Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

Geothermics 96 (2021) 102162

Contents lists available at ScienceDirect

Geothermics
journal homepage: www.elsevier.com/locate/geothermics

An approach to reconstruct the thermal history in active magmatic systems:


Implications for the Los Humeros volcanic complex, Mexico
Paromita Deb a, *, Guido Giordano b, Xiangyun Shi a, Federico Lucci b, Christoph Clauser a
a
Institute for Applied Geophysics and Geothermal Energy, RWTH Aachen University, Aachen, Germany
b
Dipartimento di Scienze, Università Roma Tre, Roma, Italy

A B S T R A C T

Reconstructing the thermal history in active volcanic complexes characterized by multiple magmatic events is challenging due to the limited knowledge of the nature
and extent of the transient heat sources. Although understanding of the geometry and architecture of a magmatic system is of prime importance for accurate
temperature assessments, it is still one of the most uncertain parameters in numerical models. In this work, we presented a methodology for thermal assessment in
active volcanic systems, whereby field-based geological, geochemical and petrological data are integrated to define the transient heat sources of a magma plumbing
system. This time-varying heat source conceptual model is applied in the Los Humeros Volcanic Complex, an active Quaternary caldera complex in the Trans Mexican
Volcanic Belt, for evaluating the thermal footprint related to the major volcanic events. The site is characterized by two caldera-forming eruptions, the Los Humeros
(164 000 years ago) and the Los Potreros (69 000 years ago) and numerous episodes of post-caldera bi-modal volcanism during Holocene period (8 000 – 3 000 years
old). The transient nature of the heat sources is implemented as time-varying temperature boundary conditions and the complete temporal evolution for a period of
182 000 years is simulated in 13 modeling stages. The thermal impact due to the voluminous caldera-forming events and the later short-lived magma pockets of
Holocene ages is simulated by emplacing heat sources in the numerical model distributed heterogeneously in space and active at different instants of time. The depth,
volume and age of the magma pockets are constrained from geochemical, petrological, geochronological and thermo-barometric analysis of erupted material. The
present temperature state obtained from this approach agrees well with the temperature data recorded in the geothermal wells. The thermal footprint of the in­
dividual volcanic events indicates that almost 80 % of the present-day thermal contribution results from the massive caldera-forming events. The post-caldera
Holocene magma pockets had additionally increased temperatures locally by 10 % - 20 % depending on the volumes and ages of the magma pockets. The
present-day thermal regime of the younger Holocene magma pockets suggests existence of super-hot resources at shallow depths in the southern part of the
geothermal field, making it a potential site for future exploration activities.

1. Introduction Miyakejima volcano in Japan (Ushioda et al., 2018), Eyjafjallajökull in


Iceland (Tarasewicz et al., 2012), and Anak Krakatau volcano in
In recent years, progress in geophysical imaging techniques, igneous Indonesia (Dahren et al., 2012) to explain the observed petrological,
petrology, geochemistry and numerical modeling have significantly geochemical and seismicity data. Caldera-forming eruptions are usually
advanced the understanding of the morphology of magmatic plumbing the result of a depletion of the shallowest level of such a complex
systems. In fact, the classical concept of a melt-dominated, single large magmatic arrangement (Ushioda et al., 2018). In fact, for some of the
magma chamber in the upper crust is increasingly questioned as it fails largest eruptions on earth such as the Snake River Plain, US and the
to explain all physical and petrological processes of magma accumula­ Taupo Volcanic Zone, New Zealand, a spectrum of processes including
tion and differentiation, and its interaction with crustal rocks (Cashman the extensive depletion of multiple layers embedded in crystal mushes is
& Giordano, 2014; Cashman et al., 2017; Magee et al., 2018). The most considered to have taken place (Ellis et al., 2010; Brown et al., 1998;
recent views of magmatic systems consider the existence of Cashman & Giordano, 2014). Caldera structures are, therefore, in­
vertically-extended arrays of magma accumulation zones or reservoirs at dications of shallow crustal heat sources (Gudmundsson, 2008; Bach­
different depths in the crust, periodically connected by feeders of mann & Bergantz, 2008) and hence present an enormous potential for
different magma flow rates (Annen et al., 2006; Caricchi & Blundy, generating geothermal energy from high-enthalpy resources within
2015; Cashman et al., 2017; Edmonds et al., 2019). This magma drillable depths.
plumbing system concept has also been suggested in numerous volcanic Large silicic calderas often host smaller nested craters related to post-
complexes such as the Campi Flegrei in Italy (Marianelli et al., 2006), caldera activity (e.g. Orsi et al., 2005; Cole et al. 2005; Wohletz &

* Corresponding author.
E-mail address: pdeb@eonerc.rwth-aachen.de (P. Deb).

https://doi.org/10.1016/j.geothermics.2021.102162
Received 16 January 2021; Received in revised form 25 May 2021; Accepted 30 May 2021
Available online 23 June 2021
0375-6505/© 2021 Elsevier Ltd. All rights reserved.
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 1. Relief map of the LHVC showing the caldera rims of Los Humeros (red dashes), Los Potreros (black dash), fault Maztaloya (NW-SE trending yellow dashes)
and fault Las Papas (E-W trending yellow dashes), geothermal wells (black dots), nearby cities (red dots) and model boundary (blue dash).

Heiken, 1992). Active volcanic complexes, subjected to multiple Humeros Volcanic Complex (LHVC), which hosts the third largest
caldera-forming and post-caldera events, are affected by numerous ep­ geothermal field of Mexico. Extensive field campaigns and geological
isodes of magma regeneration, differentiation, eruption and cooling (e. studies during the last decades revealed the nested caldera structure in
g., Claiborne et al., 2015; Lipman, 2007; Selva et al. 2019). For accurate the LHVC, formed by two major caldera-forming events – the Los
assessment of available geothermal resources, it is necessary to recon­ Humeros and the Los Potreros events and later episodes of post-caldera
struct the thermal history related to each volcanic episode (Di Renzo bimodal volcanism (Ferriz & Mahood, 1984; Carrasco-Núñez et al.,
et al., 2016). Such a methodology requires a plethora of information 2017a; 2017b; 2018; Norini et al., 2015). However, the structure of the
related to the architecture of the magmatic system, age, temperature, magmatic system under the caldera complex is still a subject of active
depth and volume of emplacement, in addition to the properties of the research. This work, therefore, represents a first order approach to
host rock formation (heat production rate, thermal conductivity, improve our understanding of the complex evolution of the magmatic
porosity, etc.) (Stimac et al., 2001; Wohletz & Heiken, 1992). heat sources in Los Humeros.
Geochemical and petrological investigations of the erupted lava Till date, the numerical models of LHVC assume the heat source to be
samples, combined with thermo-barometric models provide information a single, melt-dominated, chemically stratified magma chamber located
on the parental magma composition, depth and initial temperature of about 5 km under the caldera complex with a volume of up to 1 400 km3
the magma bodies, while geochronological studies of the erupted (Verma et al., 1990; 2011; Verma & Andaverde, 2007; Verma &
products provide insights regarding the age and residence period of the Gómez-Arias, 2013). However, recent volcanological, geochemical and
magma in the crust (Feng and Zhu, 2018; Lucci et al., 2020; Ushioda petrological studies based on thermo-barometric models (Lucci et al.,
et al., 2018; Costa 2008). However, one of the largest uncertainties that 2020), field studies and analogue modeling (Urbani et al., 2020; 2021;
prevails in numerical modeling of volcanic systems is related to the Bonini et al.,2021) do not corroborate the current presence of a large
volume and depth of the magma chamber (Smith and Shaw, 1975). melt-dominated crustal magma chamber under the Los Humeros
Geophysical imaging of subsurface can potentially fill this gap by caldera. Some of these studies provide evidence for Holocene mono­
providing information regarding the geometry and extent of the magma genetic volcanic centers on the caldera floor, with compositions ranging
plumbing system (Magee et al., 2018). Methods such as electromagnetic, from olivine-basalts to trachyte and rhyolite. Their source depths range
gravity and passive seismicity are widely-used cost-effective mapping from deeper than 30 km (basalts) to less than 3 km (trachytes) with
tools in geothermal exploration to constraint subsurface structures, but uniquely evolved compositions and complex histories of ascent at
their applicability is usually limited to certain investigation depths different depths (Lucci et al., 2020). This rise of the different mafic and
beyond which the data resolution and quality are not sufficient for felsic magmas to the caldera floor, indicate that at present the Los
resolving complex geological structures (Domra Kana et al., 2015). In Humeros caldera is not underlain by a single melt-dominated magma
addition, the efficiency of any geophysical method depends on chamber volume, but rather a complex magma plumbing system char­
site-specific geology such as temperature, presence of fluid, shallow acterized by numerous magma pockets, each evolving in a closed system
magmatic bodies, etc., (Hersir et al., 2020; Domra Kana et al., 2015). and unaffected by recharge and mixing processes (Lucci et al., 2020).
Further discussion about available geophysics data specific to the case These studies propose an alternative model for the Los Humeros magma
study is presented in Section 2.3. plumbing system, characterized by multi-layered ephemeral magma
In this work, we investigate the thermal state of one of the active pockets rather than a single, large magma chamber (Cashman & Gior­
Quaternary calderas of the Trans Mexican Volcanic Belt, the Los dano, 2014; Gualda and Ghiorso, 2013).

