Guide to Key Outcrops for Reconstruction of the Geologic-Tectonic History of Sierra de Catorce: Northeastern Mexico

Barboza-Gudiño, José Rafael1; Morales Gamez, Miguel2; Díaz-Bravo, Beatriz A.1;

Huerta González, Rosa María3; Zavala-Monsiváis, Aurora2; Jaime Rodríguez, Diego3

1Instituto de Geología, Universidad Autónoma de San Luis Potosí. Av. Dr. Manuel Nava 5, Zona Universitaria, 78240 San Luis, San Luis Potosí, México.

2 Facultad de Ingeniería, Universidad Autónoma de San Luis Potosí. Av. Dr. Manuel Nava 304, Zona Universitaria, 78210 San Luis, San Luis Potosí, México.

3 Posgrado en Geología Aplicada, Centro de Investigación y Estudios de Posgrado, Universidad Autónoma de San Luis Potosí. Dr. Manuel Nava No. 8, Col. Zona Universitaria Poniente, C.P. 78290, San Luis, San Luis Potosí, México.

* rbarboza@uaslp.mx

ABSTRACT

In the Sierra de Catorce, there is an uplifted block in the southernmost Basin and Range province in northern San Luis Potosí, where most of the oldest stratigraphic units of the region are well exposed. The succession includes Upper Triassic siliciclastics consisting of an alternation of fine-grained sandstone and shale layers, interpreted as lateral equivalents to deep marine turbidites, named the Zacatecas Formation, known from several localities to the west in the region. At Real de Catorce, the previously described deposits underlie coarse chaotic deposits which are products of flow events that change upwards in a marine marginal succession that includes the Triassic-Jurassic boundary. Upwards, the succession includes volcanic and volcaniclastic strata related to the Lower to Middle Jurassic Nazas Formation, underlying conglomerate-breccia and red sandstones of the La Joya Formation that represents an erosional or break-up unconformity related to the initial spreading in the Gulf of Mexico Basin. La Joya Formation changes upsection into limestones of the Zuloaga Formation, which resulted from the Middle-Late Jurassic marine transgression. Outcrops in the General Canyon, along the route Carretas-Real de Catorce include typical strata that allow interpretation of the tectonic-paleogeographic evolution of the Late Triassic-Early Jurassic Pacific margin of Mexico, which evolved into extensional to transtensional basins related to the Gulf of Mexico Basin and probably in part, a result of back-arc extension related to the Nazas volcanic arc. Upper Jurassic strata in the region represent the bottom of the marine sedimentary succession deposited in the Central Mexico Basin, a trough shaped subsiding paleogeographic element in Central Mexico during Late Jurassic-Cretaceous time. Finally, we describe some outcrops that illustrate structures produced by contractile deformation (Late Cretaceous-Paleogene) and post-deformational magmatism (Eocene) that occurred in the area.

Keywords: Real de Catorce, Triassic, Jurassic, stratigraphy, geochronology.

RESUMEN

En la Sierra de Catorce, hay un bloque levantado en la porción meridional de la provincia de Cuencas y Sierras en el norte de San Luis Potosí donde la mayoría de las unidades estratigráficas más antiguas de la región están bien expuestas. La sucesión estratigráfica incluye rocas siliciclásticas del Triásico Superior en forma de una alternancia de capas de arenisca de grano fino y lutitas, interpretadas como equivalentes de turbiditas marinas más profundas conocidas como Formación Zacatecas en varias localidades al oeste de esta zona. En la Sierra de Catorce la parte superior de esta alternancia de areniscas finas y lutitas pasa a depósitos caóticos gruesos que son producto de eventos de flujo masivo que evolucionan en una sucesión marina marginal que incluye al límite Triásico-Jurásico. Hacia arriba, la sucesión incluye productos volcánicos y horizontes volcanoclásticos relacionados con la Formación Nazas, del Jurásico Inferior a Medio, la cual subyace a un conglomerado o brecha y areniscas rojas de la Formación La Joya, que representa una discordancia erosional o “break-up unconformity” relacionada con extensión durante la apertura del Golfo de México. La Formación La Joya pasa hacia arriba en la secuencia a calizas de la Formación Zuloaga, como resultado de la transgresión marina ocurrida durante el Jurásico Tardío. Afloramientos en el Cañón General, a lo largo de la ruta Carretas-Real de Catorce, incluyen estratos que permiten interpretar la evolución tectónica paleogeográfica del margen Pacífico antiguo de México durante el Triásico Tardío-Jurásico Temprano, que posteriormente evolucionó en las cuencas extensionales a transtensionales relacionadas con la apertura del Golfo de México y posiblemente en parte como una extensión tras-arco, relacionada con el magmatismo del llamado arco volcánico Nazas. Los estratos del Jurásico Superior en la región representan la parte inferior de la sucesión sedimentaria marina depositada en la Cuenca Mesozoica del Centro de México, un elemento paleogeográfico subsidente durante el Jurásico Tardío-Cretácico. Finalmente se describen algunos afloramientos que ilustran las estructuras producto de la deformación contractiva (Cretácico Tardío-Paleógeno) y el magmatismo post deformación (Eoceno) ocurridos en la zona.

Palabras clave: Real de Catorce, Triásico, Jurásico, estratigrafía, geocronología.

1. Introduction

The Mesozoic sedimentary cover in Eastern Mexico overlies a peri-Gondwanan Grenvillian block known as the Oaxaquia microcontinent (Ortega-Gutierrez et al., 1995). Westward from westernmost San Luis Potosí, there is no evidence of a major block of Precambrian-Paleozoic basement. In these areas, ancient oceanic crust and overlying Triassic pelagic sediments and turbidites constitute the sole of the Jurassic-Cretaceous marine volcanic and volcano-sedimentary successions, which make up the so-called composite Guerrero terrane (Figure 1). To the east, Precambrian-Paleozoic crystalline basement underlies a folded and thrusted sedimentary cover, which consists of Jurassic-Cretaceous limestones and evaporites as well as subordinate clastic rocks in the Sierra Madre Oriental province.

According to the distribution of the Precambrian-Paleozoic basement in Mexico, the ancient westernmost Pangea paleomargin can be interpreted to extend from southern Sonora and Chihuahua, along north-central and eastern Mexico, and southeastward to Puebla and western Guerrero. During early Mesozoic times, thick Triassic siliciclastic successions were deposited on an ancient passive margin of Pangea, (Silva-Romo et al., 2000) evolving to the west in the so called “Potosí submarine fan” (Centeno-García 2005, Centeno-García et al., 2008), while coeval continental fluvial deposits in Nuevo León and Tamaulipas are known as El Alamar Formation (Barboza-Gudiño et al., 2010).

After marine regression and deformation of the Triassic rocks caused by subduction at the western margin of Pangea, a continental volcanic arc, known as the Nazas Formation or Nazas arc (Pantoja-Alor, 1972, Lawton and Molina-Garza., 2014) evolved during Early to Middle Jurassic time. Moderately deformed subaerial volcanic rocks of the Nazas arc rest unconformably on tightly folded Triassic turbidites. Volcanic materials of the Nazas Formation are partly interlayered and covered by Lower Jurassic volcaniclastic “red-beds” (La Boca Formation; Mixon et al., 1959). Finally, in the Middle Jurassic La Joya Formation (Mixon et al., 1959), there is a fining upward siliciclastic sequence, grades upward into marine evaporitic and carbonate beds of the Callovian-Oxfordian Zuloaga Group (Sandstrom, 1982) and the fossil-rich La Caja (Imlay, 1938) or La Casita Formation (Imlay, 1936) at the top of the Jurassic succession.

The Sierra de Catorce, which lies at the boundary between the Mesa Central and Sierra Madre Oriental provinces in northern San Luis Potosí, (Figure 2) as a strongly uplifted area offers the possibility to examine exposures of the oldest rocks in the region, including the above mentioned Triassic rocks and the Jurassic succession consisting of volcanogenic and non-marine sedimentary Lower to Middle Jurassic units. The Triassic-Lower Jurassic succession unconformably underlies the Upper Jurassic represented by a marine, first mostly calcareous (Zuloaga Formation) and then calcareous-siliciclastic (La Caja Formation) shelf sequence. Probably the most complete Jurassic succession in the Sierra de Catorce is exposed in Cañón General between the towns of Los Catorce and Real de Catorce, along a 5 km route in the Canyon (Figures 3, 4), where the base of the Jurassic succession is exposed east of Los Catorce, at an elevation of 2200 meters and the Jurassic-Cretaceous boundary at 2900 meters on the mountains around Real de Catorce.

