You are on page 1of 47

Accepted Manuscript

Timing the evolution of a monogenetic volcanic field: Sierra


Chichinautzin, Central Mexico

M.C. Jaimes-Viera, A.L. Martin Del Pozzo, P.W. Layer, J.A.


Benowitz, A. Nieto-Torres

PII: S0377-0273(17)30681-9
DOI: doi:10.1016/j.jvolgeores.2018.03.013
Reference: VOLGEO 6331
To appear in: Journal of Volcanology and Geothermal Research
Received date: 14 November 2017
Revised date: 1 March 2018
Accepted date: 16 March 2018

Please cite this article as: M.C. Jaimes-Viera, A.L. Martin Del Pozzo, P.W. Layer, J.A.
Benowitz, A. Nieto-Torres , Timing the evolution of a monogenetic volcanic field: Sierra
Chichinautzin, Central Mexico. The address for the corresponding author was captured
as affiliation for all authors. Please check if appropriate. Volgeo(2017), doi:10.1016/
j.jvolgeores.2018.03.013

This is a PDF file of an unedited manuscript that has been accepted for publication. As
a service to our customers we are providing this early version of the manuscript. The
manuscript will undergo copyediting, typesetting, and review of the resulting proof before
it is published in its final form. Please note that during the production process errors may
be discovered which could affect the content, and all legal disclaimers that apply to the
journal pertain.
ACCEPTED MANUSCRIPT

Timing the evolution of a monogenetic volcanic field: Sierra Chichinautzin, Central


Mexico.
1 2 3 3 1
Jaimes-Viera, M.C. , Martin Del Pozzo, A.L. , Layer, P.W. , Benowitz, J.A. , Nieto-Torres, A.

1
Posgrado en Ciencias de la Tierra, Universidad Nacional Autónoma de México, Cd. Universitaria
Coyoacán, CDMX, 04510, México.
2
Departamento de Vulcanología, Instituto de Geofísica, Universidad Nacional Autónoma de México,
Cd. Universitaria, Coyoacán, CDMX, 04510, México
3
Department of Geology and Geophysics, Geophysical Institute, University of Alaska, Fairbanks,
AK, USA

T
Contact information for the corresponding author

IP
Jaimes-Viera, Maria del Carmen.
Posgrado en Ciencias de la Tierra, Universidad Nacional Autónoma de México, Cd. Universitaria

CR
Coyoacán, CDMX, 04510, México.
e-mail: cjaimes@geofisica.unam.mx

US
AN
M
ED
PT
CE
AC
ACCEPTED MANUSCRIPT

Timing the evolution of a monogenetic volcanic field: Sierra Chichinautzin, Central Mexico.
1 2 3 3 1
Jaimes-Viera, M.C. , Martin Del Pozzo, A.L. , Layer, P.W. , Benowitz, J.A. , Nieto-Torres, A.

Abstract.
The unique nature of monogenetic volcanism has always raised questions about its origin, longevity
and spatial distribution. Detailed temporal and spatial boundaries resulted from a morphometric
study, mapping, relative dating, twenty-four new 40Ar/39Ar dates, and chemical analyses for the
Sierra Chichinautzin, Central Mexico. Based on these results the monogenetic cones were divided
into four groups: (1) Peñón Monogenetic Volcanic Group (PMVG); (2) Older Chichinautzin
Monogenetic Volcanic Group (Older CMVG); (3) Younger Chichinautzin Monogenetic Volcanic

T
Group (Younger CMVG) and (4) Sierra Santa Catarina Monogenetic Volcanic Group (SSC).

IP
The PMVG cover the largest area and marks the northern and southern boundaries of this field.
The oldest monogenetic volcanism (PMVG; 1294±36 to 765±30 ka) started in the northern part of

CR
the area and the last eruption of this group occurred in the south. These basaltic-andesite cones
are widely spaced and are aligned NE-SW (N60°E). After this activity, monogenetic volcanism
stopped for ~527 ka.

US
Monogenetic volcanism was reactivated with the birth of the Tezoyuca 1 Volcano, marking the
beginning of the second volcanic group (Older CMVG; 238±51 to 95±12 ka) in the southern part of
the area. These andesitic to basaltic andesite cones plot into two groups, one with high MgO and
Nb, and the other with low MgO and Nb, suggesting diverse magma sources. The eruption of the
AN
Older CMVG ended with the eruption of Malacatepec volcano and then monogenetic volcanism
stopped again for ~60 ka.

At ~35 ka, monogenetic volcanism started again, this time in the eastern part of the area, close to
M

Popocatépetl volcano, forming the Younger CMVG (< 35±4 ka). These cones are aligned in an E-W
direction. Geochemical composition of eruptive products of measured samples varies from basalts
to dacites with low and high MgO. The Younger CMVG is considered still active since the last
ED

eruptions took place less than 2 ka.

The SSC (132±70 to 2±56 ka) is located in the northern part of the area, in the old Chalco Lake and
is separated by faults from the rest of the volcanic groups as a different range. The SSC formed
PT

closely spaced basaltic andesites to andesitic cones oriented NE-SW (N70°E). The SSC samples
have high Zr, P2O5, and Nb, indicating a different magma source.
CE

The northern and southern spatial boundaries of the field (the surface area with monogenetic
volcanoes) became smaller with time: 78 km for PMVG, ~40 km for the Older CMVG and ~25 km
for the Younger CMVG, concentrating the volcanoes in the central part of the area. The alignment
of the cones changed progressively from NNE-SSW to NE-SW to E-W through the time, associated
AC

with the changes in the stress field which appears also to have caused the gaps.

Results suggest that the Sierra Chichinautzin is actually four different volcanic fields, some partially
overlapping, instead of one as previously considered. The differences in age, emplacement
orientation and geochemistry support this conclusion.

40 39
Keywords: Monogenetic volcanic fields. Ar/ Ar geochronology. Sierra Chichinautzin. Volcanic
geomorphology.

1. Introduction.
The origin and length of time a monogenetic volcanic field lasts is still the subject of uncertainty.
Identifying how and when such volcanism occurs, and establishing the spatial limits of the volcanic
ACCEPTED MANUSCRIPT

field through time and space, is necessary to understand the behaviour of monogenetic volcanism.
Investigating the physical–chemical characteristics of the magma as well as the external
parameters, such as groundwater, type of substrate or topography that influences the cone
architecture are also essential (Nemeth and Kereszturi, 2015). This paper deals with the evolution
of monogenetic fields and we provide new information on how the initial volcanic activity occurred
and how it has varied over time.

Monogenetic volcanism is short-lived and is formed during one eruptive period of time. It refers to

T
small-scale magmatic systems that form fields of volcanoes, which are produced by explosive and

IP
effusive eruptions and are triggered by the rise of small batches of magma (Cañón-Tapia and
Walker 2004; Smith and Németh, 2017).

CR
Monogenetic fields are found in different tectonic settings: intraplate volcanism such as the

US
Auckland field in New Zealand (Németh and Kereszturi, 2015; Le Corvec et al., 2013; Lindsay et al.,
2011) and the Eiffel field in Germany (Schmincke, 2007) and in volcanic arcs such as the San
Francisco field in the USA (Tanaka et al., 1986) and the Carrán-Los Venados field in Chile (Bucchi
AN
et al., 2015), just to mention a few (Table 1). Although these fields are generally linked to regional
extensional stress, short-lived and characterized by small amounts of magma, between 0.0001 -
3
0.10 km (Connor and Conway 2000; Németh 2010; Le Corvec et al., 2013; Németh and Kereszturi,
M

2015), they are also associated with compressional stress fields (van de Hove, et al., 2017) and the
Anatolian Volcanic Province also with transtensional systems (Pasquaré et al., 1988). Geostatistical
ED

analyses, Nearest Neighbor and Kernell models for example, have been used at the Abu, Japan;
Michoacán-Guanajuato, Mexico; Yucca Mountains, USA; Auckland, New Zealand and Tharsis,
PT

Mars monogenetic fields to establish the probability of future eruptions (Connor et al., 2000;
Alberico et al., 2002, Kiyosugi, 2010; Mazzarini et al., 2010; Bebbington, 2013; Richardson et al.,
CE

2013; Pozzobon et al., 2014). The cones appear to be tectonically controlled and may occur in
clusters, indicating that magma generation processes are similar to those controlling their
propagation through the lithosphere (Le Corvec et al., 2013). A strong correlation between fractures
AC

and cone alignment has been identified in several active volcanic fields (Mazzarini et al., 2010).

The length of time a monogenetic field lasts has to be defined by detailed dating considering that
there is no a clear evidence of a temporal break (Smith and Németh, 2017). Individual monogenetic
volcanoes can last weeks, months, years or decades, as Paricutín volcano (1943-1952) in the
Michoacán-Guanajuato Volcanic Field, Mexico, but the duration of a complete monogenetic field is
still not clearly known. There are several monogenetic fields that last up to millions of years like in
the Bakony-Balaton Highland volcanic field, Hungary (2.6-7.9 Ma; Wijbrans et al., 2007) or the
Newer Volcanic Province, in Australia (0.005-4.6 Ma; Matchan et al., 2001; Gray and McDougall,
2009), but periods of repose inside the field or how long each lasts are little known (Table 1).
ACCEPTED MANUSCRIPT

Table 1. Number of volcanoes, tectonic environment and ages for monogenetic volcanic fields.

# Tectonic Age Area


Volcanic field Country Age min Method 2 References
volcanoes enviroment max (km )

T
Ar-Ar, 0.05- Luhr et al., 1995b; Williams, 1999; Aranda-
San Quintín volcanic field Mexico 10 intraplate 22 ka 165 ka 3 4
He- He 50 Gómez et al., 2005.

I P Kermode 1992; Allen and Smith, 1994;


Edbrooke et al., 2003; Cassidy and Locke,

Auckland volcanic field


New
Zealand
49 intraplate 10 ka 0.25 Ma
K-Ar,

C
Ar-Ar,
14
C
R336
2004; Magill et al., 2005; Lindsay and
Leonard 2007; Molloy et al., 2009; Kereszturi
et al., 2010; Németh et al., 2010; Lindsay et

U S al., 2011; Le corvec et al., 2013; Németh


and Kereszturi, 2015.

Bakony-Balaton Highland
volcanic field
Hungary 50 intraplate 2.61 Ma

A N
7.96 Ma K-Ar 3500
Balogh and Németh, 2005; Wijbrans et al.,
2007; Kerezsturi et al., 2011.

La Garrotxa volcanic field Spain 54 intraplate 10


M 600 ka K-Ar 600
Martí and Mallarach, 1987; Martí et al.,
2011; Bolós et al., 2014;

Abu volcanic group Japan 56 intraplate


E D10 ka 3.3 Ma K.Ar 400
Uto and Koyaguchi, 1987; Kakubuchi et al.,
2000; Kiyosugi et al., 2010.

Carrán-Los Venados Chile 65


P T
subduction
1979 last
eruption
13.9 ka 14C 160 Bucchi et al., 2015

Lamongan volcanic field Java 91

C E intraplate
1898 last
eruption
40 ka 260 Carn, 2000; Carn and Pyle, 2001.

Valle de Bravo volcanic field

Eiffel volcanic field (East)


Mexico

A C 120

57
subduction 5 ka 40 ka

0.6-0.7
Ar-Ar

K-Ar,
3703

400
Blatter et al., 2001; Aguirre-Díaz et al., 2006.

Schmincke and Mertes, 1979; Schmincke et


Germany intraplate 0.01 Ma Ar-Ar,
Ma al., 1983; Büchel, 1993; Schmincke, 2007.
Eiffel volcanic field (West) 170 14C 600
ACCEPTED MANUSCRIPT

Brenna et al., 2011; Brenna et al., 2012;


South Ar-Ar,
Jeju Island 250 intraplate 1 ka 1.8 Ma 14 2000 Brenna et al., 2015; Cheong et al., 2007;
Korea C
Sohn et al., 2008, 2012.
White, 1991; Lefebvre, 2013; Re et al.,
Hopi Buttes volcanic field USA 300 intraplate 5 Ma 7 Ma Ar-Ar 2000

P T
2015; Muirhead et al., 2016.
McDougall et al., 1966; Wellman and
Newer Volcanics Province Australia 400 subduction 0.005 Ma 4.6 Ma K-Ar
I
25000

R
McDougall, 1974; Price et al., 1997; Hare
and Cas, 2005.

San Francisco USA

Kingdom
360 subduction
1064 A.D.
(.25 Ma)
5 Ma
C
K-Ar

S
4800 Tanaka et al., 1986

Camp and Roobol 1989; Coleman and


Harrat Rahat volcanic field of Saudi
Arabia
968 intraplate 0.05 Ma 10 Ma

N U Ar-Ar 20000 Gregory 1983; El Difrawy et al. 2013; Moufti


et al. 2013; Murcia et al., 2014.

Michoacán-Guanajuato
Mexico 1000 subduction
Paricutín
(1942-
A
5 Ma K-Ar,
14
C 40000
Hasenaka and Carmichael, 1985, 1987;
Hooper and Sheridan, 1998; Mazzarini et al.,

M
volcanic field
1953) 2010.
Karapınar Central 0.16 Ma 0.36 Ma K–Ar
Eğrikuyu
Keçiboyduran-Melendiz
Anatolian
Volcanic
819
possibly
transtentional
E D
0.06 Ma
0.22 Ma
0.55 Ma
0.65 Ma
Ar–Ar
Ar–Ar 20000
Innocenti et al., 1975; Pasquaré et al., 1988;
Ercan et al., 1992; Notsu et al., 1995;
Gençalioğlu-Kuşcu, 2011; O'Sadnick et al.,

T
Province subduction
Derinkuyu-Acıgöl 0.89 Ma 1.08 Ma U-Pb
(CAVP), related 2013; Aydin et al., 2014.
Erciyes Turkey

E P 0.8 Ma 2.6 Ma K-Ar

C C
A
ACCEPTED MANUSCRIPT

The size of monogenetic volcanic fields varies, some are small such San Quintin volcanic field with
10 volcanoes, while others are very large with up to 1000 cones, such as in the Michoacán-
Guanajuato volcanic field (Table 1).

