disassembly. The first
phase consists of Early Permian (294–274 Ma; n = 3) granitoids with geochemical sig-
natures indicative of a continental arc tectonic setting. This phase formed on the mar-
gins of Gondwana during the closure of the Rheic Ocean, prior to the final amalga-
mation of Pangea. It likely represents a lateral analogue of late Carboniferous–Early
Permian granitoids that intrude the Acatlán and Oaxacan Complexes. The second

Coombs, H.E., Kerr, A.C., Pindell, J., Buchs, D., Weber, B., and Solari, L., 2020, Petrogenesis of the crystalline basement along the western Gulf of Mexico: Postcol-
lisional magmatism during the formation of Pangea, in Martens, U.C, and Molina Garza, R.S., eds., Southern and Central Mexico: Basement Framework, Tectonic
Evolution, and Provenance of Mesozoic–Cenozoic Basins: Geological Society of America Special Paper 546, p. 29–52, https://doi.org/10.1130/2020.2546(02).
© 2020 The Geological Society of America. All rights reserved. For permission to copy, contact editing@geosociety.org.

29

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
30 Coombs et al.

phase of magmatism includes Late Permian–Early Triassic (263–243 Ma; n = 13)


granitoids with suprasubduction geochemical affinities. However, Lu-Hf isotope data
indicate that these granitoids formed from crustal anatexis, with εHf values and two-
step Hf depleted mantle model ages (TDM[Hf]) comparable to the Oaxaquia continental
crust into which they intrude. This phase of magmatism is likely related to coeval
granitoids in the Oaxaca area and Chiapas Massif. We interpret it to reflect late- to
postcollisional magmatism along the margin of Gondwana following the assembly of
Pangea. Finally, the third phase of magmatism includes Early–Middle Jurassic (189–
164 Ma; n = 2) mafic porphyries, which could be related to the synchronous suprasub-
duction magmatism associated with the Nazas arc. Overall, our results are consistent
with Pangea assembly through diachronous collision of Laurussia and Gondwana
during subduction of the Rheic Ocean. They suggest that postorogenic magmatism in
the western termination of the Rheic suture occurred under the influence of a Pan-
thalassan subduction zone, before opening of the Gulf of Mexico.

INTRODUCTION Herein, we present new whole-rock geochemical data, U-Pb


zircon geochronology, and Lu-Hf zircon isotope data from well
The late Paleozoic was characterized by the formation of cuttings from the crystalline basement of the western Gulf of
Pangea through the collision and amalgamation of the Laurus- Mexico. Jurassic plate reconstructions (Pindell and Kennan,
sia and Gondwana continents. This collision was diachronous, 2009) of western Pangea place the area of the western Gulf of
initiating in the east (Europe and Africa) during the Carbonifer- Mexico close to the Ouachita-Sonora suture zone. Based on these
ous (ca. 340–330 Ma; Bussien et al., 2011) and terminating in new data, we propose a pre- to postcollisional model for magma-
the west (Mexico) during the Early Permian (ca. 280 Ma; Pindell tism associated with the formation of Pangea along the Sonora
and Dewey, 1982; Pindell, 1985; Ross, 1986; Viele and Thomas, segment of the Pangea suture zone.
1989; Sedlock et al., 1993; Dickinson and Lawton, 2001). The
suture of this major collision zone is recorded in the Sonora-Mar- GEOLOGICAL SETTING
athon-Ouachita-Alleghanian orogenies of northern Mexico and
the United States, and the Variscan orogeny of Europe. Crystalline Basement of the Western Gulf of Mexico
Existing geochronological and geochemical data document
a complex tectonic history during the formation of Pangea. The Ages for the crystalline basement (cores) of the western
collision of Laurussia and Gondwana during the Variscan orogeny Gulf of Mexico have previously been presented in an internal
of Europe was associated with three distinct, syn- (340–330 Ma), Petróleos Mexicanos (PEMEX) report by Lopez-Ramos (1979).
late- (310 Ma), and postcollisional magmatic phases (Schalteg- The method of dating was primarily K-Ar on biotite or potas-
ger, 1997; Bussy et al., 2000; Von Raumer and Bussy, 2004; sium feldspars, and these analyses yielded a wide range of het-
Bussien et al., 2011). Further west, the Laurussia-­Gondwana erogeneous ages from the Neoproterozoic to Early Cretaceous
collision along the Ouachita-Marathon-Alleghanian orogenies (Table 1). A detailed description of the crystalline basement of
did not occur until the late Carboniferous–Early Permian, with Mexico can be found in Ortega-Gutiérrez et al. (2014); however,
arc rocks relating to the closure of the Rheic Ocean dating to we present below a brief outline of the major basement-forming
325 Ma (U-Pb zircon; Shaulis et al., 2012) and syn-, late-, and ­Proterozoic–Jurassic igneous events in Mexico.
postcollisional granitoids dating to 300–275 Ma (Walsh et al.,
2007; Heatherington et al., 2010; Mueller et al., 2014). In north- Oaxaquia
east Mexico, the Sonora segment of the Laurussia-Gondwana
collision could provide important constraints on the timing and Oaxaquia or the Oaxaquia microcontinent is associated with
mode of assembly and disassembly of Pangea. This segment a Grenvillian phase of granulite-facies metamorphism and is
includes a continuous 3000-km-long Paleozoic carbonate conti- thought to underlie the majority of Gondwana-derived Mexico.
nental shelf with consistent stratigraphic and structural character- Oaxaquia was first described in Ortega-Gutierrez et al. (1995),
istics extending from Sonora to the southern United States (Poole with outcrops found in the Oaxacan Complex, Novillo Gneiss,
et al., 2005). However, very little is known about the orogenic Huiznopala Gneiss, and Guichicovi Complex (Fig. 1A).
magmatism in this area. Yet, understanding the mechanics of the The oldest reported ages for Oaxaquia range between 1.5
collision in the Sonora segment is important because it likely and 1.4 Ga (U-Pb zircons in orthogneiss and migmatites; Solari
marks the latest phase of the Laurussia-Gondwana collision and et al., 2003; Schulze, 2011; Weber and Schulze, 2014). These
final amalgamation of Pangea. ages have been associated with an early island arc known as

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 31

TABLE 1. PREVIOUS GEOCHRONOLOGY


Area Age (Ma) Method Reference
Western Gulf of Mexico 312–916 K-Ar Lopez-Ramos (1979)
Sosola rhyolite 270.5 ± 2.5 U/Pb zircon Ortega-Obregón et al. (2014)
Cuanana pluton 310.8 ± 1.8 U/Pb zircon Ortega-Obregón et al. (2014)
Carbonera stock 272.5 ± 1.0 U/Pb zircon Ortega-Obregón et al. (2014)
Zaniza batholith 287.0 ± 1.9 U/Pb zircon Ortega-Obregón et al. (2014)
Etla granite 255.2 ± 1.0 U/Pb zircon Ortega-Obregón et al. (2014)
Honduras batholith 290.1 ± 2.2 U/Pb zircon Ortega-Obregón et al. (2014)
Xolapa Complex 272.0 ± 10 U/Pb zircon Ducea et al. (2004)
Chiapas Massif 251–258 U/Pb zircon Weber et al. (2005)
Chiapas Massif 271.9 ± 2.7 U/Pb zircon Weber et al. (2007)
Cozahuico granite 270.4 ± 2.6 U/Pb zircon Elías-Herrera et al. (2005)
Sierra Pinta granites 265–275 U/Pb zircon Arvizu and Iriondo (2011)
Totoltepec pluton 306.0 ± 1.0 U/Pb zircon Kirsch et al. (2012)
Totoltepec pluton 289.0 ± 1.0 U/Pb zircon Kirsch et al. (2012)
Totoltepec pluton 289 ± 1 U/Pb zircon Keppie (2004)
Totoltepec pluton 287 ± 1 U/Pb zircon Yañez et al. (1991)
Chichihualtepec dike 265–275 U/Pb zircon Kirsch et al. (2012)
Rabinal complex 215–270 U/Pb zircon Ratschbacher et al. (2009)

100°W 99°W 98°W 97°W 96°W 95°W 94°W


Legend
Geochronology and Geochemistry Proposed Oaxaquia
Benemerito- 245.4 ± 4.6 Ma
B
Geochemistry Extension (Keppie et Trincheras- 254.6 ± 3.4 Ma 25°N
al., 2003) Linares- 255.1 ± 4.8 Ma
Proterozoic-Triassic Basement Outcrops Proterozoic-Triassic Chaneque- 243.4 ± 2.8 Ma
Nazas Arc (Lawton & Molina Garza, 2014) Basement Complexes 24°N

Oaxaquia (Ortega-Gutiérrez et al., 1995) Suture


23°N
A Nayade- 257.0 ± 5.2 Ma
amaulipas- 247.9 ± 4.0 Ma
Tamaulipas-
USA Arenque- 293.5 ± 3.7 Ma
Jurel 22°N
Erizo- 249.8 ± 2.7Ma
Sierra Mexico
Pinta Nazas arc Marathon Pinonal- 294.1 ± 3.4 Ma
Segment & 249.8 ± 3.2 Ma
21°N
30° Sonora
Muro- 188.3 ± 4.0 Ma
Segment Magdalena
Paso de Oro- 274.2 ± 3.5 Ma Cupelado- 262.7 ± 4.5 Ma
Oaxaquia Tlapacoyan- 163.5 ± 4.7 Ma
20°N

Plan de Las Hayas- 261.0 ± 3.9 Ma


& 251.7 ± 5.2 Ma
Paso de Ovejas- 256.7 ± 5.5 Ma 19°N

Orizaba- 261.1 ± 4.9 Ma

Novillo Gneiss 18°N

Guerrero Gulf of Mexico


Composite Huiznopala Gneiss
Terrane

Totoltepec pluton Cozahuico Granite


Acatlán Complex Carbonera Stock Rabinal Complex
Pacific Ocean Cuanana pluton
Zaniza batholith Etla Granite
Guichicovi Complex
Mixtequita batholith
Honduras batholith
15°N Oaxacan Complex Chiapas Massif

N 250 km
110°W 90°
Figure 1. (A) Schematic map of Mexico displaying the extent of Oaxaquia type basement. Carboniferous–Triassic pluton exposures in Mexico
are also displayed (Elías-Herrera et al., 2005; Weber et al., 2005; Ratschbacher et al., 2009; Arvizu and Iriondo, 2011; Kirsch et al., 2012; Ortega-
Obregón et al., 2009, 2014). (B) Location of wells from which basement samples were taken in this study.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
32 Coombs et al.

p­ roto-Oaxaquia (Weber and Schulze, 2014). A later and more Venezuela, and Colombia (Vinasco et al., 2006; Cochrane et al.,
prevalent pulse of island-arc magmatism, possibly associated 2014; Spikings et al., 2015).
with back-arc magmatism, occurred between 1.3 and 1.2 Ga
(Lawlor et al., 1999; Keppie et al., 2001; Keppie, 2004; Weber et Nazas Arc
al., 2010; Weber and Schulze, 2014).
Migmatization (Olmecan event) occurred in Oaxaquia ca. Early Jurassic magmatism in Mexico is most commonly
1.1 Ga (Solari et al., 2003), which has been linked to either back- associated with the Nazas arc system (Fig. 1A; Dickinson and
arc extension (Keppie et al., 2001, 2003; Keppie, 2004) or a Lawton, 2001; Lawton and Molina Garza, 2014). Volcanic expo-
compressional event during accretion of the microcontinent to sures outcrop in northern central Mexico (Barboza-Gudiño et al.,
Amazonia. Following migmatization, an anorthosite-mangerite- 2008) and possibly extend as far south as the Chiapas Massif
charnockite-granite (AMCG) suite was emplaced ca. 1012 Ma (Godinez-Urban et al., 2011; Lawton and Molina Garza, 2014).
(Solari et al., 2003), which has also been linked to back-arc exten- Plutonic equivalents of these Jurassic arc rocks can be found in
sion (Solari et al., 2003; Keppie, 2004) or Oaxaquia-­Amazonia northeastern Mexico (Ortega-Gutíerrez et al., 2014).
collision (Weber et al., 2010).
SAMPLES AND PETROGRAPHY
Acatlán Complex
Samples for this study were supplied by PEMEX and
The Acatlán Complex of southern Mexico (Fig. 1A) is one extracted from wells on basement highs along the western Gulf
of the few areas exposed in the region that preserves Paleozoic of Mexico (Tamaulipas Arch, Tuxpan, and Santa Ana). The wells
rocks. It is composed of a complex succession of deformed occur at regular intervals, over an area of ~120,000 km2, and
high-pressure (HP) metamorphic rocks, including eclogites and therefore provide a good representation of the basement through-
blueschists, deformed granitoids, and metasedimentary rocks. out eastern Mexico (Fig. 1B).
The origin of these Paleozoic rocks is still debated. Some stud- Petrographic observations (Figs. 2A–2D) revealed that the
ies propose formation at an active continental margin along Lau- majority of the samples are phaneritic with granodiorite-granite
rentia (Talavera-Mendoza et al., 2005; Vega-Granillo et al., 2007 compositions. The main mineral assemblages include quartz-
and 2009) or Gondwana (Van der Lelij et al., 2016). Other mod- feldspars-biotite ± hornblende ± muscovite. Common accessory
els suggest that the Acatlán Complex formed along a passive minerals include zircon, apatite, titanite, and opaques. Minor
margin of Gondwana, with HP metamorphic rocks extruded into (common) to extensive (rare) hydrothermal alteration in the gran-
the upper plate during the Carboniferous (Keppie et al., 2008, itoids is evidenced by sericitized feldspars, chloritized biotite,
2012; Ortega-Obregón et al., 2009, 2014; Estrada-Carmona et calcite, and epidote. There is little evidence from the petrography
al., 2016). of significant deformation affecting the granitoids.
Regardless of how the Acatlán Complex was formed, it is A subset of samples (Muro-2 and Tlapacayon-1) displayed
generally agreed that it was emplaced along a continental margin finer phaneritic/porphyritic textures, typical of hypabyssal rocks
(either Laurussia or Gondwana). This indicates that it could have (Figs. 2E–2F). This suggests that they were emplaced at a shal-
been proximal to the Sonora segment of the collision zone during lower depth than the coarser plutonic samples described above,
the final amalgamation of Pangea. or they represent small intrusive bodies. These hypabyssal rocks
are characterized by more mafic mineral assemblages, with
Carboniferous–Triassic Magmatism minor clinopyroxene and olivine (partially altered to iddingsite)
and no quartz.
Late Carboniferous–Early Triassic granitoids can be found
across Mexico, including the Sierra Pinta, Chiapas Massif, and GEOCHRONOLOGY
intrusions into the Acatlán and Oaxacan Complexes (Fig. 1A;
Table 1). One interpretation of these igneous rocks is that they Analytical Methods
formed in a continental arc system, associated with the eastward-
dipping subduction of the Panthalassa Ocean beneath Pangea Geochronological data (U-Pb zircon) were obtained by
(Keppie, 2004; Arvizu and Iriondo, 2011; Kirsch et al., 2012; laser ablation–inductively coupled plasma–mass spectrometry
Ortega-Obregón et al., 2014). Alternatively, the igneous event (LA-ICP-MS) at the Laboratorio de Estudios Isotópicos (LEI),
could have been linked to the final amalgamation of Pangea Centro de Geociencias, Universidad Nacional Autonóma de
(Yañez et al., 1991; Elías-Herrera et al., 2005). México, Querétaro, Mexico. A Thermo X-ii quadrupole ICP-
Paleogeographic reconstructions of Pangea (Pindell and MS was used coupled with a high-resolution M050 excimer
Kennan, 2009) indicate a close proximity between Mexico and laser-ablation workstation. A cathodoluminescence (CL) study
South America at this time. This suggests that the Carbonifer- before analysis identified suitable target areas in the zircons,
ous–Triassic magmatism in Mexico may have been associated avoiding core-rim transitions, cracks, and inclusions. During
with the Permian–Triassic arc rocks that are found in Ecuador, analysis, a 30 µm laser spot was employed, with the Plešovice

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 33

A B
Chl Qtz
Ttn

Zrn Ser
Pl
Cal
500 µm 500 µm Figure 2. Photomicrographs of the analyzed
samples. Abbreviations are as follows:
C Qtz D Chl—chlorite, Cpx—­clinopyroxene, Ttn—
titanite, Qtz—quartz, Zrn—zircon, Pl—
plagioclase, Ser—sericite, Ol—­olivine, Id—­
iddingsite, Bt—biotite, Am—­amphibole,
Cal—­ calcite, Epi—epidote. (A) Hydro-
Pl Epi thermally altered granite containing quartz,
plagioclase (partially sericitized), chlorite,
titanite, and zircon as a minor phase; cross-
Bt polarized light (XPL). (B) Hydrothermally
Am altered granite with abundant secondary
calcite; XPL. (C) Granodiorite containing

Ser Qtz quartz, plagioclase, biotite, and amphibole;


XPL. (D) Hydrothermally altered granite
with epidote; XPL. (E) Mafic porphyry
500 µm 500 µm containing olivine, iddingsite, and plagio-
clase phenocrysts and a matrix of feldspars;
XPL. (F) Microdiorite containing partially
E F chloritized biotite and minor clinopyroxene.