Our goal is to describe and interpret several outcrops that illustrate the processes and geotectonic environments that have been associated over time with different lithostratigraphic units recognized and well exposed in Cañón General, along the road between Los Catorce and Real de Catorce. The present work can be taken as a field guide along a section through the mountain range, coming from the west along Cañon General and passing through Real de Catorce towards the northeast.

2. Upper Triassic Turbiditic Succession in the Mesa Central

Centeno-García (2005) gave the name Potosí fan to a widespread succession of deep-marine siliciclastic turbiditic deposits exposed in the Mesa Central in the states of San Luis Potosí and Zacatecas, and westward, in Guerrero and Michoacán. After the first report of Late Triassic fauna in the neighborhood of Zacatecas city (Burckhardt and Scalia, 1905), rocks of the Potosí fan were first named “Triassic of Zacatecas” (Gutiérrez-Amador, 1908) and later Zacatecas Formation (Martínez-Pérez, 1972). In the respective localities, different names are used for equivalent siliciclastic successions, such as La Ballena Formation proposed by Silva Romo et al. (2000) in the Sierra de Salinas, Zacatecas, while Córdoba-Méndez (1964) used the name Taray Formation in northern Zacatecas. The Zacatecas Formation outcropping in the Mesa Central province is the marine counterpart of the continental El Alamar Formation outcropping in the Sierra Madre Oriental (Barboza-Gudiño et al., 2010, 2012). Generally, the Zacatecas Formation consists of interstratified sandstones, siltstones, shales and conglomeratic sandstones.

In the Cañón General area (for example outcrops at coordinates: 23°42.1’N; 100°54.5’W) the sequence of interlayered fine-grained sandstone beds and black shales, whose base is not exposed, crops out around the town of Los Catorce (Figure 5A), where it exhibits good stratification, mostly strongly folded with development of pervasive cleavage and vertical pencil structures parallel to common steeply-dipping fold axes. A Late Triassic age for this succession is determined through U-Pb ages of detrital zircons (Barboza-Gudiño et al., 2010) that are in concordance with correlations based on lithological similarities and stratigraphic positions of equivalent fossil-bearing strata in the other above-mentioned localities from the Mesa Central. To the top of the Triassic sequence exposed in the Cañón General area, slumping structures (Figure 5B) and debris flow, as well as massive sandstone deposits containing floated pebbles (Figure 5C) to meter size blocks are very common (Figure 5D), and it appears to be a transitional change into the so-called Cerro El Mazo beds (Barboza-Gudiño et al., 2004). There are no reports of fossils in the Triassic turbidites of Sierra de Catorce and only the “Cerro el Mazo beds”, that overlie the Zacatecas Formation, considered also in part latest Triassic in age (Wengler, 2014), contain plant fossils (see below). The facies associations in the Zacatecas Formation are generally interpreted as inter-channel deposits and subordinate supra-fan, levee, and channel deposits. Hoppe (2000) measured a ~213 m thick succession at Los Catorce, which is considered an incomplete section because the base of this unit is not exposed. In addition to the strong deformation, this unit changes gradually upwards in the siliciclastic and volcanogenic layers of the informal unit termed “Cerro El Mazo beds” (Barboza-Gudiño et al., 2004), which include the Triassic-Jurassic boundary (Venegas-Rodriguez et al., 2009, Wengler, 2014). The Cerro El Mazo beds in turn change into the volcanic arc succession known as Nazas Formation in north-central to northeastern Mexico.

3. “Cerro El Mazo Beds” Triassic-Jurassic Marine Marginal Succession

At Cerro El Mazo, two kilometers west of Los Catorce and along the first half of the way between Los Catorce and Real de Catorce, Barboza-Gudiño et al. (2004) and Venegas-Rodríguez et al. (2009) described the informal unit named “Cerro El Mazo beds”. In the region, this unit is only recognized in the area of Real de Catorce and consists of more than 300 m of quartzite and conglomeratic sandstone, green, yellow and red shale, as well as interlayered andesitic “greenstone”, including massive deposits composed of conglomeratic sandstone with floating quartz pebbles to blocks of quartzite and cherty rocks in a muddy-sandy matrix. The Cerro El Mazo beds rest with erosive contact on turbidites of the Zacatecas Formation. According to U-Pb data of detrital zircons reported by Venegas-Rodríguez et al. (2009) and new data by Wengler (2014), the lower part of this succession corresponds to the latest Triassic, and the Triassic-Jurassic boundary is located in the middle to upper part of the succession. To the top, the presence of nodular shales, flasher bedding and possible paleosols indicates a marginal marine sequence that contains interlayered volcanogenic materials interpreted as the base of the Lower Jurassic Nazas Formation.

Outcrops of the informal unit “Cerro El Mazo beds” (Barboza-Gudiño et al., 2004, Venegas-Rodríguez et al., 2009) occur west and east of the town “Los Catorce”, along the road to Santa Cruz de Carretas, and along the road to Real de Catorce, for example at coordinates 23°42’N; 100°54.3 W. These sections represent, respectively, west-dipping and east-dipping flanks of a north-south trending antiform, the core of which shows along the canyon the oldest stratigraphic units known from the Mesa Central. The Cerro El Mazo beds are composed of quartzite or conglomeratic, compact litharenite which includes the Triassic-Jurassic boundary in the Cañón General succession (Figure 6A) containing remains of plants (Figure 6B), probably Cycadeoids like Zamites sp. (Bartolini et al., 1999), and interlayered red and green-yellow mudstone (Fig. 6C), as well as basaltic-andesitic dikes and lava flows that represent products of synsedimentary volcanic activity (Figure 6D). At least the upper part of the Cerro El Mazo beds is interpreted as shallow marine marginal facies. Upwards in the succession, interlayered andesitic to rhyodacitic lava flows and pyroclastic deposits represent the basal Early Jurassic volcanic succession known in the region as the Nazas Formation. On the east flank of the above-described Los Catorce Antiform, the Nazas Formation is thick, up to more than 200 m over the Cerro El Mazo beds, whereas southwest of Los Catorce, the thickness of the Lower Jurassic volcanic succession, markedly diminishes to only a few meters of green and red-purple colored tuffaceous layers. Some field verifications are required in order to determine the presence of a possible more voluminous volcanic edifice to the east of Los Catorce antiform as the origin of the thickness change, or the same as a result of a typical structure of an inverted half graben where the Nazas would have been deposited inside the sunken block and partially outside it, later this structure would have been inverted by contractive tectonics, originating the described antiform.

4. Nazas Formation: Lower Jurassic Volcanic Succession

The Nazas Formation (Pantoja-Alor, 1972) is a Lower to Middle Jurassic volcanic and volcano-sedimentary succession, composed at the type locality in the Villa Juárez Uplift, northeastern Durango of intermediary to felsic lavas, ash and tuffs, including several related epiclastic deposits. Other authors described pre-Oxfordian volcano sedimentary successions outcropping in several localities from north-central and northeastern to southeastern México (López-Infanzón, 1986; Grajales-Nishimura et al., 1992; Tristán-Gonzalez and Torres-Hernández, 1994; Jones et al., 1995; Blickwede, 2001; Bartolini et al., 2003; Barboza-Gudiño et al., 2004, 2008, 2012; Barboza-Gudiño, 2012; Godínez-Urban et al., 2011; Zavala-Monsiváis et al., 2012; Lawton and Molina-Garza, 2014). Most of these authors related the volcanic activity to the active continental margin of southwestern North America during the Early to Middle Jurassic, as well as partially to activity of a volcanic arc extending across southwestern north America and known in México as the “Nazas arc” (Bartolini et al., 2003). Martini and Ortega-Gutiérrez (2018), interpreted the volcanic activity as related to intrusion of magmas in the back-arc zone, forced by flat subduction of the Farallón plate during Early Jurassic time, and considered magmatic rocks outcropping in western Mexico as the true continuation of the western North American magmatic-arc. A recent review article by Busby and Centeno-García (2022) largely agrees with Martini and Ortega-Gutierrez (2018). U-Pb geochronology of detrital zircons shows that the volcanic activity in north central to northeastern Mexico was active for a period of ca. 30-40 Ma during the Early and Middle Jurassic, from ca. 195 Ma to 165 Ma BP (Barboza-Gudiño et al., 2008, 2012).