The morphology and type of cones found in a monogenetic field are spatter cones, scoria or cinder
cones, tuff and tephra rings, maar craters, tuff cones and a few small shield volcanoes. Internal
factors (magma composition, magmatic flux, ascent rate, viscosity, volatile contents) and external

T
factors (regional and local tectonics, topography, presence of water) define the specific morphology

IP
of the volcanic structures, their distribution pattern, and the volcanic facies (Kereszturi et al., 2011).
The type, shape and size of the pyroclasts that built up the cones are used to understand their

CR
eruptive behaviour, and can have different shapes even in the same volcano (Tsuya 1939, 1941).
The type and shape of the clasts is dependant of several factors, like the rush of air acting on a fluid

US
clot during flight, the forces acting at the moment of ejection of the magma in the conduit, and the
amount of gas expansion (Walker, 1969; Cañón-Tapia, 2017). Coarse lapilli, blocks and bombs with
spindle and ribbon shapes are more frequent in the Strombolian activity, as well as lenticular and
AN
reverse grading. In violent Strombolian activity, lapilli and ash with angular to slightly rounded clasts
and massive to planar stratification, are dominant. The stratigraphy of the deposits records the
compositional evolution and periods of magma ascent during the monogenetic activity; and the
M

textural changes are related to the eruptive dynamics (Erlund et al., 2010).
ED

In this paper, we provide relative ages established by means of cone morphometry and 24 new
40 39
Ar/ Ar dates as well as their eruptive style emphasising the spatial and temporal distribution of
PT

the volcanoes in the Sierra Chichinautzin (SC), Central Mexico ( Fig. 1). We also defined the
deposits of the representative individual volcanoes that make up the field in order to better
CE

understand the type of eruptive activity.

2. Geologic setting.
AC

The Sierra Chichinautzin (SC) is situated 280 km from the Mesoamerican Trench, in central
Mexico, and is part of the Trans Mexican Volcanic Belt (TMVB), which is made up many
monogenetic fields separated by large polygenetic volcanoes such as Popocatépetl (Fig. 1a). The
TMVB is the result of the subduction of the Cocos and Rivera plates beneath the North American
plate at the Middle American Trench since the Miocene (Fig. 1b; Nixon, 1982; Johnson and
Harrison, 1990; Pardo and Suarez, 1995; Pérez-Campos, et al., 2008; Ferrari et al., 2012).

The TMVB has a complex tectonic history. The volcanic arc is not exactly parallel to the trench and
this is because the geometry of the subducting Cocos Plate. Geophysical studies of the TMVB
ACCEPTED MANUSCRIPT

show that plate is subhorizontal and as it moves in it subducts almost vertically, reaching at a depth
of about 150 km under the Mexico City (Pérez-Campos et al., 2008). The Cocos plate has an
oblique subduction with respect to the Trench, subducting towards the N-NE (Pardo and Suárez,
1995, Ferrari et al., 1999).

The study area has more than 227 volcanic structures, including the volcanoes located to the
Southeast of Mexico City, which are Chimalhuacán, Peñón Del Marqués, Cerro de la Estrella and
the Sierra Santa Catarina. The area is limited to the west by the Sierra de Las Cruces Tertiary

T
volcanism, to the east by active Popocatépetl volcano, to the north by lake sediments and to the

IP
south by Cretaceous limestones and volcanic debris avalanches from older polygenetic volcanoes
(Fig. 2). Part of Mexico City was built on the monogenic field that is considered active because the

CR
recent eruptions were witnessed by the inhabitants (Kirianov et al., 1990, Córdoba et al., 1994;
Martin del Pozzo et al., 1997; Siebe, 2000) and therefore it still poses considerable risk to the area

US
AN
M
ED
PT
CE
AC
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT
CE
AC

Figure 1. (a) Map of the Sierra Chichinautzin monogenetic volcanoes (SC) on a Landsat 8 Thematic Mapper
(RGB 4-3-2 composite image). The white rectangle outlines the study area. CHI: Chimalhuacán, JT:
Jumiltepec; TCM: Tecmilco, PB: Peñón Del Marqués, EST: La Estrella, TLT: Tlaltizapán, TLY: Tlayacapan,
CTZ: Coatzontle, ZTL: Zitlaltépetl, ACH: Achichipilco, MT: Malacatepec, MZ: Mezontepec, AH: Aholo, CUA:
Cuautepel, QP: Quepil, TB: Tabaquillo, TZ: Tezoyuca 1 and 2, CO: La Corona, TT: Las Tetillas, 3C: Tres
Cumbres, SSC: Sierra Santa Catarina. (b) Tectonic map of the Trans Mexican Volcanic Belt (TMVB) with
location of Chichinautzin Monogenetic Volcanic Field (CMVF). Black lines: faults, hyphenated lines: inferred
faults (modified from Campos-Enríquez et al., 1997; 2015).
2
The SC has an approximate area of ~5,945 km that was previously thought to be less than 40 ka
(Martin Del Pozzo, 1982). These volcanoes are basaltic to dacitic and had Strombolian, violent
Strombolian and effusive activity (Martin Del Pozzo, 1982). The rocks of the SC show normal
ACCEPTED MANUSCRIPT

polarity, belonging to the Bruhnes Chron (Böhnel and Molina-Garza, 2002; Urrutia-Fucugauchi and
Martin Del Pozzo, 1993), which places them it at less than 777.76 ka (Hyodo and Kitaba,
14
2015).Volcanoes dated by C range from 38592 ± 3210 to 1675 ± 35 years BP but little has been
studied about how the activity began (Bloomfield, 1975; Kirianov et al., 1990, Córdoba et al., 1994;
Martin del Pozzo et al., 1997; Siebe, 2000).

The Sierra Santa Catarina (SSC) is a volcanic chain made up of seven volcanoes aligned ENE.
They are pyroclastic cones with associated lava flows that were emplaced in the northern part of the

T
study area, in what was the Chalco Lake (Fig. 1). Lugo-Hubp et al. (1994) proposed that volcanism

IP
migrated from W to E, based on geomorphological criteria. Rocks from SSC have normal and
reverse polarity, suggesting that either volcanism occurred in the Brunhes and Matuyama Chrons,

CR
or in transitional periods during magnetic excursions (Urrutia-Fucugauchi and Martin Del Pozzo,
1993). The Cerro de la Estrella has normal polarity, but has an intermediate inclination and for this

US
reason, was associated to the Matuyama Chron (Urrutia-Fucugauchi, 1995). In the case of
Chimalhuacán and Peñón Del Marqués, they have reverse polarity (Mooser et al., 1974).
AN
In the SC many cones are aligned E–W that has been associated with regional faults (Mooser et al.
1974; Bloomfield 1975; Martin Del Pozzo, 1982; Alaniz-Alvarez et al. 1998; Siebe et al. 2004). In
the study area there are north dipping normal faults toward the Chalco-Xochimilco Lake, and gravity
M

inferred shallow E-W trending faults in graben and half graben structures in the Chalco Lake
(Campos-Enríquez et al., 1997, 2000).
ED

3. Methods
PT

3.1 Morphometry of the volcanoes.


Reconnaissance mapping and image processing were carried out along with morphometric
CE

analyses for relative dating and selecting cones. Representative volcanoes were chosen according
to relative age emphasizing the older cones, eruptive styles, cone morphologies as well as their
spatial distribution in order to study them in further detail. Twenty-two cones, a thick lava flow, and
AC

two domes were chosen for mapping and sampled for dating and major and trace element
chemistry.

The morphometric parameters height (h), basal diameter (W b) and crater diameter (W cr) of the
cones were measured on Landsat 8 Thematic Mapper composite images (RGB 4-3-2), vertical
aerial photographs 1:75,000, topographic maps (INEGI 1:50,000) and on a 5 m Digital Elevation
Model (DEM). The h/W b ratio, and the slope were calculated with the formulae: =tan (2h)/(W b-
-1

2 2
W cr). The volume of the cones was calculated according to the formulae: V=1/3**h(R +r +R*r).
ACCEPTED MANUSCRIPT

The h/W b ratio and slope () were used to determine the degree of cone erosion and how much the
slope has decreased over time. Both these parameters are considered useful in assessing the
relative ages of the volcanic cones (Wood, 1980; Hooper, 1995; Inbar et al., 2011). The cones were
reclassified according to relative age, also taking into account the preservation of lava flows
features such as flow margins, thickness of soil cover and pressure ridges using the morphological
criteria of Martin Del Pozzo (1982). Also, we made a geostatistical analysis with the Spline method,
which interpolates a raster surface from points using a two-dimensional minimum curvature spline
technique. The resulting smooth surface was useful to determine the distribution of the cones in the

T
study area.

IP
Stratigraphic sections were measured and sampled to define the type of deposits that formed the

CR
cones, highlighting their differences, as well as the morphological characteristics of the cones, lava
flows and domes, and a geologic map was prepared.

US
40 39
3.2 Ar/ Ar geochronology.
Twenty-four samples were analysed at the Geochronology laboratory of the University of Alaska,
AN
Fairbanks (UAF). Samples were washed and hand-picked for phenocryst-free whole rock ground
mass chips. The monitor mineral TCR‐2 with an age of 28.619 Ma (Renne et al., 1994, 2010) was
M

used to monitor neutron flux and calculate the irradiation parameter (J) for all samples. The samples
and standards were wrapped in aluminium foil and loaded into aluminium cans of 2.5 cm diameter
ED

and 6 cm height. The samples were irradiated in position 8c of the uranium enriched research
reactor of McMaster University in Hamilton, Ontario, Canada for 20 megawatt-hours.
PT

Upon their return from the reactor, the samples and monitors were loaded into 2 mm diameter holes
in a copper tray that was then loaded in a ultra-high vacuum extraction line. The monitors were
CE

fused, and samples heated, using a 6-watt argon-ion laser following the technique described in
York et al. (1981), Layer et al. (1987) and Benowitz (2014). Argon purification was achieved using a
liquid nitrogen cold trap and a SAES Zr-Al getter at 400C. The samples were analysed in a VG-
AC

3600 mass spectrometer controlled by a Visual Basic operating program written in-house at the
Geophysical Institute, University of Alaska Fairbanks. The argon isotopes measured were corrected
for system blank and mass discrimination, as well as calcium, potassium and chlorine interference
reactions following procedures outlined in McDougall and Harrison (1999). Typical full-system 8 min
-16 40 39 -18
laser blank values (in moles) were generally 2 × 10 mol Ar, 3 × 10218 mol Ar, 9 × 10 mol
38 -18 36
Ar and 2 × 10 mol Ar, which are 10–50 times smaller than the sample/standard volume
fractions. Correction factors for nucleogenic interferences during irradiation were determined from
39 37 -4 36 37 -4
irradiated CaF2 and K2SO4 as follows: ( Ar/ Ar) Ca = 7.06×10 , ( Ar/ Ar) Ca = 2.79 × 10 and
40 39
( Ar/ Ar)K = 0.0297. Mass discrimination was monitored by running calibrated air shots and a
ACCEPTED MANUSCRIPT

zero-age glass sample. The mass discrimination during these experiments was 0.8% per mass unit.
While doing our experiments, calibration measurements were made on a weekly–monthly basis to
check for changes in mass discrimination with no significant variation seen during these intervals. A
sample is considered to have a plateau if it has 3 or more contiguous fractions constituting at least
39
50% Ar release and is significant at the 95% confidence level (as indicated by an Mean Square
Weighted Deviates; MSWD <~2.5). A sample is considered to form an isochron if it has 3 or more
contiguous fractions that form a linear array that is significant at the 95% confidence level (MSWD <
~2.5). All ages quoted to the ± 1 sigma level and calculated using the constants of Renne et al.

T
(2010). The integrated age is the age given by the total gas measured and is equivalent to a

IP
potassium-argon (K-Ar) age. The spectrum provides a plateau age if three or more consecutive gas
fractions represent at least 50% of the total gas release and are within two standard deviations of

CR
each other (MSWD < 2.5).

US
3.3 Geochemistry.
Chemical analyses were carried out at the X-Ray Fluorescence Laboratory of the Instituto de
Geología, at the Universidad Nacional Autónoma de México, Mexico City. Twenty five rock samples
AN
were crushed and pulverized into powders of 200 and were analyzed on a sequential X-ray
spectrometer (Rigaku Primus II) equipped with a rhodium tube and beryllium window of 30  for the
M

determination of major elements (Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K, P) and trace elements (Ba, Zr,
Sr, Nb and Rb). The estimated errors are <1% in major elements and <4% in trace elements.
ED

4. Results.
PT

4.1 Morphometric parameters and relative ages.


The morphometric data obtained indicate that the selected scoria and cinder cones have heights (h)
CE

between 20 and 390 m and basal diameters (W b) from 180 m to 2800 m. The volume of the cones
3 3
ranges from 0.0014 km (La Corona) to 0.24 km (Jumiltepec) and their lava flows are 1 to 11.5 km
long with thicknesses of 1 to 9 m. The h/W b ratio for the cones varies from 0.06 to 0.28 and the
AC

slope from 8º to 43º. The domes have heights of 90 to 170 m and basal diameters of 550 to 575 m.
The thick lava flow lacking a cone Tabaquillo, is 300 m thick and 6 km long (Table 2).