Id Bt
Pl
OI
Cpx
Pl

500 µm 500 µm

zircon (Sláma et al., 2008) being used as a bracketing standard. Results


NIST 610 glass was used to recalculate elemental concentra-
tions, and 29Si was used as an internal standard. Following anal- Zircons were successfully extracted and analyzed in 18 sam-
ysis, the data were reduced using an algorithm slightly modified ples from wells along the western Gulf of Mexico. Th/U ratios
from Solari and Tanner (2011) and then exported into Excel®, (Figs. 3A–3R) indicate that the zircons are dominantly magmatic
where concordia plots and weighted mean age calculations in origin, with the majority of the analyzed grains falling between
were constructed using Isoplot (Ludwig, 2012) version 4.15. 0.1 and 1. The CL study (Figs. 4A–4R) supports this interpreta-
Where necessary, a correction for common lead was applied tion, with most of the zircons displaying typical igneous textures
using the algebraic method of Andersen (2002). Further details (oscillatory and sector zoning). Furthermore, the CL images
on the equipment used and analytical procedures can be found revealed core and rim relationships and resorption features in
in Solari et al. (2010). many of the samples (indicated below when present).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
by UNAM user
box heights are 1σ box heights are 1σ box heights are 1σ
34
A. Pinonal N2F3C1 B. Arenque N15F3C2 293.5±3.7 Ma MSWD= 1.4
C. Paso de Oro N8F4C2
Biotite granitoid 370 Biotite granite 340 Lineated granite 315
Analyzed grains= 43 350 Analyzed grains= 40 Analyzed grains= 26 305
320
0.10 330 0.10 0.09 290
300 275
310
280 260
0.08 290 0.08 245
270 260 0.07 230
294.1±3.4 Ma MSWD= 1.15 293.5±3.7 Ma MSWD= 1.4 274.2±3.5 Ma MSWD= 0.77

206
206

206
250 0.06 10
215 274.2±3.5 Ma MSWD= 0.77
0.06 10 10

Pb/ Pb
Pb/ Pb

Pb/ Pb
207
207

207
420 380 340 380 340 300
300 260 1 260 1 0.05 330 310 290 270 250 1
0.04 0.04

Th/U

Th/U
Th/U

0.1 0.1 0.1

0.02 0.02 0.03


16 18 20 22 24 26 0.01
14 16 18 20 22 24
238
26 0.01
270 290 310 330 350 370 238 250 275 300 325 350 18 20 22 24 238
26 0.01250 270 290 310
data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma)

box heights are 1σ box heights are 1σ box heights are 1σ


D. Benemerito N2F7C5 E. Trincheras N3F5AC1 F. Linares N7F1C1
Granite 270 245.4±4.6 Ma MSWD= 1.8 280 Biotite granite
Biotite granite 280
Analyzed grains= 35 260 Analyzed grains= 40 260 Analyzed grains= 40
400 400
0.11 0.22 260
250 240
200 200

Age
Age (Ma)
0 240 0 220 0.18
10 100 1000 10 100 1000 240
0.12 U (ppm) 0.09 U (ppm)
230 200
0.14

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


1100 254.6±3.4 2.7
255.8±3.3 Ma MSWD= 3.4 220 253.5±5.2
255.1±4.8 Ma
254.3±4.6 Ma MSWD= 4.4
MSWD= 5.4
4.0
0.10 180

206
220 10
0.07 10 10
1400
Pb/206Pb

Pb/206Pb
0.10

Pb/ Pb
900 700
207

207

207
0.08 1
1000 1 1
500 0.06
0.05

Th/U
300 360
Th/U

Th/U
Coombs et al.

0.06 600 280


0.1 0.1 0.1
0.02 320 240
0.04 200
0 40 80
0.030
10 20 30 40 0.01 17 19 21 23 25 27
238
120206 0.010 400 800 1200 238 206 0 400 800 238
29 206 0.01220 240 260 280 300 320
data-point error ellipses are 1σ U/ Pb Age (Ma) data-point error ellipses are 1σ U/ Pb Age (Ma) data-point error ellipses are 1σ U/ Pb Age (Ma)

box heights are 1σ box heights are 1σ box heights are 1σ


G. Chaneque N6F1C4 H. Nayade N13F2C1 257.0±5.2 Ma MSWD= 3.7
I. Tamaulipas N1F2C1
290 Granodiorite 290
Biotite granitoid Biotite granite 275
Analyzed grains= 38 Analyzed grains= 35 Analyzed grains= 37
270 265 270
0.12 0.14
250 255 250
0.10 0.07 0.12
880 245
230 0.10 230
0.08 235
720 247.9±4.0 Ma MSWD= 1.8
243.4±2.8
243.4 ± 2.8
MaMSWD= 1.9
MSWD=1.9 0.06 0.08 210
206

210 200 10
0.06 340 10 560
Age

260 180

Pb/206Pb
Pb/ Pb

Pb/206Pb
0 0.06

207
300
207

207
220 400 100 1000 U (ppm) 310 270 230
0.04 1 1
0.05 10 0.04 330 290
240 250 210
Th/U

Th/U
0.02 0.1 0.02 0.1
1
Th/U

0.00 0.04 0.00


16 20 24 28 32 36 238
40 206 0.01170 190 210 230 250 270 290 0 10 20 30 40 238
50 206 0.1 18 20 22 24 26 28 238 220
30 206 0.01 240 260 280
data-point error ellipses are 1σ Age (Ma) data-point error ellipses are 1σ 200 400 600 800 data-point error ellipses are 1σ
U/ Pb U/ Pb 0 U/ Pb Age (Ma)

Figure 3. Tera-Wasserburg, Wetherill, and weighted mean diagrams for the dated samples along the western Gulf of Mexico. Th/U vs. age plots are also included, as well as U vs. age
plots where metamictization is suspected. MSWD—mean square of weighted deviates. (Continued on facing page.)
by UNAM user
box heights are 1σ box heights are 1σ 262.7±4.5 Ma
J. Erizo N19F19C1 K. Pinonal N2F11C2 275 L. Cupelado N9F1C1 274 MSWD= 0.86
Mean= 249.5±3.1 Ma
290 Granodiorite MSWD= 1.6 Biotite granite
Granite 265
Analyzed grains= 49 Analyzed grains= 31 Analyzed grains= 19 266
270 255
10 258
220 250 245

Age
0.10 0.10 235 0.3 1 250

Th/U
100 284.7±7.7 Ma
10
230 1300 300 MSWD= 0.11
100 1000 225 1400
U (ppm) 0.08 0.1 292
0.08 210 1100 900 215 0 800 1600
Mean= 250.1±3.1 Ma Age (Ma)
249.8±3.2 Ma MSWD= 1.6 284

238
249.8±2.7 Ma MSWD=
MSWD= 2.2 1.8 0.2
190 0.06 500 10
10 700

206
277

Pb/ U
0.06 1000

206
300

Pb/206Pb
260

Pb/ Pb
180 268

207
0.04

207
300 220 1 1 600
140 0.1 1550
0.04

Th/U
Th/U
0.1 0.02 0.1 1350
200
0.02 1150 1418±41
1428±48 Ma
0.01
0.00 0.0
15 25 35 45 0 10 20 238
0 1 2 4 MSWD= 0.280.21
238 100 150 200 250 300 30 0.010 400 800 1200 3 207 235
data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ Pb/ U950 box heights are 1σ

M. Plan de Las Hayas N14F1C1 box heights are 1σ N. Plan de Las Hayas N14F6C2 box heights are 1σ O. Paso de Ovejas N6F9C2 box heights are 1σ
Quartz diorite 300 Quartz monzodiorite
Granodiorite
Analyzed grains= 29 285 Analyzed grains= 30 Analyzed grains= 32 300
280
275 280
260
0.08 265 0.12 0.11 260
255 240 240
0.10
0.07 245

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


Mean= 261.0±4.1 Ma 220 0.09 220 Mean= 256.7±5.5 Ma
261.0±3.9 Ma MSWD=
MSWD= 2.8 2.8 Mean= 250.4±4.6 256.7±5.5 MSWD= 2.3
235 251.7±5.2 Ma MSWD=Ma3.0
10 0.08 MSWD= 2.5 MSWD= Ma2.3
200

206
200 10 10

Pb/206Pb
0.06 0.07
Pb/206Pb

207
Pb/ Pb
207

300 1 0.06

207
260 340 260 1 300 1
220
320

Th/U
0.05 280 300

Th/U
240 220 0.05
Th/U

0.1 0.04 340 260 220


0.1 0.1

0.04 0.02 0.03


18 20 22 24 26 28 0.01220 240 260 280 300 16 20 24 32 0.01 16 20 24 0.01
238 28238 200 250 300 28238 200 240 280 320
data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma)

285 261.1±4.9 Ma box heights are 1σ box heights are 1σ


P. Orizaba N6F1C1 MSWD= 4.4 Q. Muro-2 N12F2C1 210 R. Tlapacoyan N10F18C5
Monzo-granite 275 Micro diorite 195
Trachyte
Analyzed grains= 52 265 Analyzed grains= 21 200 Analyzed grains= 30 185
1 255
245 0.24 175
0.28
0.1 235 190 0.08 165

Th/U
0.24 1300 970±53 Ma 0.20
1100 MSWD= 1.2 155
0.01 180
0.20 0 400 800 1200 1000 0.07 145
Age (Ma) 0.16 Mean= 188.5±3.5 Ma
Petrogenesis of the crystalline basement along the western Gulf of Mexico

188.3±4.0Ma 5.4 800 163.5±4.7 Ma MSWD= 3.6


100 900 MSWD=MSWD=
5.6 135

238
0.16 10

206
170 0.06 600
800 0.12 10

Pb/ U
700
Pb/206Pb

Pb/ Pb
0.12 700 400

206
1400
207

207
1149±34 Ma 0.08 0.05 1
500 MSWD= 1.02 1 200
0.08 600
1000
Th/U

300 1250
Th/U

0.04 200 0.04 0.1


0.04 1150 0.1

0.00 1050 0.00 0.03


0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 0 10 20 0.01 0 10 20 30 40 0.01
207 235
30238 0 500 1000 1500 2000 238 0 200 400 600 800 1000
data-point error ellipses are 1σ Pb/ U 950 box heights are 1σ data-point error ellipses are 1σ U/206Pb Age (Ma) data-point error ellipses are 1σ U/206Pb Age (Ma)

Figure 3 (Continued).
35
by UNAM user
36

(1)
307±13 Ma 1266±93 Ma 253±9 Ma (Hf)
245±3 Ma
277±8 Ma 275±7 Ma 265±7 Ma
282±10 Ma 76±1 Ma (M) 245±4 Ma 258±6 Ma 288±5 Ma
283±10 Ma 247±7 Ma
122±2 Ma (M) 258±6 Ma 260±5 Ma
306±13 Ma 265±8 Ma
252±4 Ma 289±4 Ma 252±4 Ma 240±6 Ma
292±8 Ma 282±11 Ma 273±6 Ma
267±8 Ma
239±3 Ma
290±10 Ma 272±11 Ma 242±6 Ma
μ μ μ μ μ μ

246±6 Ma (2)
261±2 Ma 236±5 Ma 1391±57 Ma
238±11 Ma 235±7 Ma
241±4 Ma 206± Ma (M) 256±4 Ma
539±16 Ma 279±7 Ma 1400±74 Ma 257±5 Ma
244±5 Ma 268±6 Ma 237±7 Ma

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


265±6 Ma 258±5 Ma
231±4 Ma (M) 268±5 Ma
282±10 Ma 251±5 Ma
558±14 Ma 267±10 Ma 247±4 Ma
283±8 Ma 147±3 Ma (M) 284±7 Ma 285±6 Ma
240±11 Ma 266±7 Ma
Coombs et al.

μ μ μ μ μ μ

(1) (2)
238±8 Ma 244±11 Ma 253±5 Ma 181±2 Ma 249±5 Ma
253±6 Ma 268±4 Ma (Hf)
184±3 Ma 170±6 Ma
1251±49 Ma
188±4 Ma 167±4 Ma
1023±47 Ma 162±4 Ma
231±14 Ma 254±6 Ma (Hf) 189±3 Ma 763±10 Ma
255±8 Ma 279±10 Ma (Hf) 199±3 Ma
281±9 Ma 165±5 Ma
417±17 Ma 184±6 Ma
226±11 Ma 247±10 Ma (Hf) 250±13 Ma 170±4 Ma
186±3 Ma
μ μ 243±9 Ma μ 298±7 Ma μ 251±6 Ma μ 196±4 Ma μ 164±5 Ma

Figure 4. Cathodoluminescence images of the zircons in each of the analyzed samples. M—metamictization; Hf—zircons used for Lu-Hf isotope analysis (dilution method). We
were unable to determine the appropriate ages for the corresponding laser spots in some of the samples due to an error during data collection.
Petrogenesis of the crystalline basement along the western Gulf of Mexico 37