At the point known as La Purísima, on the road between Los Catorce and Real de Catorce, the volcanic succession of the Nazas Formation, consisting of rhyodacites and felsic pyroclastic rocks, rests on red siltstones, greenstones and quartzite of the “Cerro El Mazo beds” (Barboza-Gudiño et al., 2004; Venegas-Rodríguez et al., 2009). The pyroclastic rocks consist of ash-fall deposits or laminated ash; some locally contained welded tuff near the vent and unwelded tuff at a distance where smaller, cooler particles fell to the ground; also recorded are volcanic breccia horizons and marked pseudo-stratification at the base (Figure 7A), while bedded pyroclastic deposits grade up-section into massive deposits showing several intensely sheared zones containing sericite due to both? Dynamic metamorphism and associated hydrothermal alteration. Basaltic-andesitic lavas also occur in the volcanic succession exposed in the Sierra de Catorce. The lava contains fluidal porphyritic texture with highly altered, probable hornblende phenocrysts, scarce pyroxene, olivine, and plagioclase in a fine groundmass composed of acicular plagioclase, ferromagnesian minerals, and opaque grains (Figure 7B ). Some lavas represent auto-breccias formed during flow emplacement. Similar basaltic-andesitic lavas crop out at Sierra de Salinas and Sierra de Charcas. The volcanic units are unconformably overlain by Middle to Upper Jurassic red beds of La Joya Formation (Figures 7C and D). The exposed succession along the road to Real de Catorce forms part of the eastern flank of Los Catorce Antiform.

U-Pb (zircon) ages of the volcanic rocks indicate an Early Jurassic age for the basal part of the Nazas Formation in the area. Now, a possible volcanic arc origin is doubtful for the layers of interstratified greenstone in the Cerro El Mazo beds, considered Upper Triassic-Lower Jurassic in age. To augment previously published U-Pb ages for Real de Catorce volcanic rocks (Barboza-Gudiño et al. 2008) and other localities of the Nazas Formation in the region (Jones et al., 1995; Barboza-Gudiño et al., 2008, 2012; Zavala-Monsiváis et al., 2012; Lawton and Molina-Garza, 2014), we present a new U-Pb age (zr) of the pyroclastic rocks at coordinates: 23°41.7’N; 100°53.6’W, on the La Purísima-Real de Catorce road. The dated sample is a light green pyroclastic rock with scarce ochre to gray aphanitic subangular lithoclasts. More than 50 zircons were separated from the sample, using standard mineral separation techniques and 43 of them were analyzed by the LA-ICP-MS technique, considering analytical errors and procedures as described by Gehrels et al. (2008). The geochronology studies were conducted in the LaserChron Laboratory at the University of Arizona at Tucson.

Zircon crystals were analyzed with a VG Isoprobe multicollector ICPMS equipped with nine Faraday collectors, an axial Daly collector, and four ion-counting channels. The Isoprobe is equipped with an ArF Excimer laser ablation system, with an emission wavelength of 193 nm. The collectors allow measurement of 204Pb in the ion-counting channel while 206Pb, 207Pb, 208Pb, 232Th and 238U are simultaneously measured with Faraday detectors. The analyses were conducted in static mode with a laser beam diameter of 35–50 μm, operated with an output energy of 32 mJ (at 23 kV) and a pulse rate of 9 Hz. Each analysis consisted of one 20-s integration on peaks with no laser firing and 20 1-s integrations on peaks with the laser firing. The analysis was monitored by analyzing an in-house zircon standard, which has a concordant TIMS age of 564 ± 4 Ma (2σ). This standard was analyzed once for every five unknowns in detrital grains. The Pb isotopic ratios were corrected for common Pb, using the measured 204Pb, assuming an initial Pb composition according to Stacey and Kramers (1975).

The concordia diagram and weighted mean calculations were made using the Isoplot/Excel v. 3.0 program (Ludwig, 2003). Uncertainties on individual analyses in data table are reported at a 1σ level and weighted mean ages were calculated at 95% of confidence. Zircon grains from the sample SC12-1 have Th, U contents of 35–1477 ppm, 118–1318 ppm respectively, and Th/U ratios of 0.15-1.55 (Table 1), indicating a magmatic origin (Hoskin and Schaltegger, 2003). In the 207Pb/235U - 206Pb/238U concordia diagram, 30 analytical spots concentrate close to the concordia line, 9 analyses have 206Pb/238Pb ages between 178 – 186 Ma, but have very high uncertainties; the other 21 analyses yielded a weighted mean 206Pb/238U age of 179.4 ± 2.1 Ma (MSWD = 0.38), which is interpreted as the crystallization age of the pyroclastic rock of La Purisima-Real de Catorce. Figure 8 shows a concordia diagram and a weighted mean age plot diagram.

5. La Joya Formation: Middle Jurassic Break-up Unconformity

Imlay et al. (1948) first described the red beds exposed in the Huizachal Valley, Sierra Madre Oriental, in southern Tamaulipas, and named them Huizachal Formation. The separation in this region of two red-bed units by Mixon et al. (1959) into the older La Boca and younger La Joya formations led to a more detailed subdivision of the succession, assigning both units to the Huizachal Group. La Boca Formation is well exposed in the Huizachal-Peregrina anticlinorium in Tamaulipas, where it consists of more than 1500 m of red sandstone, siltstone, mudstone, and conglomeratic sandstone intercalated with volcanic products comparable to the Nazas arc volcanics (Barboza-Gudiño et al., 2012). Mixon et al. (1959) defined La Joya Formation as a sequence of conglomerates and red sandstones, which rest unconformably on red beds of La Boca Formation near the town of La Joya Verde, in the Huizachal Valley. La Joya represents a fining upward sequence composed of polymictic conglomerates and breccias that include clasts of volcanic, plutonic, and metamorphic rocks, as well as older sedimentary rocks (Rubio-Cisneros and Lawton; 2011, Barboza-Gudiño et al., 2012).

The La Joya Formation overlies a regional erosional unconformity (Barboza-Gudiño, 2012) and its thickness varies in northeastern Mexico from zero to over 500 m (Barboza-Gudiño, 2012; Barboza-Gudiño et al., 2014). Michalzik (1991) interprets these deposits to record an environmental change from terrestrial to marine conditions during Middle to early Late Jurassic Time, which agrees with several maximum depositional ages later reported for the La Boca and La Joya formations by Rubio Cisneros and Lawton (2011) and Barboza-Gudiño et al. (2012, 2014).

At “Puerta del Sol” west of Real de Catorce (23°41.8’N; 100°53.5’W) polymictic conglomerate (Figure 7C) and conglomeratic red sandstones of the basal member of La Joya Formation (Bathonian to Callovian) form a prominent cliff, and offer a panoramic view of the stratigraphic units exposed in the Canyon, which include, with only minor stratigraphic breaks or erosional unconformities, the most complete Jurassic succession exposed in the Mesa Central province (Figure 9). The conglomerates underlie red sandstones in the middle part of the succession (Figure 7D) that change up section to a thick mostly massive red siltstone unit. The complete La Joya section is more than 500 m in thickness. At Cerro El Mazo in the western flank of Los Catorce antiform, a 250 m thick fining upward sequence of La Joya Formation consists of a basal 80 to 100 m thick conglomerate-breccia member and an upper member of red sandstone-siltstone as much as 150 m thick.