The cones were divided into four groups: three of them according to the progressive change in
morphometric parameters with the time, by comparative measurements that reflect the relative age
of the cones. The h/W b ranged from 0.06 to 0.10 with slope angles from 7.98º to 17.45º for the older
group of relative age; 0.11 to 0.14 with slopes of 13.87º to 21.80º for the intermediate group, and
0.146 to 0.28 with slopes from 21.41º to 34.17º for the younger group. The fourth group was
separated according to its spatial position and corresponds to the SSC. The SSC is considered a
ACCEPTED MANUSCRIPT

separate group since it is offset from SC by lake sediments and normal faulting. Their h/W b ratio
ranges from 0.1364 to .2338 and the slope from 20º to 43º.The (Table 2).

Table 2. Geomorphological parameters of the volcanoes.


Lava flow (*average)
Average V cone Length Thickness
Volcano h (m) Wb (m) Wcr (m) h/Wb Slope ()
() (km3) (*km) (*m)

Tlaltizapán 80 400 100 0.2000 28.07 0.0044 1.50 2.0


Tecmilco 50 280 55 0.1786 23.96 0.0013 5.20 1.5
Younger group

Tabaquillo 300 ----- ----- ----- 6.00 -----

T
Mezontepec 200 840 75 0.2381 27.60 0.0429 4.20 2.7
28.13°

IP
Aholo 140 500 87.5 0.2800 34.17 0.0110 1.70 3.0
Cuautepel 200 860 120 0.2326 28.39 0.0449 10.00 3.0
Achichipilco 115 550 200 0.2091 33.31 0.0136 11.50 2.7

CR
Zitlaltepel 98 650 150 0.1508 21.41 0.0139 4.10 1.0
Malacatepec 125 950 200 0.1316 18.43 0.0370 2.50 1.5
Intermediate

Tlayacapan 100 900 90 0.1111 13.87 0.0235 6.40 2.8


group

Tezoyuca 2 20 180 80 0.1111 21.80 0.0062 1.25 4.0

US
18.57°
Quepil 90 575 ----- ----- ----- -----
Coatzontle 60 500 190 0.1200 21.16 0.0073 1.20 2.0
Tezoyuca 1 65 500 90 0.1300 17.59 0.0052 2.10 7.0
C. La Estrella 120 1300 150 0.0923 11.79 0.0599 1.50 5.0
AN
La Corona 30 500 72 0.0600 7.98 0.0023 2.10 3.0
Older group

Jumiltepec 218 1800 450 0.1211 17.90 0.2406 7.80 3.7


Las Tetillas 150 1700 200 0.0882 11.31 13.07° 0.1283 4.70 5.0
M

Chimalhuacan 170 1700 100 0.1000 12.00 0.1366 1.90 2.0


Peñón de
77 860 370 0.0895 17.45 0.0241 1.10 1.7
Baños
Mazatepec 180 770 380 0.2338 42.71 0.0485 2.30 3.0
ED

Tecuautzi 170 550 ----- ----- ----- -----


SSC

Yuhualixqui 110 670 100 0.1642 21.10 26.23° 0.0151 1.50 1.0
Tetecón 90 650 170 0.1385 20.56 0.0132 1.90 2.0
PT

Xaltepec 150 1100 300 0.1364 20.56 0.0640 2.40 2.8


CE

4.2 Radiometric dating.


40 39
The Ar/ Ar dates obtained in this study ranged from ~1294 ka to ~7 ka, and correlate well with
the relative age groups (Table 2 and 3).
AC

The age distribution shows two long periods of magmatic quiescence lasting several thousands of
years. In figure 2a one can see that the beginning of the earliest monogenetic activity occurred at
1294 ± 36 ka and ended at 765 ± 30 ka. We have named the group of volcanoes formed during this
period as the Peñón Monogenetic Volcanic Group (PMVG) that is the oldest group of relative age.
After the last eruption of the PMVG there was a gap in the monogenetic volcanism that lasted ~527
ka.
ACCEPTED MANUSCRIPT

Monogenetic activity began again at 238 ± 51 ka forming the Older Chichinautzin Monogenetic
Volcanic Group (Older CMVG, the intermediate group of relative age), with the last eruption dated
at 95 ± 12 ka (Fig. 2b).

40 39 14
Table 3. Ar/ Ar and C dates for the Chichinautzin monogenetic volcanic field.
Relative
age Volcano Date Method Position / type of sample Author

T
14
1,675 ± 40 yr B.P. C Charcoal in ash Siebe 2000

IP
Xitle 14 Charcoal in archeological
1975 ± 60 yr B.P. C Córdoba, et. al., 1994
material under lava
14
Chichinautzin 1835 ± 55 yr B.P. C Charcoal in ash Siebe et. al. 2004

CR
40 39
PL-1 Tecuautzi 2 ± 56 ka Ar/ Ar Whole rock Layer et al., 2009
14
Jumento 2010 ± 30 yr B.P. C Basal wet surge deposit Arce et al., 2015
14 Reworked charcoal against

US
Pelagatos >2520 ± 105 yr B.P. C Guilbaud et al., 2009
lava flows
14
Gespalapa >2835 ± 75 yr B.P. C Paleosol on scoria Siebe et al., 2004
14
Pelado 4070± 150 yr B.P. C Paleosoil betwen lava blocks Kirianov, et. al. 1990
AN
14
Tláloc >6200 ± 85 yr B.P. C Palcosoil under ash Siebe et. al. 2005
14
P3C1 >6700 ± 150 yr B.P. C Paleosoil on cone Kirianov, et. al. 1990
40 39
Tabaquillo 7±9 ka Ar/ Ar Whole rock This paper
M

14
7360 ± 120 yr B.P. C Block and ash flow deposit Siebe et. al. 2005
Younger group

Cuauhtzin 14 Lahar underneath block and


8225 ± 130 yr B.P. C Siebe et. al. 2005
ash flow deposit
ED

14
>1060 ± 70 yr B.P. C Paleosoil on cone Kirianov, et. al. 1990
La Cima 14
>10410 ± 80 yr B.P. C Paleosoil on cone Kirianov, et. al. 1990
40 39
Mezontepec 11 ± 3 ka Ar/ Ar Whole rock This paper
PT

14
13585 ± 75 yr B.P. C Ash below PWA Siebe et. al. 2005
Ocusacayo >21675 + 220/- 14
yr B.P. C Paleosol on lava flow Siebe et. al. 2005
215
CE

14
Oyameyo 13755 ± 95 yr B.P. C Charcoal below lapilli Wallace & Carmichael, 1999
14
>15570 ± 300 yr B.P. C Paleosoil on cone Kirianov, et. al. 1990
Raíces 14
>15740 ± 80 yr B.P. C Paleosoil on cone Kirianov, et. al. 1990
AC

14
P3C2 18680 ± 120 yr B.P. C Paleosoil on lava Kirianov, et. al. 1990
14
Malinale 18900 ± 600 yr B.P. C Charcoal in cone scoria Kirianov, et. al. 1990
Hijo del >20895 + 245/- 14
yr B.P. C Paleosoil on lava below PWA Siebe et. al. 2005
Cuautzin 235
40 39
V. Aholo 23 ± 15 ka Ar/ Ar Whole rock This paper
40 39
V. Cuautepel 34 ± 5 ka Ar/ Ar Whole rock This paper
40 39
Achichipilco 34 ± 6 ka Ar/ Ar Whole rock This paper
40 39
Zitlaltpel 35 ± 4 ka Ar/ Ar Whole rock This paper
Intermediate

40 39
Malacatepec 95 ± 12 ka Ar/ Ar Whole rock This paper
40 39
group

Tlayacapan 133 ± 32 ka Ar/ Ar Whole rock This paper


40 39
Tezoyuca 2 159 ± 26 ka Ar/ Ar Whole rock This paper
40 39
Quepil 181 ± 20 ka Ar/ Ar Whole rock This paper
ACCEPTED MANUSCRIPT

40 39
Coatzontle 222 ± 52 ka Ar/ Ar Whole rock This paper
40 39
Tezoyuca 1 238 ± 51 ka Ar/ Ar Whole rock This paper
40 39
C. La Estrella 765 ± 30 ka Ar/ Ar Whole rock This paper
40 39
La Corona 820 ± 11 ka Ar/ Ar Whole rock This paper
40 39
Older group

Jumiltepec 846 ± 8 ka Ar/ Ar Whole rock This paper


40 39
Las Tetillas 986 ± 46 ka Ar/ Ar Whole rock This paper
40 39
AT-1 1020 ± 160 ka Ar/ Ar Whole rock Arce et al., 2013
40 39
Chimalhuacan 1094 ± 23 ka Ar/ Ar Whole rock This paper
40 39
Peñón de 1124 ± 21 ka Ar/ Ar Whole rock This paper

T
Baños 40 39
1294 ± 36 ka Ar/ Ar Whole rock This paper
40 39
Mazatepec 23 ± 4 ka Ar/ Ar Whole rock This paper

IP
40 39
Tecuautzi 29 ± 4 ka Ar/ Ar Whole rock This paper
SSC

40 39
Yuhualixqui 62 ± 98 ka Ar/ Ar Whole rock This paper

CR
40 39
Tetecón 85 ± 32 ka Ar/ Ar Whole rock This paper
40 39
Xaltepec 132 ± 70 ka Ar/ Ar Whole rock This paper

US
After this, no monogenetic volcanism occurred for ~60 ka, with the next reactivation at 35 ± 4 ka,
AN
with the beginning of the volcanoes from the Younger Chichinautzin Monogenetic Volcanic Group
(Younger CMVG, Fig. 3b) which is the younger group of relative age. The last eruptions in this
M

group were dated at 1675 ± 40 to 1975 ± 60 yr B.P. in the already populated area (Córdoba et al.,
1994; Martin Del Pozzo et al., 1997; Siebe, 2000).
ED

For SSC, the ages vary from 132 ± 70 to 23 ± 4 ka.


PT
CE
AC
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 2. Age of the monogenetic volcanoes. (a) PMVG was active from 1294 ± 36 ka to 765 ± 30 ka, followed by
a ~527 ka gap without monogenetic volcanism. (b) The close-up of the last 240 ka showing the age of the
Older CMVG (238 ± 51 to 95 ± 12), the Younger CMVG (<~40 ka), the SSC (132 ± 70 ka to 2 ± 56 ka) and the
second ~60 ka gap. Filled symbols are from this paper; open symbols from published data (see Table 3).
PT

The four volcanic groups are shown in the geologic map (Fig. 3). The oldest volcanic group, PMVG,
CE

is located at the northern and southern edges of the area. We consider that this volcanism
continued under the more recent flows and under lake deposits as suggested by lacustrine
sediments interbeded with volcanic ash as well as by geophysical studies in the lake (Ortega-
AC

Guerrero et al., 2017; Campos-Enríquez et al., 1997, 2015; Lozano-García et al., 1993). The cones
of the Older CMVG are located at the eastern, western and southern edges of the area and their
deposits cover the PMGV volcanic group. The Younger CMVG was emplacement in the central part
of the area covering the deposits from the Older CMVG and PMVG (Fig. 3).
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED
PT
CE
AC

Fig. 3. Geologic map of the study area. White lines: faults; white hyphenated lines: inferred faults. Blue lines
show the border of the Chalco Lake modified from Bellia et al., 1992.
ACCEPTED MANUSCRIPT

The SSC is limited to the north and the south by lacustrine deposits and normal faulting (Fig. 3). It
has a different position separated from to the other volcanic groups and its cones are closely
spaced ~1.5 km apart.

In order to better understand the differences between the volcanic groups, each was treated
separately, describing the changes in the volcanism through time.

4.3 Peñón Monogenetic Volcanic Group (PMVG).

T
The earliest monogenetic volcanism in the area began in the north with the formation of Peñón Del

IP
Marqués and Chimalhuacán volcanoes, at 1294 ± 36 ka that mark the northern boundary of this
volcanic group. Volcanism then migrated towards the south to Las Tetillas, Jumiltepec and La

CR
Corona volcanoes, which are the southernmost cones in the area (Fig. 3). Deposits from this group
probably underlie recent deposits in the central part the study area. The last monogenetic activity of

US
the PMVG was the eruption of La Estrella volcano at 765 ± 30 ka, consistent with previous
paleomagnetic studies, which show it has reverse polarity (Martin Del Pozzo, 1989) and are older
than 0.7 Ma.
AN
The volcanoes of the PMVG are aligned NE-SW (N60ºE), and the distance between the northern
and southern cones is about 78 km, making it the group with the largest area. Morphologically the
M

cones are wide and low, and they are widely spaced, except for La Corona and Las Tetillas
volcanoes that are nestled in between folded Cretaceous limestone.
ED

The earliest eruption that started at Peñón Del Marqués, Chimalhuacán, La Estrella and Jumiltepec
PT

volcanoes began with lava fountaining, which produced large massive spatter blocks and welded
lava agglomerates forming the first cone-building deposits, related with gas bubbles bursting near
CE

the surface. After this, the eruptions produced reddish coarse-grained ash layers 10 to 40 cm thick,
interbedded with coarse lapilli beds with scoriaceous and dense lava blocks in variable proportions.
Coarse lapilli beds are found at the summit in beds 4 to 7 cm thick (Fig. 4). The ash and lapilli
AC

deposits indicate that the activity became mildly explosive. The lava flows that were emplacement
on each side of the volcano are short, 4 km long on average. In the case of La Corona and Las
Tetillas volcanoes, the eruptions started with ejection of massive juvenile blocks that are found at
the base of the cone. Later on, there was a gradual change from spatter blocks to a stratified
succession of reddish ash and lapilli deposits in beds 10 to 20 cm thick, interbedded with thin
scoriaceous blocks in layers 2 to 3 cm thick indicating gradual variations in fragmentation and
vesiculation of the magma, associated to a magma with higher gas content. Lava flows from Las
Tetillas, La Corona and Jumiltepec volcanoes, in the southern part of the area, reached 8 km (Fig.
4).
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 4. Stratigraphic sections showing the alternating deposits of spatter, scoria, lapilli and ash of the PMVG
ED

cones.

