Rancho Seco monogenetic volcano (Michoacán, Mexico): Petrogenesis and lava flow emplacement based on LiDAR images

https://doi.org/10.1016/j.jvolgeores.2020.107169Get rights and content

Highlights

  • Rancho Seco volcano lava flows derived from a single batch of magma that was gradually extruded.

  • Lavas viscosities of Rancho Seco volcano ranged between 105 and 109 Pa·s.

  • We suggest an eruption duration of 2 to 6 years with effusion rates of 4 to 15 m3/s.

  • Our results are relevant for future eruptive scenarios and risk assessment.

Abstract

Given the high eruption recurrence in the Michoacán-Guanajuato volcanic field (MGVF) in central Mexico, the birth of a new monogenetic volcano can be expected in the future. It is important, therefore, to reconstruct the past eruptions of its many different volcanoes, including estimates of lava flow emplacement times and their rheological properties. These studies define the range of possible future eruptive scenarios and are necessary to evaluate potential risk. The Rancho Seco monogenetic volcano, located in the central part of the MGVF (19°37′03”N, 101°28′21”W), was radiocarbon-dated at ~27,845 years BP. Its eruption initiated with a violent-Strombolian phase that produced a scoria cone and was followed by the effusion of at least seven associated andesitic lava flows, reflecting drastic changes in the eruptive dynamics. Effusive activity probably involved decreases in the magma ascent and discharge rates linked to efficient degassing in an open system. Lava chemical composition suggest an origin of partial melting of a subduction-modified hydrated heterogenous mantle wedge and textural and mineralogical analysis indicates significant crystal fractionation and minor assimilation of granodioritic basement rocks. High-resolution LiDAR imagery was used to estimate lava flow viscosities and emplacement times by following a morphology-based methodology. Results indicate that lava flow viscosities ranged from 105 to 109 Pa·s and emplacement durations between 32 and 465 days for the flow units considered (F5 and F6). The entire eruption may have lasted from 2 to 6 years with a mean effusion rate of 4 to 15 m3/s. Our results are also pertinent to archaeologists studying the architectural remains of Angamuco, a large urban pre-Hispanic site built on Rancho Seco's lava flows.

Introduction

The Trans-Mexican Volcanic Belt (TMVB) is a continental volcanic arc containing ~8000 volcanic structures (Demant, 1978; Gómez-Tuena et al., 2005), and is tectonically related to the subduction of the oceanic Cocos and Rivera plates underneath the North American continental plate along the Mesoamerican trench (Nixon, 1982; Pardo and Suárez, 1995; Gómez-Tuena et al., 2005). It is important to highlight that since pre-Hispanic times this geologic province has grown in population and economic activity (Siebe and Macías, 2006; Ferrari et al., 2012). In the states of Guanajuato and Michoacán, the TMVB overlaps the older Oligocene to middle Miocene Sierra Madre Occidental (SMO) volcanic arc and a complex prevolcanic basement (Pasquaré et al., 1991).

The Michoacán-Guanajuato Volcanic Field (MGVF) is situated in the central part of the TMVB (Fig. 1) and comprises more than 1400 Quaternary monogenetic eruptive centres (including Rancho Seco volcano) distributed over an area of 40,000 km2 (Hasenaka and Carmichael, 1985). It has one of the greatest concentrations of monogenetic volcanoes within a subduction zone in the world (Hasenaka and Carmichael, 1985; Hasenaka, 1994). It is bounded to the north by the valley of the Lerma river and to the south by the depression of the Balsas river (Fig. 1); to the east it is contained by the Tzitzio anticline (Blatter and Hammersley, 2010), and to the west by the so-called “Mazamitla Volcanic Gap” (Kshirsagar et al., 2015). The MGVF contains mainly scoria cones and associated lavas, but also isolated thick lava flows, domes, maars, and tuff rings. In addition, there are approximately 400 volcanic edifices described as small or medium-sized shield volcanoes (Hasenaka and Carmichael, 1985; Hasenaka, 1994; Chevrel et al., 2016a) and only two stratovolcanoes (Tancítaro and Patamban), both considered extinct (Ownby et al., 2007; Siebe et al., 2014). Two historical eruptions have been witnessed in this field, Jorullo (1759–1774) and Paricutin (1943–1952), hinting that the future birth of a new volcano is highly likely (e.g. Guilbaud et al., 2011; Luhr and Simkin, 1993). Recent studies of areas within the MGVF (e.g. Guilbaud et al., 2011, Guilbaud et al., 2012, Guilbaud et al., 2019; Siebe et al., 2014; Kshirsagar et al., 2015, Kshirsagar et al., 2016; Chevrel et al., 2016a, Chevrel et al., 2016b; Larrea et al., 2017, Larrea et al., 2019; Reyes-Guzmán et al., 2018; Osorio-Ocampo et al., 2018; Pérez-Orozco et al., 2018; Ramírez-Uribe et al., 2019) have shown that eruptions have not only been diverse in style, but also frequent during Late Pleistocene and Holocene time.