Depending on the age of the zircons, results of the analysis Arenque-22. The sample from the Arenque-22 well is a bio-
have been graphically presented in either a Tera-Wasserburg plot tite granite, containing subhedral zircons that are 100–300 µm
in association with a 206Pb/238U weighted-mean age calculation along their longest axis, with aspect ratios of 2:1–5:1, and CL
or Wetherill concordia diagram in association with a 207Pb/206Pb textures that reveal core and rim relationships (Fig. 4B).
weighted-mean age calculation. The 206Pb/238U weighted-mean Of the 40 analyzed zircons, 36 were selected for age calcu-
age calculation is more precise for Phanerozoic zircons, whereas lations (<20% discordant, <5% error, <5% inversely discordant;
the 207Pb/206Pb weighted-mean age calculation is more suitable Fig. 3B). One inherited core was analyzed and yielded a late Car-
for Precambrian zircons (Jackson et al., 2004; Ludwig, 2012). boniferous age of 322 ± 11 Ma. A weighted-mean calculation for
Our data reveal three apparent phases of magmatism: the main population of zircons yielded an age of 293.5 ± 3.7 Ma
(1) Early Permian (294–274 Ma; n = 3), (2) Late Permian–Early (n = 34; MSWD = 1.4; two ages rejected by the algorithm), which
Triassic (263–243 Ma; n = 13), and (3) Early–Middle Jurassic can be interpreted as the igneous crystallization age of the sample.
(188–164 Ma; n = 2; Table 2). Significantly, these results indicate Paso de Oro-101. The sample from the Paso de Oro-101
that the basement of the western Gulf of Mexico is characterized well is a granodiorite containing prismatic, euhedral zircons that
by magmatic phases more consistent than previously suggested are 100–600 µm along their longest axis, with aspect ratios of
by K-Ar dating (Lopez-Ramos, 1979; see also Table 1). The pre- 2:1–8:1, and CL textures that display no evidence of inherited
vious ages of Lopez-Ramos (1979) have been included in the text cores (Fig. 4C). The basement from this well has previously been
below for comparison, where available. dated to 258 ± 11 Ma (K-Ar biotite; Lopez-Ramos, 1979).
Of the 26 analyzed zircons, 19 were selected for age calcula-
Early Permian tions (<20% discordant, <5% error, <5% inversely discordant; Fig.
Pinonal-1 (1). The first sample from the Pinonal-1 well is a 3C). A weighted-mean calculation yielded an age of 274.2 ± 3.5 Ma
biotite granitoid that has undergone subsequent alteration. It con- (n = 18; MSWD = 0.77; one age rejected by the algorithm), which
tains subhedral zircons that are 200–800 µm along their longest can be interpreted as the igneous crystallization age of the sample.
axis, with aspect ratios of 2:1–5:1, and CL textures that display
no evidence of inherited cores (Fig. 4A). Late Permian–Early Triassic
Of the 43 analyzed zircons, 36 were selected for age cal- Benemerito-1. The sample from the Benemerito-1 well is
culations (<20% discordant, <5% error, and <5% inversely a granite containing prismatic, euhedral zircons that are 80–
discordant; Fig. 3A). A weighted-mean calculation for the zir- 400 µm along their longest axis, with aspect ratios of 2:1–6:1.
cons yielded an age of 294.1 ± 3.4 Ma (n = 35; mean square The CL textures revealed core and rim relationships and evi-
of weighted deviates [MSWD] = 1.15; one age rejected by the dence for partial resorption of the zircons (Fig. 4D). The base-
algorithm), which can be interpreted as the igneous crystalliza- ment from this well has previously been dated to 916 ± 35 Ma
tion age of the sample. and 203 ± 10 Ma (method unknown; Lopez-Ramos, 1979).
Relationships between U concentrations and 238U/206Pb
ratios indicate that partial metamictization of the zircons has
TABLE 2. GEOCHRONOLOGY OF THIS STUDY occurred (Fig. 3D). Therefore, zircons with high U concentra-
Sample Age (Ma) tions were disregarded.
Early Permian Of the 35 analyzed zircons, 11 were selected for age calcula-
Pinonal (1) 294.1 ± 3.4 tions (<20% discordant, <5% error, <5% inversely discordant,
Arenque 293.5 ± 3.7 and <1000 ppm U; Fig. 3D). One inherited core was analyzed
Paso de Oro 274.2 ± 3.5 and yielded a Mesoproterozoic age of 1266 ± 93 Ma. A weighted-
Late Permian—Early Triassic mean calculation for the youngest population of zircons yielded
Cupelado 262.7 ± 4.5 an age of 245.4 ± 4.6 Ma (n = 10; MSWD = 1.8), which can be
Orizaba 261.1 ± 4.9
interpreted as the igneous crystallization age of the sample.
Plan de Las Hayas (1) 261.0 ± 3.9
Nayade 257.0 ± 5.2
Trincheras-1. The sample from the Trincheras-1 well is a
Paso de Ovejas 256.7 ± 5.5 biotite granite, containing euhedral to subhedral zircons that are
Trincheras 254.6 ± 3.4 100–300 µm along their longest axis, with aspect ratios of 2:1–
Linares 255.1 ± 4.8 5:1, and CL textures that reveal core and rim relationships (Fig.
Erizo 249.8 ± 2.7 4E). The basement from this well has previously been dated to
Plan de Las Hayas (2) 251.7 ± 5.2
147 ± 5 Ma (K-Ar biotite; Lopez-Ramos, 1979).
Pinonal (2) 249.8 ± 3.2
Tamaulipas 247.9 ± 4.0 Relationships between U concentrations and 238U/206Pb
Benemerito 245.4 ± 4.6 ratios indicate that partial metamictization of the zircons has
Chaneque 243.4 ± 2.8 occurred (Fig. 3E). Therefore, zircons with high U concentra-
Jurassic tions were disregarded.
Muro 188.3 ± 4.0 Of the 40 analyzed zircons, 33 were selected for age calcula-
Tlapacoyan 163.5 ± 4.7 tions (<20% discordant, <5% error, <5% inversely discordant,

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
38 Coombs et al.

and <1000 ppm U; Fig. 3E). Four inherited cores were analyzed population of zircons yielded an age of 257 ± 5.2 Ma (n = 13;
and yielded Proterozoic (1018 ± 82 Ma and 755 ± 8 Ma) and MSWD = 3.7), which can be interpreted as the igneous crystal-
Early Permian (288 ± 5 Ma and 289 ± 4 Ma) ages. A weighted- lization age of the sample.
mean calculation for the youngest population of zircons yielded Tamaulipas-103. The sample from the Tamaulipas-103
an age of 254.6 ± 3.4 Ma (n = 29; MSWD = 3.4; three ages well is a granodiorite containing subhedral zircons that are 150–
rejected by the algorithm), which can be interpreted as the igne- 400 µm along their longest axis, with aspect ratios of 2:1–6:1,
ous crystallization age of the sample. and CL textures that reveal core and rim relationships (Fig. 4I).
Linares-1. The sample from the Linares-1 well is a grano- Of the 37 analyzed zircons, 28 were selected for age calcu-
diorite, containing euhedral to subhedral zircons that are 150– lations (<20% discordant, <5% error, <5% inversely discordant;
250 µm along their longest axis, with aspect ratios of 2:1–7:1, Fig. 3I). Two inherited cores were analyzed and yielded an Early
and CL textures that reveal core and rim relationships (Fig. 4F). Permian age (282 ± 10 Ma and 279 ± 8 Ma; Appendix DR1). A
The basement from this well has previously been dated to 112 ± weighted-mean calculation for the youngest population of zircons
5 Ma (K-Ar biotite; Lopez-Ramos, 1979). yielded an age of 247.9 ± 4.0 Ma (n = 26; MSWD = 1.8), which
Of the 40 analyzed zircons, 31 were selected for age calcula- can be interpreted as the igneous crystallization age of the sample.
tions (<20% discordant, <5% error, <5% inversely discordant; Erizo-1. The sample from the Erizo-1 well is a granite that
Fig. 3F). One inherited core was analyzed and yielded a late Car- has undergone subsequent hydrothermal alteration. It contains
boniferous (304 ± 8 Ma) age (see Appendix DR1 in the GSA subhedral zircons that are 100–300 µm along their longest axis,
Data Repository1). A weighted-mean calculation for the concor- with an aspect ratio of 2:1–5:1, and CL textures that display no
dant grains yielded an age of 254.3 ± 4.6 Ma (n = 29; MSWD = evidence of inherited cores (Fig. 4J).
4.0; one age was rejected by the algorithm), which can be inter- It is apparent from relationships between U concentrations
preted as the igneous crystallization age of the sample. and the 238U/206Pb ratios that partial metamictization of the zir-
Chaneque-1. The sample from the Chaneque-1 well is a cons has occurred (Fig. 3J). Therefore, those with high U concen-
granitoid that has undergone subsequent hydrothermal alteration. trations were not considered in age calculations.
It contains subhedral zircons that are 150–300 µm along their Of the 49 analyzed zircons, 38 were selected for age calcula-
longest axis, with aspect ratios of 3:1–7:1, and CL textures that tions (<20% discordant, <5% error, <5% inversely discordant,
display no evidence of inherited cores (Fig. 4G). The basement and <800 ppm U; Fig. 3J). A weighted-mean calculation for the
from this well has been previously dated to 133 ± 5 Ma (K-Ar concordant grains yielded an age of 249.8 ± 2.7 Ma (n = 37;
K-feldspar; Lopez-Ramos, 1979). MSWD = 1.8; one age rejected by the algorithm), which can be
Of the 38 analyzed zircons, 28 were selected for age calcula- interpreted as the igneous crystallization age of the sample.
tions (<20% discordant, <5% error, <5% inversely discordant; Pinonal-1 (2). The second sample from the Pinonal-1 well is
Fig. 3G). A weighted-mean calculation for the zircons yielded a granodiorite that has undergone subsequent hydrothermal altera-
an age of 243.4 ± 2.8 Ma (n = 26; MSWD = 1.9; two ages were tion. It contains euhedral to subhedral zircons that are 200–400 µm
rejected by the algorithm), which can be interpreted as the igne- along their longest axis, with aspect ratios of 2:1–10:1, and CL tex-
ous crystallization age of the sample. tures that reveal core and rim relationships (Fig. 4K).
Nayade-1. The sample from the Nayade-1 well is a biotite Of the 31 analyzed zircons, 27 were selected for age calcula-
granite that has undergone subsequent hydrothermal alteration. tions (<20% discordant, <5% error, <5% inversely discordant;
It contains subhedral zircons that are 100–400 µm along their Fig. 3K). Three inherited cores were analyzed and yielded Meso-
longest axis, with aspect ratios of 1:1–5:1, and CL textures that proterozoic (1235 ± 153 Ma), Cambrian (509 ± 7.8 Ma), and
reveal core and rim relationships (Fig. 4H). Ordovician (478 ± 6.8 Ma) ages (Appendix DR1). A weighted-
It is apparent from relationships between U concentrations mean calculation for the youngest population of zircons yielded
and 238U/206Pb ratios that partial metamictization of the zircons an age of 249.8 ± 3.2 Ma (n = 22; MSWD = 1.6; two ages rejected
has occurred (Fig. 3H). Therefore, zircons with high U concen- by the algorithm), which can be interpreted as the igneous crys-
trations were disregarded. tallization age of the sample.
Of the 35 analyzed zircons, 16 were selected for age calcula- Cupelado-1. The sample from the Cupelado-1 well is a
tions (<20% discordant, <5% error, <5% inversely discordant, biotite granitoid, containing euhedral to subhedral zircons that
and <1000 ppm U; Fig. 3H). Three inherited cores were analyzed are 100–300 µm along their longest axis, with aspect ratios of
and yielded Neoproterozoic (558 ± 4 Ma), Cambrian (538 ± 1:1–5:1. The CL images revealed distinct populations of igneous
6 Ma), and Devonian (397 ± 6 Ma) ages (Appendix DR1 [see zircons that were defined by their luminescence, with evidence of
footnote 1]). A weighted-mean calculation for the youngest resorption in some of the grains (Fig. 4L).
Of the 19 analyzed zircons, 17 were selected for age calcula-
tions (<20% discordant, <5% error, <5% inversely discordant;
Fig. 3L). These 14 ages can be subdivided into three groups on
1
Data Repository Item, Geochronology appendix. Please visit https://doi
.org/10.1130/SPE.S.12296039 to access the supplemental material, and contact the basis of their age. The weighted-mean calculation for the
editing@geosociety.org with any questions. oldest population of zircons yielded a Mesoproterozoic age of

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 39

1418 ± 41 Ma (n = 7; MSWD = 0.28; one age rejected by the Fig. 3P). The weighted-mean calculation for the oldest popula-
algorithm). A second population of zircons yielded an Early tion of inherited cores revealed a Mesoproterozoic age of 1149
Permian age of 284.7 ± 7.7 Ma (n = 4; MSWD = 0.11). The ± 34 Ma (n = 9; MSWD = 1.02), characterized by CL textures
youngest population of zircons yielded a Late Permian age of typical of igneous zircons (Fig. 4P; oscillatory zoning). A sec-
262.7 ± 4.5 Ma (n = 5; MSWD = 0.86), which can be interpreted ond population of inherited cores yielded a Neoproterozoic age
as the igneous crystallization age of the sample. of 970 ± 53 Ma (n = 9; MSWD = 1.2) and CL textures more
Plan de Las Hayas-1 (1). The first sample from the Plan de typical of metamorphic zircons (Fig. 4P; homogeneous, with no
Las Hayas-1 well is a granodiorite, containing subhedral, pris- zoning). A weighted mean calculation for the youngest popula-
matic zircons that are 80–200 µm along their longest axis, with tion of zircons yielded an age of 261 ± 4.9 Ma (n = 17; MSWD =
aspect ratios of 1:1–4:1, and CL textures that display no evidence 4.4), which can be interpreted as the igneous crystallization age
of inherited cores (Fig. 4M). The basement associated with this of the sample.
well has been previously dated to 312 ± 25 Ma (K-Ar K-feldspar;
Lopez-Ramos, 1979). Jurassic
Of the 29 analyzed zircons, 24 were selected for age cal- Muro-2. The sample from the Muro-2 well is a micro-
culations (<20% discordant, <5% error, <5% inversely discor- diorite that contains euhedral-subhedral zircons that are 50–
dant; Fig. 3M). A weighted-mean calculation for the concor- 120 µm along their longest axis, with aspect ratios of 2:1–7:1.
dant grains yielded an age of 261.0 ± 4.1 Ma (n = 24; MSWD The CL images revealed core and rim relationships and evi-
= 2.8), which can be interpreted as the igneous crystallization dence of resorption features (Fig. 4Q). Basement samples from
age of the sample. this well have previously been dated to 153 ± 11 Ma and 178 ±
Plan de las Hayas-1 (2). The second sample from the Plan 11 Ma (K-Ar biotite; Lopez-Ramos, 1979).
de Las Hayas-1 well is a quartz diorite, containing euhedral, pris- Of the 21 analyzed zircons, 13 were selected for age calcula-
matic zircons that are 150–250 µm along their longest axis, with tions (<20% discordant, <5% error, <5% inversely discordant;
aspect ratios of 2:1–10:1. The CL images revealed complex tex- Fig. 3Q). Two inherited cores were analyzed and yielded Meso-
tures, with evidence of the zircons being partially resorbed but proterozoic (1430 ± 24 Ma) and Devonian (398 ± 3.2 Ma) ages
no inherited cores (Fig. 4N). The basement associated with this (Appendix DR1). A weighted-mean calculation for the youngest
well has previously been dated to 312 ± 25 Ma (K-Ar K-feldspar; population of zircons yielded an age of 188.3 ± 4.0 Ma (n = 11;
Lopez-Ramos, 1979). MSWD = 5.4), which can be interpreted as the igneous crystal-
Of the 30 zircons analyzed, 25 were selected for age calcula- lization age of the sample. The high MSWD recorded in this
tions (<20% discordant, <5% error, <5% inversely discordant; sample was likely caused by common lead contamination along
Fig. 3N). A weighted-mean calculation for the zircons yielded an fractures, as evidenced by the bright areas of the zircons (along
age of 251.7 ± 5.2 Ma (n = 24; MSWD = 3.0; one age rejected by their c axis) in the CL images (Fig. 4Q).
the algorithm), which can be interpreted as the igneous crystal- Tlapacoyan-1. The sample from the Tlapacoyan-1 well is
lization age of the sample. a trachyte that contains euhedral to subhedral zircons that are
Paso de Ovejas-2. The sample taken from the Paso de 80–200 µm along their longest axis, with aspect ratios of 2:1–7:1,
­Ovejas-2 well is a quartz monzodiorite. It contained elongated, and CL textures that reveal core and rim relationships (Fig. 4R).
euhedral, prismatic zircons that were 100–300 µm along their lon- The basement from this well has previously been dated to 179 ±
gest axis, with aspect ratios of 2:1–10:1. The CL images revealed 14 Ma (K-Ar on biotite; Lopez-Ramos, 1979).
core and rim relationships and resorption features (Fig. 4O). Of the 30 analyzed zircons, 19 were selected for age calcula-
Of the 30 analyzed zircons, 23 were selected for age calcula- tions (<20% discordant, <5% error, <5% inversely discordant;
tions (<20% discordant, <5% error, <5% inversely discordant; Fig. 3R). Two inherited cores were analyzed and yielded Neo-
Fig. 3O). Two inherited cores were analyzed and yielded an Early proterozoic (838 ± 60 Ma) and Late Permian (263 ± 7 Ma) ages
Permian age (297 ± 9 Ma and 298 ± 7 Ma; Appendix DR1). A (Appendix DR1). A weighted mean calculation for the youngest
weighted-mean calculation for the youngest population of zir- population of zircons yielded an age of 163.5 ± 4.7 Ma (n = 16;
cons yielded an age of 256.7 ± 5.5 Ma (n = 21; MSWD = 2.3; MSWD = 3.6; one age rejected by the algorithm), which can be
two ages rejected by the algorithm), which can be interpreted as interpreted as the igneous crystallization age of the sample.
the igneous crystallization age of the sample.
Orizaba-1. The sample from the Orizaba well is a monzo- MAJOR- AND TRACE-ELEMENT GEOCHEMISTRY
granite that contains euhedral to subhedral zircons that are 100–
400 µm along their longest axis, with aspect ratios of 2:1–5:1. Analytical Methods
The CL images revealed zircons that are characterized by igne-
ous rims and cores that are both igneous and metamorphic in ori- Sample preparation and analysis were carried out at Car-
gin (Fig. 4P). diff University, Wales. Veins and weathered surfaces were first
Of the 52 analyzed zircons, 35 were selected for age calcula- removed, and then the samples were crushed using a steel jaw
tions (<20% discordant, <5% error, <5% inversely discordant; crusher and powdered using an agate Tema® mill. Powdered