2
P. Deb et al. Geothermics 96 (2021) 102162

We integrated the latest 3D geological model (Calcagno et al., 2018; more than 123 million tons of geothermal fluids (Arellano et al., 2015),
2020; 2021) and geochemical information (Carrasco-Núñez et al., 2018; it is Mexico’s third largest geothermal field.
Cavazos & Carrasco-Núñez, 2020; Lucci et al., 2020) to propose a con­
ceptual model for the evolution of the heat sources under the Los
Humeros caldera. This conceptual model is used for evaluating the 2.1. Geology
contribution of different heat sources to the current thermal state of the
LHVC. Based on the published results of Los Humeros and Los Potreros The heat source of the LHVC hydrothermal system is related to the
and new data from thermo-barometric analysis of Holocene lavas, we young magmatic complex, which is characterized by multiple volcanic
defined time-varying heat sources under the Los Humeros caldera, episodes, with eruptive products ranging from olivine-basalts to rhyo­
which represents the different magma bodies at different instants of lites (Ferriz & Mahood, 1984). The regional crystalline basement below
time. These heat sources are implemented in the numerical model as the LHVC consists of Late-Paleozoic to Mesozoic intrusive and meta­
time-varying boundary conditions using SHEMAT Suite (Clauser, 2003; morphic rocks about 250 million years old (Carrasco-Núñez et al.,
Keller et al., 2020) for reconstructing the thermal history of the LHVC. 2017b). The crystalline basement is partially covered by less than 3 km
This numerical model tests the hypothesis of multi-layered, localized thick Triassic-Cretaceous series of limestones and terrigenous sedimen­
and ephemeral magma pockets under the Los Humeros caldera to tary rocks (Norini et al., 2015; Norini et al., 2019) deformed during the
explain the present-day anomalous temperature distribution observed in Late Cretaceous-Palaeocene Laramide orogeny (Campos-Enríquez &
different parts of the geothermal field. Temperature data from the Garduño-Monroy, 1987; Fitz-Díaz et al., 2018; Suter, 1987; Roure et al.,
geothermal wells at Los Humeros document a highly variable temper­ 2009). As a result of 15 million years old Tertiary intrusions of grano­
ature distribution at short distances within the geothermal field. A dif­ diorite and syenite (Yañez & García,1982), the Cretaceous limestones
ference of more than 100 K in the bottom hole temperatures (BHTs) is are locally metamorphosed to marble, hornfels and skarn (Carra­
recorded at similar depths in wells within one kilometer distance. The sco-Núñez et al. 2017b; Norini et al. 2015).
influence of localized heat sources in the Los Humeros geothermal field The Cretaceous limestone basement is overlain by an andesite
is also reflected in the productivity characteristics of the geothermal sequence, emplaced during the Miocene - Pleistocene volcanic period
wells, which drastically changes between the northern and the southern lasting from 11 million years to 1.46 million years ago (Carrasco-Núñez
wells of the geothermal field (Aguilar et al., 2019; Arellano et al., 2015). et al., 2017a, 1997; Gómez-Tuena & Carrasco-Núñez, 2000;
The thermal modeling is performed for a period of about 182 000 López-Hernández, 1995). This forms the Pre-caldera volcanic group and
years representing the different volcanic phases of Los Humeros: pre- is mainly composed of (i) the Miocene Cuyoaco and Alseseca andesites
caldera, caldera and post-caldera. The pre-caldera phase is modeled (Yañez & García,1982) and (ii) the Plio-Pleistocene Teziutlan andesite
by performing a steady-state simulation and represents the initial con­ (Carrasco-Núñez et al., 2017b; Ferriz & Mahood, 1984; Yañez & García,
ditions in the model (Section 4.3). The caldera phase, represented by the 1982). This naturally fractured andesite sequence varies in thickness
Los Humeros and the Los Potreros caldera-forming events, are modelled between 900 m to more than 1 500 m, and is the primary geothermal
separately to evaluate their individual thermal impact on the present reservoir of the Los Humeros field (Carrasco-Núñez et al., 2017b;
thermal regime of the geothermal field (Sections 5.1 and 5.2). The Cedillo-Rodriguez, 2000; Norini et al., 2015).
localized and short-lived volcanic events of the post-caldera Holocene The next sequence of rocks observed in the wells belongs to the
resurgent phase is modeled by incorporating shallow magma storage Xaltipan and Zaragoza ignimbrite erupted during the caldera phase.
zones under the caldera complex (Section 5.3). The modeling parame­ These represent the Los Humeros and the Los Potreros caldera-forming
ters are constrained using the results of the published studies (Carra­ events, respectively. In a recent study, Carrasco-Núñez et al., (2018)
sco-Núñez et al., 2018; Cavazos & Carrasco-Núñez, 2020; Lucci et al., determined much younger ages for both these ignimbrite deposits by
2020). Temperature data from the geothermal wells in this region are U/Th and 40Ar/39Ar dating of zircon and plagioclase, respectively.
used as a reference for comparing the results of the simulated model Accordingly, the Xaltipan ignimbrite related to Los Humeros caldera
temperatures. It is important to emphasize that we do not intend to event is now estimated to be about 164 000 years old, while the Zar­
deterministically reproduce the observed temperature distribution in agoza ignimbrite related to Los Potreros event is of much younger age of
the LHVC. Rather, the approach is to test whether the modeling of approximately 69 000 years. The Los Humeros event resulted from a
multi-phase, transient magmatic events using a field-based conceptual voluminous eruption of 290 km3 (Dense Rock Equivalent) of Xaltipan
model of the magmatic heat source can capture the most prominent ignimbrite, containing rhyolitic magma with more than 70 % silica,
architectural features of the geothermal reservoir and explain the creating a caldera of dimension 21 km × 15 km (Fig. 1) (Cavazos &
anomalous temperature distribution observed in the geothermal wells. Carrasco-Núñez, 2020; Ferriz & Mahood, 1984). The Los Potreros
In Section 2, we will present an overview of the volcanological history caldera, on the other hand, which is sized 9 km × 10 km and nested
of the LHVC and briefly discuss the available information from temper­ within the Los Humeros caldera, resulted from the eruption of 10 km3 –
ature and geophysical sources. In Section 3, we discuss the latest findings 15 km3 of Zaragoza Tuff, a non-welded ignimbrite of rhyodacitic to
regarding the volumes and ages of the erupted events from geochemical andesitic composition (Carrasco-Núñez et al., 2012; Carrasco-Núñez &
and petrographic studies and their applications for constraining the pa­ Branney, 2005; Ferriz & Mahood, 1984)
rameters of the thermal model. In Section 4 we present the methodology The fourth and last stage is dated from about 50 000 years ago up to
and the numerical constraints for different stages of modeling. Results the present day, and includes an Upper Pleistocene resurgent phase,
and their discussion are presented in Sections 5 and 6, respectively. followed by the youngest and still active Holocene ring-fracture and
bimodal phase (Carrasco-Núñez et al., 2018; Norini et al., 2015, 2019).
2. LHVC background The latter involves the post-caldera stage of volcanism of the LHVC,
which is characterized by smaller explosive and effusive eruptions, with
Located around 260 km east of Mexico City and at an elevation of products ranging from basaltic to trachytic lava flows, fed by scattered
around 2 800 m a.s.l. (Fig. 1), the LHVC hosts an active, high-enthalpy monogenetic volcanic centers (Norini et al., 2015; Carrasco-Núñez et al.,
geothermal field with more than 50 geothermal wells and has been 2017a). Latest petrographic and field studies on Holocene lava samples
commercially exploited by the Comisión Federal de Electricidad (CFE) suggest that each magma pocket associated with post-caldera volcanic
for the last 30 years (Cedillo-Rodriguez, 2000; Lorenzo-Pulido, 2008). events can be related to scattered, small volume, short-lived (less than 1
With an operational capacity of 93.9 MW (IEA, 2018) and production of 000 years) heat sources under the Los Humeros caldera (Lucci et al.,
2020; Urbani et al., 2020). Details of the petrographic studies, including

3
P. Deb et al. Geothermics 96 (2021) 102162

Table 1
Summary of previous thermal modeling work in the LHVC and comparison of parameters with the present study (V: magma chamber volume; z: depth of magma
chamber top below surface; N: numerical time steps; FD: Finite Difference, FE: Finite Element, CV: Control Volume),* indicates the depth of the model domain where
the heat flow boundary conditions were applied.
Source Model type and objective/ Method Model V z N Mesh size Δx
duration × Δy × Δz
(years) (km3 ) (km) (m × m × m)

Prol Ledesma and 2D heat conduction/ FD - 100 5 100 500 × 500


Gonzalez-Moran, 1982
Campos-Enríquez and 2D steady-state heat conduction/ FE - 100 - - -
Duran, 1986
Verma et al., 1990 2D heat conduction and advection/ FD 500 000 1 500 5 100 500 × 500
Castillo-Roman et al., 1991 2D heat conduction; test of sensitivity of thermal regime to 600 000, 1 500 4,5,6 - -
magma chamber at different depths and different ages/ FD 100 000
Verma and Andaverde, 3D heat conduction and chemical modeling of magma chamber/ 460 000 1 500 5 Emplacement- 40, 250 × 250 ×
2007 CV Cooling- 1 000 250
Verma et al., 2011 3D heat conduction; magma chamber emplacement, test of 30 000 1 000 5– 250 250 × 250 ×
sensitivity of thermal regime to magma chamber volume and –1 400 10 250
position/ CV
Verma & Gómez-Arias, 3D heat conduction and advection/ CV 460 000 1 400 5 10 200 × 200 ×
2013 200
Deb et al., 2019a 3D steady-state heat conduction/advection FD - - 7.5* - 250 × 250 ×
50
Deb et al., 2019b 3D steady-state heat conduction/advection FD - - 6.5* - 50 × 50 × 50
This study 3D heat conduction/ FD 182 000 1 100 5 variable (see 100 × 100 ×
Appendix D.) 50

mineral chemistry and major-element bulk composition, are discussed in integrated to construct a 3D geological model (Calcagno et al., 2020),
Lucci et al. (2020). which is utilized in the current study.
In Los Humeros, MT and TEM have been acquired and 3D joint
2.2. Temperature data from geothermal wells inversion of the data have been performed for detection of deep re­
sistivity anomalies (Benediktsdóttir et al., 2019; Hersir et al., 2020). MT
The Los Humeros geothermal field has been exploited by more than data investigates the resistivity of the subsurface and therefore, are
50 geothermal wells since the first commercial production began in sensitive to magma or hydrothermal fluids. In Los Humeros, however,
1990 (Arellano et al., 2015). Most of the wells in the field produce a the MT data does not provide any conclusive evidence of deep-seated
high-enthalpy (more than 2 500 kJ/kg) geothermal fluid with a steam low resistivity anomalies. This is partly due to the limited resolution
fraction greater than 90 % at separation conditions (Arellano et al., of the MT soundings below 3 km depth (Hersir et al., 2020). Another
2015). While super-hot conditions were encountered by many wells in reason of the low resolution of MT data is because Los Humeros exhibits
this area, the production is limited to only a few wells due to insufficient a low-resistivity (smectite) cap at a very shallow depth of around 2 600
permeability in the other wells (Aguilar et al., 2019; Arellano et al., m.a.s.l along NNW-SSE, the main tectonic trend in this area (Bene­
2003). Permeability in Los Humeros is secondary in nature and is, diktsdóttir et al., 2019). Such low resistivity caps are formed in locations
therefore, localized and restricted to damaged zones surrounding faults characterized by high temperature and permeability and has also been
and fractures (Cedillo-Rodriguez, 2000; Lorenzo-Pulido, 2008). In fact, observed in other high temperature geothermal areas such as Krafla (Lee
the low permeability of the field and inadequate natural recharge are the et al., 2019). Presence of such conductive caps at a shallow level reflects
primary reasons for the pressure depletion, resulting in boiling and a large portion of the energy of the electromagnetic waves, leaving little
steam gain (Arellano et al., 2015). These processes have also affected the energy to penetrate deeper, thereby reducing the efficiency of the MT
temperature and pressure data acquired in the geothermal wells of Los data to resolve deeper anomalies (Hersir et al., 2020). Nevertheless, the
Humeros over time. orientation of the smectite cap in the direction of NNW-SSE trend sug­
To establish the initial temperature conditions from well data, we gests presence of heat sources aligned in this direction.
analyzed the temperature and pressure data from more than 50 Interestingly, gravity data of Los Humeros indicates areas of high-
geothermal wells provided by CFE. After reviewing the quality of the density anomalies at different depths, also aligned in the direction of
available temperature data, we selected a subset of 32 wells for cor­ the main tectonic NNW-SSE trend (Cornejo et al., 2019). The density
recting the transient bottom-hole temperature data series using Horner’s distribution map at the depth of the geothermal reservoir (1 800 m
method (Dowdle & Cobb, 1975) to obtain static temperatures at bottom below surface) is characterized by patches of high-density anomalies in
hole depth in each well. Details regarding the correction, input data and the direction of NW-SE, while low density anomalies orient in the
corrected temperature are provided in Table A. in Appendix A. NE-SW to E-W direction. Further down, at 2 800m and at 3 800 m below
surface, the low-density patches combine to form a single structure
2.3. Geophysical constraints aligned in the NW-SE direction. The location and the depth of the small
patches of high-density anomalies are consistent with the proposed
Within the framework of H2020 project GEMex, different types of small, ephemeral magma pockets in the shallow crust, acting as heat
geophysical data have been acquired which include resistivity (magne­ sources for the Los Humeros geothermal system. However, as empha­
totelluric (MT) and transient electromagnetics (TEM)), passive seis­ sized in Cornejo et al., 2019, the 3D inversion results should be inter­
micity and gravity (Hersir et al., 2020). Apart from the newly acquired preted with caution due to coarse measurement points away from the
data, CFE also provided four active 2D seismic lines and additional central area and limited vertical resolution at deeper depths. The density
passive seismic data. Results of the geophysical surveys, well data, distribution maps from 3-D inversion of the Los Humeros area by Cor­
geochemistry, analogue modeling and surface geological map were nejo et al., 2019 is presented in Fig. B in the Appendix section.