6. Upper Jurassic Marine Transgression

The upper part of La Joya Formation grades into shallow marine deposits of the Zuloaga Formation (Imlay, 1938). The Zuloaga Formation consists of medium to dark gray limestone and local interlayered evaporites and marls. The limestones are frequently thick bedded, mudstones to grainstones with stylolites and rare gray to brown chert nodules or bands, with a few fossils of Nerinea sp. and corals. The Zuloaga represents shallow marine sedimentation in an Oxfordian sea during the initial marine transgression in the area and probably represents a pelagic lime mud facies (Bacon, 1978; Oivanki, 1974). At Real de Catorce, the Zuloaga Formation is more than 300 m thick and, as all the overlying Upper Jurassic-Cretaceous rocks, is intensely folded, showing an abrupt upper contact concordant with the pink to gray thin bedded fine-grained marine clastics and marls of La Caja Formation. The La Caja Formation in San Luis Potosí and Zacatecas is an offshore basinal equivalent of the Kimmeridgian–Tithonian shallow-marine clastic sediments of La Casita Formation in Nuevo León and Tamaulipas.

7. Cretaceous–Paleogene Deformation

The Triassic–Early Jurassic sedimentary and volcanic rocks of the Sierra de Catorce exhibit folds and pervasive cleavage related to ancient pre-Late Jurassic compressive deformation. Late Jurassic–Early Cretaceous carbonates in turn are detached from the older units at the base of the Zuloaga Formation. Finally, in their external structure, the Sierra de Catorce is an uplifted block bounded by regional north-south-trending post-Laramide extensional faults. The extensional north-south faults cut northwest-trending sinistral faults and poor developed northeast-trending dextral faults. In turn, north-south and oblique faults are cut by west-northwest trending normal faults.

The Jurassic–Cretaceous carbonate sequence shows intense folding and imbrication (Figure 10). Cretaceous limestones west of Los Catorce rest direct upon La Joya Formation as a result of nappe thrusting (Hoppe, 2000; Barboza-Gudiño et al., 2004) showing strong shearing or tectonically depressed layers and omission of parts of the Upper Jurassic–Lower Cretaceous stratigraphic units such as the Oxfordian Zuloaga Formation and the Tithonian La Caja Formation.

In all the studied outcrops in the study area, the lower part of the Zuloaga Limestone consists of a mylonitic rock with strong foliation and extensional lineation, as a well-developed S-C fabric or sometimes appearing as a white colored horizon along the contact with the underlying fine grained red sandstone and siltstone beds of La Joya Formation (Figure 11A). In all the cases, tectonic transport was to the east-northeast (Gutierrez-Navarro, 2017). Cretaceous carbonate layers above the detachment show also folds and intraformational shear zones with a general transport to the East-Northeast.

The point known as “El Mirador” (23°42’N; 100°51.5’W) close to the entrance to the Ogarrio tunnel, coming from the road Cedral-Vanegas, north of the Sierra de Catorce, offers a good overview of the strongly deformed rocks of the detached Jurassic–Cretaceous cover. Cretaceous limestones, as well as Oxfordian medium-bedded limestones of the Zuloaga Formation (Imlay, 1938) and marls of the Kimmeridgian–Berriasian La Caja Formation (Imlay, 1938; Olóriz, et al., 1999), are well exposed along the road to Real de Catorce. They are intensely folded, showing several minor thrust faults formed by Laramide shortening (Figure 12), which also produced uplift of Sierra de Catorce and the detachment of its mostly calcareous Upper Jurassic-Cretaceous cover. The detachment allowed independent deformation of the cover into north-northwest trending folds (Figure 11B). The age of folding in the area is constrained between Campanian–Maastrichtian folded sediments and undeformed middle Eocene quartz monzonite and granodiorite porphyritic intrusions dated at 48.6 ± 0.8 Ma (Huerta-Gonzalez, 2017) and 44.6 ± 0.1 Ma (Díaz-Bravo et al., 2021) (Figure 11C). 40Ar/39Ar dating of authigenic illite in detachment surfaces, shear zones and slip surfaces on the flanks of folds indicates three deformation episodes during the Cenomanian–Turonian (96 – 90 Ma), Campanian– Maastrichtian (80 – 69 Ma) and Paleocene–Lower Eocene time (62 – 52 Ma) (Gutiérrez-Navarro, 2017; Gutierrez-Navarro et al., 2021).

8. Magmatism and Mineralization

The described Mesozoic stratigraphic units were intruded by Eocene granodioritic dikes, which are considered apophyses of a main intrusion in Potrero de Catorce-Real de Catorce area, like a second stock exposed in Real de Maroma, in the southern Sierra de Catorce, consisting of a major stock with several radial associated dikes. Polymetallic Ag, Pb, Zn, Sb and Hg mineralization in the area is spatially and likely genetically related to such hydrothermal activity related to the magmatism. The rock shows idiomorphic white phenocrysts of feldspar, plagioclase and minor quartz in a microcrystalline gray-green to yellow matrix, usually very altered throughout the sierra with the exception of some outcrops ca. 3 km to the northeast of Potrero de Catorce town to the left (east) of the road to Cedral (Fig. 11D).

According to their geochemical and petrographic characteristics, the intrusives are of calc-alkaline character. They display the geochemical signature of a continental volcanic arc associated with a subduction environment; absence of deformation indicates emplacement during or after the final compressive phase of the orogeny. U-Pb (zr) ages from Mascuñano et al., (2013) vary between 40.8 ± 0.3 to 44.6 ± 0.8 Ma. In addition, Huerta-Gonzalez (2017) reported a U-Pb age (Zr by LA-MC-ICPMS) of 48.6 ± 0.8 for the intrusion outcropping north of Potrero de Catorce (23°44’N; 100°50.2’W).

In the epithermal districts Real de Catorce at the north and Real de Maroma at the south of the Sierra de Catorce, mineralization occurred in form of veins, stock works and strata bound bodies. Silver occurs in the oxidation zone as cerargirite-browargirite and minor cerussite; malachite and iron and mangan oxides are common. Deeper in the sulfide zone, lead, zinc and copper sulfides are present. Antimony is more common in the Tierras Negras District, 10 km south of Real de Catorce, in the form of strata-bound calcite, corps, with stibnite or in many cases in the oxidized zone as radial, stibnite pseudomorphs of cervantite and valentinite aggregates.

Discussion and Conclusions

The recognized stratigraphy of Sierra de Catorce includes Upper Triassic turbidites of the Zacatecas Formation; uppermost Triassic to Lower Jurassic slope up to marine marginal strata with interlayered andesitic greenstone dikes and flows of the informally named “Cerro El Mazo beds”; rhyolitic, dacitic, and andesitic volcanic rocks of the Lower to Middle Jurassic Nazas Formation; continental to shallow marine conglomerate and red beds of La Joya Formation; limestones of the Zuloaga Formation, and the uppermost Jurassic beds of the La Caja Formation, the top of which represents the Jurassic–Cretaceous boundary. The transition between red beds of La Joya Formation into limestone of the Zuloaga Formation is a detachment surface, or zone, where limestones of the Zuloaga Formation are strongly mylonitized and appear as a whitish zone, microbrecciated, with numerous shear bands.

In order to facilitate use of the present article as a field guide, in Figure 13 we present a detail of the geological map of the area, on which the most illustrative outcrops of the lithological structures and units described are indicated by the numbers corresponding to the respective figures throughout the text.

In the Sierra de Catorce, Upper Triassic strata of the Zacatecas Formation record submarine fan facies sedimentation that change upwards from turbiditic sandstones and shale into channelized conglomeratic sandstones and quartzite, with interlayered siltstone and volcanic-volcaniclastic rocks. The occurrence of slump structures and chaotic deposits with several intrabasinal floating clasts and blocks in the succession indicates mass flow events close to an ancient continental margin.

Upward in the section exposed near Real de Catorce, the presence of shales with carbonate nodules, flaser bedding and glauconite as well as possible paleosols indicates a shallowing of the basin and a transition from deep marine to a marginal-marine facies with red sandstones interlayered with volcanogenic materials that mark the bottom of the Lower Jurassic Nazas Formation.

All pre-Jurassic rocks are strongly deformed suggesting that a subduction zone was active in a low stress stage during latest Triassic time, producing deformation of Zacatecas Formation strata. In contrast, sedimentary and volcanogenic layers of the Nazas are notably less deformed than Zacatecas beds, related probably to a subsequent high-stress stage of subduction that provoked also the volcanic arc or subduction magmatism.