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
40 Coombs et al.

samples were then digested by fusion in platinum crucibles on a Trace Elements


Claisse Fluxy automated fusion system using 0.1 ± 0.0005 g of Chondrite-normalized rare earth element (REE) and normal
sample with 0.4 ± 0.0005 g of lithium tetraborate flux. Major-ele- mid-ocean-ridge basalt (N-MORB)–normalized multi-­element
ment abundances were determined using a JY Horiba Ultima 2 diagrams are shown in Figure 6. The Early Permian, Late ­Permian–
inductively coupled plasma–optical emission spectrometer (ICP- Early Triassic, and undated crystalline basement of the western
OES), while trace-element abundances were measured using a Gulf of Mexico displayed enriched REE trends (27×, 31×, and
Thermo X7 series ICP-MS. Accuracy and precision of the data 40× chondrite, respectively), with enriched chondrite-normalized
were assessed using the international reference material NIM-G. light (L) REE signatures when compared to heavy (H) REE signa-
Further details on analytical procedures can be found in McDon- tures (La/SmCh = 2.6–7.8). The middle (M) REEs were commonly
ald and Viljoen (2006). defined by concave trends (Dy/Dy* = 0.42–0.80), and the HREEs
were flat to positively sloping (Ho/LuCh 0.54–3.19). Enrichment
Results in Th, as well as negative anomalies in Nb (Nb/Nb* = 0.03–0.26),
Ta (Ta/Ta* = 0.05–0.39), and Ti (Ti/Ti* = 0.14–0.53), were preva-
The geochronological results revealed the more mafic, hyp- lent throughout the samples.
abyssal basement samples from the Muro-2 and Tlapacayon-1 The basement samples from the western Gulf of Mexico
wells to be Jurassic in age. It is therefore apparent that they are were compared in tectonic discrimination diagrams (Fig. 7) with
unrelated to the intermediate-felsic intrusive rocks of the west- Carboniferous–Triassic granitoids from the Totoltepec Pluton,
ern Gulf of Mexico, which yielded Permian–Triassic ages. The Cozahuico granite, and La Carbonera stock of southern Mexico
geochemical analysis of this study focused on the more prevalent (Fig. 1; Kirsch 2012), the Chiapas Massif of southeastern Mexico
Permian basement, but it is possible that the few Jurassic samples (Fig. 1; Weber et al., 2005; Estrada-Carmona et al., 2012), and the
were related to coeval volcanism associated with the Nazas arc Colombian Andes (Vinasco et al., 2006; Cochrane et al., 2014).
discussed above. Significant overlap can be observed between the Late Perm-
ian–Early Triassic crystalline basement of the western Gulf of
Element Mobility Mexico and granitoids of comparable ages in the Chiapas Massif,
Petrographic observations indicated that the majority of the indicating similar sources and possible tectonomagmatic setting.
samples were relatively unaltered. These observations were sup- The Early Permian crystalline basement displays similar trends,
ported by low loss of ignition (LOI) values that ranged between although there are some examples that appear to have more of an
0.60 and 4.17 wt% (Table 3). The samples that displayed signs of affinity to the Carboniferous–Permian granitoids of the Totoltepec
hydrothermal alteration in thin section can be associated with the Pluton (Fig. 7). These older granitoids plot firmly in the volcanic
highest LOI values (5.42–6.84 wt%). arc fields in all of the discrimination plots, whereas the younger
Assessment of mobility using conventional element mobil- Permian–Triassic granitoids plot closer to the syncollisional
ity plots (i.e., an immobile element such as Zr plotted against boundary. The Permian–Triassic granitoids from the Colombian
other elements) was not suitable in our study. These plots work Andes do not show such strong relationships with the western
under the assumption of a homogeneous source region, but our Gulf of Mexico and Chiapas, but some overlap is still observed.
samples have a large spatial distribution and temporal span that The tectonic discrimination diagrams suggest an affiliation
make the validity of this assumption unlikely. However, the to volcanic arc magmatism with potentially a minor syncolli-
petrographic observations coupled with low LOI values suggest sional component in some of the younger Permian–Triassic gran-
that there has not been substantial element mobility in most of itoids. However, these diagrams are not effective in defining all
the studied samples. tectonic settings, e.g., granitoids that form in postcollisional envi-
ronments (Pearce et al., 1984), due to the heterogeneous nature
Major Elements and Classification of the magmas that form in these settings. This will be discussed
Major-element compositions for the majority of the Early in more detail below.
Permian, Late Permian–Early Triassic, and undated samples were
similar and typical of intermediate-evolved granitoids, with SiO2, ZIRCON ISOTOPE GEOCHEMISTRY
TiO2, and MgO values ranges of 61.1–74.4 wt%, 0.1–0.8 wt%,
and 0.1–4.2 wt%, respectively (Table 3). This is supported by Analytical Methods
the quartz–alkali feldspar–plagioclase–feldspathoid (QAPF) and
total-alkali silica (TAS) classification diagrams (Fig. 5), which Chemical preparation and element separation were car-
show that the samples are predominantly granodiorite-granite in ried out on PicoTrace® clean benches at Laboratorio Ultra-
composition. Two of the Late Permian granitoids that yielded the limpio de Geología Isotópica, Departamento de Geología,
highest LOI values (Chaneque N6F1C2 and Pinonal N2F11C2) Centro de Investigación Científica y de Educación Superior de
showed anomalous major-element values e.g., SiO2 (Fig. 5). Ensenada (CICESE, Baja California). Individual zircon grains,
This indicates that the major elements in these samples had been previously dated by LA-ICP-MS (Fig. 4), were removed from
mobilized during subsequent alteration. mounts under a stereomicroscope with a needle and weighed into

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 41
TABLE 3. MAJOR AND TRACE ELEMENT GEOCHEMISTRY
Sample Plan de las Hayas (1) Arenque Paso de Oro Benemerito Trincheras Pinonal (1) Linares Paso de Ovejas
Latitude 19°46'15.60"N 24°57'2.07"N 24°43'14.62"N 25°40'45.91"N 24°57'2.07"N 25°40'45.91"N 24°43'14.62"N 19°15'33.17"N
Longitude 96°37'18.83"W 99°26'29.23"W 99°22'23.28"W 99°51'43.91"W 99°26'29.23"W 99°51'43.91"W 99°22'23.28"W 96°24'54.03"W
Major elements (wt%)
SiO2 62.73 66.94 65.05 71.01 68.78 56.03 67.02 67.19
TiO 0.78 0.47 0.52 0.21 0.50 0.90 0.19 0.54
2

Al2O3 16.10 15.04 16.06 14.40 15.66 13.43 11.82 15.32


Fe2O3 5.60 3.68 4.71 1.52 3.62 7.38 2.27 3.98
MnO 0.08 0.04 0.06 0.04 0.06 0.08 0.03 0.03
MgO 2.36 1.01 2.03 0.66 1.07 4.11 0.80 1.61
CaO 2.58 2.65 4.07 1.09 1.92 4.31 9.35 2.07
Na2O 4.53 3.87 2.95 3.62 3.77 3.47 3.02 3.21
K2O 2.54 5.63 3.43 5.64 3.28 4.43 3.90 4.61
P2O5 0.21 0.12 0.10 0.08 0.20 0.10 0.07 0.15
LOI 3.54 1.77 2.28 1.43 1.48 6.84 1.98 2.36
Total 101.05 101.22 101.25 99.69 100.32 101.07 100.45 101.09
Trace elements (ppm)
V 106.16 86.26 101.49 24.01 46.84 207.32 23.07 71.52
Rb 51.75 99.90 74.64 212.68 130.98 90.17 93.35 78.75
Sr 688.34 337.77 314.68 312.93 383.62 138.89 337.55 404.50
Y 15.06 18.64 20.99 11.55 19.732 12.62 20.60 10.69
Zr 165.91 145.67 83.25 83.70 50.98 151.93 89.96 130.09
Nb 6.36 8.76 5.73 7.52 11.73 2.93 1.93 8.85
Cs 5.34 4.78 3.26 6.89 13.14 3.53 1.01 1.86
Ba 414.91 923.83 775.31 1539.42 1048.80 315.81 486.60 739.97
La 18.58 23.05 14.933 19.51 43.28 7.73 15.70 18.58
Ce 37.84 44.92 0.09 39.56 83.40 16.43 30.84 41.26
Pr 4.95 5.28 4.05 4.78 9.71 2.04 4.09 5.07
Nd 19.82 18.76 15.88 16.67 34.67 8.01 15.52 18.71
Sm 4.04 3.61 3.65 3.36 6.11 2.01 3.30 3.54
Eu 1.12 1.05 1.21 0.82 1.66 0.67 0.93 0.99
Gd 3.78 3.39 3.56 2.70 5.22 2.14 3.21 2.92
Tb 0.48 0.48 0.55 0.35 0.65 0.36 0.47 0.38
Dy 2.66 2.85 3.49 1.87 3.32 2.32 3.01 1.95
Ho 0.46 0.56 0.67 0.32 0.60 0.46 0.60 0.34
Er 1.40 1.79 2.04 0.98 1.83 1.42 1.99 1.06
Tm 0.20 0.29 0.32 0.16 0.28 0.24 0.31 0.16
Yb 1.34 1.90 2.07 1.06 1.75 1.57 2.05 1.12
Lu 0.20 0.33 0.33 0.18 0.30 0.26 0.32 0.18
Hf 4.36 4.33 2.34 2.71 6.81 4.37 2.60 3.59
Ta 0.53 0.85 0.47 0.48 0.99 0.26 0.17 0.74
Pb 4.59 7.70 8.82 48.14 21.09 4.36 12.57 8.74
Th 2.92 4.95 3.95 15.80 14.99 2.91 5.01 2.56
U 1.38 2.14 1.31 5.18 3.60 0.83 1.38 0.72
Sample Chaneque Erizo Nayade Tamaulipas Cupelado Plan de las Hayas (2) Orizaba Muro
Latitude 24°34'19.97"N 22°10'12.03"N 22°45'16.58"N 22°24'31.66"N 20°18'43.29"N 19°46'15.60"N 18°46'17.46"N 20°50'38.52"N
Longitude 98°38'2.21"W 97°31'6.47"W 97°36'14.99"W 97°57'13.29"W 97° 2'0.53"W 96°37'18.83"W 97° 4'43.88"W 97°26'46.19"W
Major elements (wt%)
SiO2 51.44 74.43 70.93 69.69 67.34 61.54 72.91 56.13
TiO2 1.36 0.24 0.33 0.49 0.31 0.68 0.12 0.77
Al2O3 16.72 11.88 13.31 15.08 14.21 14.28 13.46 18.17
Fe2O3 8.88 0.77 2.71 3.29 2.47 4.43 1.26 7.76
MnO 0.08 0.03 0.04 0.05 0.04 0.07 0.02 0.11
MgO 4.15 0.12 0.37 1.26 0.74 1.81 0.24 3.53
CaO 3.21 2.43 2.09 3.07 2.85 4.60 1.12 6.55
Na2O 3.37 2.11 2.26 4.09 3.62 5.31 3.88 3.25
K2O 5.59 5.08 4.67 1.91 3.99 2.48 4.76 1.54
P2O5 0.29 0.04 0.12 0.12 0.10 0.20 0.03 0.14
LOI 6.28 3.63 3.58 0.60 3.47 5.42 1.01 2.64
Total 101.36 100.76 100.40 99.64 99.13 100.81 98.81 100.59
(Continued)

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
42 Coombs
TABLE 3. MAJOR AND et al. GEOCHEMISTRY (Continued)
TRACE ELEMENT
Sample Chaneque Erizo Nayade Tamaulipas Cupelado Plan de las Hayas (2) Orizaba Muro
Latitude 24°34'19.97"N 22°10'12.03"N 22°45'16.58"N 22°24'31.66"N 20°18'43.29"N 19°46'15.60"N 18°46'17.46"N 20°50'38.52"N
Longitude 98°38'2.21"W 97°31'6.47"W 97°36'14.99"W 97°57'13.29"W 97° 2'0.53"W 96°37'18.83"W 97° 4'43.88"W 97°26'46.19"W
Trace elements (ppm)
V 205.18 24.79 33.68 54.59 13.24 68.30 16.79 173.27
Rb 137.92 133.06 88.36 55.80 89.96 53.29 109.39 17.83
Sr 419.40 115.32 314.02 470.37 303.09 183.47 288.25 382.21
Y 10.90 15.48 9.43 13.42 10.72 14.38 6.93 17.42
Zr 150.16 117.75 129.03 166.74 135.72 206.38 90.81 113.22
Nb 6.18 9.90 7.82 9.62 6.19 8.31 7.26 3.92
Cs 15.69 6.97 5.11 0.89 12.20 3.76 1.62 0.64
Ba 1125.66 1468.05 1081.40 1586.14 863.20 334.76 522.18 459.26
La 21.31 24.28 7.81 51.23 22.18 25.88 9.34 9.76
Ce 43.32 47.55 14.90 99.77 40.56 54.42 20.05 21.49
Pr 5.66 5.44 1.74 11.91 5.16 7.24 2.48 2.93
Nd 21.51 18.04 6.23 41.84 19.34 28.28 8.89 12.23
Sm 4.08 3.19 1.46 6.21 3.49 5.35 1.69 2.86
Eu 1.39 0.89 0.73 1.75 1.17 1.45 0.59 1.00
Gd 3.46 2.99 1.51 4.76 3.10 4.52 1.45 2.90
Tb 0.43 0.39 0.24 0.50 0.36 0.52 0.18 0.45
Dy 2.23 2.16 1.45 2.36 1.80 2.56 1.08 2.80
Ho 0.35 0.42 0.29 0.40 0.30 0.43 0.19 0.52
Er 0.95 1.35 0.90 1.23 0.91 1.18 0.63 1.64
Tm 0.13 0.24 0.16 0.19 0.13 0.17 0.12 0.27
Yb 0.72 1.64 1.05 1.15 0.84 1.05 0.82 1.71
Lu 0.1
1 0.28 0.19 0.19 0.13 0.17 0.16 0.28
Hf 3.68 3.77 3.54 4.27 3.39 5.25 2.98 2.89
Ta 0.38 0.88 0.74 0.54 0.53 0.70 0.55 0.24
Pb 7.31 54.86 13.66 13.64 11.62 5.26 1.71
1 5.62
Th 7.16 1.51
1 7.49 9.81 3.46 4.66 3.14 1.1
1
U 2.70 1.67 1.75 2.19 0.59 1.59 0.91 0.31