4
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 2. Conceptual heat source model used for


reconstructing the thermal history of the LHVC:
(a) Los Humeros magma chamber at about 164
000 years ago before the Los Humeros event (b)
Los Potreros magma chamber at about 69 000
years ago before the Los Potreros event, (c)
Post-caldera Holocene magma pockets with
approximate ages between 8 000 years and 3
000 years (red – 8 000 years, blue – 7 000 years,
green (not visible here) – 3 800 years and yel­
low – 2 800 years). Black vertical lines are the
geothermal wells in the Los Humeros field. The
brown and green surfaces are the interfaces
between the Miocene-Pliocene andesite and the
Cretaceous limestone and between the Creta­
ceous limestones and the granitic basement,
respectively. Vertical depth is relative to mean
sea level (0 m), and the green arrow indicates
the north direction. The magma bodies shown
in figure do not represent separate geological
entities, but rather the volumes of the numeri­
cal model to which the time-varying tempera­
ture boundary conditions are applied.

3. Development of a conceptual model for heat source under at geothermal reservoir depth were significantly underestimated
LHVC (mismatch of 50 ◦ C and 100 ◦ C) as compared to the temperatures
measured in the wells (Verma & Gómez-Arias, 2013). Similarly,
Several numerical studies have been performed over the past three steady-state 3D conductive/ convective thermal models reported in Deb
decades that provide important information regarding the evolution of et al. (2019a; 2019b) could not reproduce the anomalously high tem­
the Los Humeros Volcanic Complex (LVHC) (Table 1). Early attempts on peratures observed in geothermal wells close to large regional faults. In
thermal modeling were made in 2D (Prol Ledesma and Gonzalez-Moran, these models, heat flow boundary conditions were applied at the bottom
1982; Campos-Enriquez and Duran, 1987; Verma et al., 1990; Cas­ of the model domain, and convective scenarios with different perme­
tillo-Roman et al., 1991), while several later thermal and chemical ability values for faults and rock formations were investigated.
models are in 3D (Verma & Andaverde, 2007; Verma et al., 2011; Verma In this study, we integrated the latest geochronological, geochemical
& Gómez-Arias, 2013). These numerical studies considered the Los and petrographic findings, proposing an alternative configuration for
Humeros volcanic event to be approximately 500 000 years old based on the heat source under the Los Humeros caldera. Our conceptual model
the old K-Ar dating of the Xaltipan ignimbrites (Ferriz & Mahood, 1984). hypothesizes a magma plumbing system under the LHVC, characterized
The magma chamber volume considered in the numerical studies was by multiple magma reservoirs or storage zones at different depths in the
approximated from the volume of the erupted ignimbrite. In their early crust active at different instants of time (Fig. 2). Despite being exten­
field work, Del Rio (1982) and Ferriz and Mahood (1984, 1987) esti­ sively studied petrologically and geophysically, the volumes of the
mated the volume of the erupted Xaltipan ignimbrite at 100 km3 to 115 magma chambers responsible for the Los Humeros and the Los Potreros
km3, respectively. Simultaneously, Verma (1984) and Verma (1985b), eruptions remain a subject of critical debate. The available literature
based on their fractional crystallization models proposed a magma reports a spectrum of possibilities for the Los Humeros magma chamber
chamber volume in the order of 1 400 km3 for Los Humeros and around volumes, from a study which speculates 2 900 km3 (Cavazos & Carra­
100 km3 for Los Potreros. These volumes were supported by the con­ sco-Núñez, 2020, based on the new estimation of 290 km3 Dense-Rock
ventional 1:10 ratio of erupted to intruded volumes (Smith et al., 1978; Equivalent (DRE) for Xaltipan Ignimbrite) to other studies which sug­
Smith & Shaw, 1975), based on which Verma et al. (2011) and Verma gest a range between 1 400 km3 – 1 500 km3 (Carrasco-Núñez et al.,
and Gómez-Arias (2013) conducted their numerical investigations. They 2018; Verma et al., 1990; Verma et al., 2011) based on the old, erupted
assumed the volume of the magma chamber to be approximately be­ DRE estimates of 115 km3.
tween 1 000 km3 - 1 500 km3 with an initial emplacement temperature To obtain reasonable constraints for the volumes of the magma
of 1350 ◦ C in their numerical models. In spite of the unrealistically high chamber, we performed an analysis of magma intrusion rates (km3/ yr)
initial emplacement temperature of 1350 ◦ C assumed for a single volu­ using the numerical modeling results of Annen (2009) (refer Figure 6 in
minous stratified magma chamber encompassing from basaltic to Annen, 2009). Annen (2009) specified conditions required for devel­
rhyolite compositions, the temperature estimates from their simulations oping eruptible magma (crystal fraction less than 40 %, Lejeune and

5
P. Deb et al. Geothermics 96 (2021) 102162

Table 2
Properties of magma bodies used in this study. V: intrusive volumes; T: temperature of magma reservoirs; c: volumetric heat capacity; h: specific enthalpy of crys­
tallization (Sources: Carrasco-Núñez et al., 2017b; 2018; Carrasco-Núñez & Branney, 2005; Cavazos & Carrasco-Núñez, 2020; Lesher & Spera, 2015; Liu & Lowell,
2011; Lucci et al., 2020; Verma, 1985b, Urbani et al., 2020)
Magma Description Age Groups Top V T c h
bodies Depth
(years) (km) (km3 ) (∘ C) (MJ m− 3
K− 1 ) (kJ kg− 1 )

Los Humeros Rhyolitic magma 164 000 - 5 1 000 850 3.28 270
Los Potreros Rhyodacitic to andesitic 69 000 - 5 100 900 2.33 300
LH26 Trachyandesite, intermediate lavas from Los Potreros caldera 8 000 G1 2.5 0.10 1 050 2.40 300
LH27-2 Trachyandesite, Maxtaloya andesites from the rim walls of Xalapasco 8 000 G1 3 1.30 1 050 2.40 300
crater
LH4 Trachyandesite, San Antonio Las Chapas lavas 8 000 G1 3 0.40 1 050 2.40 300
LH13 Trachyandesite, intermediate lavas from Los Potreros caldera 7 000 G2 1 0.01 1 050 2.40 300
LH11 Rhyolite, Obsidian dome, at Los Humeros fault 7 000 G2 1 0.03 800 3.28 270
LH21-2 Trachyandesite free of Orthopyroxene, Sarabia lava flow from extra- 7 000 G2 6 1.30 1 050 2.40 300
caldera
LH15 Trachyandesite free of Orthopyroxene, El Limon lava flow from extra- 7 000 G2 6 1.30 1 050 2.40 300
caldera
LH27-1 Olivine-Basalt, mafic lavas inside the Xalapasco crater 3 800 G3 2.5 0.05 1 100 3.78 420
LH5-2 Olivine-Basalt, mafic lavas from Los Humeros town 3 800 G3 1.5 0.30 1 100 3.78 420
LH5-1 Trachyte, Chocomiapa-Los Parajes felsic lavas 3 800 G3 0.5 0.01 950 2.40 300
LH6-1 Trachyte, Felsic lavas from El Pajaro unit 2 800 G4 3 0.15 950 2.40 300
LH17 Orthopyroxene trachyandesite, Tepeyahualco lava flow from extra-caldera 2 800 G4 3 1.30 1 050 2.40 300