4.4 Older Chichinautzin Monogenetic Volcanic Group (Older CMVG).


PT

The Older CMG started forming at 238 ± 51 ka, after a ~527 ka period with no recorded
monogenetic volcanism. It began in the southern part of the Older CMVG with Tezoyuca 1 volcano,
CE

and then volcanism shifted northward to form the Coatzontle cone and Quepil dome. After this,
Tezoyuca 2 and Tlayacapan volcanoes were born. The last eruption of this group was dated at 95 ±
12 ka and occurred on the west, at Malacatepec volcano (Fig. 3). Coatzontle volcano formed the
AC

northern boundary; while Tezoyuca 1 and Tezoyuca 2 volcanoes mark the southern boundary of
the field 40 km away. The cones are spaced 9 km apart.

The first eruption of Tezoyuca 1 and Coatzontle volcanoes started with the emission of large lava
blocks and spatter bombs that formed the base of the cone, and after that, the eruptions became
more explosive; producing thick deposits of fine and medium-grained ash (50-70 cm) interbedded
with scoriaceous blocks in thin beds (3-4 cm). To the summit, the deposits are graded beds of
coarse to fine-grained ash, suggesting that the energy in the eruptive column decreased (Fig. 5).
Lava flows from these volcanoes surround the cones.
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M
ED

Fig. 5. Stratigraphic sections from the volcanoes belong to the OCMVF showing the alternating deposits.
PT

In the case of Tezoyuca 2, Tlayacapan and Malacatepec volcanoes the eruptions started with a
central vent that produced a basal cone formed by ribbon bombs (1.5 m), spindle bombs (10 cm)
CE

and black scoriaceous lapilli (3-7 cm) randomly distributed at the base of the cone (Fig. 5). Covering
these deposits, there is an alternating sequence of beds 5 to 7 cm thick of reddish–black fine and
AC

medium-grained ash interbedded with thicker beds of spatter blocks and bombs (70 to 100 cm).
These deposits reflect changes related with gas bubbles bursting near the surface, increasing
magma fragmentation to produce the ash (Fig. 5). Magma was also emplaced through lava domes
such as Quepil dacitic dome, located in the western part of the area. The monogenetic activity in
the Older CMGF finished at 95 ka.

4.5 Younger Chichinautzin Monogenetic Volcanic Group (Younger CMVG)


The first eruptions of this volcanic group occurred after ~60 ka without monogenetic volcanism, in
the eastern part of the area at 35 ± 4 ka with the formation of Zitlaltépel volcano, followed by
Achichipilco, Cuautepel and Aholo (Fig. 3). Monogenetic volcanism then migrated to the center and
ACCEPTED MANUSCRIPT

the west of the area with the eruption of Mezontepec and other dated volcanoes (Table 3). The last
eruptions dated in this volcanic group occurred at 1675 ± 40 to 1975 ± 60 years B.P. so this field is
still considered active (Table 3; Córdoba et al., 1994; Martin Del Pozzo et al., 1997; Siebe, 2000).
The volcanoes from this volcanic group are located in the central part of the area with a distance
between the northern and southern cones of about 25 km. Many of the volcanoes are aligned E-W
and some form cone clusters.

The eruptions in the eastern part of the area began with the ejection of massive lava spatter blocks

T
(20 cm), lava agglomerates and bombs (10 cm) at Zitlaltépel, Achichipilco, Cuautepel and Aholo

IP
volcanoes. These deposits formed the basal cones and are covered by beds of scoriaceous bombs
and vesicular lapilli and ash layers (5 to 20 cm thick), indicating more fragmentation of the magma

CR
(Fig. 6). After this, the eruptions became less explosive with less fragmentation and produced thick
horizons of massive spindle bombs. This alternating between thick coarse-grained and thin finer-
grained beds through the cone is considered related to pulses in the eruptive column in an open

US
vent. To the summit of these volcanoes there are reworked deposits from Popocatépetl that was
dated in ~23 ka (Tochimilco pumice, also called White Pumice; Espinasa-Pereña and Martin Del
AN
Pozzo, 2006, Siebe, 2017). At Mezontepec volcano, to the west of the study area, volcanism also
started with the ejection of andesitic spindle bombs near the vent forming the basal cone. Covering
the basal deposits, there are thick (40-70 cm) layers of coarse vesicular lapilli and medium-grained
M

ash interbeded with thin layers (4-10 cm) of massive spatter lava blocks. The eruption ended with
scoriaceous bomb layers at the summit of the cone (Fig. 6). Tecmilco volcano, located in the central
ED

part of the area, also started with an open conduit eruption, with the ejection of basaltic welded lava
agglomerates, ribbon bombs and massive lava blocks (Fig. 6). These layers are interbedded with
PT

thin horizons of black to reddish coarse lapilli fallout and its lava flow was emplacement to the south
and reached 5 km.
CE

The Tabaquillo lava is thick flow with no cone that travelled to the south for 6 km, and has an
incipient soil cover with well preserved pressure ridges. Younger volcanoes such as Pelado, Xitle
AC

and Chichinautzin have voluminous lava flows with little ash cover and evident flow features many
of these volcanoes form cone clusters (Martin Del Pozzo, 1982; Martin Del Pozzo et al., 1997;
Siebe et al., 2005).
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 6. Stratigraphic sections for the Younger CMVG volcanoes.


ED

As we mentioned before, the Younger CMVG is considered still active and poses considerable ash
hazard for Mexico City.
PT

4.6 Sierra Santa Catarina Monogenetic Volcanic Group (SSC).


CE

These volcanoes grew in the Chalco Lake as a separate range. The first eruptions began at 132 ±
70 ka in the western part of the volcanic chain (Xaltepec volcano), then, volcanism migrated to the
East with the eruption of Tetecón volcano. The next eruption formed the Yuhualixqui volcano,
AC

followed by the emplacement of the Tecuautzi dome and Mazatepec volcano, dated at 23 ± 4 ka
(Fig. 7). However, the youngest volcano is Guadalupe on the eastern edge of the SSC since its
deposits cover those of Mazatepec volcano. The youngest age reported for this volcanic chain is 2
± 56 ka (Layer et al., 2009). The seven cones that form the SSC are closely spaced (<1.5 km),
oriented N70ºE and were emplacement over a NE-SW trending fault.
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
Fig. 7. Geologic map of Sierra Santa Catarina
M

Volcanism started in the Chalco Lake with hydromagmatic activity at Xaltepec volcano, followed by
Yuhualixqui and Mazatepec volcanoes. The lake was 15-20 m deep in the 1300-1500 A.D., Aztec
ED

period (Bellia et al., 1992) but the lake level fluctuated for the last 150kyr, at times turning into a
marsh ( Ortega-Guerrero et al., 2017). At the base of the Xaltepec cone there are basaltic-andesite
medium to coarse-grained wet pyroclastic surges with juvenile glassy scoraceous fragments and
PT

accrecionary lapilli, associated endurated massive fine ash surges and diatomite beds (Fig. 8). In
these volcanoes, there were no diatreme structures, volcanic breccias nor country rock fragments in
CE

the layers that form the cones, confirming a shallow interaction with the lake water. This contrasts
with the Souana crater in Japan which has units of chaotic massive breccia with blocks from lava
wall rocks with juvenile bombs that are elongated upward. a pile of steeply dipping volcanic
AC

breccias, breccia with angular lithic blocks of host rocks several meters in diameter and a unit of
stratified coarse grained volcanic breccias rich in angular blocks of lava, covered by a horizontally
stratified tuff of fine-grained volcanic ash, lapilli and lithic sand (Geshi et al., 2011). In diatremes or
maars successive explosions produce a mixture of host deposits (Graettinger et al., 2011) that are
not found in the SSC.

The SSC cones are tuff cones that evolve into scoria cones. They do not have the morphology
associated with the interaction of groundwater with magma (tuff rings when the crater floor level is
above the preeruptive topography and maars when it is below; Brand and Heiken, 2009). However,
ACCEPTED MANUSCRIPT

the cones in the SSC were formed in the early stages by the interaction of magma and a shallow
lake that formed the base of the tuff cones with abundant juvenile glassy fragments, massive
surges and lapilli deposits. These characteristics are similar to those of the large Table Rock tuff
cone that has massive to diffusely-stratified lapilli tuff, with a finely-laminated cap of fine-grained
sediment (Brand and Clark; 2009) and to those of Jeju Island where the base of the cone is made
up of bedded pyroclastic surge and fall deposits, associated with hydromagmatic activity in shallow
seawater (Sohn et al., 2011).

T
Once the magma-lake water interaction decreased, volcanism turned subaerial, forming spatter,

IP
spindle bombs and coarse lapilli horizons, such as at Tetecón volcano. In other volcanoes, like
Mazatepec, there are layers of coarse scoriaceous lapilli fallout deposits and alternating sequences

CR
of wet surges, large amounts of fine ash (90 cm thickness), lapilli (10 cm thick) and spatter blocks in
horizons 5 cm thick, related to fluctuating lake levels (Fig. 8). This change from hidromagmatic to

US
magmatic activity was also identified at the GVF (Cataluña, Spain) where activity began with
hydromagmatic phases forming wet surges and ash and lapilli deposits that changed to more
magmatic with welded scoria, bombs and even lava flows (Martí and Mallarach, 1987; Bolós et al.,
AN
2014) and at Maungauika cone (AVF) where coarse ash, coarse lapilli, fine lapilli, bombs and
juvenile lithics formed the base of the volcano during hydromagmatic activity and then, turned
Strombolian when water no longer entered the vent, leading to the magmatic phase with
M

spatter/scoria (Agustín-Flores et al., 2015).


ED

The last activity at the Sierra Santa Catarina formed the lava dome that was extruded in the central
part of the range (Tecuautzi dome; Fig. 8).
PT
CE
AC
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
M

Fig. 8 Stratigraphic sections of SSC.


ED

5. Geochemistry.
The geochemical data shown in Table 4 indicates that the samples range from basalts to dacites.
PT

However, there are differences between each volcanic group: PMVG is mainly basaltic-andesite,
the Older and Younger CMVF are basalts to dacites and plot on the medium K calc alkaline field.
CE

The SSC is mostly basaltic-andesite and plots in the high and medium K calc alkaline fields (Fig. 9).
The PMVG samples have similar values for SiO2 (Fig. 10a), wide dispersion for P2O5, Zr and Rb
(Figs. 10b, 10d, 10e) and the Nb concentrations increase with age (Fig. 10f). The samples from the
AC

Older CMVG plot into two groups one with high MgO and Nb and the other with low MgO and Nb
(Fig. 10b and 10f). The rocks from the Younger CMVG also have low and high MgO (Fig. 10b). Our
samples have low Nb, but those from the youngest cones like the ones from Chichinautzin and
Suchiooc volcanoes also have higher Nb (Straub et al., 2006, 2009). The samples from Younger
CMVG have a wide range in Rb concentrations (Fig. 10e). The SSC samples plot distinctly
separated from the other volcanoes, with higher concentrations of P 2O5, Zr and Nb, and
intermediate Rb values (Figs. 10b, 10d, 10f, 10e).
ACCEPTED MANUSCRIPT

Table 4. Geochemistry of the four volcanic groups. Major elements are reported in %wt and trace elements in
ppm
Volcanic
Volcano Sample SiO2 TiO2 CaO K2O P2O5 Rb Zr
group Al2O3 Fe2O3t MnO MgO Na2O Nb

Tlaltizapán JJ5 53.30 0.97 14.96 8.24 0.13 9.65 8.30 3.19 1.11 0.34 26 171 4
Younger CMVG

Tecmilco CV3 49.79 1.85 16.49 10.40 0.15 8.37 8.58 3.38 0.90 0.34 19 181 16
Mezontepec MPA4 62.98 0.94 15.51 5.74 0.09 3.16 4.92 4.07 1.84 0.28 55 240 10
Aholo AJ2 64.13 0.71 15.67 5.04 0.08 3.01 4.70 4.13 1.85 0.16 60 190 3
Cuautepel AJ1 58.34 0.89 17.54 6.48 0.11 5.02 5.77 3.76 1.48 0.21 43 182 5
Achichipilco CT5 61.34 0.67 15.18 5.44 0.08 4.26 5.19 3.66 2.04 0.18 71 199 4

T
Zitlaltepetl CT2 63.33 0.64 17.46 4.54 0.06 2.89 4.05 3.96 2.26 0.14 92 228 6
Malacatepec MPA6 62.43 0.95 15.41 5.82 0.09 3.56 5.09 4.23 1.83 0.28 54 233 9

IP
Malacatepec MPA7 60.32 0.87 15.73 6.12 0.09 4.86 5.85 3.60 1.72 0.18 45 210 4
Older CMVG

Tlayacapan CT4 50.76 1.51 17.26 9.09 0.13 8.04 7.90 3.25 1.10 0.34 26 198 17

CR
Tezoyuca 2 CV-9 51.74 1.96 15.62 9.42 0.15 8.27 7.79 3.44 1.21 0.46 20 198 17
Quepil MPA5 68.18 0.58 14.65 3.91 0.07 1.67 4.06 4.56 1.92 0.15 29 219 3
Coatzontle MPA1 62.18 0.70 17.38 5.14 0.07 2.89 4.97 4.30 1.34 0.17 30 198 4
Tezoyuca 1 CV2 51.99 1.64 15.79 9.76 0.14 6.95 8.06 3.57 1.14 0.45 27 202 16

US
La Corona CV4 59.53 0.94 17.47 6.20 0.09 3.14 6.16 4.24 1.60 0.21 42 207 4
Jumiltepec JJ1 54.28 0.93 16.99 7.52 0.11 7.14 7.83 3.47 1.33 0.27 30 175 4
PMVG