In accordance with our observations and specifically in the studied key outcrops of the Cañón General area, the Triassic to Lower Jurassic stratigraphic units represent remnants of paleogeographic elements that evolved close to the paleo-Pacific margin of western Pangea during early Mesozoic time. To the west, parts of the sedimentary pile consisting of deep marine turbidites were likely deposited on the continental slope and the adjacent oceanic floor. To the east, the continent Pangea began to disperse, and coeval fluvial and alluvial red beds of the Triassic El Alamar Formation (Barboza-Gudiño et al., 2010) and the Lower Jurassic La Boca Formation (Mixon et al., 1959), known also as the Huizachal Formation (Imlay et al., 1948, Carrillo-Bravo, 1961), were deposited on the edge of the continent. Westwards, such continental Lower Jurassic sediments are interstratified with the Early Jurassic volcanogenic rocks of the Nazas Formation.

The La Joya Formation represents the sedimentation following development of a regional erosional unconformity recognized in several localities of northeastern Mexico. This regional unconformity was related to the opening of the Gulf of Mexico basin and thus represents a break-up unconformity (Michalzik, 1988). The La Joya Formation overlies in many localities the volcanic rocks of the Nazas arc, and a large proportion of its clastic components are eminently characterized as a product of subduction volcanism, which places the La Joya Formation and the Gulf of Mexico itself in a controversial back-arc position (Stern and Dickinson, 2010).

In addition to Early Jurassic volcanic rocks in San Luis Potosí and northeastern Mexico, correlative calk-alkaline granitoids are exposed in Sonora, Baja California, Marias Islands and in southern México State, suggesting also a Lower–Middle Jurassic subduction in this region more than 600 km far away at the actual pacific coast. Martini and Ortega-Gutiérrez (2018) proposed regional extension imposed by North America-South America divergence that allowed the emplacement of magmas that were clearly influenced by the subduction of the Farallon Plate from the Pacific. Those authors interpreted the Nazas arc as a hybrid domain that reflects the superposition of the Atlantic and Pacific tectonic processes.

The Sierra de Catorce stratigraphy, and the nature and regional distribution of the several described Triassic and Jurassic lithostratigraphic units, suggest the occurrence of a continental margin in central Mexico that evolved from a passive to an active margin during Late Triassic-Early Jurassic time.

Late Jurassic Early Cretaceous carbonates in the Sierra de Catorce are detached from the older units at the base of the Zuloaga Formation. Cretaceous limestones west of Los Catorce structurally overlie the Middle Jurassic La Joya Formation because of nappe thrusting and tectonically omitted Upper Jurassic–Lower Cretaceous layers in the succession. The age of folding in the area is constrained between Campanian–Maastrichtian folded sediments and undeformed middle Eocene quartz monzonite, and granodiorite porphyritic intrusions (Mascuñano et al., 2013). Mineralization in the area is spatially and probably genetically related to hydrothermal activity associated with the same magmatism (Huerta-Gonzalez, 2017).

References

Bacon, R. W., 1978, Geology of northern Sierra de Catorce, San Luis Potosí, México: University of Texas at Arlington, Master thesis, 109 p.

Barboza-Gudiño, J. R., 2012, Sedimentary Tectonics and Stratigraphy: The Early Mesozoic Record in Central to Northeastern Mexico, in Ömer, E. (ed.), Stratigraphic Analysis of Layered deposits, 255–278, InTech. http://dx.doi.org/10.5772/35219

Barboza-Gudiño, J. R., Hoppe, M., Gómez-Anguiano, M., Martínez-Macías, P. R., 2004, Aportaciones para la interpretación estratigráfica y estructural de la porción noroccidental de la Sierra de Catorce, San Luis Potosí, México: Revista Mexicana de Ciencias Geológicas, 21, 299–319. https://www.redalyc.org/articulo.oa?id=57221301

Barboza-Gudiño, J. R., Orozco-Esquivel, M. T., Gómez-Anguiano, M., Zavala-Monsiváis, A., 2008, The Early Mesozoic volcanic arc of western North America in northeastern Mexico: Journal of South American Earth Sciences, 25, 49–63. https://doi.org/10.1016/j.jsames.2007.08.003

Barboza-Gudiño, J. R., Zavala-Monsiváis, A., Venegas-Rodríguez, G., Barajas-Nigoche, L. D., 2010, Late Triassic stratigraphy and facies from northeastern Mexico: Tectonic setting and provenance: Geosphere, 6(5), 621–640. https://doi.org/10.1130/GES00545.1

Barboza-Gudiño, J. R., Molina-Garza, R. S., Lawton, T. F., 2012, Sierra de Catorce: Remnants of the ancient western equatorial margin of Pangea in central Mexico in Aranda-Gómez, J. J., Tolson, G., Molina-Garza, R. S., (eds.), The Southern Cordillera and Beyond: Geological So­ciety of America. https://doi.org/10.1130/2012.0025(01).

Barboza-Gudiño, J. R., Ocampo-Díaz, Y. E., Zavala-Monsiváis, A., López-Doncel, R. A., 2014, Procedencia como herramienta para la subdivisión estratigráfica del Mesozoico temprano en el noreste de México: Revista Mexicana de Ciencias Geológicas, 31(3), 301–324. https://doi.org/10.22201/cgeo.20072902e.2014.3.236

Bartolini, C., Lang, H., Spell, T., 2003, Geochronology, geoche­mistry, and tectonic setting of the Mesozoic Nazas arc in north-central Mexico, and its continuation to northern South America, in Bartolini, C., Buffler, R.T., Blickwede, J.F. (eds.), The Circum-Gulf of Mexico and the Caribbean: Hydrocarbon Habitats, Basin Formation and Plate Tectonics: American Association of Petroleum Geo­logists Memoir, 79, 427–461. https://doi.org/10.1306/M79877C20

Bartolini, C., Lang, H., Stinnesbeck, W., 1999, Volcanic rock outcrops in Nuevo León, Tamaulipas and San Luis Potosí, Mexico: Remnants of the Permian-Early Triassic magmatic arc?, in Bartolini, C., Wilson, J. L., Lawton, T. F. (eds.), Mesozoic Sedimentary and Tectonic History of North-Central Mexico, Boulder Colorado, Geological Society of America, Special Paper 340, 347–356. https://doi.org/10.1130/0-8137-2340-X.347

Blickwede, J. F., 2001, The Nazas Formation: A detailed look at the early Mesozoic convergent margin along the western rim of the Gulf of Mexico Basin, in Bartolini, C., Buffler, R. T., Cantú-Chapa, A. (eds.), The Western Gulf of Mexico Basin: Tectonics, Sedimentary Basins, and Petroleum Systems: American Association of Petroleum Geologists Memoir 75, 317–342. https://doi.org/10.1306/M75768C13

Burckhardt, C., Scalia, S., 1905, La faune marine du Trias Supérieur de Zacatecas: Instituto de Geología de México Boletín, 21, 44 p.

Busby C. J., Centeno-García E., 2022, The “Nazas Arc” is a continental rift province: Implications for Mesozoic tectonic reconstructions of the southwest Cordillera, U.S. and Me­xico: Geosphere, 18 (10), 1–23. https://doi.org /10.1130/GES02443.1

Carrillo-Bravo, J., 1961, Geología del Anticlinorio Huizachal-Peregrina, al NW de Ciudad Victoria, Tamaulipas: Boletín de la Asociación Mexicana de Geólogos Petroleros, 13 (1-2), 1–98.

Centeno-García, E., 2005, Review of Upper Paleozoic and Mesozoic stratigraphy and depositional environments of central and west Mexico: Constraints on terrane analysis and paleogeography, in Anderson, T. H., Nourse, J. A., McKee, J. W., Steiner, M.B. (eds.), The Mojave-Sonora Megashear hypothesis: Development, assessment and alternatives: Geological Society of America Special Paper, 393, 233–258. https://doi.org/10.1130/0-8137-2393-0.233

Centeno-García, E., Guerrero-Suastegui, M., Talavera-Mendoza, O., 2008, The Guerrero Composite Terrane of western Mexico: Collision and subsequent rifting in a supra-subduction zone: Geological Society of America Special Paper, 436, 279–308. https://doi.org/10.1130/2008.2436(13)

Córdoba-Méndez, D. A., 1964, Geology of Apizolaya Quadrangle (east half), northern Zacatecas, Mexico, Austin, The University of Texas, Master thesis, 111 p.