Richet, 1995) in large magma chambers of diameter 10 km and 20 km about 69 000 years ago.
(representative of Los Humeros and Los Potreros calderas) for variable In the models, the top of the Los Humeros (Xaltipan) and the Los
magma intrusion rates. Note that the magma intrusion rate used in this Potreros (Zaragoza) magma reservoirs are assumed to be at 5 km. This
study is calculated from the ratio of erupted volume to residence time assumption is based on geobarometric data provided in Ferriz and
and is not sensu stricto as it considers only the eruptible magma (Costa, Mahood (1987) and in agreement with the general considerations about
2008). For the Xaltipan ignimbrite, we obtained a very high volumetric the depth of crustal magma chamber tops in silicic calderas (Bachmann
magma intrusion rate of 0.06 km3/yr corresponding to its very short and Bergantz, 2008; Wohletz & Heiken, 1992; Verma et al., 1990). In
residence time of 5 000 years estimated from magmatic zircon U-Th ages Table 2, we present the list of parameters used for constraining different
(Carrasco-Núñez et al., 2018) and the recently updated DRE volume of heat sources under the Los Humeros caldera.
290 km3 (Cavazos & Carrasco-Núñez, 2020). This intrusion rate is Fig. 2c shows the conceptual heat source model for the last 8 000
coherent with the magma flux estimates for large eruptions obtained years. During this time, the Los Potreros caldera floor was subjected to
from the analysis of Zircon ages (Caricchi et al., 2014), which suggested several small magma eruptions, ranging from olivine-basalts to tra­
that magmatic reservoirs feeding large volcanic eruptions might have chytes (Carrasco-Núñez, et al., 2017a), derived from variable depths as
formed over a restricted time with high magma intrusion rates. Thus, constrained by geobarometric data (Lucci et al., 2020). The eruption of
assuming the calculated magma intrusion rate as a proxy for the average the denser and increasingly mafic magma during the post-caldera stage
magma flux required to build the Los Humeros magma chamber, the represented the absence of a large magma chamber capable of receiving
maximum eruptible portion in the magma chamber could be as high as and homogenizing recharge magmas (Ferriz and Mahood, 1984, 1987).
65 % (Annen, 2009). This leads to a minimum total volume of 450 km3 Classical thermo-barometric modeling applied to the post-caldera Ho­
for the chamber (including the crystallized part) if the eruption emptied locene lava samples by Lucci et al. (2020) (also refer to Appendix C.)
the entire 65 % fraction of eruptible magma. However, previous esti­ provided estimates of the mean pre-eruptive pressure-temperature
mates of the amount of erupted material have been calculated at about conditions of the magma pockets feeding the Holocene eruptive activity.
1/3 to 1/10 of the partly molten magma reservoir below calderas larger In consideration of their smaller erupted volumes, the intrusive volumes
than 10 km (Smith and Shaw, 1975; Carrasco-Núñez et al., 2018). Based have been assumed to be twice the erupted volume to account for
on this reasoning, we assumed conservatively an estimate of 1000 km3 magmas left within the feeding system. The integration of these results
for the Los Humeros magma chamber considering the lower end of the combined with (i) the updated Los Humeros geological map (Carra­
ratio proposed by Smith and Shaw (1975) i.e., three times the erupted sco-Núñez, et al., 2017a), (ii) the intra-caldera subsurface geology
volume of Xaltipan Ignimbrite. Fig. 2a represents the heat source con­ (Carrasco-Núñez et al., 2017b; Cavazos & Carrasco-Núñez, 2020; Cav­
dition under LHVC, right before the eruption of the Xaltipan Ignimbrites azos-Alvarez et al., 2020) (iii) the fieldwork and the analogue modeling
and collapse of the Los Humeros caldera (164 000 years ago). by Urbani et al. (2020; 2021), allowed first-order estimates of the top
The Los Potreros caldera-forming event occurred almost 100 000 depths, volumes and temperatures of mafic to felsic magma batches
years after the Los Humeros episode, resulting in an eruption of almost (Table 2; Table C.1 in Appendix section). The geometry of the magma
10 km3 – 15 km3 of Zaragoza ignimbrite (Carrasco-Núñez et al., 2018; batches has been conservatively assumed as axisymmetric and with a
Ferriz & Mahood, 1984), less than 5 % of the amount released during the high aspect ratio, resulting in a discoidal shape. This is in agreement
previous eruption. Assuming the residence period of Zaragoza ignim­ with the existing literature concerning bore-hole stratigraphies and the
brites to be in the same range as Xaltipan (5 000 years), we obtained a viscous rheology of erupted magmas (Carrasco-Núñez et al., 2017a), and
magma intrusion rate of 0.003 km3/ yr. At this rate, the eruptible with the deformation induced along the caldera floor (Urbani et al.
portion of magma in a chamber of diameter 10 km is approximately 2020; 2021). These parameters are assumed as a data-supported, first
around 10 % (refer Figure 6 in Annen, 2009). Furthermore, based on the order scenario for the magmatic configuration of the shallower portion
fractional crystallization model, Verma (1985b) and Verma et al., of the LHVC post-caldera stage magmatic plumbing system.
(1990) reported the amount of minerals left behind after Los Potreros Again, using a similar approach, as for the Los Humeros and the Los
eruption to be around 84 km3. Therefore, the total volume of the Los Potreros magma chambers, we computed the volumetric magma flow
Potreros magma storage zone was assumed to be around 100 km3. rates required for emplacing the small magma pockets. Following Pyle
Fig. 2b shows the configuration of the Los Potreros magma chamber at (1992), Cathles et al., (1997), and Nabelek et al., (2012), we assumed

6
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 3. 3D geological model of the LHVC showing the present day main lithostratigraphic units (left) (Calcagno et al., 2020); cross-section showing the present-day
structures and the boundary conditions applied at the top and bottom of the numerical model. Note that the top and bottom boundary conditions remain unchanged
during all the transient simulation stages. The time-varying boundary conditions are applied in the magma bodies shown in Fig. 2.

Table 3
Broad lithostratigraphic groups of the Los Humeros Volcanic Complex (LHVC) and the average petrophysical properties (porosity ϕ, thermal conductivity λ_m,
volumetric heat capacity c and heat production rate H for each group) (Sources: Calcagno et al., 2020; Weydt et al., 2020; Bär & Weydt, 2019; Rybach, 1976; Schön, 2015)
Lithostratigraphic group Dominant rock type ϕ λm c H
% W m− 1
K− 1
MJ m− 3
K− 1
μW m− 3

Post-caldera Andesitic and basaltic lava flows, tuff, pumice 18.59 1.41 1.68 0.31
Caldera Ignimbrites and Tuff 35.04 2.03 1.21 1.48
Pre-caldera Andesites (Teziutlian, Cuyoaco) 6.66 1.02 1.03 1.08
Basement Limestone, Marble 1.76 2.73 2.14 0.62
Granite, Granodiorite 2.00 2.73 1.96 2.45

their residence period in the crust before eruption to be a maximum of 1 conductivity and volumetric heat capacity) for different lithological
000 years. This assumption is reasonable in light of similar volcanic groups are calculated using laboratory measurements on different
events (Isaia et al., 2019). For example, magma chambers of the East outcrop samples from Los Humeros area (Bär & Weydt, 2019; Weydt
Pacific Rise are considered to have residence times of 1 000 – 10 000 et al., 2020). As each lithological group is composed of different rock
years (Pyle, 1992). Another example is the deformation pattern at Campi types, the properties assigned to each group are weighted averages,
Flegrei, which indicate that the inflation before the eruption in the year reflecting the contributions of different rock types within each group
1538 began around 150 years earlier and after a long period of subsi­ (Deb et al., 2019a). The weights used for each rock type are calculated
dence (Di Vito et al., 2016). Considering the above reasonings, we as­ based on the lithology of the geothermal wells in this area (Calcagno
sume that a maximum residence time of 1 000 years for Los Humeros et al., 2018; 2020). For detailed petrophysical report, please refer (Bär &
Holocene magmas is a reasonable assumption for numerical modeling. Weydt, 2019; Deb et al., 2019a; Weydt et al., 2020). The in situ matrix
Furthermore, longer residence times of such small magma pockets (refer thermal conductivity λm (W m− 1 K− 1 ) at temperature T (K) is calculated
Table 2 for volumes) would result in cooling of magma below solidus according to (Sekiguchi, 1984), which is considered to provide the best
preventing the eruptions of the lava, which were analysed in Lucci et al. fit for the temperature dependency of thermal conductivity for igneous
(2020). and metamorphic rocks (Lee & Deming, 1998). The relationship is
implemented as
4. Methodology ( )
( ) 1
λm = 1.8418 + λm,0 − 1.8418 − 0.2485 , (1)
0.002732 T + 0.7463
The numerical simulation is performed using SHEMAT-Suite, an
open-source, finite-difference code for simulating coupled heat transfer
where λm,0 is a measured matrix conductivity value at room temperature
and fluid flow in porous media (Clauser, 2003; Keller et al., 2020). The
(Chiozzi et al., 2017; Pasquale et al., 2014; Pasquale et al., 2017). The
area covered in this study is indicated by the model boundary (blue
effective in situ rock thermal conductivity λ with porosity ϕ filled with
dash) in Fig. 1. The structural framework and the present day main
fluid of thermal conductivity λf is then calculated according to the
lithostratigraphic units is shown in Fig. 3 (Calcagno et al., 2018; 2020).
geometric mean (e.g., Clauser, 2003)
The model extends 28 km in E-W direction and 22 km in N-S direction
and has a vertical extent of 11 km (7 km below mean sea level). The size λ = λϕf λ(1−
m
ϕ)
. (2)
of the numerical cells is 100 m × 100 m × 50 m yielding a total of 14
million model cells. It should be noted that no magma-specific thermal properties are
assigned to the cells representing the magma reservoir zones. The
emplacement and eruption of the magma pockets are modelled by
4.1. Thermal properties applying specific temperature and heat flow boundary conditions,
respectively. The magma pockets are located within the granite forma­
Based on the evolutionary history, the stratigraphy of the LHVC is tion, and hence their petrophysical properties in the model are similar to
broadly classified into four main groups: Basement, Pre-caldera, that of granite (Table 3).
Caldera, and Post-Caldera (Calcagno et al., 2018; 2020) (Table 3).The
thermal and petrophysical properties (porosity, matrix thermal

7
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 4. Temporal chart of simulation stages (top row), the numbers in the bottom of the chart indicate the geologic time duration (in years) of each phase (refer
section 3 for details). The model outputs used to evaluate the temperature contributions of different volcanic events in Section 6 are marked in Italics.

4.2. Heat transport equation implying that the groundwater level and, therefore, the hydraulic head
coincides with the topographic surface.
The heat transport is described by the transient heat conduction At the bottom of the model domain (at about 10 km below ground
equation which calculates the change in heat content dq in a control surface), a constant specific heat flow of 52 mW m− 2 is applied as a
volume dV during a time interval dt by Neumann boundary condition following Giordano et al., (2019) and
Davies (2013), which report regional heat flow values between
∂q ( )
= ∇⋅ λ ∇T − ρf cf Tv + H, (3) 85 mW m− 2 and 35 mW m− 2 in the study area (Fig. 3). It is assumed
∂t
that any influence from the previous volcanic activity (Miocene –
where λ (W m− 1 K− 1 ) is the effective thermal conductivity of the rock- Pleistocene andesitic volcanism) has subsided and that the volcanic
fluid mixture, (ρ c)f is the thermal capacity of the fluid (J m− 3 K− 1 ), v complex has returned to the thermal regime prior to the Los Humeros
event.
is the Darcy velocity (ms− 1 ) and H is the radiogenic heat production rate
Additional time-varying Dirichlet and Neumann boundary condi­
(Wm− 3 ) (Clauser, 2003; Clauser 2020a). For different lithology groups,
tions, applied to the numerical cells representing the magma bodies
H was assigned based on published values for dominant rock types
during the different stages of the simulation, are discussed in detail in
(Rybach, 1976; Schön, 2015). However, our simulations consider only
4.5.
heat conduction and therefore the advective term does not contribute to
the total heat transfer. While it may be argued that heat advection by
fluid flow, in principle, contributes to the total heat transport, it is 4.5. Time-varying boundary conditions
neglected here as a first approximation, because of (i) insufficient in­
formation on rock permeability and (ii) lack of information on the dis­ For modeling the different phases of emplacement, eruption, and
tribution, geometry and aperture of fluid conducting fractures. cooling for the two caldera-forming episodes and the following post-
The heat released by the magmatic source due to eruption is modeled caldera magmatic phase, the simulation was performed in 13 stages.
according to Newton’s law of cooling (Davidzon, 2012) This multi-stage approach was adopted to reduce the complexity of
( ) computationally intensive simulations and at the same time to overcome
q = α Tmag − Te , (4) the impossibility of changing the nature of the boundary condition
applied to a specific cell, which is necessary for modeling the
where, q is the heat flux (Wm− 2 ), α is the heat transfer coefficient emplacement and eruption. Except for the first stage, in which the
(Wm− 2 K− 1 ), Tmag and Te are the magma and the external temperatures steady-state model is used to represent the pre-caldera thermal regime,
(K), respectively. for each magma chamber (i) the emplacement and (ii) the eruption/
cooling are modelled in two separate simulation steps (Fig. 4). The
4.3. Initial conditions caldera phase, therefore, requires four stages to simulate the Los
Humeros (stages 2 and 3, Fig. 4) and the Los Potreros events (stages 4
In our multi-stage cascade simulation approach (see 4.5), the output and 5, Fig. 4). The post-caldera stage is similarly simulated by
of each simulation feeds the following by providing the proper initial emplacement, followed by eruption and cooling of the mafic-to-felsic
conditions that ensure unconditional continuity in the temperature field. small-volume magma batches considered as representative for the
A different approach is used during the first stage (pre-caldera), in which pattern of the Holocene LHVC magma storage system. These post-
the lack of previous data is overcome by performing a steady-state caldera magma pockets are grouped according to their estimated ages
simulation (stage 1, Fig. 4) with boundary conditions defined in 4.4. (Carrasco-Núñez et al., 2017a). The first group erupting at about 8 000
years ago, the second one at about 7 000 years ago, and the third and
4.4. Constant boundary conditions fourth ones at 3 800 and 2 800 years ago, respectively (Table 2). Thus,
the emplacement of the four post-caldera groups and their subsequent
The upper boundary condition is the temperature of the topographic eruption and cooling requires eight simulation stages (stages 6 – 13 in
surface (Ts, Fig. 3), calculated according to a wet adiabatic air temper­ Fig. 4). For details regarding time discretization and simulation run
ature gradient of 0.00491 Km− 1 (Clauser, 1984). The average annual air times, refer to Table D.1 in the Appendix section.
temperature of 12◦ C at Los Humeros climate station, located at an The emplacement of the magma chamber in most thermal modeling
elevation of 2 862 m.a.s.l. is used as a reference temperature. At the studies is instantaneous to avoid the mechanisms related to growth and
topographic surface, the pressure is assumed to be atmospheric, emplacement (Di Renzo et al., 2016; Verma, et al., 2011; Fjeldskaar