Las Tetillas * LT3* 53.30 0.96 16.60 8.10 0.13 8.15 8.23 3.59 0.92 0.19 18 134 5
AN
Chimalhuacán CH2 54.93 1.12 18.83 7.23 0.11 4.53 7.24 3.54 1.44 0.35 35 222 9
Peñón de los
54.42 1.34 17.13 8.18 0.11 5.42 7.72 3.75 1.33 0.41
Baños CH3 32 227 12
Mazatepec CM1 57.22 1.41 16.92 8.84 0.13 3.03 6.65 3.90 1.61 0.44 40 284 16
M

Mazatepec CM-11 56.78 1.46 17.16 8.61 0.14 3.37 6.58 4.08 1.67 0.46 38 281 16
Tecuautzi CM3 59.01 1.24 16.40 8.42 0.14 2.90 6.10 4.06 1.61 0.40 44 280 15
SSC
Tetecón CM-16b 56.79 1.34 15.69 8.32 0.14 4.34 6.87 4.07 1.89 0.65 40 372 22
ED

Xaltepec CM-5 56.63 1.34 15.94 8.27 0.14 4.33 6.82 3.99 1.88 0.66 41 375 22
Tecuautzi CM-18 56.49 1.40 16.02 8.33 0.14 4.71 6.35 3.97 1.93 0.60 39 354 21
* Straub et al., 2008
PT
CE
AC

Fig. 9 Chemical classification diagrams of the samples from the study area indicating composition varies from
(a) basaltic to dacitic, (b) samples plot on the medium potassium and (c) calc alkaline and field.
ACCEPTED MANUSCRIPT

T
IP
CR
US
AN
Fig. 10. Concentrations of MgO, P2O5, SiO2, Nb, Zr, and Rb with age.
M

6. Discussion.
ED

6.1. Morphometry of the cones.


The cones in the youngest volcanic group are higher and narrower than older volcanoes; this
PT

indicates that the erosion of the cinder cones is associated with the amount of time it is exposed to
degradation processes. The slope and h/W b ratio show good correlation with age groups. The
average slope for each volcanic group increases according with the age, 13º for PMVG; 18º for the
CE

Older CMVG, 28º for the Younger CMVG and 26º for the SSC. The h/W b ratio for the recent cones
is up to 0.28 and for the older ones as low as 0.06. This is consistent with published data for several
AC

areas in the world in different climatic conditions (Hooper 1995; Wood, 1980; Inbar and Risso,
2001, Inbar et al., 2011). There is a clear relationship between slope angle and h/W b ratio for our
data and those calculated with other volcanic zones like the Etna cones and San Francisco
Volcanic Field or Tolbachic Volcanic Field (Fig. 11). This supports the utility of relative dating since
the most recent cinder cones are associated with the largest slopes and h/W b ratios. The relative
dating obtained in this paper for the cones from the four volcanic groups was corroborated with the
radiometric dates.
ACCEPTED MANUSCRIPT

T
IP
CR
Fig. 11. Cone slope angles and h/Wb ratios for cinder cones in the study area. Predicted curve is from

US
weathering and mass wasting model by Wood (1980).

The older cones have larger volumes whereas the youngest volcanoes have smaller cones but with
AN
voluminous lava flows (Fig. 12).
M
ED
PT
CE
AC

Fig. 12. Volume of the cones. Note larger cones volumes are for PMVG and the lowest values for Younger
CMVG

6.2 Eruptive activity.


The initial eruptions in the PMVG at Peñón Del Marqués, Chimalhuacán, La Estrella and Jumiltepec
volcanoes, were characterized by the ejection of abundant clots of magma with low gas content that
formed large massive blocks, welded agglomerates, ribbon and spindle bombs, and lapilli. Minor
amounts of ash and scoriaceous bombs were present. The pyroclasts are found in inversely graded
beds. The pyroclast morphology indicates that the PMVG clasts experienced little gas expansion
ACCEPTED MANUSCRIPT

and magma fragmentation near the surface with low velocity at which the particle was expelled from
the conduit. The size (1-10 cm), the aerodynamic shapes (ribbon and spindle) of the pyroclasts and
the inversly graded beds are related with Strombolian activity. The eruptions in the PMVG were
characterized by weak explosions with the release of pressurized gas, forming bombs and clasts. In
addition, we found a few ash and fine-lapilli layers that were associated with more magma
fragmentation during the eruptions. The PMVG was formed mainly by basaltic-andesite Strombolian
eruptions.

T
The second volcanic group (Older CMVG) started with the Tezoyuca 1 and Coatzontle volcanoes.

IP
These cones were predominantly formed by scoriaceous lapilli and coarse-grained ash in several
normal graded beds. Massive spatter bombs and ribbon and spindle bombs are sparse. The clasts

CR
are angular to sub angular and moderately to highly vesicular. The size of the finer volcanic material
(ash and lapilli) is in average < 1 cm. More explosive activity produced larger amounts of lapilli and

US
ash due to increased gas content and fragmentation, features associated with violent Strombolian
eruptions. The Older CMVG produced basaltic andesites-andesites.
AN
The Younger CMVG started with Zitlaltepetl, Achichipilco, Cuautepel and Aholo volcanoes and
ejected massive spatter lava blocks and escoriaceous bombs (1-10 cm) that formed massive,
spindle, ribbon and rounded small clasts, features belonging to the Strombolian activity. This is
M

related to low velocities and less gas expansion. Afterwards the eruptive dynamics changed
producing blocky and angular to slightly rounded clasts, coarse vesicular lapilli scoria, and fine-
ED

grained ash (<1 cm), indicating it was violent Strombolian activity. These ash-rich deposits are
associated to high velocities and more gas expansion. The presence of glassy crusts and cracks in
PT

the pyroclasts increases in the younger cones is not frequent in the older ones. The youngest
volcanoes in this group have extensive lava flows that are less than 2 ka (Martin Del Pozzo et al.,
CE

1997). Some of the cones have normal and inversely graded beds, associated with a fluctuating
eruptive columns and changes between Strombolian and violent Strombolian eruptive activity. The
Younger CMVG rocks show a wider composition range, from basalts to dacites.
AC

The SSC is basaltic andesite to andesite, and the initial eruptions in the Xaltepec, Yuhualixqui, and
Mazatepec volcanoes have thin to thick, planar to low angle cross-bedding, with massive to graded
layers with ballistic impact sags. These features are associated with hydromagmatic activity, in the
shallow lake that varied in depth (from 15-20 m to a marshy playa Bellia et al., 1992; Ortega
Guerrero et al., 2017), when the magma- lake water interaction was high and formed fine-grained
ash in thick massive beds and wet surges deposits (<0.5 cm). This hydromagmatic phase changed
to subaerial volcanism when the lake water no long entered the conduit, producing welded
agglomerates, massive lava blocks and escoriaceous spindle bombs (2-15 cm). Clasts with glassy
crusts, scoriaceous bombs and scoriaceous lapilli were dominant. The pyroclastic sequence
ACCEPTED MANUSCRIPT

observed in the SSC cones, with its clear differences in the hydromagmatic and magmatic deposits,
is similar to those identified at other volcanoes fields as at the Catalán PlaSa Ribera and Puig d’Àdri
volcanoes that have wet surges at the base of the cone produced by hydromagmatic activity and a
scoria cone built up in a Strombolian phase (Martí et al., 2011; Bolós et al., 2014; Martí and
Mallarach, 1987).

According to the results, we can see that, although in each volcanic groups individual cones were
built up by specific eruptive activity, there seems to be a predominant type: (1) the PMVG was

T
mainly Strombolian; (2) the Older CMVG was dominantly violent Strombolian style, (3) the Younger

IP
CMVG varied from Strombolian, violent Strombolian to effusive, and (4) the SSC was
hydromagmatic that turned to Strombolian.

CR
6.3 Longevity and controlling factors on cone emplacement.

US
Monogenetic volcanism lasted ~529 ka (PMVG), ~143 ka (Older CMVG), at least ~35 ka (Younger
CMGF) and ~130 ka (SSC). Two temporal gaps were identified; the first one was ~527 ka long
(between the PMVG and Older CMVG) and the second one ~60 ka (between Older CMVG and
AN
Younger CMVG), that mark the end of the activity in every group. These time gaps are much longer
than those for any time documented between individual events in other volcanic fields, like in the
Campi Felgrei, Italy (3 ka) and in the Newer Volcanics Province, Australia (9.7 ka; De Vito et al.,
M

2001; Hare and Cass, 2005).


ED

The early monogenetic cones in the Chichinautzin area show a NE-SW (N60°E) alignment. This
volcanism seems to be contemporaneous with a transcurrent episode that generated several faults
PT

trending NE in the area to the west studied by García-Palomo et al. (2000). After the last eruption in
the PMVG, magma ascent stopped probably because of the re-orientation of the stress field in the
CE

area.

Cone orientation progressively changed in the younger volcanic groups, from N60ºE to N70ºE. This
AC

N70ºE orientation is the alignment of the SSC cones, indicating that the emplacement was
controlled by NE-SW faults. García-Palomo et al. (2000) and Belloti et al. (2006), mention that
during the early Pleistocene there was a fault system with NE-SW orientation in the area. We
consider that the monogenetic volcanism in the Pleistocene (Older CMVG and SSC) is related to
that faults system. Monogenetic volcanism stopped again after the last eruption of the Older CMVG
by ~60 ka.

The stress pattern changed again after the second gap (~60 ka), and monogenetic activity was
reactivated producing a dominate E-W orientation in the new cones. Transtensional faulting with
this same orientation appears to control this emplacement. This volcanic reactivation could be
ACCEPTED MANUSCRIPT

associated to faults and fractures that were produced by the Tenango Fault System (Bloomfield,
1975, García-Palomo et al., 2000; Belloti et al., 2006) that are considered active at present as
suggested by seismic swarms (Yamamoto and Mota, 1988). The youngest volcanoes are
concentrated in the central part of the area which could be due to increased fracturing.
Cone orientation appears to be fault controlled also indicating changes in the stress field with time.
This behavior is common in other monogenetic fields, and reveals how sensitive the magma
migration patterns are to the regional and/or local tectonics (Martí et al., 2016).

T
The geostatistical analyses show the distribution of the cones and how the area with monogenetic

IP
volcanoes shrinks with the age. The central part corresponds to the Younger CMVG area, which is
covering the deposits of the previous volcanic groups (Fig, 13).

CR
US
AN
M
ED
PT
CE
AC

Fig. 13 Spatial distribution of the areas for each volcanic group.


ACCEPTED MANUSCRIPT

There are several differences among the volcanic groups. The first one is their age, separated by 2
important temporal gaps. The second is the orientation of the volcanoes, with cones that changed
their alignment, from N60°E to N70°E to E-W, and covered smaller areas with time. The third is the
geochemistry, although the samples are medium K, calc alkaline andesites, there are differences
among the groups. The existence of the high MgO and Nb and low MgO and Nb groups in the
Older and Younger CMVG, and the high Zr, P2O5 and Nb concentrations in the SSC samples,
suggest differences in the source related to changes in the stress orientation.

T
IP
With these features, we propose that these volcanic groups are, actually, different monogenetic
volcanic fields.

CR
7. Conclusions

US
This study indicates that the Sierra Chichinautzin is actually four different monogenetic volcanic
fields, some partially overlapping, instead of one as previously considered. These four volcanic
fields are: PMVG (1294 ± 36 to 765 ± 30 ka); Older CMVG, (238 ± 51 to 95 ± 12 ka); Younger
AN
CMVG (< 35 ± 4 ka), and the SSC (132 ± 70 to 2 ± 56 ka). Each one of them shows distinct
temporal, lineation and chemical characteristics.
M

Monogenetic volcanism in the study area started in the north, with the PMVG forming widely-space
large cones aligned NE-SW (N70ºE) controlled by NE-SW faults. After the first gap of ~527 ka,
ED

monogenetic volcanism formed the Older CMVG that ended about 60 ka. After this second gap in
the monogenetic activity, the fault system changed from NE-SW to E-W producing the Younger
PT

CMVG cones aligned E-W, with closer spaced volcanoes. The stopping of monogenetic volcanism
could be due to changes in the stress orientation. The SSC was emplaced NE-SW with closely
CE

spaced cones.

The northern and southern boundaries of the monogenetic volcanoes in the area got smaller
AC

through time: ~78 km for PMVG, ~40 km for Older CMVG and ~25 km for Younger CMVG; but the
E-W limits were constant. This shrinking in area of the monogenetic fields may be linked to the
changing fault systems from NNE-SSW to NE-SW to E-W.

The geochemical differences between the volcanic groups could be related to the changes in the
stress fields and different magma sources.

Eruptive styles changed from Strombolian (PMVG) to Violent Stromboliant in the volcanoes of the
Older CMVG; while the Younger CMVG had more Strombolian, violent Strombolian and effusive
eruptions. The SSC was a dominantly hydromagmatic and was following by Strombolian activity.
ACCEPTED MANUSCRIPT

Future eruptions will most likely be Strombolian to violent Strombolian since the lake has been dried
out, although ash fall in the Mexico City and nearby area as well as lava effusion could impact the
area.

The differences in age, emplacement orientation and chemistry support the idea that these volcanic
groups are actually four different volcanic fields. However, the Older and the Younger CMVG could
belong to the same volcanic field, due to the similarities in the geochemistry, but were separated
because of the temporal gap (60 ka) in monogenetic volcanism that is longer than any previously

T
known in other areas.

IP
This paper shows that with more detailed dating and geomorphological analyses, large volcanic

CR
fields may actually be made up of several fields. This could also be the case of other large
monogenetic volcanic fields throughout the world.