Díaz-Bravo, B. A., Barboza-Gudiño, J. R., Ortega-Obregón, C., Morales-Gámez, M. 2021, Late Cretaceous to Oligocene overlapping plutonic magmatism episodes in the eastern Mesa Central province of Mexico: International Geology Review. https://doi.org/10.1080/00206814.2021.1871866

Godínez-Urban, A., Lawton, T. F., Molina-Garza, R. S., Iriondo, A., Weber, B., López-Martínez, M., 2011, Jurassic volcanic and sedimentary rocks of La Silla and Todos Santos formations, Chiapas: Record of Nazas arc magmatism and rift-basin formation prior to opening of the Gulf of Mexico: Geosphere, 7, 121–144. https://doi:10.1130/GES00599.1

Gehrels, G. E., Valencia, V. A., Ruiz, J., 2008, Enhan­ced precision, accuracy, efficiency, and spa­tial resolution of U-Pb ages by laser ablation–multicollector–inductively coupled plasma–mass spectrometry: Geochemistry, Geophysics, Geosystems, 9, 1–13. https://doi.org/10.1029/2007GC001805

Grajales-Nishimura, J. M., Terrell, D. J., Damon, P.J., 1992, Evidencias de la prolongación del arco magmático cordillerano del Triásico Tardío-Jurásico en Chihuahua, Durango y Coa­huila: Boletín de la Asociación Mexicana de Geó logos Petroleros, 42, 1–18.

Gutiérrez-Amador, M., 1908, Las capas cárnicas de Zacatecas: Boletín de la Sociedad Geológica Mexicana, 4, 29–35.

Gutiérrez-Navarro, R., 2017, Historia de deformación por acortamiento de la Sierra de Catorce, San Luis Potosí, México, Universidad Nacional Autónoma de México, Master tesis, 82 p.

Gutiérrez-Navarro, R., Fitz-Díaz, E., Barboza-Gudiño, J. R., Stockli, D. F., 2021, Shortening and exhumation of Sierra de Catorce in northeastern Mexico, in light of 40Ar/39Ar illite dating and (U-Th) /He zircon thermochronology: Journal of South American Earth Sciences, 111 103334. https://doi.org/10.1016/j.jsames.2021.103334

Hoppe, M., 2000, Geologische Kartierung (1:10 000) im Gebiet Ojo de Agua, nordwestliche Sierra de Catorce und sedimentpetrologische Untersuchungen an prä-oberjurassischen Sedimenten (“Zacatecas Formation”), Technische Universität Clausthal, Germany, master’s thesis, 235 p.

Hoskin, P. W. O., Schaltegger, U., 2003, The composition of zircon and igneous and metamorphic petrogenesis: Reviews in Mineralogy and Geochemistry, 53, 27–62. https://doi.org/10.2113/0530027

Huerta-González, R. M., 2017, Emplazamiento de pórfidos eocénicos y su posible relación con zonas mineralizadas en la Sierra de Catorce, S.L.P., Universidad Autónoma de San Luis Potosí, México, Master thesis, 88 p.

Imlay, R. W., 1936, Evolution of the Coahuila Peninsula, Mexico: Part IV, Geology of the western part of the Sierra de Parras: Geological Society of America Bulletin, 47, 1091–1152. https://doi.org/10.1130/GSAB-47-1091

Imlay, R. W., 1938, Studies of the Mexican geosyncline: Geological society of America Bulletin, 49, 1651–1694. https://doi.org/10.1130/GSAB-49-1651

Imlay, R. W., Cepeda, D. L. C. E., Álvarez, M., Díaz, G. T., 1948, Stratigraphic Relations of Certain Jurassic Formations in Eastern México: American Association of Petroleum Geologist Bulletin, 2 (9), 1750–1761.

Jones, N. W., McKee, J. W., Anderson, T. H., Silver, L. T., 1995, Jurassic volcanic rocks in northeastern Mexico: A possible remnant of a Cordilleran magmatic arc, in Jacques-Ayala, C., González-León, C., Roldán-Quintana, J. (eds.), Studies on the Mesozoic of Sonora and adjacent areas: Geological Society of America Special Paper, 301, 179–190, https://doi:10.1130/0-8137-2301-9.179

Lawton, T. F., Molina-Garza, R. S., 2014, U-Pb geochronology of the type Nazas Formation and superjacent strata, northeastern Durango, Mexico: Implications of a Jurassic age for continental-arc magmatism in north-central Mexico: Geological Society of America Bulletin, 126 (9/10), 1181–1199. https://doi.org/10.1130/B30827.1

López-Infanzón, M., 1986, Estudio petrogenético de las rocas ígneas en las Formaciones Huizachal y Nazas: Boletín de la Sociedad Geológica Mexicana, 47(2), 1–42. http://dx.doi.org/10.18268/BSGM1986v47n2a1

Ludwig, K. R., 2003, Isoplot 3.00: A geochronological toolkit for Microsoft Excel: Berkeley Geochronology Center Special Pub, No. 4, 74 p.

Martínez-Pérez, J., 1972, Exploración geológica del área El Estribo–San Francisco, San Luis Potosí: Boletín de la Asociación Mexicana de Geólogos Petroleros, 24(7-9), 327–402.

Martini, M., Ortega-Gutierrez, F., 2018, Tectono-stratigraphic evolution of eastern Mexico during the break-up of Pangea: A review: Earth-Science Reviews, 183, 38-55. http://dx.doi.org/10.1016/j.earscirev.2016.06.013

Mascuñano, E., Levresse, G., López, E. C., Cambra, J. T., Esquivel, R. C., Meyzen, C., 2013, Post-Laramide, Eocene magma­tic activity in Sierra de Catorce, San Luis Potosí, México: Revista Mexicana De Ciencias Geológicas, 30, 299–311.

Michalzik, D., 1988, Trias bis tiefste unter-Kreide der nordöstliche Sierra Madre Oriental, Mexiko—Fazielle Entwicklung eines passiven Kontinentalrandes, Ph.D. tesis, Technische Hochschule Darmstadt, 247 p.

Michalzik, D., 1991, Facies sequence of Triassic-Jurassic red beds in the Sierra Madre Oriental (NE Mexico) and its relation to the early opening of the Gulf of Mexico: Sedimentary Geology, 71(3-4), 243–259. https://doi.org/10.1016/0037-0738(91)90105-M

Mixon, R. B., Murray, G. E., Diaz, T. G., 1959, Age and correlation of Huizachal Group (Mesozoic), state of Tamaulipas, Mexico: The American Association of Petroleum Geologists Bulletin, 43, 757–771. https://doi.org/10.1306/0BDA5ED3-16BD-11D7-8645000102C1865D

Oivanki, S. M., 1974, Paleodepositional environments in the upper Jurassic Zuloaga Formation (Smackover) northeastern Mexico: Gulf Coast Association of Geological Societies, 24, 258–278. https://doi.org/10.1306/83D91AF6-16C7-11D7-8645000102C1865D

Olóriz, F., Villaseñor, A. B., González-Arreola, C., Westermann, G. E. G., 1999, Ammonite biostratigraphy and correlations in the Latest Jurassic–Earliest Cretaceous La Caja Formation of North-Central Mexico (Sierra de Catorce, San Luis Potosí), in Olóriz, F., Rodríguez-Tovar, F. J. (eds.), Advancing Research on Living and Fossil Cephalopods: Plenum Press, London, 463–492. https://doi:10.1007/978-1-4615-4837-9_31

Ortega-Gutiérrez, F., Ruiz, J., Centeno-Garcia, E., 1995, Oaxaquia, a Proterozoic microcontinent accreted to North America during the late Paleozoic: Geology, 23, 1127–1130. https://doi.org/10.1130/0091-7613(1995)023<1127:OAPMAT>2.3.CO;2

Pantoja-Alor, J., 1972, La Formación Nazas del Levantamiento de Villa Juárez, Estado de Durango: Segunda Convención Nacional de la Sociedad Geológica Mexicana, Memorias, 25–31.