8
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 5. Schematic figure showing the application of time-varying boundary conditions, T0 is the initial temperature of the magma sill, N is the total number of steps
and Δtem is the time discretization between two emplacement steps, Δter is the eruption duration or time during which the heat flow boundary condition is active

Fig. 6. Temperature evolution through the center of Los Humeros magma chamber during a period of 182 000 years: the numbers indicate the different simulation
stages:1a - emplacement of the Los Humeros magma reservoir, 1b- eruption, 1c- cooling, 2a- emplacement of the Los Potreros magma chamber, 2b- eruption, 2c-
cooling; Depth in y axis is relative to sea level (0 m), the top of the magma reservoir is at -2 000 m below sea level, ground surface is at approx. 2800 m.a.s.l.

et al., 2008; Xiao-Yin et al., 2014). However, as shown in Galushkin disk-shaped structures or sills (Annen, 2009). The horizontal extent of
(1997), instantaneous emplacement overestimates the thermal influence the sills is assumed to be approximately equivalent to the dimension of
of intrusions. In addition, geochronological investigations of large the caldera collapse on ground surface (Roche & Druitt, 2001), while the
intrusive complexes have indicated that the formation of such plutons thickness or the vertical extent of the sills is decided according to the
occurs over thousands of years by amalgamation of many small inter­ numerical time steps N and total chamber volume V to be emplaced (also
acting pulses of magma (Caricchi & Blundy, 2015; Coleman et al., 2004; refer Table 1). The time steps or time discretization N for each simula­
Michel et al., 2008. Therefore, we created a numerical workflow to tion stage is a choice based on simulation approach and computational
model the thermal influence due to the progressive magmatic pulses capacity and is presented in Appendix D (Table D.1). The grid cells
based on a conceptual approach that more adequately reflects the dy­ representing the magma pulse or sill are assigned an initial temperature
namics evidenced by geochronological studies. Instead of instanta­ T0 as a Dirichlet temperature boundary condition for a period, corre­
neously emplacing a large magma chamber with initial temperature T0, sponding to one third of each sill’s emplacement time, after which the
we implemented a workflow, where the magma chamber is formed by boundary conditions are turned off (Fig. 5). The magma pulses (or the
the aggregation of many small magma pulses implemented as thermal boundary conditions) are applied at a rate proportional to the

9
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 7. E-W Temperature cross-sections through the center of the Los Humeros magma chamber: (I) at 164 000 years ago, after emplacement and just before eruption
of the Los Humeros magma reservoir, (II) at 100 000 years ago, after 64 000 years of cooling following Los Humeros eruption, (III) at 69 000 years ago, after the
emplacement of the Los Potreros magma chamber, (IV) at 9 000 years ago, after 60 000 years of cooling following Los Potreros eruption.

magma intrusion rate (discussed in Section 3). While the vertical simulate the post-eruption transient cooling by heat conduction only.
stacking continues, the temperature of each emplaced sill decays grad­ However, heat advection might quicken the process of cooling provided
ually due to cooling and diffusion of heat into the host rock, until a fresh the magma’s melt fraction exceeds 60 % (Annen, 2009). However, the
pulse of magma arrives. There are two major assumptions at this stage emplacement process of our magma reservoirs by incremental accretion
(i) during the aggregation process, we do not account for heat released of sills reduces the temperature difference between the magma and the
or added due to magma mixing, differentiation, or crystallization and host rock, due to gradual heating of the host rocks. This reduces the
(ii) the cooling of the magma bodies takes place through heat conduc­ effectivity of convection to some extent.
tion only.
Following the emplacement of the magma reservoirs, the next step is 5. Results
to model the eruption and cooling. Both eruption (removal of heat) and
crystallization is modelled by applying heat flow boundary condition to 5.1. Los Humeros caldera volcanism
the grid cells of the magma chamber. As we do not perform any
geochemical modeling, which is required to account for the latent heat The time required for accumulation of magma before large caldera
during the process of crystallization, we supplied an additional heat to eruptions is an active research area (Cashman & Giordano, 2014 and
the system equivalent to the specific enthalpy of crystallization of references therein). We used the magma intrusion rate of 0.06 km3 /yr
magma (Table 2, Clauser, 2020b). The latent heat effects are calculated (calculated in Section 3) to approximate an emplacement duration of 18
assuming that all the remnant magma chamber volume (total volume – 000 years for a chamber volume of 1 000 km3 (refer Fig. 4). Therefore,
erupted volume) is crystallized. This workflow is not a replacement of our transient simulation began 18 000 years before the eruption. The
fractional crystallization models, but rather an intermediate solution time discretization N used for emplacement of Los Humeros magma
adopted to replace the instantaneous emplacement of intrusive bodies chamber is 159 (Table D.1). Thus, every time step of numerical simu­
with more realistic emplacement steps, implemented in a simplified lation indicates approximately 113 years in geologic time (Δtem ) in this
manner. phase. Stage 1a in Fig. 6 represents the vertical stacking of sills starting
The eruption is modeled by removing heat from the system using the from the top at 5 km below ground surface down to the last sill at around
standard heat transfer Eq. (4) and properties of magma bodies presented 10 km below ground surface. The initial sills get emplaced within an
in Table 2. For avoiding numerical instabilities, the heat due to eruption environment which is originally cooler and hence gets solidified faster.
is not instantaneously removed from the system but distributed over With the upper sills cooling and crystallizing, the temperature at the top
several months according to a Gaussian profile, where 68 % of the mass of the magma chamber increases up to 500 ◦ C at the end of the
erupts within three months. This duration is not unreasonable consid­ emplacement stage (Stage 1a, Fig. 6), while the center of the magma
ering the reported data on the duration of a volcanic eruption. In a study chamber is around 900 ∘ C. Fig. 7(I) shows a vertical E-W temperature
presented in Siebert et al., 2010, 90 % of the 3 929 eruptions studied cross-section passing through the center of the Los Humeros magma
ended within three months. Nevertheless, this duration is like an instant reservoir just before the eruption event. The stacking of the sills is
for the simulation period of 182 000 years, and it does not affect the indistinguishable and appears as continuous in Stage 1a in Fig. 6 due to
temperature evolution. Following eruption, the simulated temperature large number of sills getting emplaced in a very short time. However, it
inside the magma chamber drops below 850 ◦ C at which the heat can be discerned in the emplacement of the Los Potreros magma reser­
transfer is primarily controlled only by heat diffusion through the solid voir (Stage 2a in Fig. 6).
mass. Following the emplacement, we modeled the Los Humeros event by
As stated before, we neglect any effects of heat advection and removing 8 × 1020 J of heat from the system corresponding to the

10
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 8. Location map of the magma pockets: color-coded and


sized according to their top depth relative to sea level and their
approximate radial extent respectively (the ages of the pockets
are shown in Table 2); geothermal wells are shown as triangles,
color-coded according to the bottom depth of the wells; the
outer, cropped outline is the boundary of the Los Humeros
magma chamber, while the inner elliptical outline indicates the
boundary of the Los Potreros magma chamber. Figures 9c and
9d continued in the next page.