The Rancho Seco monogenetic volcano is an andesitic scoria cone associated with lava flows situated at the eastern margin of the Pátzcuaro lake basin (Fig. 1) in the state of Michoacán and dated at 27,845 +445/−425 years BP (before present) (Ramírez-Uribe et al., 2019). The availability of LiDAR imagery provides the means to precisely map the lava flows, determine their volumes, highlight their main morphological features in great detail and measure their geometric parameters. We, then, apply morphology-based methods to estimate flow emplacement duration and effusion rates. In addition, chemical and petrological analyses of distinct lava units were used to constrain the magma source, storage conditions, and lava rheology, and to propose an eruptive model. These parameters are crucial for constraining future eruptive scenarios and evaluating potential volcanic risk.

Section snippets

Background

Lava flow hazards depend on several factors, namely the rheology of the lava and its effusion rate, the topography of the existing terrain, and the distance reached by the flow front before it entirely solidifies. The interpretation of remote observations of ancient flows can be helpful in understanding their emplacement dynamics. Flow geometry, structure, and surface morphology offer hints about eruption rates and lava rheology (Hulme, 1974; Griffiths, 2000). The evolution of a lava flow

GIS and high-resolution LiDAR topography: Map and lava volume estimates

The lava flow sequence map in Fig. 3 was developed using QGIS software and topographic chart information, digital elevation models with a resolution of 5 m from INEGI (2019), and Google Earth (2002–2018) satellite images. A 50-cm-high-resolution LiDAR image acquired for the “Legacies of Resilience”: The Lake Pátzcauro Basin Archaeological Project (LORE-LPB) (e.g., Fisher and Leisz, 2013; Fisher et al., 2017) was used to further improve a previously published map in Ramírez-Uribe et al. (2019).

Lava flow morphology

Despite being covered by an oak/pine forest, the lava morphology of Rancho Seco volcano is remarkably well-preserved. The high-resolution LiDAR image allowed for more accurate mapping and has established a new emplacement sequence (Fig. 3) relative to Ramírez-Uribe et al. (2019).

The Rancho Seco volcano is a monogenetic scoria cone with a maximum elevation of ~2520 m a.s.l. (above sea level) and a height of ~200 m above surrounding ground, and an almost circular base with a diameter of ~880 m.

Magma source and evolution

The andesitic rocks of Rancho Seco volcano originated probably from a primary basaltic melt (as postulated for other volcanoes in the MGVF, e.g. Rasoazanamparany et al., 2016; Corona-Chávez et al., 2006; Osorio-Ocampo et al., 2018; Ramírez-Uribe et al., 2019) that derived from the partial melting of a heterogeneous mantle wedge (e.g. Larrea et al., 2019) metasomatized by aqueous fluids (our studied rocks reveal marked enrichments in LILE with respect to HFSE, Fig. 11B) from the subducting

Archaeological aspects and future volcanic hazards

The Michoacán lake basins enabled early human settlement because they offered favourable aquatic and riparian conditions with abundant resources (fish, birds, amphibians, etc.), which promoted the development of agriculture in the region. The latter was of great relevance for the development of the Tarascan Empire, which flourished from ~AD 1250 until the conquest by the Spaniards in AD 1530 (Pollard, 2008; Carot, 2013). In Michoacán, several urban archaeological sites have been discovered on

Conclusions

This study provides new data for better understanding the effusive aspects of monogenetic volcanism in the MGVF. Andesite lava flows emitted from Rancho Seco volcano are relatively homogeneous in composition, indicating that probably one single batch of magma was gradually extruded. The lavas are the evolved products from a basaltic melt originally derived from the partial melting of the mantle wedge metasomatized by aqueous fluids from the subducting plate. Its subsequent evolution during

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgments

Field and laboratory costs were defrayed by project DGAPA-UNAM-IN103618 (Dirección General de Asuntos del Personal Académico, UNAM) and DGAPA-UNAM-IN104221 granted to C. Siebe. Israel Ramírez benefitted from a CONACYT graduate-student fellowship (2017-2019), while Oryaelle Chevrel's work in Mexico was financed by a DGAPA-UNAM postdoctoral fellowship. We thank Noemí Salazar and Giovanni Sosa for help with the electron-microprobe at Instituto de Geofísica, Unidad Michoacán, UNAM. The LiDAR image

References (110)

  • D. Giordano et al.