US
Acknowledgements
Funding for this study was provided by DGAPA-PAPIIT (UNAM) and SECITI (CDMX). We thank
AN
Alan Rodríguez Moreno, Crisóforo Torres Romero and Andrés Salinas for their assistance in the
field. The authors would also thank to Dr. Karoly Németh and anonymous reviewer, for their
comments that considerably improved the paper.
M

List of Tables and Figures.


ED

Table 1. Number of volcanoes, tectonic environment and ages for monogenetic volcanic fields.
PT

Table 2. Geomorphological parameters of the volcanoes.

40 39 14
CE

Table 3. Ar/ Ar and C dates for the Chichinautzin monogenetic volcanic field.

Table 4. Geochemistry of the four volcanic groups. Major elements are reported in %wt and trace
AC

elements in ppm.

Figure 1. (a) Map of the Sierra Chichinautzin monogenetic volcanoes (SC) on a Landsat 8 Thematic
Mapper (RGB 4-3-2 composite image). The white rectangle outlines the study area. CHI:
Chimalhuacán, JT: Jumiltepec; TCM: Tecmilco, PB: Peñón Del Marqués, EST: La Estrella, TLT:
Tlaltizapán, TLY: Tlayacapan, CTZ: Coatzontle, ZTL: Zitlaltépetl, ACH: Achichipilco, MT:
Malacatepec, MZ: Mezontepec, AH: Aholo, CUA: Cuautepel, QP: Quepil, TB: Tabaquillo, TZ:
Tezoyuca 1 and 2, CO: La Corona, TT: Las Tetillas, 3C: Tres Cumbres, SSC: Sierra Santa
Catarina. (b) Tectonic map of the Trans Mexican Volcanic Belt (TMVB) with location of
ACCEPTED MANUSCRIPT

Chichinautzin Monogenetic Volcanic Field (CMVF). Black lines: faults, hyphenated lines: inferred
faults (modified from Campos-Enríquez et al., 1997; 2015).

Fig. 2. Age of the monogenetic volcanoes. (a) PMVG was active from 1294 ± 36 ka to 765 ± 30 ka,
followed by a ~527 ka gap without monogenetic volcanism. (b) The close-up of the last 240 ka
showing the age of the Older CMVG (238 ± 51 to 95 ± 12), the Younger CMVG (<~40 ka), the SSC
(132 ± 70 ka to 2 ± 56 ka) and the second ~60 ka gap. Filled symbols are from this paper; open
symbols from published data (see Table 3).

T
IP
Fig. 3. Geologic map of the study area. White lines: faults; white hypenathed lines: inferred faults

CR
Fig. 4. Stratigraphic sections showing the alternating deposits of spatter, scoria, lapilli and ash of
the PMVG cones.

US
Fig. 5. Stratigraphic sections from the volcanoes belong to the OCMVF showing the alternating
deposits.
AN
Fig. 6. Stratigraphic sections for the Younger CMVG volcanoes.
M

Fig. 7. Geologic map of Sierra Santa Catarina


ED

Fig. 8 Stratigraphic sections of SSC.


PT

Fig. 9 Chemical classification diagrams of the samples from the study area indicating composition
varies from (a) basaltic to dacitic, (b) samples plot on the medium potassium and (c) calc alkaline
CE

and field.

Fig. 10. Concentrations of MgO, P2O5, SiO2, Nb, Zr, and Rb with age.
AC

Fig. 11. Cone slope angles and h/Wb ratios for cinder cones in the study area. Predicted curve is
from weathering and mass wasting model by Wood (1980).

Fig. 12. Volume of the cones. Note larger cones volumes are for PMVG and the lowest values for
Younger CMVG

Fig. 13 Spatial distribution of the areas for each volcanic group.


ACCEPTED MANUSCRIPT

References.
Aguirre-Díaz, G. J., Jaimes-Viera, M. C., and Nieto-Obregón, J., 2006, The Valle de Bravo Volcanic Field.
Geology and geomorphometric parameters of a Quaternary monogenetic field at the front of the Mexican
Volcanic Belt, in Neogene-Quaternary continental margin volcanism: A perspective from Mexico, Siebe, C.,
Macías, J.L., Aguirre-Díaz, G.J., eds., Geological Society of America Special Paper 402, 125-140.

Agustin-Flores, J., Nemeth, K., Cronin, S.J., Lindsay, J.M. Kereszturi, G., 2015. Construction of the North Head
(Maungauika) tuff cone: a product of Surtseyan volcanism, rare in the Auckland Volcanic Field, New Zealand.
Bull Volcanol, 77, 2.

T
IP
Alaníz-Álvarez, S., Nieto-Samaniego, A.F., Ferrari, L., 1998. Effects of strain rate in the distribution of
monogenetic and polygenetic volcanism in the Transmexican Volcanic Belt. Geology, 26, 591-594.

CR
Alberico, I., Lirer, L., Petrosino, P., and Scandone, R. 2002. A methodology for the evaluation of long-term
volcanic risk from pyroclastic flows in Campi Flegrei (Italy): J Volcanol Geoth Res; 116, 63-78

US
Allen, S. and Smith, I. 1994, Eruption styles and volcanic hazard in the Auckland Volcanic Field, New Zealand,
in Geoscience Reports of Shizuoka University, Shizuoka Univ., Shizuoka, Japan. 5–14
AN
Aranda–Gómez, J.J., Luhr, J.F., Housh, T., Valdez–Moreno, G., Chávez–Cabello, G., 2005, El volcanismo tipo
intraplaca del Ceno oico tard o en el centro y norte de ico una revisi n: Boletín de la Sociedad Geológica
M

Mexicana, 57, 187–225.


ED

Arce, J.L., Layer, P.W., Lassiter, J.C., Benowitz, J.A., Macias, J.L., Ramírez-Espinosa, J., 2013b, 40Ar/39Ar
dating, geochemistry, and isotopic analyses of the quaternary Chichinautzin volcanic field, south of Mexico
City: implications for timing, eruption rate, and distribution of volcanism: Bull. Volcanol., 75, 774. doi:
PT

10.1007/s00445-013-0774-6
CE

Arce, J.L., Muñoz-Salinas, E., Castillo, M., Salinas, I. 2015. The ~2000 yr BP Jumento volcano, one of the
youngest edifices of the Chichinautzin Volcanic Field, Central Mexico, J Volcanol Geoth Res; 308, 1, 30-38.
AC

Aydın, F., Schmitt, A.K., Siebel, W., Sönme , ., Ersoy, Y., Lermi, A., Dirik, K., Duncan, R. 2014. Quaternary
bimodal volcanism in the Niğde Volcanic Comple (Cappadocia, central Anatolia, Turkey) age, petrogenesis
and geodynamic implications. Contrib. Mineral. Petr., 168, 1078-1102, 10.1007/s00410-014-1078-3

Balogh, K., and Németh, K. 2005. Evidence for the Neogene small-volume intracontinental volcanism in
Western Hungary: K/Ar geochronology of the Tihany Maar Volcanic Complex. Geologica Carpathica, 56, 1, 91-
99.

Benowitz, J. A., Layer, P. W., Vanlaningham, S., 2014, Persistent long-term (c. 24 Ma) exhumation in the
Eastern Alaska Range constrained by stacked thermochronology. Geological Society, London, Special
Publications, 378, 1, 225-243.
ACCEPTED MANUSCRIPT

Bebbington M. 2013. Assessing spatio-temporal eruption forecast in a monogenetic volcanic field. J Volcanol
Geoth Res, 252, 14-28.

Bellotti, F., Capra, L., Groppelli, G., Norini, G., 2006. Tectonic evolution of the central eastern sector of Trans
Mexican volcanic belt and its influence on the eruptive history of the Nevado de Toluca Volcano (Mexico). J
Volcanol Geoth Res,158, 21–36 Special Issue.

Blatter, D., Carmichael, I.S.E., Deino, A., and Renne, P., 2001, Neogene volcanism at the front of the central

T
Mexican Volcanic Belt: Basaltic andesites to dacites, with contemporaneous shoshonites and high-TiO2 lava:

IP
Geological Society of America Bulletin, 113, 1324–1342. doi: 10.1130/0016-
7606(2001)113<1324:NVATFO>2.0.CO;2.

CR
Bloomfield, K., 1975, A late-Quaternary monogenetic volcano field in central Mexico: Geol Rundsch, 64, 476–
497.

US
Böhnel, H. and Molina-Garza, R. 2002. Secular variation in Mexico during the last 40,000 years, Phys. Earth
Planet. Int., 4150, 1-11.
AN
Bolós, X., Planagumà, L., Martí, J., 2014. Volcanic stratigraphy of the Quaternary La Garrotxa Volcanic Field
(north-east Iberian Peninsula). J. Quat. Sci. 29 (6), 547–560.
M

Brand, B.D. Clarke, A.B., 2009. The architecture, eruptive history, and evolution of the Table Rock Complex,
ED

Oregon: From a Surtseyan to an energetic maar eruption. J Volcanol Geoth Res,, 180(2-4): 203-224.

Brand, B.D. Heiken, G., 2009. Tuff cones, tuff rings, and maars of the Fort Rock-Christmas Valley basin,
PT

Oregon; exploring the vast array of pyroclastic features that record violent hydrovolcanism at Fort Rock and the
Table Rock Complex. GSA Field Guide, 15: 521-538.
CE

Brenna, M., Cronin, S.J., Németh, K., Smith, I.E.M. and Sohn, Y.K., 2011. The influence of magma plumbing
complexity on monogenetic eruptions, Jeju Island, Korea. Terra Nova, 23, 2, 70-75.
AC

Brenna, M., Cronin, S.J., Smith, I.E.M., Maas, R. and Sohn, Y.K., 2012. How Small-volume Basaltic Magmatic
Systems Develop: a Case Study from the Jeju Island Volcanic Field, Korea. Journal of Petrology, 53, 5, 985-
1018.

Brenna, M., Nemeth, K., Cronin, S.J., Sohn, Y.K., Smith, I.E.M. and Wijbrans, J., 2015. Co-located
monogenetic eruptions similar to 200 kyr apart driven by tapping vertically separated mantle source regions,
Chagwido, Jeju Island, Republic of Korea. Bull. Volcanol., 77, 5.

Bucchi, F., Lara, L.E., and Gutiérrez, F. 2015. The Carran-Los Venados volcanic field and its relationship with
coeval and nearby polygenetic volcanism in an intra-arc setting. J Volcanol Geoth Res, 308, 70–81.
ACCEPTED MANUSCRIPT

Büchel, G., 1993, Maars of the Westeifel, Germany, in Negendank, J.F.W., and Zolitschka, B., eds.,
Paleolimnology of European Maar Lakes: Berlin, Heidelberg, Springer-Verlag, 1–13

Camp, V.E., and Roobol, M.J.1989. The Arabian continental alkali basalt province: Part I. Evolution of Harrat
Rahat, Kingdom of Saudi Arabia. Geol Soc Am Bull 101, 71-95. doi: 10.1130/0016-7606

Campos-Enríquez, J.O., Lermo-Samaniego, J.F., Antayhua-Vera, Y.T., Chavacán, M., Ramón-Márquez,


V.M. 2015. The Aztlán Fault System: control on the emplacement of the Chichinautzin Range volcanism,

T
southern Mexico Basin, Mexico. Seismic and gravity characterization. Boletín de la Sociedad Geológica

IP
Mexicana 67, 315-335.

CR
Campos Enríquez, J.O., Delgado Rodríguez, O., Chávez Segura, R., Gómez Contreras, P., Flores Márquez,
E.L. And F.S. Birch, 1997. The Subsurface Structure of The Chalco Subbasin (Mexico City) Inferred from
Geophysical Data, Geophysics, 62, 23-35.

US
Cañón-Tapia, E. and Walker, G.P.L. 2004. Global aspects of volcanism the perspectives of ‘plate tectonics’
and ‘volcanic systems’. Earth-Science Reviews, 66, 163–182.
AN
Carn, S.A., 2000. The Lamongan volcanic field, East Java, Indonesia: physical volcanology, historic activity
and hazards. J Volcanol Geoth Res. 95, 81–108
M

Carn, S.A. and Pyle, D.M. 2001. Petrology and geochemistry of the Lamongan volcanic field, East Java,
ED

Indonesia: Primitive sunda arc magmas in an extensional tectonic setting? Journal of Petrology, 42, 9, 1643-
1683.
PT

Cassidy, J., and Locke, C.A., 2004. Temporally linked volcanic centres in the Auckland Volcanic Field: New
Zealand. Journal of Geology and Geophysics, 47, 287–290.
CE

Cheong, C.S., Choi, J.H., Sohn, Y.K., Kim, J.C. and Jeong, G.Y., 2007. Optical dating of hydromagmatic
volcanoes on the southwestern coast of Jeju Island, Korea. Quaternary Geochronology, 2, 1-4, 266-271.
AC

Coleman RC, Gregory R (1983) Cenozoic volcanic rocks. US Geological Survey Professional Paper, 287–287

Connor, C.B., and Conway, F.M., 2000. Basaltic volcanic fields. In: Sigurdsson H (ed) Encyclopedia of
volcanoes. Academic Press, San Diego, 331–343

Córdoba, C., Martin Del Pozzo, A.L., López Camacho, J., 1994. Paleolandforms and volcanic impact on the
environment of prehistoric Cuicuilco, southern Mexico City. Journal of Archaeological Science 21, 585-596.

Di Vito, M. A., Isaia, R., Orsi, G., Southon, J., de Vita, S., D’Antonio, ., Pappalardo, L. and Piochi . 1999,
Volcanism and deformation in the past 12 ka at the Campi Flegrei caldera (Italy), J. Volcanol. Geotherm. Res.,
91, 221–246, doi:10.1016/S0377-0273(99)00037-2.
ACCEPTED MANUSCRIPT

Edbrooke, S.,Mazengarb, C. and Stephenson, W. 2003, Geology and geological hazards of the Auckland
urban area, New Zealand, Quat. Int., 103, 3–21, doi:10.1016/S1040-6182(02)00129-5.