Sample Tlapocoyan Magdalena Benemerito Jurel Trincheras Standards


Latitude 19°59'49.64"N 20°23'40.72"N 25°40'45.91"N 22° 9'4.42"N 24°57'2.07"N NIM-G NIM-G r.s.d.
97°38'16.60"W official measured
Longitude 97° 8'47.80"W 97° 4'19.43"W 99°51'43.91"W 99°26'29.23"W
Major elements (wt%)
SiO2 61.34 47.17 69.55 51.52 69.73 75.70 75.75 0.04
TiO2 0.44 1.66 0.51 0.84 0.44 0.09 0.10 6.71
Al2O3 16.23 13.02 14.50 18.73 14.82 12.08 12.27 1.08
Fe2O3 2.50 8.36 2.09 4.79 3.37 2.02 2.04 0.77
MnO 0.05 0.12 0.05 0.15 0.07 0.02 0.02 4.07
MgO 0.94 7.72 0.82 4.25 1.00 0.06 0.02 73.24
CaO 4.03 9.65 1.24 5.70 1.89 0.78 0.76 1.82
Na2O 4.90 1.81 3.67 3.94 3.41 3.36 3.25 2.33
K2O 6.13 6.29 5.26 2.16 3.73 4.99 5.03 0.50
P2O5 0.23 1.48 0.1
1 0.38 0.18 0.01 0.01 0.62
LOI 3.99 4.17 1.40 6.59 1.95
Total 100.79 101.45 99.21 99.05 100.57
Trace elements (ppm)
V 53.14 183.21 58.79 138.64 47.89 2.00 0.03 137.65
Rb 175.35 80.69 45.67 36.85 160.76 320.00 326.27 1.37
Sr 617.70 1551.75 348.02 759.47 354.30 10.00 7.52 20.01
Y 14.83 35.73 10.10 31.82 17.12 143.00 145.26 1.11
Zr 14.27
1 142.29 245.90 256.75 191.07 300.00 279.51 5.00
Nb 10.24 30.01 4.71 11.64 10.09 53.00 48.08 6.88
Cs 6.92 10.62 1.33 1.73 11.33 N.D. 0.87 10.06
Ba 1084.14 5427.39 544.82 1175.15 1531.09 120.00 105.82 8.88
La 25.23 242.04 15.78 36.32 43.13 109.00 111.73 1.75
Ce 50.10 491.43 31.25 79.89 82.87 195.00 200.55 1.98
Pr 6.03 61.06 3.76 10.25 9.54 N.D. N.D. N.D.
Nd 21.20 222.73 13.31 39.12 32.96 72.00 71.05 0.94
Sm 4.15 33.95 2.43 7.44 5.56 15.80 14.49 6.14
Eu 0.83 8.82 0.98 2.12 1.51 0.35 0.33 4.55
Gd 3.27 24.15 2.21 7.05 4.81 14.00 14.99 4.83
(Continued)

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 43

TABLE 3. MAJOR AND TRACE ELEMENT GEOCHEMISTRY (Continued)


Sample Tlapocoyan Magdalena Benemerito Jurel Trincheras Standards
Latitude 19°59'49.64"N 20°23'40.72"N 25°40'45.91"N 22° 9'4.42"N 24°57'2.07"N NIM-G NIM-G r.s.d.
official measured
Longitude 97° 8'47.80"W 97° 4'19.43"W 99°51'43.91"W 97°38'16.60"W 99°26'29.23"W
Tb 0.43 2.13 0.29 0.95 0.57 3.00 2.54 11.63
Dy 2.30 7.85 1.62 5.27 3.04 17.00 17.94 3.80
Ho 0.40 1.10 0.30 0.99 0.53 N.D. N.D. N.D.
Er 1.23 3.07 0.89 3.05 1.62 N.D. N.D. N.D.
Tm 0.21 0.33 0.15 0.48 0.25 2.00 2.15 5.15
Yb 1.37 1.96 0.99 2.92 1.51 14.20 14.04 0.79
Lu 0.23 0.29 0.17 0.49 0.26 2.00 2.09 3.19
Hf 3.79 2.81 5.66 6.36 5.13 12.00 11.57 2.56
Ta 0.75 2.36 0.31 0.80 0.97 4.50 4.07 7.16
Pb 38.30 45.54 8.45 17.74 17.70 40.00 30.87 18.21
Th 14.04 23.64 1.92 3.64 11.90 50.00 47.93 2.99
U 3.95 5.59 1.02 0.97 3.50 15.00 15.86 3.93
LOI—loss on ignition; r.s.d.—relative standard deviation.

­ icrocapsules. The zircons were then washed several times with


m ratio of JMC-475 measured over the last 3 yr during a total of six
warm 7 M HNO3, and then with cold, concentrated HNO3. Next, analytical sessions is 0.282149 ± 0.000025 (n = 41). Data reduc-
a 180Hf-176Lu spike was pipetted into the microcapsules before tion was carried out using IsotopeHf®, an R-based data reduction
adding ~0.5 mL of concentrated HF. Microcapsules were heated software package that transforms raw mass spectrometry data into
with HF as a pressure medium in a Parr® bomb for 6 d at 180 meaningful isotopic ratios, including all the necessary corrections
°C. After digestion, the samples were dried down on a hotplate. for spiked samples (González-Guzmán et al., 2016).
The closed microcapsules were heated overnight in 6 M HCl, and
then the drying process was repeating to facilitate sample-spike Results
equilibration. Sample residues were then dissolved in ~0.5 mL
of 1 M HCl and loaded to microcolumns filled with ~160 µL of Five zircon grains from Permian plutonic rocks that had been
Eichrom Ln-spec® resin. Lu + Yb, and Hf were eluted follow- dated by LA-ICP-MS were chosen for isotope-dilution MC-ICP-
ing a single-column separation procedure adopted from Nebel- MS analyses. These included two zircons from the Orizaba-1
Jacobsen et al. (2005). sample, two from the Paso de Ovejas-2 sample, and one from
The determination of Lu and Hf isotope ratios was carried the Linares-1 sample. The analyzed zircons were characterized
out on a Thermo Neptune Plus® multicollector (MC) ICP-MS by typical igneous textures (oscillatory and sector zoning), with
installed at the Centro de Geociencias, Universidad Nacional no indication of inherited cores. However, all three samples did
Autonóma de México, in Juriquilla, Querétaro, México, on Far- contain zircons, not used for Hf analysis, that displayed inherited
aday cups in static mode (González-Guzmán et al., 2016). The cores (see Geochronology Results section).
sample solutions were introduced to the plasma via an Aridus® Paso de Ovejas-2 zircons yielded the lowest initial eHf values
desolvating sample-introduction system using an Ar carrier gas of –12.1 to –10.4, with corresponding two-step Hf depleted mantle
and a blended Ar + N2 sweep gas. The Hf fraction was taken up model ages, TDM(Hf), of 1.9 and 1.82 Ga (Table 4). The Orizaba-1
with 1 mL of 0.56 M HNO3–0.24 M HF solution, and the Lu frac- sample had slightly higher eHf(t) of –9.3 and –6.0, with correspond-
tion was taken up with 0.6 mL of 0.1 M HNO3 solution. For Lu ing TDM(Hf) of 1.75 and 1.56 Ga (Table 4). Differences observed in
isotope data acquisition, one block of 40 cycles with 4 s integra- the Lu-Hf isotopic signatures between the two analyzed zircons
tion time each was performed. The 177Hf intensity was measured from the Orizaba sample likely reflect hetero­geneities in the source,
to monitor for isobaric interference of 176Hf on the 176Lu signal. as evidenced by the diverse populations of inherited zircons found
For the mass bias correction, each sample was doped with ~10 ppb in the sample (Geochronology Results section). Zircons from the
of Re, and the masses 185 and 187 were measured simultane- Lineras-1 sample yielded highest eHf(t) values of –4.6, with corre-
ously. For Hf isotope data acquisition, eight blocks with 10 cycles sponding TDM(Hf) of 1.47 Ga (Table 4).
per block and an integration time of 4 s per cycle were measured.
Isobaric interferences of 176Yb and 176Lu on the 176Hf signal were DISCUSSION
monitored by measuring 172Yb and 175Lu. Moreover, 181Ta and
182
W were measured to monitor for isobaric interferences of 180Ta I-Type versus S-Type
and 180W on the spiked isotope 180Hf. To examine the accuracy of
Hf isotope measurement, a 50 ppb JMC-475 Hf standard solution The alumina saturation index (ASI) of Chappell and
was measured after every 4–5 unknowns. The average 176Hf/177Hf White (1974) distinguishes between metaluminous (I-type) and

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
44 Coombs et al.

A Q 10 B Syenite
Chan-1
Syenite Alkali-
Granite
1 Pin- 1 (1)
Granite
8 Syeno-diorite

Na2O+K2O (wt%)
2 Ijolite
Gabbro
6 Tonalite/
Granodiorite
Gabbro
Diorite
4
4

lic
Al c
b- ali
ka
A B 5

Su Alk
Gabbro
2
3 6

0
35 40 45 50 55 60 65 70 75 80
11 7 SiO2 (wt%)
10 9 Pin- 1 (1) 8
12 13 14 15 16
A P
21 20 19 Chan-1 18 17

22 25
23 24

26

Early Permian
Late Permian
F No age

Figure 5. Classification diagrams for the basement core samples of the western Gulf of Mexico. (A) Quartz–alkali f­ eldspar–
plagioclase–feldspathoid (QAPF) diagram. CIPW norms were first calculated, and then Q, A, P, F percentages were re-
calculated to 100%. 1—quartzolite; 2—quartz-rich granitoids; 3—alkali feldspar granite; 4—granite (A—­syenogranite,
B—monzogranite); 5—granodiorite; 6—tonalite;, 7—quartz diorite/quartz gabbro/quartz anorthosite; 8—quartz monzo-
diorite/quartz monzogabbro; 9—quartz monzonite; 10—quartz syenite; 11—quartz–alkali feldspar–­syenite; 12—alkali
feldspar–syenite; 13—syenite; 14—monzonite; 15—monzodiorite/monzogabbro; 16—diorite/­gabbro/anorthosite; 17—
foid-bearing diorite, foid-bearing gabbro, foid-bearing anorthosite; 18—foid-bearing monzodiorite, foid-bearing monzo-
gabbro; 19—foid-bearing monzonite; 20—foid-bearing syenite; 21—foid-bearing alkali feldspar syenite; 22—foid sy-
enite; 23—foid monzosyenite; 24—foid monzodiorite, foid monzogabbro; 25—foid diorite, foid gabbro; 26—­foidolite.
(B) Total alkali–silica (TAS) diagram of Le Maitre et al. (1989). Field boundaries were taken from Wilson (1989).

p­ eraluminous (S-type) granitoids. The ASI values for the samples The ASI results indicate that the basement of the western
have been plotted against their U-Pb age in Figure 8. The Early Gulf of Mexico is predominantly I-type (metaluminous) in
Permian (n = 3) and Late Permian–Early Triassic samples (n = composition, with minor S-type (peraluminous) influence. The
13) that display no zircon inheritance yielded ASI values typical observation that the more peraluminous samples preserve zir-
of I-type granitoids (0.73–1.01 and 0.45–1.08, respectively). The con cores, whereas the metaluminous samples do not, is con-
Late Permian–Early Triassic samples with inherited zircon cores sistent with the zircon solubility model of Watson and Harrison
yielded higher ASI values (0.92–1.18), but still mainly I-type. (1983) for I- and S-type magmas. It is unclear from the ASI
The higher ASI values in the samples with inherited zircon cores values if the I-type granites are primary melts from juvenile
indicate a contribution from a more peraluminous (e.g., crustal mantle, or if they were derived from the remelting of preexist-
sedimentary) source. ing crustal igneous rocks.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
1000 Petrogenesis of the crystalline basement
100
along the western Gulf of Mexico 45
Early Permian Early Permian
Late Permian-Early Triassic Late Permian-Early Triassic
Not dated Not dated

100 10

10 1

1
A 0.1
B
La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Th Nb Ta La Ce Sr Nd Sm Hf Zr Eu Ti Gd Tb Dy Y Ho Er Tm Yb

Figure 6. (A) Chondrite-normalized rare earth element (REE) and (B) normal mid-ocean-ridge basalt (N-MORB)–normalized multi-element plot
for the basement samples from the western Gulf of Mexico. Chondrite normalizing values were taken from Sun and McDonough (1989), and
N-MORB normalizing values were taken from McDonough and Sun (1995).

500 100
WITHIN Early Permian
PLATE Permo-Triassic
GRANITES No age WITHIN
P-T Chiapas PLATE
P-T Colombia Cordillera GRANITES
Totoltepec (306-287 Ma) 10
Cozahuico Granite (270 Ma)
Carbonera Stock (275-273 Ma)
50 VOLCANIC SYN-COLLISIONAL
ARC GRANITES GRANITES
Nb (ppm)

Ta (ppm)

& SYN-COLLISIONAL
GRANITES
1

OCEAN OCEAN
5 RIDGE RIDGE
GRANITES VOLCANIC GRANITES
0.1 ARC GRANITES

0.5 0.01
1 10 Y (ppm) 100 1000 0.1 1 10 100
Yb (ppm)

1000 1000 SYN-COLLISIONAL


SYN-COLLISIONAL GRANITES
GRANITES
WITHIN
WITHIN PLATE
PLATE GRANITES
GRANITES
100 100
Rb (ppm)

Rb (ppm)

10 VOLCANIC VOLCANIC
ARC GRANITES OCEAN 10 ARC GRANITES OCEAN
RIDGE RIDGE
GRANITES GRANITES

1 1
1 10 100 1000 0.1 1 10 100
Y+Nb (ppm) Yb+Ta (ppm)
Figure 7. Nb-Y, Ta-Yb, Rb-(Y+Nb), and Rb-(Yb+Ta) tectonic discrimination diagrams (Pearce et al., 1984). P-T Chiapas—Permian–Triassic
Chiapas (Weber et al., 2005; Estrada-Carmona et al., 2012); P-T Colombia—Permian–Triassic Colombia (Vinasco et al., 2006; Cochrane et al.,
2014); Totoltepec Pluton, Cozahuico Granite, and La Carbonera Stock (Kirsch, 2012).

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
46 Coombs et al.