Rubio-Cisneros, I. I., Lawton, T. F., 2011, Detrital zircon U-Pb ages of sandstones in continental red beds at Valle de Huizachal, Tamaulipas, NE Mexico: Record of Early-Middle Jurassic arc volcanism and transition to crustal extension: Geosphere, 7, 159–170. https://doi.org/10.1130/GES00567.1

Sandstrom, M., 1982, Stratigraphy and environments of deposition of the Zuloaga group, Victoria, Tamaulipas. México, in The Jurassic of the Gulf rim, Gulf Coast Section, SEPM, 3er. Annual Research Conference, Program and Abstracts, 94–97.

Silva-Romo, G., Arellano-Gil, J., Mendoza-Rosales, C., Nieto-Obregon, J., 2000, A submarine fan in the Mesa Central, Mexico: Journal of South American Earth Sciences, 13, 429–442. https://doi:10.1016/S0895-9811(00)00034-1

Stern, R. J., Dickinson, W. R., 2010, The Gulf of Mexico is a Jurassic backarc basin: Geosphere, 6(6), 739–754. https://doi.org/10.1130/GES00585.1

Stacey, J., Kramers, J, 1975, Approximation of terrestrial lead isotope evolution by a two-stage model: Earth and Planetary Science Letters, 26, 207 – 221. https://doi.org/10.1016/0012-821X(75)90088-6

Tristán-González, M., Torres-Hernández, J. R., 1994, Geología de la Sierra de Charcas, Estado de San Luis Potosí, México: Revista Mexicana de Ciencias Geologicas, 11, 117–138.

Venegas-Rodríguez, G., Barboza-Gudiño, J. R., López-Doncel, R. A., 2009, Geocronología de circones detríticos en capas del Jurásico Inferior de las áreas de la Sierra de Catorce y El Alamito en el estado de San Luis Potosí: Revista Mexicana de Ciencias Geológicas, 26(2), 466–481.

Wengler, M., 2014, Provenance Analysis of Triassic and Jurassic sediments in NE Mexico, University Göttingen, Germany, Master thesis, 139 p.

Zavala-Monsiváis, A., Barboza-Gudiño, J. R., Velasco-Tapia, F., García-Arreola, M. E., 2012, Sucesión Volcánica Jurásica en el Área de Charcas, San Luis Potosí: Contribución al entendimiento del Arco Nazas en el noreste de México. Boletín de la Sociedad Geológica Mexicana, 64(3), 277–293. http://dx.doi.org/10.18268/BSGM2012v64n3a2

Cómo citar / How to cite: Barboza-Gudiño, J. R.; Morales Gamez, M.; Díaz-Bravo, B. A.; Huerta González, R. M.; Zavala-Monsiváis, A.; Jaime Rodríguez, D. (2023). Guide to Key Outcrops for Reconstruction of the Geologic-Tectonic History of Sierra de Catorce: Northeastern Mexico. Boletín del Instituto de Geología 126, 1–22.

Manuscrito recibido: Diciembre 2, 2022.

Manuscrito corregido recibido: Marzo 6, 2023.

Manuscrito aceptado: Marzo 14, 2023.

Figure 1. Major paleogeographic elements in the state of San Luis Potosí, north-central to northeastern Mexico and location of the Sierra de Catorce.

Figure 2. Regional map and access to the Sierra de Catorce, modified from Barboza-Gudiño et al., 2012.

Figure 3. Geological map of the northwestern Sierra de Catorce (from Barboza-Gudiño et al., 2014, 2012).

Figure 4. Stratigraphic column of the Sierra de Catorce (after Barboza-Gudiño et al., 2004, 2012).

Figure 5. A. Deformed sandstone interlayered with black shale horizons in the Upper Triassic turbiditic succession (Zacatecas Formation) of Los Catorce; B. Slump folds in fine-grained sandstones layers of the Zacatecas Formation at Los Catorce; C and D floating quartz pebbles and carbonate cemented sandstone and cherty blocks in massive deposits corresponding to the uppermost part of the Triassic Succession of the informal unit named “Cerro El Mazo beds”, exposed in the neighborhood of the town of Los Catorce.

Figure 6. A. Thick bedded quartzite or conglomeratic litharenite showing internal cross-lamination in the Cerro El Mazo beds along the road from Los Catorce to Real de Catorce; B. Remains of plants, probably Cycadeoids similar to Zamites sp. in the basal sequence of Cerro El Mazo beds, east of Los Catorce. Coin is 2 cm in diameter; C. Thick bedded quartzite, and litharenite interlayered with red and yellowish mudstone layers; D. Chloritized products of mafic synsedimentary volcanic activity are interlayered in the marine marginal facies of the Cerro El Mazo beds. In the microphotography under XPL (magnification ×5) the lower half of the view is composed of laths of fine plagioclase and some carbonate patches and the upper half displays a clastic deposit with abundant fine-to medium-grained quartz fragments.

Figure 7. A. Pyroclastic layered rock at Socavón La Purísima, road Los Catorce-Real de Catorce; B. Mafic to intermediary lava flow in the neighborhood of Real de Catorce, shows under the microscope several possible pyroxene and olivine phenocrysts, according to the external form of crystals, because they are replaced by opaque minerals (XPL, 5X magnification); C. Conglomerate of the lower part of La Joya Formation west of Real de Catorce; D. Red siltstone and fine-grained sandstone layers in the upper part of La Joya Formation east of Real de Catorce.

Table 1. LA-ICPMS, U-Th-Pb analytical data for zircons of sample SC12-1 from Stop 3.

Isotopic Rates

Aparents Ages

Zircon

Th (ppm)

U (ppm)

Th/U

207Pb/

207P

206Pb

207Pb/

± (Ma)

207Pb/

± (Ma)

206Pb/

± (Ma)

Best Age

± (Ma)