El Difrawy, .A., Runge, .G., oufti, .R., Cronin, S.J., Bebbington, ., 2013. A first ha ard analysis of the
Quaternary Harrat Al- adinah volcanic field, Saudi Arabia. J. Volc. Geotherm. Res. 267, 39–46.

Ercan, T., Tokel, S., Matsuda, J.I., Ui, T., Notsu, K., Fujitani, T. 1992. New geochemical, isotopic and
radiometric data of the Quaternary volcanism of Hasandağı-Karacadağ (Central Anatolia) TJK Bülteni, 7, 8-21

T
IP
Erlund, E.J., Cashman, K.W., Wallace, P.J., Pioli, L., Rosi, M., Johnson, E. and Granados-Delgado, H. 2010.
Compositional evolution of magma from Paricutin Volcano, Mexico: the tephra record. J Volcanol Geoth Res,

CR
197, 167–187.

Espinasa-Pereña, R., and Martin-Del Pozzo, A.L. 2006. Morphostratigraphic evolution of Popocatépetl

US
volcano, México, in Siebe, C., Macías, J.L., and Aguirre-Díaz, G.J., eds., Neogene-Quaternary continental
margin volcanism: A perspective from Mexico: GSA Sp Paper 402, 101–123, doi: 10.1130/2006.2402(05).
AN
Ferrari, L., Lopez-Martinez, M., Aguirre-Díaz, G., Carrasco-Núñez, G., 1999. Space–time patterns of Cenozoic
arc volcanism in central Mexico: from the Sierra Madre Occidental to the Mexican Volcanic Belt. Geology 27,
303–306.
M

Ferrari, L., Orozco-Esquivel, T., Manea, V., Manea, M., 2012. The dynamic history of the Trans-Mexican
ED

Volcanic Belt and the Mexico subduction zone. Tectonophysics 522–523, 122–149.

Ferrari L., Orozco-Esquivel T., Bryan S.E., López-Martínez M., Silva Fragoso A., 2017. Cenozoic magmatism
PT

and extension in western Mexico: linking the Sierra Madre Occidental Silicic Large gneous Province and the
Comond Group with the Gulf of California rift. Earth Science Reviews, DOI: 10.1016/j.earscirev.2017.04.006
CE

García-Palomo, A., Macías, J.L., and Garduño, V.H., 2000, Miocene to Recent structural evolution of the
Nevado de Toluca Volcano region, central Mexico: Tectonophysics, Special Volume, Post-Laramide
AC

magmatism and tec-


tonics in Mexico plate interaction, 318, 281–302

García -Palomo, A., Macías, J.L., Arce, J.L., Capra, L., Garduño, V.H., Espindola, J.M., 2002. Geology of
Nevado de Toluca Volcano and surrounding areas, central Mexico. Geol. Soc. Am. Map and Chart Series
MCH089, 1– 48.

Gençalioğlu-Kuşcu. G. 2011. Geochemical characterization of a Quaternary monogenetic volcano in Erciyes


Volcanic Complex: Cora Maar (Central Anatolian Volcanic Province, Turkey) Int. J. Earth Sci., 100, 1967-1985
Geshi, N., Nemeth, K., Oikawa, T., 2011. Growth of phreatomagmatic explosion craters: A model inferred from
Suoana crater in Miyakejima Volcano, Japan. J. Volcanol. Geotherm. Res., 201, 1-4, 30-38.
ACCEPTED MANUSCRIPT

Graettinger, A.H., Valentine, G.A., Sonder, I., Ross, P.S., White, J.D.L., Taddeucci, J., 2014. Maar-diatreme
geometry and deposits: Subsurface blast experiments with variable explosion depth. Geochemistry
Geophysics Geosystems, 15, 3, 740-764.

Gray, C.M., McDougall, I. 2009. K-Ar geochronology of basalt petrogenesis, Newer Volcanic Province, Victoria.
Australian Journal of Earth Sciences 56.2; 245-258.

Guilbaud, M.-N., Siebe, C., Agustín-Flores, J., 2009. Eruptive style of the young high-Mg basaltic-andesite

T
Pelagatos scoria cone, southeast of México City. Bull. Volcanol. 71, 8, 859–880

IP
Hare, A.G. and Cas, R.A.F. 2005. Volcanology and evolution of the Werribee Plains intraplate, basaltic lava

CR
flow-field, Newer Volcanics Province, southeast Australia. Australian Journal of Earth Sciences, 52:1, 59-78,
DOI: 10.1080/08120090500100051

US
Hasenaka, T., Carmichael, I.S.E.1985. A compilation of location, size, and geomorphological parameters of
volcanoes of the Michoacan-Guanajuato volcanic field, central Mexico. Geofis Int 24, 4, 577–608
AN
Hasenaka T, Carmichael ISE. 1987. The cinder cones of Michoacán-Guanajuato, Central Mexico: petrology
and chemistry. J Petrol 28, 2, 241–269.
M

Hooper, D.M., 1995, Computer-simulation models of scoria cone degradation in the Colima and Michoacán-
Guanajuato Volcanic Field, Mexico: Geofísica Internacional, 34, 321–340.
ED

Hooper, D.M., Sheridan, M.F., 1998. Computer-simulation models of scoria cone degradation. J. Volcanol.
Geotherm. Res. 83, 3–4, 241–267.
PT

Hyodo, M. and Kitaba, I. 2015. Timing of the Matuyama–Brunhes geomagnetic reversal: Decoupled thermal
CE

maximum and sea-level highstand during Marine Isotope Stage 19. Quaternary International. 383.
10.1016/j.quaint.2015.01.052.
Inbar, M., Risso, C., 2001. A morphological and morphometric analysis of a high density cinder cone volcanic
AC

field Payun Matru, south-central Andes, Argentina. Zeitschrift für Geomorphologie 45, 321–343.
Inbar, M., Gilichinsky, M, Melekestsev, M., Melnikov. D. and Zaretskaya, N. 2011. Morphometric and
morphological development of Holocene cinder cones: A field and remote sensing study in the Tolbachik
volcanic field, Kamchatka, J Volcanol Geoth Res, 201, 301–311.

Innocenti. F., Mazzuoli, G., Pasquare, F., Radicati Di Brozola F., Villari, L. 1975. The Neogene calc-alkaline
volcanism of Central Anatolia: geochronological data on Kayseri-Niğde area. Geol. Mag., 112, 349-360

Johnson, C.A., Harrison, C.G.A., 1990. Neotectonics in Central Mexico. Phys. Earth Planet. Inter. 64, 187–210.
ACCEPTED MANUSCRIPT

Kakubuchi, S., Nagao, T., and Nagao, K., 2000, K-Ar ages and magmatic history of the Abu Monognetic
Volcano Group: Japanese Magazine of Mineralogical and Petrological Sciences, 29, 191-198

Kereszturi G, Németh K, Csillag G, Balogh K, Kovács J., 2011 The role of external environmental factors in
changing eruption styles of monogenetic volcanoes in a Mio/Pleistocene continental volcanic field in western
Hungary. J Volcanol Geotherm Res 201, 1–4, 227–240

Kereszturi, G., Csillag, G., Németh, K., Sebe, K., Balogh, K., Jáger, V., 2010. Volcanic architecture, eruption
mechanism and landform evolution of a Plio/Pleistocene intracontinental basaltic polycyclic monogenetic

T
volcano from the Bakony-Balaton Highland Volcanic Field, Hungary. Cent. Eur. J. Geosci. 2, 3, 362–384.

IP
doi.org/10.2478/v10085-010-0019-2.

CR
Kermode, L., 1992. Geology of the Auckland urban area. Scale 1: 50,000. Institute of Geological and Nuclear
Sciences Geological Maps 2. Institute of Geological and Nuclear Sciences, Lower Hutt, New Zealand 1–63

US
Kirianov, V.Y., Koloskov, A.B., De la Cruz, S., Martin del Pozzo, A.L., 1990. The major stages of manifestation
of recent volcanism in the Chichinautzin zone. USSR Academy of Sciences, Geological series 311, 432-434.
AN
Kiyosugi, K., Connor, C.B., Zhao, D., Connor, L.J., Tanaka, K., 2010. Relationships between volcano
distribution, crustal structure, and P-wave tomography: an example from the Abu Monogenetic Volcano Group,
SW Japan. Bull. of Volcanol. 72:331–340 DOI 10.1007/s00445-009-0316-4
M

Layer, P.W., Macias, J.L., Arce, J.L., García-Tenorio, F., 2009, Late Pleistocene-Holocene volcanism of the
ED

Mexico Basin and assessment of volcanic hazards in one of the world’s largest cities AGU, Fall eeting, 1,
1951
PT

Layer, P.W., Hall, C.M. & York, D., 1987.The derivation of 40Ar/39Ar age spectra of single grains of
hornblende and biotite by laser step heating, Geophys. Res. Lett., 14, 757-760.
CE

Le Corvec, N., Menand, T., and Lindsay, J. 2013a, Interaction of ascending magma with pre-existing crustal
fractures in monogenetic basaltic volcanism: an experimental approach, J. Geophys. Res. Solid Earth, 118,
AC

968–984, doi:10.1002/jgrb.50142.

Le Corvec N, Spoerli KB, Rowland J, Lindsay J., 2013b. Spatial distribution and alignments of volcanic centers:
clues to the formation of monogenetic volcanic fields. Earth Sci Rev 124, 96–114
Lindsay, J., and Leonard, G., 2007, Volcanic Hazard Assessment for the Auckland Volcanic Field: State of the
Play: Geological Society of New Zealand Miscellaneous Publication 123A, 88.

Lefebvre, NS (2013) Volcanology of maar-diatreme volcanic vent complexes, Hopi Buttes Volcanic Field,
Navajo Nation, Arizona, USA. PhD Thesis, University of Otago, New Zealand.
ACCEPTED MANUSCRIPT

Le a, J., Seghedi, ., N meth, K., S akács, A., Koneĉny, V., P cskay, Z., Fülöp, A. and Kovacs, ., 2010.
Neogene-Quaternary volcanic forms in the Carpathian-Pannonian Region: a review. Central European Journal
of Geosciences, 2, 3, 207-270.

Lindsay, J., Leonard, G.S., Smid, E.R., Hayward, B.W., 2011. Age of the Auckland volcanic field: a review of
existing data. N. Z. J. Geol. Geophys., 54, 4, 379-401

Lozano-García, M.S., Ortega-Guerrero, B., Caballero-Miranda, M., Urrutia-Fucugauchi, J., 1993. Late
Pleistocene and Holocene paleoenvironments of Chalco lake, central Mexico. Quat. Res. 40, 3, 332-342

T
IP
Lugo-Hubp, J., Mooser, F., Pérez-Vega, A., Zamorano-Orozco, J., 1994, Geomorfología de la Sierra de Santa
Catarina, D.F., México: Revista Mexicana de Ciencias Geológicas, 11, 43–52.

CR
Luhr, J. F., Aranda-Gómez, J. J., Housh, T. B., 1995, San Quintín Volcanic Field, Baja California Norte,
México: Geology, Petrology and Geochemistry: Journal of Geophysical Research, 100, B7, 10353-10380.

US
Magill, C.R., McAneney, K.J., and Smith, I.E.M., 2005, Probabilistic assessment of vent locations for the next
Auckland volcanic field event: Mathematical Geology, 37, 227–242, doi: 10.1007/s11004-005-1556-2
AN
Martí J, López C, Bartolini S, Becerril Land Geyer A. 2016. Stress Controls of Monogenetic Volcanism: A
Review. Front. Earth Sci. 4:106. doi: 10.3389/feart.2016.00106
M

Martí, J., Planagumà, L., Geyer, A., Canal, E., and Pedrazzi, D. 2011. Complex interaction between
ED

Strombolian and phreatomagmatic eruptions in the Quaternary monogenetic volcanism of the Catalan Volcanic
Zone (NE of Spain). J Volcanol Geoth Res, 201, 178–193.
PT

Martin del Pozzo, A.L., 1989. Geoquímica y paleomagnetismo de la Sierra Chichinautzin. PhD thesis,
Universidad Nacional Autónoma de México, DF, México, 148.
CE

Martin Del Pozzo, A.L., 1982, Monogenetic volcanism in Sierra Chichinautzin, Mexico: Bull Volcanol, 45, 9–24.
AC

Martin Del Pozzo, A.L., Córdoba, C., López, J., 1997. Volcanic impact on the southern basin of Mexico during
the Holocene. Quat. Int. 43r44, 181–190.

Márquez, A., Verma, S.P., Anguita, F., Brandle, J.L., y Oyarzum, R., 1999c, Tectonics and volcanism of Sierra
Chichinautzin– extension at the front of the central Trans-Mexican Volcanic Belt: J Volcanol Geoth Res, 93,
125-150.

40 39
Matchan, E., Phillips, D. 2011. "New Ar/ Ar ages for selected young (<1 Ma) basalt flows of the Newer
Volcanic Province, southeastern Australia." Quaternary Geochronology 6.3; 356-368.
ACCEPTED MANUSCRIPT

Mazzarini, F., Ferrari, L., Isola, I., 2010. Self-similar clustering of cinder cones and crust thickness in the
Michoacan-Guanajuato and Sierra de Chichinautzin volcanic fields, Trans-Mexican Volcanic Belt.
Tectonophysics, 486, 55–64.

40 39
McDougall, I., Harrison, T.M., 1999. Geochronology and Thermochronology by the Ar/ Ar method, 2nd ed.
Oxford University Press, New York. 269.
McDougall, I., Allsopp, H. L. and Chamalaun, F. H. 1966. Isotopic dating of the Newer Volcanics of Victoria,
Australia, and geomagnetic polarity epochs. Journal of Geophysical Research, 71, 6107 – 6118.