TABLE 4. Lu-Hf DATA FOR ZIRCONS


Lu-Hf isotopic ratios Age corrected† Model age§
Lu Hf 176
Hf/ 176
Hf/ ±2 s.e. ×106 εHf* Age Hf/
176 εHf (t) TDM(Hf)
Sample
(ppm) (ppm) 177
Hf 177
Hf (2σm) (Ma) (Ga)
number Hf (t)
177

Orizaba
N6F1C1-6 87 9257 0.001342 0.282449 5 -11.4 278 0.282442 -6.0 1.56
N6F1C1-35 57 6876 0.001181 0.282361 7 -14.5 267 0.282355 -9.3 1.75
Linares
N7F1C1-6 41 11651 0.000500 0.282497 11 -9.7 253 0.282495 -4.6 1.47
Paso de Ovejas
N6F9C2-11 23 8359 0.000398 0.282334 8 -15.5 254 0.282332 -10.4 1.82
N6F9C2-4 85 8701 0.001384 0.282295 5 -16.9 247 0.282288 -12.1 1.90
*Epsilon Hf is the deviation of 176Hf/177Hf of the sample relative to the chondritic uniform reservoir (CHUR) × 104. For the calculations present-day
CHUR values 176Hf/177HfCHUR(0) = 0.282785 and 176Lu/177HfCHUR = 0.0336 (Bouvier et al., 2008).
†176
Hf/177Hf (t) and εHf(t) were calculated using the 206Pb/238U age of the zircons.
§
Two-stage crustal residence model ages were calculated from the following equations:
tDM =(1/λ)ln(1+m)
m= {(176Hf/177Hf)DM - [(176Hf/177Hf)i,zinc +(176Lu/177Hf)avg. crust (eλt -1)]} / {(176Lu/177Hf)DM - (176Lu/177Hf)avg. crust}
Assumptions made: (176Lu/177Hf)avg. crust = 0.015 (Condie et al., 2005).
Present-day depleted mantle (DM) model is based on (176Hf/177Hf)DM = 0.283224 (Vervoort et al., 2000) and (176Lu/177Hf)DM = 0.03826
(Weber et al., 2010).

Mantle-Derived Melts versus Crustal Anatexis ing eHf(t) of –4.6 and TDM(Hf) of 1.47 Ga, is the only analysis that
might include a slightly younger juvenile component. It is likely
Although there were only a few samples analyzed, the that the Late Permian–Early Triassic granitoids from this study
strongly negative eHf(t) values (–12 to –6) of the Late Permian– formed through the fusion of Oaxaquia continental crust, because
Early Triassic granitoids of the western Gulf of Mexico are a their Hf isotope evolution trend, as expressed from TDM(Hf) model
significant result that indicates that the magmas from which the ages, is indistinguishable from those of Mesoproterozoic Oaxa-
zircons crystallized were mainly formed by continental anatexis. quia microcontinent into which they emplaced (Fig. 9; Weber et
The TDM(Hf) model ages of these samples suggest that this crust al., 2010; Weber and Schulze, 2014).
was probably juvenile in the early Mesoproterozoic or late Paleo- However, it is important to highlight the fact that this con-
proterozoic. The zircon analyzed from the Linares sample, yield- clusion is based on the Hf isotope analysis of only five zircon
grains from a study area spanning several hundreds of kilome-
ters. A more comprehensive Lu-Hf isotopic study of the base-
ment along the western Gulf of Mexico is needed to confirm
1.4 this anatectic origin.
Late Permian–Early Triassic granitoids from the Sierra Pinta,
1.2 S-type granite northern Mexico (Arvizu and Iriondo, 2011), Oaxaca (Ortega-
Obregón et al., 2014), and Colombia and Ecuador (Cochrane et
I-type granite al., 2014) have Lu-Hf isotope signatures that are similar to those
1.0
reported here (Fig. 9). However, they also display evidence for
mantle-crustal mixing, with some TDM(Hf) model ages that are
0.8
younger than the Paleoproterozoic Laurentian crust (Sierra Pinta)
and Mesoproterozoic Oaxaquia crust (Oaxaca) into which they
0.6 were emplaced.
E. Permian (with cores) Late Carboniferous–Early Permian granitoids of the Oaxaca
0.4 E. Permian (no cores) area (Ortega-Obregón et al., 2014) show clear evidence for man-
L. Permian (no cores)
L. Permian (with cores) tle-crust mixing and appear unrelated to the Late Permian–Early
0.2 Triassic granitoids of the western Gulf of Mexico (Fig. 9). The
200 220 240 260 280 300 geochronological similarities and the spatial proximity of the late
Carboniferous–Early Permian granitoids of the Oaxaca area and
Figure 8. Alumina saturation index (ASI) vs. age for the base- Early Permian granitoids of the western Gulf of Mexico suggest
ment cores of the western Gulf of Mexico. they may be analogous.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 47

Permian Sonora (Arvizu & Iriondo, 2011) Late Carboniferous-Early Permian Oaxaca
(Ortega-Obregón et al., 2014)
Permo-Triassic magmatism in Colombia
Permo-Triassic, Oaxaca
and Ecuador (Cochrane et al., 2014) (Ortega-Obregón et al., 2014)
Paleoproterozoic Sonora (Arvizu & Iriondo, 2011) "Typical Oaxaquia" (Weber et al., 2010)
16 Figure 9. eHf(t) vs. age plot for western
Gulf of Mexico samples selected for
Depleted Hf isotope analysis on zircons (thermal
12 Mantle Mo
del Age ionization mass spectrometry). Our re-
sults are compared to other laser abla-
8 tion–inductively coupled plasma–mass
a
7G spectrometry (LA-ICP-MS) data from
1.4 around Mexico, Colombia, and Ecuador.
4 a Intercept between eHf(t) and the depleted
a 5G
.56
G 1.7 Ga
mantle curve is based on a 176Lu/177Hf
CHUR Model age 1 1.9 average crustal ratio of 0.015 (Griffin et
0
ε

al., 2002). Values for chondritic uniform


reservoir (CHUR) are from Bouvier et
a
-4 2G al. (2008). The fields represent zircon
1.8 depleted mantle model ages, TDM(Hf),
from Permian–Triassic granitoids in
-8 Mexico, Ecuador, and Colombia, as
Orizaba 279±10Ma well as for older igneous events such as
Orizaba 268±4Ma the Paleoproterozoic Sierra Pinta and
-12 Linares 253±9Ma Oaxaquia continental crust.
Paso de Ovejas 247± 10Ma
Paso de Ovejas 254± 7Ma
-16
200 400 600 800 1000 1200 1400 1600 1800 2000

PETROGENESIS and Acatlán area (Kirsch, 2012) indicate that they formed in a
continental arc setting. This interpretation is consistent with the
The geochemistry of the granitoids from the western Gulf of Lu-Hf isotopic signatures for the late Carboniferous–Early Perm-
Mexico is similar to that of plutonic rocks of similar age intrud- ian granitoids of the Oaxaca area (Ortega-Obregón et al., 2014),
ing the Acatlán Complex (Kirsch, 2012), Oaxacan Complex
(Ortega-Obregón et al., 2014), and Chiapas Massif (Weber et al.,
2005). The probability density diagram of Figure 10 suggests that This Study Weber et al., 2005
Carboniferous–Triassic magmatism in Mexico was intermittent. Elías-Herrera et al., 2005 Kirsch et al., 2012
The earliest phase of magmatism is clearly documented between Ducea et al., 2004 Ortega-Obregon et al., 2014
311 and 286 Ma, but inherited cores reported in this study indi-
Formation of Pangea

cate that this magmatism may have initiated in the Mississippian Post-collisional
10 magmatism
(ca. 326 Ma; Supplementary Material 1 [footnote 1]). A second
and more prevalent phase of magmatism occurred between 274
and 243 Ma. This magmatism inherited zircons from the late 8 Relative probability
Carboniferous–Late Permian event, as well as from Oaxaquia
(Grenville).
Number

6
Jurassic reconstructions of western Pangea place Mexico Gondwanan arc
close to the Laurussia-Gondwana suture (Pindell and Kennan,
4
2009). This, along with geochronological similarities with Car-
boniferous–Permian magmatism in the Variscan and Alleghanian
orogenies and geochemical constraints presented in this study, 2
suggests that the Permian–Triassic magmatism of Mexico was
related to the final stages of the formation of Pangea.
0
230 250 270 290 310 330
Late Carboniferous–Early Permian Arc Age (Ma)

Figure 10. Probability density plot for Carboniferous-Triassic gran-


Trace-element signatures of the late Carboniferous–Early itoids from this study, Chiapas Massif, Sierra Pinta, Acatlán, and
Permian granitoids from the western Gulf of Mexico (this study) Oaxaca areas.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
48 Coombs et al.

which indicate mantle-derived melts mixing with continental and Thomas, 1989; Sedlock et al., 1993; Dickinson and Law-
crust. This arc system intruded into Oaxaquia continental crust, ton, 2001) and the latest magmatism in the late Carboniferous–
suggesting a Gondwanan affinity. Early Permian arc (286 Ma; Ortega-Obregón et al., 2014). We
In Ortega-Obregón et al. (2014), this continental arc system therefore propose that the late Carboniferous–Early Permian arc
was explained by the eastward-dipping subduction of the Pan- formed from the subduction of the Rheic Ocean on the margins
thalassa Ocean beneath Pangea. However, there is a strong cor- of Gondwana, prior to the final amalgamation of Pangea (Figs.
relation between the timing of the final amalgamation of Pan- 11A and 11C). After the closure of the Rheic Ocean, and col-
gea along the Ouachita-Marathon-Sonora suture ca. 290–280 lision between Laurussia and Gondwana had occurred, the arc
Ma (Pindell and Dewey, 1982; Pindell, 1985; Ross, 1986; Viele system shut down.

A- Late Carboniferous B- Permo-Triassic


Depth (km)

Suture Anatectic
Carboniferous arc (Blueschists and eclogites magmatism
of the Acatlán Complex?)
Oaxaquia Oaxaquia Crust
Rheic Ocean
0
Laurussia Gondwana Laurasia Gondwana

50 Lithospheric
Mantle Lithospheric Mantle

100
Asthenosphere
Asthenosphere
150
D- Early Jurassic?
Appalachian
Permo-Triassic
Pacific Arc SW Laurasia Alleghanian Suture
Ouachita Oaxaca displaced
to the south during
initial phase of
Marathon Pangea rifting.
Late Carboniferous-Early Permian
Gondwanan arc Oaxaca 300 km
Sonora
NW Gondwana E- Late Jurassic
Mexican post-collisional "anatectic
province", including basement of the West GoM
transform
D&E Western Gulf of Mexico (Tamaulipas
arch, Tuxpan, Santa Ana), Chiapas
Massif and Oaxaca
Panthalassa
Ocean
Oaxaca
Chiapas
Chiapas displaced to the
250 km N C-Pangea (Permo-Triassic) south and Western GoM
lineated. 300 km

Figure 11. (A) Late Carboniferous and (B) Permian–Triassic tectonic models for the formation of the granitoids along the western Gulf of
Mexico (GoM). (C) Schematic paleogeographic reconstruction of the western flank of the Pangea collision zone. The green stars represent
magmatism related to a continental arc in Sonora (Laurentian side) and Colombia (Gondwanan side). (D–E) Phases of shearing in the ­Permian–
Triassic anatectic province.

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 49

Late Permian–Early Triassic Postcollisional Magmatism Subsequent Tectonic Activity

The widespread Late Permian–Early Triassic granitoids of In many models of Pangea, overlap is observed between
the western Gulf of Mexico (this study), Oaxaca area (Ortega- central and southern Mexico and northwest South America (Pin-
Obregón et al., 2014), and Chiapas Massif (Weber et al., 2005; dell and Dewey, 1982; Pindell, 1985; Handschy et al., 1987; Pin-
Estrada-Carmona et al., 2012) display trace-element signatures dell and Kennan, 2001). Therefore, the terranes of central and
that are consistent with formation in a continental arc environ- southern Mexico are likely to be allochthonous in origin. This
ment. However, depleted Lu-Hf isotope signatures indicate that implies that the terranes into which the Carboniferous–Triassic
the dominant process of magma generation was by crustal ana- plutons of the Oaxaca area, Chiapas Massif, and western Gulf of
texis. This type of magmatism is more commonly associated Mexico were intruded had migrated from elsewhere during the
with continental collision environments. breakup of Pangea.
The Late Permian–Early Triassic granitoids of the Chiapas
Massif, western Gulf of Mexico, and Oaxaca area postdate the Opening of the Gulf of Mexico
final amalgamation of Pangea (ca. 290–280 Ma), and so they are A possible mechanism for terrane migration occurred during
unlikely to represent a syncollisional magmatic event. The gran- the Early Jurassic in association with the early stages of the open-
itoids may instead have been associated with a phase of post- ing of the Gulf of Mexico, along the Mojave-Sonora megashear
collisional magmatism (Figs. 11B and 11C) occurring during (Anderson and Schmidt, 1983; Böhnel, 1999; Pindell and Ken-
a period of thermal relaxation, after the final amalgamation of nan, 2001). This shear zone is proposed to have accommodated
Pangea, e.g., as seen in the Alleghanian orogeny ca. 300–275 Ma 700 km of left-lateral motion along the southwestern flank of the
(Heatherington et al., 2010; Mueller et al., 2014). Postcollisional North American plate (Anderson and Schmidt, 1983; Pindell and
granitoids of the western Gulf of Mexico postdate equivalents in Kennan, 2001, 2009). In this scenario, the terranes of central and
the Alleghanian orogeny by ~30 m.y. This lag period is consis- southern Mexico (including the Acatlán Complex and Oaxacan
tent with the diachronous assembly of Pangea, with Laurentia-­ Complex) originated in northeast Mexico (Fig. 11D).
Gondwana collision occurring in the Alleghanian orogeny In plate reconstructions of Pangea, the Yucatan block is posi-
ca. 335 Ma (Wortman et al., 1998) and along the Ouachita-­ tioned adjacent to Florida, and it was then displaced to the south
Marathon-Sonora segment ca. 290 Ma (Pindell and Dewey, 1982; in a rotational motion during the main phase of the opening of the
Pindell, 1985; Ross, 1986; Viele and Thomas, 1989; Sedlock et Gulf of Mexico in the Late Jurassic (Pindell and Dewey, 1982;
al., 1993; Dickinson and Lawton, 2001). Pindell, 1985; Schouten and Klitgord, 1994; Dickinson and Law-
Granitoids that form in such postcollisional settings often ton, 2001; Jacques et al., 2004; Bird et al., 2005; Imbert, 2005;
inherit arc-like trace-element signatures from previous subduc- Imbert and Phillippe, 2005; Pindell et al., 2005). This movement
tion events (e.g., Pearce et al., 1984; Grimes et al., 2015). There- likely affected the postcollisional granitoids of the western Gulf
fore, the subduction-related trace-element signatures observed in of Mexico and Chiapas Massif, shearing the basement and dis-
the Late Permian–Early Triassic granitoids of Chiapas Massif, placing it to the south (Fig. 11E).
western Gulf of Mexico, and the Oaxaca area may well have been This tectonic activity associated with the breakup of Pangea
inherited from the late Carboniferous–Early Permian arc (dis- provides a mechanism for migration of the late Carboniferous–
cussed above), as well as Oaxaquia, which comprises the lower Early Permian arc granitoids and Late Permian–Early Triassic
continental crust. This hypothesis is supported by the widespread postcollisional granitoids: The rocks were displaced to the south,
Oaxaquia and late Carboniferous–Early Permian zircon inheri- away from the Laurentian-Gondwanan suture from which they
tance observed in the Late Permian–Early Triassic igneous event originated (Figs. 11D and 11E).
(this study).
CONCLUSIONS
Late Permian–Early Triassic Arc
This contribution has shown that the crystalline basement of
Late Permian–Early Triassic granitoids found in the Sierra the western Gulf of Mexico formed as a result of three distinct
Pinta, Sonora, northeast Mexico (Arvizu and Iriondo, 2011) and magmatic events, replacing and reworking preexisting Oaxaquia
in Colombia and Ecuador (Cochrane et al., 2014) appear distinct continental crust on the margins of Gondwana. These magmatic
from granitoids of comparable age in the western Gulf of Mexico events have been interpreted to be related to the following tec-
and southern Mexico. The TDM(Hf) model ages in the Sierra Pinta, tonic settings:
Colombia, and Ecuador granitoids display evidence for mantle- (1) An Early Permian (ca. 294 Ma) continental arc that
crustal mixing, which is more typical of continental arcs. We formed in response to the subduction of the Rheic
therefore agree with the interpretations of previous authors that Ocean under the northern margins of Gondwana, prior
granitoids of northeast Mexico, Colombia, and Ecuador repre- to the final amalgamation of Pangea: We propose that
sent a Late Permian–Early Triassic arc associated with the sub- these Early Permian granitoids of the western Gulf of
duction of the paleo-Pacific beneath Pangea (Fig. 11C). Mexico are related to coeval late Carboniferous–Early