206Pb

/235U

/238U

206Pb

235U

238U

SC12-_02

148

274

0.54

0.050

1.548

0.197

2.839

0.028

1.603

214

35.5

182

4.7

180

2.8

180

2.8

SC12-_03

324

264

1.23

0.05

1.404

0.272

2.702

0.040

1.579

195

32.3

244

5.8

250

3.9

250

3.9

SC12-_05

191

216

0.88

0.049

1.588

0.188

2.969

0.028

1.766

147

36.8

175

4.8

178

3.1

178

3.1

SC12-_06

129

161

0.8

0.050

1.787

0.188

3.145

0.028

1.764

169

41.2

175

5

175

3.1

175

3.1

SC12-_07

189

216

0.88

0.053

1.865

0.210

3.480

0.029

2.206

339

41.7

194

6.1

182

4

182

4

SC12-_08

513

460

1.12

0.051

1.420

0.207

2.704

0.029

1.561

259

32.3

191

4.7

186

2.9

186

2.9

SC12-_09

111

136

0.81

0.052

5.078

0.207

7.094

0.029

2.656

294

112

191

12.3

183

4.8

183

4.8

SC12-_10

684

494

1.38

0.051

1.346

0.197

2.591

0.028

1.472

224

30.8

183

4.3

180

2.6

180

2.6

SC12-_11

234

314

0.75

0.052

1.281

0.284

2.538

0.040

1.473

280

29.1

254

5.7

252

3.6

252

3.6

SC12-_12

270

256

1.06

0.051

1.579

0.202

2.884

0.029

1.635

216

36.1

187

4.9

185

3

185

3

SC12-_15

510

467

1.09

0.052

1.347

0.202

2.667

0.029

1.600

262

30.6

187

4.5

181

2.9

181

2.9

SC12-_18

59

131

0.45

0.052

2.038

0.205

3.372

0.029

1.745

288

45.9

189

5.8

182

3.1

182

3.1

SC12-_19

477

443

1.08

0.052

1.251

0.280

2.492

0.039

1.437

274

28.4

251

5.5

249

3.5

249

3.5

SC12-_20

423

562

0.75

0.05

1.303

0.196

2.562

0.029

1.483

194

30

182

4.3

181

2.7

181

2.7

SC12-_22

232

301

0.77

0.051

1.469

0.206

2.798

0.029

1.65

257

33.4

190

4.8

185

3

185

3

SC12-_23

201

1318

0.15

0.073

1.082

1.551

2.497

0.154

1.650

1019

21.8

951

15.3

923

14.2

923

14.2

SC12-_24

209

309

0.68

0.050

1.609

0.196

2.955

0.029

1.712

191

37

182

4.9

181

3.1

181

3.1

SC12-_26

122

166

0.73

0.051

2.437

0.201

3.803

0.029

1.831

234

55.3

186

6.4

182

3.3

182

3.3

SC12-_27

1477

954

1.55

0.051

1.061

0.269

2.319

0.038

1.479

259

24.2

241

5

240

3.5

240

3.5

SC12-_28

508

481

1.06

0.056

1.058

0.534

2.220

0.070

1.322

442

23.4

435

7.8

433

5.5

433

5.5

SC12-_29

97

150

0.65

0.050

2.013

0.198

3.260

0.029

1.62

212

46

184

5.5

181

2.9

181

2.9

SC12-_31

134

306

0.44

0.050

1.464

0.197

2.739

0.029

1.624

200

33.6

182

4.6

181

2.9

181

2.9

SC12-_32

237

253

0.94

0.049

1.712

0.189

2.907

0.028

1.514

130

39.8

176

4.7

179

2.7

179

2.7

SC12-_33

516

575

0.9

0.052

1.078

0.281

2.242

0.039

1.333

286

24.5

252

5

248

3.2

248

3.2

SC12-_34

215

274

0.78

0.049

1.440

0.190

2.584

0.028

1.388

164

33.3

177

4.2

178

2.4

178

2.4

SC12-_35

111

153

0.72

0.052

1.797

0.202

3.047

0.028

1.622

297

40.5

187

5.2

178

2.8

178

2.8

SC12-_36

35

201

0.18

0.078

1.013

2.092

2.179

0.194

1.313

1155

20

1146

14.9

1142

13.7

1142

13.7

SC12-_37

160

452

0.35

0.050

1.277

0.194

2.431

0.028

1.374

197

29.4

180

4

179

2.4

179

2.4

SC12-_38

123

178

0.69

0.05

1.559

0.194

2.789

0.028

1.562

197

35.8

180

4.6

179

2.8

179

2.8

SC12-_39

234

636

0.37

0.052

1.119

0.334

2.260

0.046

1.308

302

25.3

292

5.7

291

3.7

291

3.7

SC12-_40

436

598

0.73

0.050

1.207

0.193

2.420

0.028

1.457

187

27.9

179

4

178

2.6

178

2.6

SC12-_42

87

134

0.65

0.050

1.830

0.197

3.084

0.028

1.632

207

41.9

182

5.1

180

2.9

180

2.9

SC12-_43

112

187

0.6

0.050

1.484

0.190

2.683

0.028

1.497

186

34.2

177

4.3

176

2.6

176

2.6

SC12-_44

255

511

0.5

0.060

1.028

0.540

3.043

0.065

2.485

600

22.1

438

10.8

408

9.8

408

9.8

SC12-_46

538

847

0.63

0.054

1.339

0.266

2.563

0.036

1.511

380

29.8

239

5.4

225

3.3

225

3.3

SC12-_47

77

125

0.61

0.052

1.892

0.204

3.153

0.029

1.647

284

42.7

189

5.4

181

2.9

181

2.9

SC12-_48

151

525

0.29

0.050

1.190

0.190

2.339

0.027

1.343

209

27.4

177

3.8

174

2.3

174

2.3

SC12-_49

153

228

0.67

0.056

1.184

0.564

2.318

0.073

1.314

451

26.1

454

8.4

455

5.8

455

5.8

SC12-_50

420

837

0.5

0.051

1.154

0.201

2.479

0.029

1.616

230

26.4

186

4.2

182

2.9

182

2.9

SC12-_51

624

619

1.01

0.049

1.175

0.199

2.350

0.029

1.383

159

27.3

184

3.9

186

2.5

186

2.5

SC12-_52

151

118

1.28

0.074

1.072

1.733

2.273

0.169

1.393

1048

21.5

1021

14.5

1009

13

1009

13

SC12-_54

136

292

0.47

0.050

1.565

0.191

2.748

0.028

1.477

184

36

177

4.5

177

2.6

177

2.6

SC12-_55

93

130

0.72

0.052

2.128

0.201

3.499

0.028

1.859

263

48.1

186

5.9

180

3.3

180

3.3

Figure 8. A. Wetherill concordia diagram (U-Pb, zr. LA-MC-ICP-MS technique) for sample SC12-1; B. Detail of youngest population in A, showing the most concordant zircons used for the age calculation as black-line ellipses and discordant zircons as gray-dashed line ellipses; C. The most concordant U-Pb zircon analyses used for the 206Pb/238U age calculation (n = 21) are plotted in a weighted mean age diagram with used 206Pb/238U ages as gray bars.

Figure 9. View of the Cañón General outcrops and stratigraphic column of the exposed units, which include the most complete Jurassic succession exposed in the Mesa Central province: Upper Triassic turbidites of the Zacatecas Formation and uppermost Triassic to Lower Jurassic marine marginal strata of the “Cerro El Mazo beds” (uTr); volcanic rocks of the Lower to Middle Jurassic Nazas Formation (lJ); continental to shallow marine conglomerate and red beds of La Joya Formation (mJ); limestones of the Zuloaga Formation and the uppermost Jurassic beds of the La Caja Formation (uJ), which includes also the Jurassic-Cretaceous boundary at the top. The boundary between red beds of La Joya Formation into limestone of the Zuloaga Formation is a detachment surface, developed in the limestones of the base of the Zuloaga Formation, which appear as a strongly foliated and scratched whitish zone (My).

Figure 10. View of the northern wall of Cañón General, western flank of Los Catorce Antiform in front of Cerro El Mazo. Cretaceous limestones are resting directly on red beds of the Middle Jurassic La Joya Formation as a result of a lack of the Upper Jurassic Zuloaga and La Caja formations as well as the lowest Cretaceous strata. This stratigraphic relation is a result of a nappe structure that evolved during the earliest horizontal laramidic movements which were stopped probably against an uplifted block in the central part of the Sierra de Catorce for this time, resulting in thrusting and imbrications and folding within the calcareous Upper Jurassic-Lower Cretaceous cover in the region.

Figure 11. A. Detachment zone at the bottom of the limestones of the Zuloaga Formation against shale of the top of La Joya Formation. This weak zone allowed independent deformation of the mostly calcareous Upper Jurassic–Cretaceous cover, producing north-northwest-trending folds commonly asymmetric to recumbent to the east. B. Recumbent fold in the Zuloaga Formation at El Mirador, Potrero de Catorce; C. Outcrop of a strongly eroded granodioritic dike cutting cretaceous limestones of the Tamaulipas Inferior Formation close to El Mirador point; D. Fresh sample of a rock similar to the granodioritic dike of figure 10C is a porphyry, composed of phenocrysts of alkali feldspar, plagioclase and quartz in a fine-grained holocrystalline matrix.

Figure 12. In A) a profile of the section is shown on the road La Luz-Túnel de Ogarrio, the view is from the north to the south and the section is East-West, it can be seen that towards the east part of it. There is a series of systems of inverse faults which have a vergence towards the Northwest, the inverse kinematics is observed in drag folds in the traces of the fault in yellow. B) The stratification is distorted by systems of inverse and normal faults. The geometry of the strata can be observed in C), which represents soft folds that were later deformed by systems of inverse faults with vergence towards the West (Cretaceous?) and extensional systems (Paleogene?). Normal faults form more recent conjugated systems in D); the conjugated main planes can be observed in dotted lines. The shortening deformation stage (Cretaceous?) is represented by reclined folds that present their vergence towards the Northwest as shown by the stereogram in E).

Figure 13. Geological map of the area, showing location of the most illustrative outcrops of lithology, structures and stratigraphic units. Numbers correspond to respective figures in the text.