eruption of 290 km3 of rhyolitic magma at a temperature of 850 ∘ C post-eruption magma chamber (Stage 2c in Fig. 6) until the emplace­
(Verma et al., 1990). The eruption, simulated as heat removal, is shown ment of the post-caldera magma reservoirs began. The thermal state of
in the temperature profile as a sharp drop in temperature (Stage 1b in the magma chamber before the emplacement of the post-caldera magma
Fig. 6). Following the eruption, we simulated the cooling of the pockets (9 000 years ago) is shown in Fig. 7-IV. The numerical time steps
post-eruption magma chamber until the emplacement of the magma for different phases of emplacement and eruption/ cooling are shown in
reservoir of the Los Potreros event (Stage 1c in Fig. 6). Fig. 7-II shows the Table D.1 in Appendix D.
temperature state through the magma chamber after 60 000 years of
cooling i.e., 100 000 years ago. At this instant of time, the center of the 5.3. Post-caldera Holocene volcanism
magma chamber cooled down to around 700 ◦ C.
The post-caldera volcanic activity is characterized by multiple
5.2. Los Potreros caldera volcanism shallow, small volumes of individual magma pockets representing
mafic, intermediate, and acidic magma. It is modeled similarly in two
A similar approach was followed for emplacement of Los Potreros phases (i) emplacement followed by (ii) eruption and cooling. The
magma chamber. Again, a magma volume of 100 km3 (see Section 3) magma pockets are divided into four groups according to their estimated
with a temperature of 1 050 ◦ C (Fig. 7-III) is emplaced step by step in sill eruption ages (Table 2). They were emplaced as discoidal bodies at an
layers at a rate of 0.003 km3 /yr (Ferriz and Mahood, 1987) (Stage 2a in influx rate determined from their respective volumes and an assumed
Fig. 6). It can be noticed from Fig. 6 that despite the fact that the Los residence period of 1 000 years (Pyle, 1992), wherein each magma pulse
Potreros magma chamber is ten times smaller than that of Los Humeros, is represented by an initial temperature applied as a Dirichlet temper­
the top of the magma chamber (around 2 km below sea level) is hotter ature boundary condition. This is followed by eruption of almost 50 % of
(around 700 ◦ C) at the end of the emplacement of Los Potreros. This is the pockets (modeled by removing heat from the system corresponding
because of a higher energy density, that protracts for a longer time in an to half of the pocket volume) and cooling of the magma pockets until the
already warm surrounding in the case of Los Potreros. This is a result of next magma group is emplaced. The location of the different magma
the smaller vertical extension of the Los Potreros magma chamber (2.5 pockets is shown in Fig. 8.
times less than the Los Humeros), due to which the latter emplaced sills Fig. 9 (a-d) show the temperature profiles through the center of the
in case of Los Potreros are located at a much shallower level (- 4 km) magma pockets for different instants in time for the 4 different groups.
than in the case of Los Humeros (- 7.5 km). Thus, the heat released in Los The tops of the first group of magma pockets, which erupted 8 000 years
Potreros remains confined in the proximity of the top of the magma ago, LH26, LH27-2 and LH4, are at approximately 2.5 km – 3 km from
chamber (- 2 km). After the emplacement, which lasted 33 000 years, the the surface. The temperature profile for 9 000 years ago shows the
eruption of Zaragoza ignimbrite is modelled by removing 1019 J of heat thermal cooling regime after the Los Humeros and Los Potreros events,
from the system corresponding to 15 km3 of andesitic magma at 900 ◦ C but prior to the emplacement of the post-caldera shallow magma
(Carrasco-Núñez et al., 2012) (Stage 2b in Fig. 6). This eruption does not pockets. Following the emplacement of the three trachy-andesitic
drastically lower the temperature of the system like in the Los Humeros magma pockets with corresponding volumes of approximately 0.1
event, since the heat removed during the Los Potreros event is two or­ km3, 1.3 km3 and 0.4 km3 the temperature at the center of the magma
ders of magnitude lower than that of the former. The eruption is fol­ reservoirs reaches around 1000 ∘ C just before eruption (indicated by the
lowed by diffusion of heat into the host rock and cooling of the temperature profile for 8 000 years ago). But after eruption and cooling,

11
P. Deb et al. Geothermics 96 (2021) 102162

the present-day temperature profiles at the center of all the magma LH4, due to their considerably smaller volumes, show an increase of
pockets have almost returned to their undisturbed states except for only around 40 K and 45 K, respectively, at the center of the magma
LH27-2 (Fig. 9a). At a depth of approximately 2.5 km from the surface pockets (see Fig. 9a).
and with volume greater than 1 km3 , LH27-2 cools down to approxi­ Magma pockets of the second group, LH21-2 and LH15, with vol­
mately 470 ∘ C at the centre of the magma pocket after 8 000 years. umes of around 1.3 km3 and magma reservoir top depth at about 6.5 km
However, without the magma replenishment during the post-caldera below ground surface, were emplaced at 1050 ◦ C, but cooled down to
period, the temperature at this depth would be less than 320 ∘ C as 538 ◦ C after almost 7 000 years (Fig. 9b). In absence of post-caldera
indicated by the temperature profile for 9 000 years ago. Therefore, the magma replenishment, the temperature at the center of both the
post-caldera magma activity of LH27-2 enhanced the temperature at this magma pockets would be less than 410◦ C. Therefore, a temperature
depth by at least 150 K, an increase of almost 40 %. However, LH26 and increase of almost 35 % is recorded at this depth after 7 000 years of

Fig. 9. Temperature profiles through the centre of the magma pockets (a) Group 1, (b) Group 2, (c) Group 3 and (d) Group 4, at different instants of time (see
legend); the profile with the temperature peak marked as E in the legend represents the thermal condition in the centre of the magma pocket just before eruption.
Figures 9c and 9d continued in the next page.

12
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 9. (continued).

13
P. Deb et al. Geothermics 96 (2021) 102162

cooling. LH13 and LH11, however with very small volumes of 0.01 km3 25 % - 30 % (Fig. 10b) (wells 23, 24 and 25) are located in the eastern
and 0.03 km3, respectively, results in a temperature increase of not more part of the geothermal field (Fig. 8). In general, all the wells located in
than 10 K at their centers (see Fig. 9b). the zone surrounded by Los Potreros fault scarp to the east, Las Papas
In spite of the relatively small volumes of around 0.05 km3, 0.3km3 fault to the north and Maztaloya fault to the south-southwest (wells 14,
and 0.01 km3 for LH27-1, LH5-2 and LH5-1 (Fig. 9c), these magma 18, 23, 24, 25, 26, 27), have turned out to be non-productive and were
pockets, which erupted 3 800 years ago, led to a temperature increase at abandoned (Aguilar et al., 2019). Interestingly, these wells also suffered
the centers of the magma pockets by around 135 K , 126 K and 55 K, an from high fluid circulation losses, between 6 m3h-1 and 20 m3h-1
increase of almost 54 %, 68 % and 70 %, respectively. Similarly, the (Aguilar et al., 2019). This might explain the overestimation of tem­
fourth and the youngest group of magma pockets, LH6-1 and LH17 perature as this zone might be cooled by the localized infiltration of cold
(Fig. 9d), with volumes of 0.15 km3 and 1.3 km3, show an increase of water through the Los Potreros fault scarp and other faults (Maztaloya
120 K and 382 K respectively. For LH17, with its largest volume and and Las Papas) located in the vicinity of the wells, the effect of which is
youngest age, the temperature excess at a depth of 2 000 m below not considered in our simulation.
ground surface is almost 200 %. Apart from the few wells in the east, where simulated temperature is
In Fig. 10(a), we compare the results of the last simulation stage, higher than the measured values, the mismatch in 40 % of the wells is
representing the present-day thermal regime, with the corrected static generally due to an underestimation of temperature (Fig. 10a). Please
the bottom-hole temperatures (BHT) in the geothermal wells. These are recall that the volumes of the shallow magma pockets used in this study
the temperatures at the greatest depth of each well obtained from are only estimates which are constrained by integration of field studies
Horner’s correction (Horner, 1951) applied to the transient temperature and geochemical modeling. The real volumes might vary largely,
recorded at this depth (see Appendix A. for details). The simulated resulting in corresponding increase or decrease of the estimated tem­
temperature at different bottom hole-depths agrees well with the tem­ perature. Also, we cannot rule out the possibility of such shallow and
perature data from the geothermal wells: for more than 50 % of the BHT small magma intrusions at other locations within the Los Humeros
(17 out of 32 data points), the simulated temperatures are in error by geothermal area. For example, at a depth of 2 300 m, the static tem­
less than 10 % (Fig. 10a). The error at the other 40 % of the wells ranges perature in well 8 is around 394 ◦ C, while well 10, only 500 m apart
between 15 % – 20 %, while at the remaining wells (4 out of 32), the from well 8, reaches a temperature of just around 174 ◦ C at the same
mismatch is greater than 25 %. The error in most of the wells is due to an depth. Similarly, at Loma Blanca, the measured temperature in well 4 at
underestimation by the simulation compared to measured data. For just a depth of about 1 000 m is around 300 ◦ C (Torres-Rodriguez, 1995;
3 out of 32 wells (wells 23, 24 and 25), the error is large (greater than 25 Prol-Ledesma, 1988, 1998; Martı ́nez-Serrano, 2002), while at well 43,
%) and is due to an overestimation by the simulation. which is less than 1 000 m west from well 4, the measured temperature is
about 170 ◦ C at similar depth (Lorenzo-Pulido, 2008). Such anomalous,
6. Discussion but extremely localized temperature distributions could be caused by
localized intrusions at depths of less than 3 km as suggested in Urbani
To isolate the contribution of the different volcanic events on the et al. (2020, 2021).
measured temperature of the wells, we performed two additional tran­ As shown in Fig. 10a, the current conceptual model is consistent with
sient models (Fig. 4), corresponding to the undisturbed cooling of Los the majority of the recorded well temperatures. The remaining
Humeros (Caldera LH) and Los Potreros (Caldera LP), whereas the main mismatch in temperature of 15 % ‒ 20 % may be due to several reasons,
simulation output including the cumulative effect of the post-caldera such as, advective heat transfer (which is not considered in the simu­
magma pockets is indicated as Holocene in Fig. 4. In ‘Caldera LH’, lation) or so-far undetected shallow magma pockets. Fluid flow through
following the emplacement and eruption of Los Humeros, the system is faults and fractures and its interaction with the shallow magma pockets
allowed to freely evolve until the present day without any thermal might create hotspots, increasing the temperature locally. Convective
disturbance due to the latter volcanic events. This model, therefore, zones can also quicken the process of cooling of the small magma
shows only the thermal contribution of the Los Humeros event and is, pockets due to infiltration of cold water resulting in a lower temperature
thus, conceptually comparable to the previous thermal models (Verma than obtained by our present-day temperature simulations (Cathles,
et al., 1990; Verma et al., 2011; Verma & Gómez-Arias, 2013). Similarly, 1977). However, as mentioned before the objective of this study was not
in ‘Caldera LP’, we simulated emplacement and eruption of Los Potreros, to deterministically match all the temperatures observed in the
followed by free cooling until the present day, without including the geothermal wells, but rather to use this information as a guide to test the
Holocene post-caldera shallow magma pockets. applicability of the proposed conceptual heat source model. Including
The result of these additional simulations in terms of temperature the effects of fluid flow and heat advection in future simulations will
contribution due to each volcanic event is presented in Fig. 10(b) require more detailed information on several parameters including fault
normalized with respect to the static BHT of each geothermal well. The permeability, thickness of the damaged zone around the fault, fault age
pre-caldera period contribution is given by the steady-state initial model and hierarchy, nature of fault sealing, fracture distribution, geometry
(blue) contributing to the extent of 25 % - 35 % to the temperatures and thickness.
observed in the wells. The voluminous Los Humeros event is responsible
for almost 30 % – 50 % of the excess temperature in the present-day 7. Conclusions
thermal regime, whereas the smaller Los Potreros event contributes to
an increase of 10 % – 20 % in the wells. The temperature excess due to In this paper, we presented a comprehensive approach to reconstruct
the post-caldera magma pockets referred to as Holocene phase is the thermal history of active volcanic systems characterized by multi-
observed only in few wells due to their localized nature. For example, a stage discrete heat sources. This transient modeling approach is
temperature increase of at least 15 % – 20 % is observed in wells 6, 12, applied in Los Humeros volcanic field to test if the conceptual model of a
39, 41 and 42, which are in the vicinity of LH 27-2. Well H-1, in the poly-phased caldera evolution can better reproduce its geothermal
vicinity of LH 4 and LH 13 (refer to Fig. 8 for locations) with volumes of configuration, and to quantify the relative contribution of deep and
0.4 km3 and 0.01 km3, respectively, shows a negligible temperature shallow magma heat sources across the lifetime of the magma plumbing
increase of less than 10 K due to post-caldera volcanism. In wells 2 and system. The simulation outcome indicates that a very high background
18, which are located within 1 km from the intrusive body LH 27-2 ( heat flow in Los Humeros is associated with the major caldera-forming
volume greater than 1 km3), a temperature excess of 30 K remains even events. However, the small and frequent magma eruptions during the
after 7 000 years of cooling. post-caldera period led to a temperature increase in localized hotspots,
The three wells in which the temperature is overestimated by almost providing a better explanation to the temperature anomalies observed in