    Viscosity of magmatic liquids: a model

    Earth Planet. Sci. Lett.

    (2008)
  • A. Gómez-Tuena et al.

    Geochemical and petrological insights into the tectonic origin of the Transmexican Volcanic Belt

    Earth Sci. Rev.

    (2018)
  • M.G. Gómez-Vasconcelos et al.

    The Sierra de Mil Cumbres, Michoacán, México: Transitional volcanism between the Sierra Madre Occidental and the Trans-Mexican Volcanic Belt

    J. Volcanol. Geoth. Res.

    (2015)
  • M.N. Guilbaud et al.

    Geology, geochronology, and tectonic setting of the Jorullo Volcano region, Michoacán, México

    J. Volcanol. Geoth. Res.

    (2011)
  • A.J.L. Harris et al.

    Lava flows and rheology

  • T. Hasenaka

    Size, distribution, and magma output rate for shield volcanoes of the Michoacán-Guanajuato volcanic field, Central Mexico

    J. Volcanol. Geoth. Res.

    (1994)
  • C.J. Hawkesworth et al.

    A geochemical study of island-arc and back-arc tholeiites from the Scotia Sea

    Earth Planet. Sci. Lett.

    (1977)
  • P. Kshirsagar et al.

    Late Pleistocene Alberca de Guadalupe maar volcano (Zacapu basin, Michoacán): Stratigraphy, tectonic setting, and paleo-hydrogeological environment

    J. Volcanol. Geoth. Res.

    (2015)
  • P. Kshirsagar et al.

    Geological and environmental controls on the change of eruptive style (phreatomagmatic to Strombolian-effusive) of Late Pleistocene El Caracol tuff cone and its comparison with adjacent volcanoes around the Zacapu basin (Michoacán, México)

    J. Volcanol. Geotherm. Res.

    (2016)
  • P. Larrea et al.

    Compositional and volumetric development of a monogenetic lava flow field: the historical case of Paricutin (Michoacán, Mexico)

    J. Volcanol. Geoth. Res.

    (2017)
  • P. Larrea et al.

    A re-interpretation of the petrogenesis of Paricutin volcano: Distinguishing crustal contamination from mantle heterogeneity

    Chem. Geol.

    (2019)
  • W.F. McDonough et al.

    The composition of the Earth

    Chem. Geol.

    (1995)
  • S. Osorio-Ocampo et al.

    The eruptive history of the Pátzcuaro Lake area in the Michoacán Guanajuato Volcanic Field, Central México: Field mapping, C-14 and 40Ar/39Ar geochronology

    J. Volcanol. Geoth. Res.

    (2018)
  • S. Ownby et al.

    Volcán Tancítaro, Michoacán, Mexico, 40Ar/39Ar constraints on its history of sector collapse

    J. Volcanol. Geoth. Res.

    (2007)
  • J.D. Pérez-Orozco et al.

    Felsic-intermediate magmatism and brittle deformation in Sierra del Tzirate (Michoacán Guanajuato Volcanic Field)

    J. S. Am. Earth Sci.

    (2018)
  • H. Pinkerton

    Factor affecting the morphology of lava flows

    Endeavour

    (1987)
  • H. Pinkerton et al.

    The 1975 sub-terminal lavas, Mount Etna: a case history of the formation of a compound lava field

    J. Volcanol. Geoth. Res.

    (1976)
  • L. Pioli et al.

    Explosive dynamics of violent Strombolian eruptions: the eruption of Parícutin Volcano 1943–1952 (Mexico)

    Earth Planet. Sci. Lett.

    (2008)
  • L. Pioli et al.

    Controls on the explosivity of scoria cone eruptions: Magma segregation at conduit junctions

    J. Volcanol. Geoth. Res.

    (2009)
  • C. Rasoazanamparany et al.

    Temporal and compositional evolution of Jorullo volcano, Mexico: Implications for magmatic processes associated with a monogenetic eruption

    Chem. Geol.