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
50 Coombs et al.

and Planetary Science Letters, v. 273, p. 48–57, https://doi.org/10.1016/j


Permian granitoids that intrude the Acatlán Complex .epsl.2008.06.010.
and Oaxaca area. Bussien, D., Bussy, F., Magna, T., and Masson, H., 2011, Timing of Palaeo-
(2) Late Permian–Early Triassic (263–243 Ma) late- to post- zoic magmatism in the Maggia and Sambuco nappes and paleogeographic
implications (central Lepontine Alps): Swiss Journal of Geosciences,
collisional anatectic magmatism that formed from oro- v. 104, p. 1–29, https://doi.org/10.1007/s00015-010-0049-6.
genic collapse in the Marathon-Sonora section of the Bussy, F., Hernandez, J., and Von Raumer, J., 2000, Bimodal magmatism
Pangea collision zone: We propose that they are coeval as a consequence of the post-collisional readjustment of the thickened
Variscan continental lithosphere (Aiguilles Rouges–Mont Blanc Massifs,
with Late Permian–Early Triassic granitoids of the Oax- western Alps), in Barbarin, B., Stephens, W.E., Bonin, B., Bouchez, J.-L.,
aca area and Chiapas Massif. Clarke, D.B., Cuney, M., and Martin, H., eds., The Fourth Hutton Sympo-
(3) An Early Jurassic (188–164 Ma) continental arc, which sium on the Origin of Granites and Related Rocks: Geological Society of
America Special Paper 350, p. 221–233, https://doi.org/10.1130/0-8137
is likely to be a part of the Nazas arc now exposed in -2350-7.221.
northern central Mexico. Chappell, B.W., and White, A.J.R., 1974, Two contrasting granite types: Pacific
These granitoids were subsequently displaced from their Geology, v. 8, p. 173–174.
Cochrane, R., Spikings, R., Gerdes, A., Ulianov, A., Mora, A., Villagómez, D.,
original positions during the breakup of Pangea. The initial Putlitz, B., and Chiaradia, M., 2014, Permo-Triassic anatexis, continen-
phase of migration occurred in the Early Jurassic during the tal rifting and the disassembly of western Pangaea: Lithos, v. 190–191,
early phases of the opening of the Gulf of Mexico, which dis- p. 383–402, https://doi.org/10.1016/j.lithos.2013.12.020.
Condie, K.C., Beyer, E., Belousova, E., Griffin, W.L., and O’Reilly, S.Y., 2005,
placed late Carboniferous–Early Permian and Late Permian– U-Pb isotopic ages and Hf isotopic composition of single zircons: The
Early Triassic granitoids of the Acatlán and Oaxaca areas to search for juvenile Precambrian continental crust: Precambrian Research,
the south along the Mojave-Sonora megashear. Displacement v. 139, p. 42–100, https://doi.org/10.1016/j.precamres.2005.04.006.
Dickinson, W.R., and Lawton, T.F., 2001, Carboniferous to Cretaceous
of the Late Permian–Early Triassic granitoids of the Chiapas assembly and fragmentation of Mexico: Geological Society of Amer-
Massif and western Gulf of Mexico occurred in the Late Juras- ica Bulletin, v. 113, p. 1142–1160, https://doi.org/10.1130/0016-7606
sic, during the main phase of Gulf of Mexico opening as the (2001)113<1142:CTCAAF>2.0.CO;2.
Ducea, M.N., Gehrels, G.E., Shoemaker, S., Ruiz, J., and Valencia, V.A., 2004,
Yucatan block rotated counterclockwise. Geologic evolution of the Xolapa Complex, southern Mexico: Evidence
from U-Pb zircon geochronology: Geological Society of America Bul-
ACKNOWLEDGMENTS letin, v. 116, p. 1016–1025, https://doi.org/10.1130/B25467.1.
Elías-Herrera, M., Ortega-Gutiérrez, F., Sánchez-Zavala, J.L., Macías-Romo,
C., Ortega-Rivera, A., and Iriondo, A., 2005, La falla de Caltepec: Raíces
We would like to thank Juan Pablo Zapata, Ricardo Milián, expuestas de una frontera tectónica de larga vida entre dos terrenos conti-
Roberto Molina, Diego Villagómez, Maria Isabel Sierra Rojas, nentales del sur de México: Boletín de la Sociedad Geológica Mexicana,
v. 57, p. 83–109, https://doi.org/10.18268/BSGM2005v57n1a5.
Rodrigo León, and Kurt Wogau for their help during sample Estrada-Carmona, J., Weber, B., Martens, U., and Lopez-Martinez, M., 2012,
preparation. Additionally, a huge “thank you” must be extended Petrogenesis of Ordovician magmatic rocks in the southern Chiapas Mas-
to Uwe Martens for all of his help and guidance over the past sif complex: Relations with the early Palaeozoic magmatic belts of north-
western Gondwana: International Geology Review, v. 54, p. 1918–1943,
few years. Finally, we would like to thank the reviewers for https://doi.org/10.1080/00206814.2012.685553.
their constructive suggestions and comments. Estrada-Carmona, J., Weber, B., Scherer, E.E., Martens, U., and Elías-Herrera,
M., 2016, Lu-Hf geochronology of Mississippian high-pressure metamor-
phism in the Acatlán Complex, southern México: Gondwana Research,
REFERENCES CITED v. 34, p. 174–186.
Godinez-Urban, A., Lawton, T.F., Molina Garza, R.S., Iriondo, A., Weber,
Andersen, T., 2002, Correction of common lead in U-Pb analyses that do not B., and Lopez-Martinez, M., 2011, Jurassic volcanic and sedimentary
report 204Pb: Chemical Geology, v. 192, p. 59–79, https://doi.org/10.1016/ rocks of the La Silla and Todos Santos Formations, Chiapas: Record of
S0009-2541(02)00195-X. Nazas arc magmatism and rift-basin formation prior to opening of the
Anderson, T.H., and Schmidt, V.A., 1983, The evolution of middle America Gulf of Mexico: Geosphere, v. 7, p. 121–144, https://doi.org/10.1130/
and the Gulf of Mexico Caribbean Sea region during Mesozoic time: GES00599.1.
Geological Society of America Bulletin, v. 94, p. 941–966, https://doi González-Guzmán, G., Weber, B., Tazzo-Rangel, M.T., and Solari, L.A., 2016,
.org/10.1130/0016-7606(1983)94<941:TEOMAA>2.0.CO;2. Validation of digestion and element separation methods and a new data
Arvizu, H.E., and Iriondo, A., 2011, Hf isotopic study of zircons from Permian reduction program (IsotopeHf®) for Lu-Hf isotope dilution analysis by
granitoids of NW Mexico: Evidence for mixing of magmas derived from MC-ICP-MS: Revista Mexicana de Ciencias Geológicas, v. 33, p. 254–
multiple crustal sources: Revista Mexicana de Ciencias Geológicas, v. 28, 269.
p. 493–518. Griffin, W.L., Wang, X., Jackson, S.E., Pearson, N.J., O’Reilly, S.Y., Xu, X., and
Barboza-Gudiño, J.R., Orozco-Esquivel, M.T., Gomez-Anguiano, M., and Zhou, X., 2002, Zircon chemistry and magma mixing, SE China: In-situ
Zavala-Monsivais, A., 2008, The early Mesozoic volcanic arc of western analysis of Hf isotopes, Tonglu and Pingtan igneous complexes: Lithos,
North America in northeastern Mexico: Journal of South American Earth v. 61, p. 237–269, https://doi.org/10.1016/S0024-4937(02)00082-8.
Sciences, v. 25, p. 49–63, https://doi.org/10.1016/j.jsames.2007.08.003. Grimes, C.B., Wooden, J.L., Cheadle, M.J., and John, B.E., 2015, “Fingerprint-
Bird, D.E., Burke, K., Hall, S.A., and Casey, J.F., 2005, Gulf of Mexico tectonic ing” tectono-magmatic provenance using trace elements in igneous zir-
history: Hotspot tracks, crustal boundaries, and early salt distribution: con: Contributions to Mineralogy and Petrology, v. 170, p. 46, https://doi
American Association of Petroleum Geologists Bulletin, v. 89, p. 311– .org/10.1007/s00410-015-1199-3.
328, https://doi.org/10.1306/10280404026. Handschy, J.W., Keller, G.R., and Smith, K.J., 1987, The Ouachita system in
Böhnel, H., 1999, Paleomagnetic study of Jurassic and Cretaceous rocks from northern Mexico: Tectonics, v. 6, p. 323–330, https://doi.org/10.1029/
the Mixteca terrane (Mexico): Journal of South American Earth Sciences, TC006i003p00323.
v. 12, p. 545–556, https://doi.org/10.1016/S0895-9811(99)00038-3. Heatherington, A.L., Mueller, P.A., and Wooden, J.L., 2010, Alleghanian plu-
Bouvier, A., Vervoort, J.D., and Patchett, P.J., 2008, The Lu-f and Sm-–Nd iso- tonism in the Suwannee terrane, USA: Implications for late Paleozoic
topic composition of CHUR: Constraints from unequilibrated chondrites tectonic models, in Tollo, R.P., Bartholomew, M.J., Hibbard, J.P., and
and implications for the bulk composition of terrestrial planets: Earth Karabinos, P.M., eds., From Rodinia to Pangea: The Lithotectonic Record

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
Petrogenesis of the crystalline basement along the western Gulf of Mexico 51

of the Appalachian Region: Geological Society of America Memoir 206, McDonald, I., and Viljoen, K.S., 2006, Platinum-group element geochemistry
p. 607–620, https://doi.org/10.1130/2010.1206(24). of mantle eclogites: A reconnaissance study of xenoliths from the Orapa
Imbert, P., 2005, The Mesozoic opening of the Gulf of Mexico: Part 1. Evidence kimberlite, Botswana: Transactions of the Institution of Mining and Met-
for oceanic accretion during and after salt deposition, in Transactions of allurgy Section B–Applied Earth Science, v. 115, p. 81–93.
the 25th Annual GCSSEPM Research Conference: Petroleum Systems of McDonough, W.F., and Sun, S.-s., 1995, The composition of the Earth: Chemi-
Divergent Continental Margins: Gulf Coast Section, Society for Sedimen- cal Geology, v. 120, p. 223–253, https://doi.org/10.1016/0009-2541
tary Geology (SEPM), p. 1119–1150. (94)00140-4.
Imbert, P., and Phillippe, Y., 2005, The Mesozoic opening of the Gulf of Mex- Mueller, P.A., Heatherington, A.L., Foster, D.A., Thomas, W.A., and Wooden,
ico: Part 2. Integrating seismic and magnetic data into a general opening J.L., 2014, The Suwannee suture: Significance for Gondwana-Laurentia
model, in Transactions of the 25th Annual GCSSEPM Research Confer- terrane transfer and formation of Pangaea: Gondwana Research, v. 26,
ence: Petroleum Systems of Divergent Continental Margins: Gulf Coast p. 365–373, https://doi.org/10.1016/j.gr.2013.06.018.
Section, Society for Sedimentary Geology (SEPM), p. 1151–1189. Nebel-Jacobsen, Y., Scherer, E.E., Münker, C., and Mezger, K., 2005, Separa-
Jackson, S.E., Pearson, N.J., Griffin, W.L., and Belousova, E.A., 2004, The tion of U, Pb, Lu, and Hf from single zircons for combined U-Pb dating
application of laser ablation–inductively coupled plasma–mass spectrom- and Hf isotope measurements by TIMS and MC-ICPMS: Chemical Geol-
etry to in situ U-Pb zircon geochronology: Chemical Geology, v. 211, ogy, v. 220, p. 105–120, https://doi.org/10.1016/j.chemgeo.2005.03.009.
p. 47–69, https://doi.org/10.1016/j.chemgeo.2004.06.017. Ortega-Gutierrez, F., Ruiz, J., and Centeno-Garcia, E., 1995, Oaxaquia, a
Jacques, J.M., Price, A.D., and Bain, J.E., 2004, Digital integration of poten- Proterozoic microcontinent accreted to North America during the late
tial fields and geologic data sets for plate tectonic and basin dynamic Paleozoic: Geology, v. 23, p. 1127–1130, https://doi.org/10.1130/0091
modeling—The first step toward identifying new play concepts in the -7613(1995)023<1127:OAPMAT>2.3.CO;2.
Gulf of Mexico Basin: The Leading Edge, v. 23, p. 384–389, https://doi Ortega-Gutiérrez, F., Elias-Herrera, M., Moran-Zenteno, D.J., Solari, L., Luna-
.org/10.1190/1.1729229. Gonzalez, L., and Schaaf, P., 2014, A review of batholiths and other
Keppie, J.D., 2004, Terranes of Mexico revisited: A 1.3 billion year odys- plutonic intrusions of Mexico: Gondwana Research, v. 26, p. 834–868,
sey: International Geology Review, v. 46, p. 765–794, https://doi.org/ https://doi.org/10.1016/j.gr.2014.05.002.
10.2747/0020-6814.46.9.765. Ortega-Obregón, C., Duncan Keppie, J., Murphy, J.B., Lee, J.K.W., and Ortega-
Keppie, J.D., Dostal, J., Ortega-Gutierrez, F., and Lopez, R., 2001, A Grenvil- Rivera, A., 2009, Geology and geochronology of Paleozoic rocks in west-
lian arc on the margin of Amazonia: Evidence from the southern Oaxacan ern Acatlán Complex, southern Mexico: Evidence for contiguity across an
Complex, southern Mexico: Precambrian Research, v. 112, p. 165–181, extruded high-pressure belt and constraints on Paleozoic reconstructions:
https://doi.org/10.1016/S0301-9268(00)00150-9. Geological Society of America Bulletin, v. 121, p. 1678–1694, https://doi
Keppie, J.D., Dostal, J., Cameron, K.L., Solari, L.A., Ortega-Gutierrez, F., and .org/10.1130/B26597.1.
Lopez, R., 2003, Geochronology and geochemistry of Grenvillian igne- Ortega-Obregón, C., Solari, L., Gomez-Tuena, A., Elias-Herrera, M., Ortega-
ous suites in the northern Oaxacan Complex, southern Mexico: Tectonic Gutiérrez, F., and Macias-Romo, C., 2014, Permian–Carboniferous arc
implications: Precambrian Research, v. 120, p. 365–389, https://doi.org/ magmatism in southern Mexico: U-Pb dating, trace element and Hf isoto-
10.1016/S0301-9268(02)00166-3. pic evidence on zircons of earliest subduction beneath the western margin
Keppie, J.D., Dostal, J., Murphy, B.J., and Nance, D., 2008, Synthesis and of Gondwana: International Journal of Earth Sciences, v. 103, p. 1287–
tectonic interpretation of the westernmost Paleozoic Variscan orogen 1300, https://doi.org/10.1007/s00531-013-0933-1.
in southern Mexico: From rifted Rheic margin to active Pacific mar- Pearce, J.A., Harris, N.B.W., and Tindle, A.G., 1984, Trace-element discrimina-
gin: Tectonophysics, v. 461, p. 277–290, https://doi.org/10.1016/j.tecto tion diagrams for the tectonic interpretation of granitic-rocks: Journal of
.2008.01.012. Petrology, v. 25, p. 956–983, https://doi.org/10.1093/petrology/25.4.956.
Keppie, J.D., Nance, R.D., Dostal, J., Lee, J.K.W., and Ortega-Rivera, A., 2012, Pindell, J., and Dewey, J.F., 1982, Permo-Triassic reconstruction of western
Constraints on the subduction erosion/extrusion cycle in the Paleozoic Pangea and the evolution of the Gulf of Mexico Caribbean region: Tecton-
Acatlán Complex of southern Mexico: Geochemistry and geochronol- ics, v. 1, p. 179–211, https://doi.org/10.1029/TC001i002p00179.
ogy of the type Piaxtla Suite: Gondwana Research, v. 21, p. 1050–1065, Pindell, J., and Kennan, L., 2001, Kinematic evolution of the Gulf of Mexico
https://doi.org/10.1016/j.gr.2011.07.020. and Caribbean, in Transactions of the GCSSEPM Foundation 21st Annual
Kirsch, M., 2012, Continental Arc Development along the Periphery of Pangea: Research Conference: Petroleum Systems of Deep-Water Basins: Gulf
Late Paleozoic Pluton Emplacement and Basin Evolution in a Transten- Coast Section, Society of Sedimentary Geology (SEPM), p. 193–220.
sional Setting, Eastern Acatlán Complex, Mexico [Ph.D. thesis]: Univer- Pindell, J., Kennan, L., Maresch, W.V., Stanek, K.P., Draper, G., and Higgs, R.,
sidad Nacional Autónoma de México. 2005, Plate-kinematics and crustal dynamics of circum-Caribbean arc-
Kirsch, M., Keppie, J.D., Murphy, J.B., and Solari, L.A., 2012, Permian–­ continent interactions: Tectonic controls on basin development in proto-
Carboniferous arc magmatism and basin evolution along the western Caribbean margins, in Avé Lallemant, H.G., and Sisson, V.G., eds., Carib-
margin of Pangea: Geochemical and geochronological evidence from the bean–South American Plate Interactions, Venezuela: Geological Society
eastern Acatlán Complex, southern Mexico: Geological Society of Amer- of America Special Paper 394, p. 7–52, https://doi.org/10.1130/0-8137
ica Bulletin, v. 124, p. 1607–1628, https://doi.org/10.1130/B30649.1. -2394-9.7.
Lawlor, P.J., Ortega-Gutiérrez, F., Cameron, K.L., Ochoa-Camarillo, H., Lopez, Pindell, J.L., 1985, Alleghenian reconstruction and subsequent evolution of the
R., and Sampson, D.E., 1999, U-Pb geochronology, geochemistry, and Gulf of Mexico, Bahamas, and proto-Caribbean: Tectonics, v. 4, p. 1–39,
provenance of the Grenvillian Huiznopala Gneiss of eastern Mexico: https://doi.org/10.1029/TC004i001p00001.
Precambrian Research, v. 94, p. 73–99, https://doi.org/10.1016/S0301 Pindell, J.L., and Kennan, L., 2009, Tectonic evolution of the Gulf of Mexico,
-9268(98)00108-9. Caribbean and northern South America in the mantle reference frame: An
Lawton, T.F., and Molina Garza, R.S., 2014, U-Pb geochronology of the type update, in James, K.H., Lorente, M.A., and Pindell, J.L., eds., The Origin
Nazas Formation and superjacent strata, northeastern Durango, Mexico: and Evolution of the Caribbean Plate: Geological Society, London, Spe-
Implications of a Jurassic age for continental-arc magmatism in north- cial Publication 328, p. 1–55.
central Mexico: Geological Society of America Bulletin, v. 126, p. 1181– Poole, F.G., Perry, W.J., Madrid, R.J., and Amaya-Martínez, R., 2005, Tectonic
1199, https://doi.org/10.1130/B30827.1. synthesis of the Ouachita-Marathon-Sonora orogenic margin of southern
Le Maitre, R.W., Streckeisen, A., Zanettin, B., Le Bas, M.J., Bonin, B., Bate- Laurentia: Stratigraphic and structural implications for timing of defor-
man, P., Bellieni, G., Dudek, A., Efremova, S., Keller, J., Lameyre, J., mational events and plate-tectonics model, in Anderson, T.H., Nourse,
Sabine, P.A., Schmid, R., Sørensen, H., and Woolley, A.R., 1989, A Clas- J.A., McKee, J.W., and Steiner, M.B., eds., The Mojave-Sonora Mega-
sification of Igneous Rocks and Glossary of Terms: Oxford, UK, Black- shear Hypothesis: Development, Assessment, and Alternatives: Geo-
well, 193 p. logical Society of America Special Paper 293, p. 543–596, https://doi
Lopez-Ramos, E., 1979, Geología de México: México D.F., Edición Privada, .org/10.1130/0-8137-2393-0.543.
454 p. Ratschbacher, L., Franz, L., Min, M., Bachmann, R., Martens, U., Stanek, K.,
Ludwig, K.R., 2012, User’s Manual for Isoplot 3.75; A Geochronological Tool- Stübner, K., Nelson, B.K., Herrmann, U., Weber, B., López-Martínez, M.,
kit for Microsoft Excel: Berkeley Geochronology Center Special Publica- Jonckheere, R., Sperner, B., Tichomirowa, M., McWilliams, M.O., Gordon,
tion 5, 75 p. M., Meschede, M., and Bock, P., 2009, The North American–­Caribbean