14
P. Deb et al. Geothermics 96 (2021) 102162

Fig. 10. (a) Simulated versus bottom-hole temperature (BHT) from geothermal wells; absolute error (in green) is the difference between BHT and the simulated
temperature (in blue) at bottom hole depths (see Fig. 8 for location and bottom hole depth of wells); (b) Contribution of different volcanic events towards the present-
day temperature observed in (numbered) wells

the wells. The expected present-day temperature of the younger magma initial temperature, thermal properties, age of intrusion, intrusion rate,
pockets, modeled on the basis of the eruptive history of the last ten recharge versus withdrawal, convection. To test the exact geometry of
thousand years, suggests the existence of super-hot resources at shallow the magma bodies and the sensitivity of the reconstructed temperature
depths. In particular, the southern part of the geothermal field indicates with respect to variation in a set of parameters, separate sensitivity
potentially unexplored opportunities, but the presence of many similar studies must be performed in small-scale models. For these reasons, we
non-erupted intrusive bodies within drillable depths in other parts of the must remind the reader that, while the input data used in our modeling
geothermal field cannot be excluded. Future exploration surveys, are derived from the best available and published volcanological,
including monitoring of surface deformation, seismic activities and petrologic and age datasets, they are also grounded on a series of as­
high-resolution geophysical surveys, may provide hints to their exis­ sumptions, for which the reconstructed intrusive heat sources should be
tence and potentially widen the future exploitation area. considered a viable, data-constrained, first-order scenario rather than an
The factors not accounted for in this study, but which might affect exact reconstruction. Nevertheless, our results show that adopting first
the final predicted temperature, are the processes related to magma order approximations for heat generation and transport phenomena, can
differentiation and mixing, convection inside the magma body, or con­ provide good results for the regional reconstruction of the caldera-
vection related to hydrothermal circulation within faults or fractures. related geothermal fields while limiting at the same time the numeri­
The first two would require new petrologic works on Xaltipan and cal complexity to an acceptable level.
Zaragoza ignimbrites for performing new fractional crystallization
models in accordance with the modern view of the magma plumbing CRediT authorship contribution statement
system. Such studies can be useful for constraining the volumes, depths,
and temperature of the Los Humeros and Los Potreros reservoirs, which Paromita Deb: Conceptualization, Methodology, Software, Valida­
are still based on older studies. The latter requires more information tion, Writing – original draft, Writing – review & editing, Visualization,
regarding the distribution of faults and fractures, density, extent, and Project administration. Guido Giordano: Conceptualization, Investi­
permeability. gation, Writing – review & editing. Xiangyun Shi: Methodology, Soft­
The biggest challenge of this approach lies in conceiving, planning ware, Formal analysis. Federico Lucci: Investigation, Data curtion,
and implementing such a temporally-varying conceptual model for the Writing – review & editing. Christoph Clauser: Writing – review &
evolution of the magma plumbing system, in contrast to the simplified editing, Project administration, Funding acquisition.
single heat source/ magma chamber models that instantaneously
emplace the body in the crust at averaged conditions (volume, depth, Declaration of Competing Interest
temperature). Our approach requires information of the heat sources in
time and space and, although conceptually closer to reality, it carries the The authors declare no competing interests.
risk of propagating errors and uncertainties associated with the char­
acteristics assigned to each heat source: volume, shape, depth range,

15
P. Deb et al. Geothermics 96 (2021) 102162

Acknowledgments geological models for this study. Additionally, we thank Eva Schill and
Natalia Cornejo from KIT, Germany for allowing us to reprint figures
We thank the European Commission for funding the project GEMex (Appendix B) from Deliverable 5.6 (Cornejo et al., 2019) of GEMex
through European Union’s Horizon 2020 research and innovation pro­ project. The grant to the Department of Science, Roma Tre University
gram under grant agreement No. 727550. The authors also thank the (MIUR-Italy Dipartimenti di Eccellenza, ARTICOLO 1, COMMI 314-337
Mexican electric company and operator of Los Humeros geothermal LEGGE 232/2016) is gratefully acknowledged. Last but not the least, we
field, the Comision Federal di Electricidad (CFE), and espesially Mr. thank the entire GEMex consortium, comprising Mexican and European
Heber Diez and Mr. Miguel Ramirez, for providing data and information partners, for the cooperation and unpublished inputs on conceptual
regarding their geothermal wells. We would also like to thank Philippe ideas which were available for this study.
Calcagno (BRGM, France) and WP-3 team of GEMex for providing us the

Appendix A. Correction of transient temperature data

Data for more than 50 geothermal wells were provided by the CFE. The temperature and pressure surveys were recorded using mechanical Kuster
temperature and pressure sensors (Probe, 2020). Unfortunately, no information is available on the type and accuracy of the equipment used for the
measurements. However, an accuracy of ± 0. 25 ∘ C is quoted for the Kuster K10 Geothermal high temperature and pressure recording tool (Sword,
2020). Surveys were performed at different intervals of time (generally between 6 and 30 hours after drilling) or during the heating-up surveys (after
several days, weeks or months after shut-in). The measurements were usually performed with a very coarse sampling intervals varying between 200 m
– 300 m for the first 1 000 m, followed by 50 m – 100 m sampling intervals down to bottom-hole depth. For reliably estimating the stable formation
temperature, measurements performed during the heating-up surveys are better suited than the temperatures recorded immediately after drilling, as
the latter are usually disturbed by different factors, such as, mud circulation, drilling technology, borehole breakouts (Goutorbe et al., 2007; Kutasov &
Eppelbaum, 2005). However, in case of Los Humeros wells, most of the temperature data taken during heating-up surveys suffer additionally due to
boiling and phase separation, and interflow between different feed zones (Arellano et al., 2003; Arellano et al., 2015). Therefore, for estimating the
stable formation temperature from Los Humeros wells, we utilized temperatures measured between 6 and 34 hours after drilling, and did not use the
heating-up surveys (Table A.1). Temperatures from some wells which reached a steady-state within 30 h – 36 h are directly used as stabilized
temperatures (denoted with * in Table A.1)
Methods for estimating static or stable temperatures from measurements after drilling have been investigated in many studies (Andaverde et al.,
2005; Goutorbe et al., 2007). After reviewing the available data quality of the transient temperature series, we selected data from 32 wells and applied
Horner’s method of correction to the measured bottom hole temperatures (Dowdle & Cobb, 1975; Horner, 1951). Static reservoir temperatures for Los
Humeros were previously determined using the spherical radial heat-flow method (García-Gutiérrez et al., 2002; Arellano et al., 2003). This method
considers the thermally perturbed zone around the borehole as a spherical region and calculates the static temperature from the intercept of a plot of T
√̅̅
vs 1/ t , where T is the temperature at different instants of time t (Ascencio et al., 1994). However, later studies comparing different correction
methods suggested that this approach is physically invalid and oversimplified and hence, not recommended for geothermal applications (Andaverde
et al., 2005; Wong-Loya et al., 2015). Horner’s method, therefore, remains a widely used method for correcting bottom-hole temperatures in
geothermal wells. It considers the temperature perturbation due to mud circulation in a well as caused by an infinitely long heat source or sink
(Andaverde et al., 2005) according to
(t + t )
c
Tws = Ti − cln ,
t

where Tws is the shut-in temperature at time t, Ti is the static formation temperature at infinite shut-in time, and tc is the mud circulation time. This
model describes a straight line with slope c and intercept Ti obtained by extrapolation to infinite shut-in time. It has been however, suggested that
Horner’s analysis of temperature build-up always underestimates the static formation temperature and can only be applied in cases of short circulation
times (less than 24 hours) (Dowdle & Cobb, 1975; Kutasov & Eppelbaum, 2005).
One limitation of this method is the input mud circulation time, which is generally not meticulously recorded in most drilling operations

Table A.1
Measured temperature and corrected BHT for different wells: depth, temperature Tws at time t (hours after shut-in), mud circulation times tc, static formation tem­
perature corrected using Horner’s method Ti; * indicates temperature in wells, which stabilized after time t and does not require correction.
Well Depth (m) t (h) Tws (◦ C) tc (h) Ti (◦ C) Well Depth (m) t (h) Tws (◦ C) tc (h) Ti (◦ C)

1 1400 12, 18, 24 195, 199, 202 6 201 19 2250 12, 17, 24 208, 216, 227 6 249
2 2300 15, 19, 23 257, 263, 265 6 282 20 2389 18, 24, 30 227, 250, 271 6 345
3 1660 12, 18, 24 220, 240, 251 6 289 21 2214 12, 18, 24 239, 248, 258 6 278
5 1845 30 - - 212* 22 1539 12, 18, 24 223, 237, 245 6 272
6 2540 12, 24, 27 275, 296, 300 6 323 23 2450 24 - - 198*
7 2281 38, 54, 60 259, 271, 273 6 301 24 3263 18, 24, 36 211, 222, 231 6 255
8 2300 14, 18, 25, 34 241, 253, 266, 274 2 394 25 2283 12, 18, 24 150, 162, 172 6 197
9 2450 6, 12, 18 227, 247, 251 6 269 26 2530 12, 18, 24 323, 330, 341 6 361
11 2376 30 - - 356* 28 2558 24 - - 360*
12 2984 14 - - 362* 30 1900 12, 18, 24 193, 204, 211 6 234
13 2400 6, 18, 24, 36 237, 274, 281, 290 6 303 31 1915 18, 24, 32 281, 291, 298 6 323
14 1373 12, 18, 24 93, 103, 107 6 125 32 2190 6, 12, 24 251, 287, 307 6 332
15 1958 6, 12, 18 178, 200, 211 6 234 38 1390 6, 12, 18 120, 140, 150 6 169
16 2038 6, 12, 18, 24 210, 254, 273, 285 6 319 39 2495 12, 18, 24 208, 220, 232 259
17 2214 12, 19, 24 206, 219, 227 6 251 41 2200 36 290*
18 2885 12, 18, 24 234, 251, 263 6 297 42 2060 12, 18, 24 215, 236, 237 4.15 268

16
P. Deb et al. Geothermics 96 (2021) 102162

Table A.2
Table showing the impact of unknown mud circulation times on the static formation temperature at bottom-hole depth of a well
corrected using Horner’s method
Mud circulation time Static formation temperature
(hours) (◦ C)

2 299.54
6 300.84
10 302.07
14 303.25
20 304.91

(Andaverde et al., 2005). It is also lacking for most of the wells of Los Humeros. For those wells with this information, it varied between 2 hours to 10
hours (personal communication with CFE, Mexico). For wells, without this information, we considered an average circulation time of 6 hours for
calculating the correction. Table A.1 shows the data used for our Horner corrections and the corrected static formation temperature. Table A.2
provides a rough estimate of the uncertainty of the calculated static formation temperatures due to the uncertainty of the mud circulation times. We
believe that our correction yields one of the best possible estimates of static formation temperature in Los Humeros wells with an uncertainty of
± 10 K .