    (2016)
  • C. Siebe

    Age and archaeological implications of Xitle volcano, southwestern basin of Mexico-City

    J. Volcanol. Geoth. Res.

    (2000)
  • C. Siebe et al.

    Geochemistry, Sr-Nd isotope composition, and tectonic setting of Holocene Pelado, Guespalapa, and Chichinautzin scoria cones, south of Mexico-City

    J. Volcanol. Geoth. Res.

    (2004)
  • P. Beattie

    Olivine-melt and orthopyroxene-melt equilibria

    Contrib. Mineral. Petrol.

    (1993)
  • D.L. Blatter et al.

    Plagioclase-free andesites from Zitacuaro (Michoacan), Mexico: petrology and experimental constraints

    Contrib. Mineral. Petrol.

    (1998)
  • P. Carot

    La larga historia purépecha

  • K.V. Cashman et al.

    How lava flows: New insights from applications of lidar technologies to lava flow studies

    Geosphere

    (2013)
  • A. Castruccio et al.

    Evolution of crust- and core-dominated lava flows using scaling analysis

    Bull. Volcanol.

    (2013)
  • A.F. Chase et al.

    Geospatial revolution and remote sensing LiDAR in Mesoamerican archaeology

    Proc. Nat. Acad. Sci.

    (2012)
  • M.O. Chevrel et al.

    The AD 1250 El Metate shield volcano (Michoacán): Mexico’s most voluminous Holocene eruption and its significance for archaeology and hazards

    The Holocene

    (2016)
  • M.O. Chevrel et al.

    The AD 1250 effusive eruption of El Metate shield volcano (Michoacán, Mexico): Magma source, crustal storage, eruptive dynamics, and lava rheology

    Bull. Volcanol.

    (2016)
  • M.O. Chevrel et al.

    Measuring the viscosity of lava in the field: a review

    Earth Sci. Rev.

    (2019)
  • C. Cimarelli et al.

    Rheology of magmas with bimodal crystal size and shape distributions: Insights from analog experiments

    Geochem. Geophys. Geosyst.

    (2011)
  • P. Corona-Chávez et al.

    Asimilación de xenolitos graníticos en el Campo Volcánico Michoacán-Guanajuato, el caso de Arócutin Michoacán

    México. Rev. Mex. Cienc. Geol.

    (2006)
  • A. Costa et al.

    A model for the rheology of particle-bearing suspensions and partially molten rocks

    Geochem. Geophys. Geosyst.

    (2009)
  • S. Couch et al.

    Experimental constraints on the conditions of formation of highly calcic plagioclase microlites at the Soufrière Hills volcano

    Montserrat. J. Petrol.

    (2003)
  • S.M. Crabtree et al.

    Complex phenocryst textures and zoning patterns in andesites and dacites: evidence of degassing-induced rapid crystallization? J

    Petrol.

    (2011)
  • J. Crisp et al.

    Crystallization history of the 1984 Mauna Loa lava flow

    J. Geophys. Res.

    (1994)
  • N.D. Deardorff et al.

    Emplacement conditions of the c. 1,600-year bp Collier Cone lava flow, Oregon: a LiDAR investigation

    Bull. Volcanol.

    (2012)
  • N.I. Deligne et al.

    Holocene volcanism of the upper McKenzie River catchment, Central Oregon Cascades, USA

    Geol. Soc. Am. Bull.

    (2016)
  • A. Demant

    Características del Eje Neovolcánico Transmexicano y sus problemas de interpretación

    Rev. Mex. Cienc. Geol.

    (1978)
  • Cited by (10)

    • Reassessing the paleointensities of three Quaternary volcanic structures of the -Michoacán-Guanajuato Volcanic Field (Mexico) through a multimethodological analysis

      2022, Physics of the Earth and Planetary Interiors
      Citation Excerpt :

      The available data indicates that the site was occupied from the Classic period (∼300 CE) to the Spanish conquest (1530 CE) (Cohen, 2016), hence associated with the Tarascan empire (which dominated the western region of Mexico from ∼1250 CE to the arrival of the Spaniards) and with their ancestors. Thanks to LiDAR images analysis, the seven lava flows emplaced during the effusive phase of the eruption, which was preceded by the violent Strombolian activity that formed the cinder cone (Fig. 1b) were mapped; it has been estimated that the entire eruption lasted between 2 and 6 years (Ramírez-Uribe et al., 2021). The studied volcano that is intermediate in age is part of the El Jabalí cone cluster (19°27′00“N, 102° 06’08”W).

    View all citing articles on Scopus
    View full text