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user
52 Coombs et al.

plate boundary in Mexico-Guatemala-Honduras, in James, K.H., Lorente, the Acatlán Complex (southern Mexico): Implications for the evolution of
M.A., and Pindell, J.L., eds., The Origin and Evolution of the Caribbean the Iapetus and Rheic Oceans: Geological Society of America Bulletin,
Plate: Geological Society, London, Special Publication 328, p. 219–293, v. 119, p. 1249–1264, https://doi.org/10.1130/B226031.1.
https://doi.org/10.1144/SP328.11. Vega-Granillo, R., Calmus, T., Meza-Figueroa, D., Ruiz, J., Talavera-Mendoza,
Ross, C.A., 1986, Paleozoic evolution of southern margin of Permian Basin: O., and Lopez-Martinez, M., 2009, Structural and tectonic evolution of
Geological Society of America Bulletin, v. 97, p. 536–554, https://doi the Acatlán Complex, southern Mexico: Its role in the collisional history
.org/10.1130/0016-7606(1986)97<536:PEOSMO>2.0.CO;2. of Laurentia and Gondwana: Tectonics, v. 28, no. 4, TC4008, https://doi
Schaltegger, U., 1997, Magma pulses in the central Variscan belt: Episodic melt .org/10.1029/2007TC002159.
generation and emplacement during lithospheric thinning: Terra Nova, Vervoort, J.D., Patchett, P.J., Albarede, F., Blichert-Toft, J., Rudnick, R., and
v. 9, p. 242–245, https://doi.org/10.1111/j.1365-3121.1997.tb00021.x. Downes, H., 2000, Hf-Nd isotopic evolution of the lower crust: Earth and
Schouten, H., and Klitgord, K.D., 1994, Mechanistic solutions to the open- Planetary Science Letters, v. 181, p. 115–129, https://doi.org/10.1016/
ing of the Gulf of Mexico: Geology, v. 22, p. 507–510, https://doi. S0012-821X(00)00170-9.
org/10.1130/0091-7613(1994)022<0507:MSTTOO>2.3.CO;2. Viele, G.W., and Thomas, W.A., 1989, Tectonic synthesis of the Ouachita oro-
Schulze, C., 2011, Petrología y Geoquímica de las Rocas del Área de Pluma genic belt, in Hatcher, R.D., Jr., Thomas, W.A., and Viele, G.W., eds.,
Hidalgo, Oaxaca, e Implicaciones Tectónicas para el Proterozoico de The Appalachian-Ouachita Orogen in the United States: Boulder, Colo-
Oaxaquia [Ph.D. thesis]: México, D.F., Universidad Autónoma Nacional rado, Geological Society of America, Geology of North America, v. F-2,
de México, 311 p. p. 695–728.
Sedlock, R.L., Ortega-Gutiérrez, F., and Speed, R.C., 1993, Tectonostrati- Vinasco, C.J., Cordani, U.G., Gonzalez, H., Weber, M., and Pelaez, C., 2006,
graphic Terranes and Tectonic Evolution of Mexico: Geological Soci- Geochronological, isotopic, and geochemical data from Permo-Triassic
ety of America Special Paper 278, 142 p., https://doi.org/10.1130/ granitic gneisses and granitoids of the Colombian Central Andes: Jour-
SPE278-p1. nal of South American Earth Sciences, v. 21, p. 355–371, https://doi.org/
Shaulis, B.J., Lapen, T.J., Casey, J.F., and Reid, D.R., 2012, Timing and rates 10.1016/j.jsames.2006.07.007.
of flysch sedimentation in the Stanley Group, Ouachita Mountains, Okla- Von Raumer, J.F., and Bussy, F., 2004, Mont Blanc and Aiguilles Rouges: Geol-
homa and Arkansas, U.S.A.: Constraints from U-Pb zircon ages of sub- ogy of their polymetamorphic basement (External Massifs, Western Alps,
aqueous ash-flow tuffs: Journal of Sedimentary Research, v. 82, p. 833– France-Switzerland): Mémoires de Géologie 42 (Lausanne, Université de
840, https://doi.org/10.2110/jsr.2012.68. Lausanne, Section des Sciences de la Terre), 204 p.
Sláma, J., Kosler, J., Condon, D., Crowley, J., Gerdes, A., Hanchar, J., Walsh, G.J., Aleinikoff, J.N., and Wintsch, R.P., 2007, Origin of the Lyme
Horstwood, M., Morris, G., Nasdala, L., Norberg, N., Schaltagger, U., Dome and implications for the timing of multiple Alleghanian deforma-
Schoene, B., Tubrett, M.N., and Whitehouse, M.J., 2008, Plešovice zir- tional and intrusive events in southern Connecticut: American Journal of
con—A new natural reference material for U-Pb and Hf isotopic micro- Science, v. 307, p. 168–215, https://doi.org/10.2475/01.2007.06.
analysis: Chemical Geology, v. 249, p. 1–35, https://doi.org/10.1016/j Watson, E.B., and Harrison, T.M., 1983, Zircon saturation revisited—­
.­chemgeo.2007.11.005. Temperature and composition effects in a variety of crustal magma
Solari, L.A., and Tanner, M., 2011, UPb age, a fast data reduction script for LA- types: Earth and Planetary Science Letters, v. 64, p. 295–304, https://doi.
ICP-MS U-Pb geochronology: Revista Mexicana de Ciencias Geológicas, org/10.1016/0012-821X(83)90211-X.
v. 28, p. 83–91. Weber, B., and Schulze, C.H., 2014, Early Mesoproterozoic (>1.4 Ga) ages
Solari, L.A., Keppie, J.D., Ortega-Gutiérrez, F., Cameron, K.L., Lopez, R., and from granulite basement inliers of SE Mexico and their implications on
Hames, W.E., 2003, 990 and 1100 Ma Grenvillian tectonothermal events the Oaxaquia concept—Evidence from U-Pb and Lu-Hf isotopes on zir-
in the northern Oaxacan Complex, southern Mexico: Roots of an oro- con: Revista Mexicana de Ciencias Geológicas, v. 31, p. 377–394.
gen: Tectonophysics, v. 365, p. 257–282, https://doi.org/10.1016/S0040 Weber, B., Cameron, K.L., Osorio, M., and Schaaf, P., 2005, A Late Perm-
-1951(03)00025-8. ian tectonothermal event in Grenville crust of the southern Maya terrane:
Solari, L.A., Gómez-Tuena, A., Bernal, J.P., Pérez-Arvizu, O., and Tanner, M., U-Pb zircon ages from the Chiapas Massif, southeastern Mexico: Inter-
2010, U-Pb zircon geochronology with an integrated LA-ICP-MS micro- national Geology Review, v. 47, p. 509–529, https://doi.org/10.2747/0020
analytical workstation: Achievements in precision and accuracy: Geostan- -6814.47.5.509.
dards and Geoanalytical Research, v. 34, p. 5–18, https://doi.org/10.1111/ Weber, B., Iriondo, A., Premo, W.R., Hecht, L., and Schaaf, P., 2007, New
j.1751-908X.2009.00027.x. insights into the history and origin of the southern Maya block, SE Mex-
Spikings, R., Cochrane, R., Villagómez, D., Van Der Lelij, R., Vallejo, C., ico: U-Pb-SHRIMP zircon geochronology from metamorphic rocks of the
Winkler, W., and Beate, B., 2015, The geological history of northwest- Chiapas Massif: International Journal of Earth Sciences, v. 96, p. 253–
ern South America: From Pangaea to the early collision of the Carib- 269, https://doi.org/10.1007/s00531-006-0093-7.
bean large igneous province (290–75 Ma): Gondwana Research, v. 27, Weber, B., Scherer, E.E., Schulze, C., Valencia, V.A., Montecinos, P., Mezger,
p. 95–139, https://doi.org/10.1016/j.gr.2014.06.004. K., and Ruiz, J., 2010, U-Pb and Lu-Hf isotope systematics of lower crust
Sun, S.-s., and McDonough, W.F., 1989, Chemical and isotopic systematics of from central-southern Mexico—Geodynamic significance of Oaxaquia in
oceanic basalts: Implications for mantle composition and processes, in a Rodinia Realm: Precambrian Research, v. 182, p. 149–162, https://doi.
Saunders, A.D., and Norry, M.J., eds., Magmatism in the Ocean Basins: org/10.1016/j.precamres.2010.07.007.
Geological Society, London, Special Publication 42, p. 313–345, https:// Wilson, M., ed., 1989, Igneous Petrogenesis: Dordrecht, Netherlands, Springer,
doi.org/10.1144/GSL.SP.1989.042.01.19. 416 p., https://doi.org/10.1007/978-1-4020-6788-4.
Talavera-Mendoza, O., Ruiz, J., Gehrels, G.E., Meza-Figueroa, D.M., Vega- Wortman, G.L., Samson, S.D., and Hibbard, J.P., 1998, Precise U-Pb zircon
Granillo, R., and Campa-Uranga, M.F., 2005, U-Pb geochronology of the timing constraints on the kinematic development of the Hyco shear zone:
Acatlán Complex and implications for the Paleozoic paleogeography and Implications for the central Piedmont shear zone, southern Appalachian
tectonic evolution of southern Mexico: Earth and Planetary Science Let- orogen: American Journal of Science, v. 298, p. 108–130.
ters, v. 235, p. 682–699, https://doi.org/10.1016/j.epsl.2005.04.013. Yañez, P., Ruiz, J., Patchett, P.J., Ortega-Gutiérrez, F., and Gehrels, G.E., 1991, Iso-
van Der Lelij, R., Spikings, R., Ulianov, A., Chiaradia, M., and Mora, A., 2016, topic studies of the Acatlan Complex, southern Mexico: implications for
Palaeozoic to Early Jurassic history of the northwestern corner of Gond- Paleozoic North American tectonics: Geological Society of America
wana, and implications for the evolution of the Iapetus, Rheic and Pacific Bulletin, v. 103, p. 817–828, https://doi.org/10.1130/0016-7606(1991)
Oceans: Gondwana Research, v. 31, p. 271–294, https://doi.org/10.1016/j 103<0817:ISOTAC>2.3.CO;2.
.gr.2015.01.011.
Vega-Granillo, R., Talavera-Mendoza, O., Meza-Figueroa, D., Ruiz, J., Geh-
rels, G.E., Lopez-Martinez, M., and De La Cruz-Vargas, J.C., 2007, Manuscript Accepted by the Society 4 November 2019
­Pressure-temperature-time evolution of Paleozoic high-pressure rocks of Manuscript Published Online 14 May 2020

Printed in the USA

Downloaded from http://pubs.geoscienceworld.org/gsa/books/book/2217/chapter-pdf/5572821/spe546-02.pdf


by UNAM user

You might also like