T
Molloy, C., Shane, P., and Augustinus, P. 2009. Eruption recurrence rates in a basaltic volcanic field based on

IP
tephra layers in maar sediments: Implications for hazards in the Auckland volcanic field. Geol Soc Am Bull,
121, 11-12, 1666-77.

CR
Mooser, F., Nairm, A.E.M., and Negendank, J.F.W., 1974. Paleomagnetic investigations of the Tertiary and
Quaternary igneous rocks: VIII, a paleomagnetic and petrologic study of volcanics of the Valley of Mexico:

US
Geologische Rundschau, 63, 451-483.

40 39
Moufti, M.R., Moghazi, A.M., Ali, K.A., 2013. Ar/ Ar geochronology of the Neogene-Quaternary Harrat Al-
AN
adinah intercontinental volcanic field, Saudi Arabia implications for duration and migration of volcanic
activity. J. Asian Earth Sci. 62, 253–268
M

Muirhead, J.D., Van Eaton, A.R., Re, G., White, J.D.L., Ort, M. 2016. Monogenetic volcanoes fed by
interconnected dikes and sills in the Hopi Buttes volcanic field, Navajo Nation, USA. Bull Volcanol., 78, 11.
ED

https://doi.org/10.1007/s00445-016-1005-8

Murcia, H., Nemeth, K., Moufti, M.R., Lindsay, J.M., El-Masry, N., Cronin, S.J., Qaddah, A. and Smith, I.E.M.,
PT

2014. Late Holocene lava flow morphotypes of northern Harrat Rahat, Kingdom of Saudi Arabia: Implications
for the description of continental lava fields. Journal of Asian Earth Sciences, 84: 131-145.
CE

Németh, K., Cronin, S.J., Charley, D., Harrison, M. and Garae, E., 2006. Exploding lakes in Vanuatu -
"Surtseyan-style" eruptions witnessed on Ambae Island. Episodes, 29, 2, 87-92.
AC

Németh, K., Goth, K., Martin, U., Csillag, G. and Suhr, P., 2008. Reconstructing paleoenvironment, eruption
mechanism and paleomorphology of the Pliocene Pula maar, (Hungary). J. Volcanol. Geotherm. Res., 177, 2,
441-456.
Németh, K., 2010, Monogenetic volcanic fields: Origin, sedimentary record, and relationship with polygenetic
volcanism, in Cañón-Tapia, E., and Szakács, A., eds., What Is a Volcano?: Geological Society of America
Special Paper 470, 43–66, doi: 10.1130/2010.2470(04).

Németh, K., and Kereszturi,G. 2015. Monogenetic volcanism: personal views and discussion. Int J Earth Sci
(Geol Rundsch). 104:2131–2146. doi: 10.1007/s00531-015-1243-6
ACCEPTED MANUSCRIPT

Nixon, G.T., 1982. The relationship between quaternary volcanism in Central Mexico and the seismicity and
structure of subducted ocean lithosphere. Geol. Soc. Am. Bull. 93, 514–523.

Norini, G., Gropelli, G., Lagmay, A.M.F., Capra, L., 2006, Recent left-oblique slip faulting in central Trans-
Mexican Volcanic Belt: Seismic hazard and geodynamic implications: Tectonics, 25.

Notsu, K., Fujitani, T., Ui, T., Matsuda, J., Ercan, T. 1995. Geochemical features of collision-related volcanic
rocks in central and eastern Anatolia, Turkey J. Volcanol. Geotherm. Res., 64,171-192

T
O'Sadnick, L.G., Reid, M.R., Cline, M.L., Cosca, M.A., Kuscu, G. 2013. Linear diffusion model dating of cinder

IP
cones in Central Anatolia, Turkey AGU Fall Meeting, 9-13 December 2013, San Francisco, USA, Abstracts.

CR
Ortega-Guerrero, B., Lozano-García, S., Herrera-Hernández, D., Caballero, M., Beramendi-Orosco, L., Bernal,
J.P., Torres-Rodríguez, E. and Avendaño-Villeda, D. 2017. Lithostratigraphy and physical properties of
lacustrine sediments of the last ca. 150 kyr from Chalco basin, central México. Journal of South American

US
Earth Sciences, 79, 507-524

Pasquare, G., Poli, S.. Vezzoli, L. and Zanchi, A., 1988. Continental arc volcanism and tectonic setting in
AN
Central Anatolia, Turkey. In: F.-C. Wezel (Editor), The Origin and Evolution of Arcs. Tectonophysics. 146: 217-
230.
M

Pardo, M., Suarez, G., 1995. Shape of the subducted Rivera and Cocos plates in southern Mexico: seismic
and tectonic implications. J. Geophys. Res. 100, B7, 12357–12373.
ED

Pérez-Campos, X., Kim, Y., Husker, A., Davis, P.M., Clayton, R.W., Iglesias, A., Pacheco, J.F., Singh, S.K.,
Manea, V.C., Gurnis, M., 2008. Horizontal subduction and truncation of the Cocos Plate beneath central
PT

Mexico. Geophys. Res. Lett. 35, L18303.


CE

Price, R.C., Gray, C.M. and Frey, F.A. 1997. Strontium isotopic and trace element heterogeneity in the plains
basalts of the Newer Volcanic Province, Victoria, Australia. Geochimica et Cosmochimica Acta 61, 171 – 192.
AC

Re, G., White, J.D.L., Ort, M.H. 2015. Dikes, sills, and stress-regime evolution during emplacement of the
Jagged Rocks Complex, Hopi Buttes Volcanic Field, Navajo Nation, USA. J Volcanol Geoth Res, 295, 65–79.

40
Renne, P.R., Mundil, R., Balco, G., Min, K., Ludwig, K., 2010, Joint determination of K decay constants and
40 40 40 39
the Ar*/ K for the Fish Canyon sanidine standard and improved accuracy for Ar/ Ar geochronology:
Geochim Cosmochim Acta, 74, 5349–5367.

Renne, P. R., Deino, A. L., Walter, R. C., Turrin, B. D., Swisher, C. C., Becker, T. A., Curtis, G.H., Sharp. W.D.,
and Jaouni, A. R., 1994, Intercalibration of astronomical and radioisotopic time. Geology, 22, 9, 783-786.
ACCEPTED MANUSCRIPT

Schmincke, H.-U. & Mertes, H., 1979. Pliocene and Quaternary volcanic phases in the Eifel volcanic fields. -
Naturwiss. 66, 614-615.

Schmincke, H.U., Lorenz, V. et al. 1983. The Quaternary Eifel Volcanic Fields. Springer, Berlin.

Schmincke, H.-U., 2007. The Quaternary volcanic fields of the East and West Eifel (Germany). Mantle Plumes:
A Multidisciplinary Approach, pp. 241–322.

14
Siebe, C., Sergio, S., Arana-Salinas, L., Macías, J.L., Gardner, J., Bonasia, R. 2017. The ~23,500y C BP

T
White Pumice Plinian eruption and associated debris avalanche and Tochimilco lava flow of Popocatépetl

IP
volcano, México. J Volcanol Geoth Res. 10.1016/j.jvolgeores.2017.01.011.

CR
Siebe, C., 2000, Age and archaelogical implications of Xitle volcano, southwestern Basin of Mexico City: J
Volcanol Geoth Res, 104, 45–64.

US
Siebe, C., Arana-Salinas, L., Abrams, M., 2005, Geology and radiocarbon ages of Tláloc, Tlacotenco,
Cuauhtzin, Hijo del Cuauhtzin, Teuhtli, and Ocusacayo monogenetic volcanoes in the central part of the Sierra
Chichinautzin, México: J Volcanol Geoth Res, 141, 225–243.
AN
Siebe, C., Macías, J.L., 2004, Volcanic hazards in the Mexico City metropolitan area from eruptions at
Popocatépetl, Nevado de Toluca, and Jocotitlán stratovolcanoes and monogenetic scoria cones in the Sierra
M

Chichinautzin Volcanic Field: Field Guide, Penrose Conference, Neogene-Quaternary Continental Margin
Volcanism, January 2004, State of Puebla, México: Boulder, Colorado, Geological Society of America, 77.
ED

Smith, I.E.M. and Németh, K. 2017. Source to surface model of monogenetic volcanism: a critical review.
From: Németh,K., Carrasco-Núñez, G., Aranda-Gómez, J. J. & Smith, I. E. M. (eds) Monogenetic Volcanism.
PT

Geological Society, London, Special Publications, 446, https://doi.org/10.1144/SP446.1


CE

Sohn, Y.K., Cronin, S.J., Brenna, M., Smith, I.E.M., Nemeth, K., White, J.D.L., Murtagh, R.M., Jeon, Y.M. and
Kwon, C.W., 2012. Ilchulbong tuff cone, Jeju Island, Korea, revisited: A compound monogenetic volcano
involving multiple magma pulses, shifting vents, and discrete eruptive phases. GSA Bull., 124, 3-4, 259-274.
AC

Sohn, Y.K., Park, K.H. and Yoon, S.-H., 2008. Primary versus secondary and subaerial versus submarine
hydrovolcanic deposits in the subsurface of Jeju Island, Korea. Sedimentology, 55, 4, 899-924

Straub, S. M., LaGatta, A. B., Martin-Del Pozzo, A. L., and Langmuir, C. H. 2008. Evidence from high-Ni
olivines for a hybridized peridotite/pyroxenite source for orogenic andesites from the central Mexican Volcanic
Belt, Geochem. Geophys. Geosyst., 9, Q03007, doi:10.1029/2007GC001583.

Tanaka, K.L., Shoemaker, E.M., Ulrich, G.E. and Wolfe, E.W. 1986. Migration of Volcanism in the San-
Francisco Volcanic Field, Arizona. Geological Society of America Bulletin, 97, 129–141.
ACCEPTED MANUSCRIPT

Tsuya, H. 1939. On the form and structure of volcanic bombs, with special reference to the origin of the
basaltic bombs from volcano Huzi (Fuji). Bull Earthq Res Inst Univ Tokyo 17:809–825

Urrutia-Fucugauchi, J., 1995. Constraints on Brunhes low-latitud paleosecular variation-Iztaccíhuatl


stratovolcano, basin of Mexico., Geofísica Internacional, 34, 3, 253-262.

Urrutia-Fucugauchi, J., Martin Del Pozzo, A.L., 1993. Implicaciones de los datos paleomagnéticos sobre la
edad de la Sierra de Chichinautzin, cuenca de México. Geofis Int, 32, 3, 523-533.

T
Uto, K., and Koyaguchi, T., 1987, K-Ar Ages of Three Alkaline Basalts from the Abu Monogenetic Volcano

IP
Group, Southwest Japan: Bulletin of the Volcanological Society of Japan, v. 32, p. 263-267

CR
Van den Hove, J, Grose, L., Betts, P.G., Ailleres,L., Van Otterloo, J., Cas, R.A.F. 2017. Spatial analysis of an
intra-plate basaltic volcanic field in a compressional tectonic setting: South-eastern Australia. J Volcanol Geoth
Res, 335, 35–53

US
Walker, G.P.L. 1969. The breaking of magma. Geol Mag 106:166–173
AN
Wallace, P.J., Carmichael, I.S.E., 1999, Quaternary volcanism near t e Valley of Mexico: Implications for
subduction zone magmatism and the effects of crustal thickness variations on primitive magma compositions:
Contrib Mineral Petrol 135, 291–314.
M

Wellman, P., McDougall, I., 1974. Cenozoic igneous activity in eastern Australia. Tectonophysics 23, 1–2, 49–
ED

65.

White, J.D.L. and Ross, P.-S., 2011. Maar-diatreme volcanoes; a review. J Volcanol Geoth Res,, 201(1-4): 1-
PT

29.
CE

White, J.D.L., 1991, Maar-diatreme phreatomagmatism at Hopi Buttes, Navajo Nation (Arizona), USA: Bull
Volcanol, 53, 239–258, doi: 10.1007/BF00414522.
AC

40 39
Wijbrans, J., Németh, K., Martin, U. and Balogh, K., 2007. Ar/ Ar geochronology of Neogene
phreatomagmatic volcanism in the western Pannonian Basin, Hungary. J Volcanol Geoth Res, 164, 4, 193-
204.

Williams, W. J. W., 1999, Evolution of Quaternary intrplate mafic lavas detailed using 3He surface exposure
and 40Ar/39Ar dating, and elemental and He, Sr, Nd, and Pb isotopic signatures; Potrillo volcanic field, New
Mexico, U.S.A. and San Quintin volcanic field, Baja California Norte, Mexico: El paso, Texas, University of
Texas at El Paso, PhD Thesis,195 p.

Wood, C.A., 1980, Morphometric evolution of cinder cones: J Volcanol Geoth Res, 7, 387–413
ACCEPTED MANUSCRIPT

Yamamoto, J., Mota, R., 1988, La secuencia de temblores del Valle de Toluca, México, de agosto 1980:
Geofísica Internacional, 27, 279–298.

York, D., Hall, C.M., Yanase, Y., Hanes, J.A. & Kenyon, W.J., 1981, 40Ar/39Ar dating of terrestrial minerals
with a continuous laser, Geophys. Res. Lett., 8, 1136-1138.

T
IP
CR
US
AN
M
ED
PT
CE
AC
ACCEPTED MANUSCRIPT

Highligths
Monogenetic volcanism started in the northern part of the Sierra Chichinautzin.
The Sierra Chichinautzin is formed by four different monogenetic volcanic fields.
Each monogenetic field shows distinct temporal and chemical characteristics.
Chemical differences could be related to changes in the stress field.
The rocks from the four volcanic fields are associated to different magma source.

T
IP
CR
US
AN
M
ED
PT
CE
AC

You might also like