Appendix B. Density distribution from 3-D inversion of gravity data of the Los Humeros area (Figures reprinted from Cornejo et al., 2019
with permission)

Figure B

Fig. B. Density distribution obtained from 3-D inversion of gravity data of the Los Humeros area at (a) 1000 m a.s.l., (b) at sea level and (c) 1000 m b.s.l. (Cornejo
et al., 2019) Densities are provided as Δρ (g cm− 3) with respect to the Bouguer density of 2670 kg m− 3; the white dots in the center of the gravity maps are gravity
stations measured during GEMex project, additional measurement points used for the 3D inversion is presented in Figure 21c of Cornejo et al., 2019; (d) locations of
the various magma pockets: the circles are color-coded and sized according to their top depth relative to sea level and their approximate radial extent respectively.
17
P. Deb et al. Geothermics 96 (2021) 102162

Appendix C. Methodology for the reconstruction of the Holocene shallow intrusive scenario at Los Potreros caldera (Los Humeros
Volcanic Complex, Mexico)

This section presents the workflow for the determination of the temperature, depth and volume input data for the geothermal modeling of the
Holocene magma pockets. The Holocene monogenetic post-caldera volcanism and its wide range of geochemical compositions related to sources
scattered along the entire crust (Lucci et al 2020) indicate that the two main magma chambers associated with the caldera-forming eruptions Xaltipan
(164 000 years) and Zaragoza (69 000 years) have now cooled below solidus. Shallower magma intrusions associated with the Holocene post-caldera
stage (LHPCS) erupted products within the Los Potreros caldera (“Ring-fissure and bimodal phase” in Carrasco-Nunez et al. 2018) are intended to
represent a spectrum of possible scenarios of discrete, short-lived, small and shallow magma heat sources as constrained by the location of the eruption
points of the LHPCS erupted products (Carrasco-Nunez et al. 2017) and related petrologic data published by Lucci et al. (2020).

Table C.1
Temperature and depths selected for the modeling of the LHPCS magma heat sources.
Rock type (a) Sample Temperature(c)(◦ C ± 1σ) Pressure(c)(MPa ± 1σ) Well-log data (d)
Depth (km) from
the surface to the
top

AB LH5-2 (b) 1100 ± 50 50 - 100 ± 100 A ca. 200 m thick basalt layer at 1.5
Estimates resulting from thermometry Minimum estimates from barometry 1300-1500 m below the surface in
models applied to Cpx3, Cpx4 and Cpx5 models applied to Cpx4 and Cpx5 well H5.
populations. populations.
AB LH27-1 (b) 1100 ± 50 50 -100 ± 100 A ca. <200 m thick mafic lava layer at 2.5
Estimates from thermometry models Minimum estimates from barometry 2300-2500 m below the surface in
applied to Cpx3, Cpx4 and Cpx5 models applied to Cpx4 and Cpx5 well H26.
populations. populations.
TA LH15 (b) 1050 ± 50 140 ± 80 - 6
Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx3 and Cpx4 models applied to Cpx4 cluster.
populations. Absence of Opx.
TA LH21 (b) 1050 ± 50 140 ± 80 - 6
Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx3 and Cpx4 models applied to Cpx4 cluster.
populations. Absence of Opx.
Opx-TA LH4 (b) 1050 ± 50 100 ± 60 - 3
Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx2, Cpx3, Cpx4 and models applied to Cpx4 and Opx
Opx populations. populations.

Opx-TA LH13 (b) 1050 ± 50 100 ± 60 - 3


Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx2, Cpx3, Cpx4 and models applied to Cpx4 and Opx
Opx populations. populations.

Opx-TA LH17 (b) 1050 ± 50 100 ± 60 - 3


Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx2, Cpx3, Cpx4 and models applied to Cpx4 and Opx
Opx populations. populations.

Opx-TA LH26 (b) 1050 ± 50 100 ± 60 A ca. <200 m thick mafic lava layer at 2.5
Estimates resulting from thermometry Minimum estimates from barometry 2300-2500 m below the surface in
models applied to Cpx2, Cpx3, Cpx4 and models applied to Cpx4 and Opx well H26.
Opx populations. populations.

Opx-TA LH27-2 (b) 1050 ± 50 100 ± 60 - 3


Estimates resulting from thermometry Minimum estimates from barometry
models applied to Cpx2, Cpx3, Cpx4 and models applied to Cpx4 and Opx
Opx populations. populations.

TR LH5-1 (b) 950 ± 50 100 ± 90 A 50-100 m thick rhyolite layer at 0.5


Estimates resulting from thermometry Minimum estimates from barometry 500 m below the surface in well H5.
models applied to Ol, Cpx3, Cpx4 and Opx models applied to Cpx4 and Opx A ca. 400 m thick rhyolite layer at
populations. populations. 500 m below the surface in well H20.
TR LH6-1 (b) 950 ± 50 100 ± 90 - 3
Estimates resulting from thermometry Minimum estimates from barometry
models applied to Ol, Cpx3, Cpx4 and Opx models applied to Cpx4 and Opx
populations. populations.

R LH11(e) 800 - A positive thermal anomaly measured <1


Average temperature estimates for rhyolite Analogue modeling from during drilling at ca. 1000 m below
melts from literature. Urbani et al. (2020) the surface in well H4.

(a) Rock types as classified in Lucci et al. (2020), AB: alkali-basalt; TA: Opx-free trachyandesite; Opx-TA: Opx-bearing trachyandesite; TR: trachyte; R: Rhyolite.
(b) Samples investigated for thermobarometry modeling in Lucci et al. (2020).
(c) Temperature, pressure, and their 1σ error values are derived from thermobarometry results presented in Lucci et al. (2020)
(d) Literature data (see Carrasco-Nunez et al. (2018), Urbani et al. (2020, 2021), Cavazos-Alvarez et al. (2020) and references therein).
(e) Sample presented in Urbani et al. (2020, 2021).

18
P. Deb et al. Geothermics 96 (2021) 102162

For these intrusions, the workflow is based on the integration of: (i) textural observations, mineral chemistry and thermobaric estimates for each
major LHPCS eruptive product published in Lucci et al. (2020), (ii) Holocene evolution of ground deformation of the Los Potreros caldera floor
(Urbani et al. 2020; 2021), (iii) the most recent Los Humeros geological map (Carrasco-Núñez et al., 2017a) and (iv) stratigraphy of intracaldera wells
(Carrasco-Núñez et al., 2017b; Cavazos & Carrasco-Núñez, 2020; Cavazos-Álvarez et al., 2020), (v) recent literature focusing on the shallow magma
storage zone feeding the Los Humeros Holocene volcanism (Dávila-Harris and Carrasco-Núñez, 2014; Creon et al., 2018).
The shallowest magma storage zone of the Holocene plumbing system is developed at a pressure of ca. 1-1.5 kbar (100 – 150 MPa) (mean values
from Lucci et al., 2020) corresponding to depths less than 4 - 6 km below the caldera floor. This shallowest portion of the plumbing system has been
interpreted by Lucci et al. (2020) as a heterogeneous plexus made up of “a system of small magma volumes, distributed in locally interconnected pockets and
batches”. Furthermore, Lucci et al., (2020) documented in all the LHPCS studied samples a general absence of lithics, xenocrysts and phenocrysts with
disequilibrium textures and dissolution patterns, which is interpreted as the evidence of magma batches evolving as nearly closed system where the
effects of magma mixing/recharge and assimilation processes are negligible. Moreover, considering the existing literature on magma recharge and
evacuation processes (e.g., DePaolo, 1981; Karlstrom et al., 2010; Hofmann, 2012; O’Neill and Jenner, 2012; Lee et al., 2014; Moghadam et al., 2020),
Lucci et al. (2020) described the Holocene LHPCS activity being compatible with a scenario dominated by (i) a high evacuation rate and/or eruptive
efficiency and (ii) an evacuation rate higher than the recharge. This scenario is also interpreted as capable of shortening the residence time of melts in
the ephemeral magma pockets reducing the effect of crystallization (e.g., Annen, 2009; Lee et al., 2014). The monogenetic and scattered style of the
intracaldera Holocene volcanic activity supports the absence of significant recharge events on pre-stored magma, justifying the absence of evidence for
mixing-mingling processes. In view of the above the maximum residence time for magmas feeding the Holocene volcanic activity has been assumed as
1000 years.
With respect to the modeling proposed in this work, the following constraints are then assumed for the Holocene LHPCS erupted magmas:

• the depth and the temperature of the magma batches are mainly constrained through the thermobaric estimates obtained from rims and unzoned
phenocrysts (clinopyroxene, olivine and orthopyroxene) as presented in Lucci et al. (2020). A summary is given in Table C.1 presented below.
○ In particular, concerning the mean temperatures chosen for the modeling: (i) the T value of 1100 ◦ C for intracaldera Ol-basalts is obtained from

convergence of thermometric estimates calculated for clinopyroxene populations; (ii) the T value of 1050 ◦ C for trachyandesites derives from the
convergence of thermometric estimates calculated for pyroxenes, plagioclase and olivine phenocrysts; and (iii) the temperature 950 ◦ C for
trachytes results from the integration of thermometric estimates obtained for pyroxenes and olivine phenocrysts.
○ The depths of modelled magma batches are derived by convergence of the minimum estimates obtained from clinopyroxene and orthopyroxene

barometry models (Lucci et al. 2020).


○ where available, the depth of the magma intrusions is integrated with information derived from well data. For example- the LH5-1 trachyte

sample and the neighbour H5 geothermal well (Carrasco-Núñez et al., 2017b; Cavazos-Alvarez et al., 2020; Urbani et al., 2020, 2021). Please
note that with this we do not imply that possible intrusive rocks identified in the stratigraphy of well H5 correspond necessarily to the intrusions
related to the eruption of sample LH5-1, but only an indication that the available well information converge with the intrusive scenario
numerically modelled in the present study.
• the intrusive magma volumes are assumed to be twice as the erupted volumes as determined (first order approximation) by the integration of field
observations and published geological maps, and in consideration of their effusive eruptive styles.
○ pre-eruptive intrusion depths for the obsidian domes are constrained by field data and analogue modeling (Urbani et al., 2020, 2021), with a

mean temperature as derived by existing literature on rhyolitic melts (e.g., Shaw, 1963; Magaritz and Hofmann, 1978; Castro et al., 2005;
Hamada et al., 2010; Lucci et al., 2018) –
○ the modelled magma intrusions are arbitrarily assumed to be located at depth, vertically below the equivalent eruptive rocks.

Appendix D. Simulation statistics

The simulations were performed using the High-Performance Computing systems CLAIX (Cluster Aix-la-Chapelle) at RWTH Aachen University and
were completed in 13 stages. The time discretization and the core hours for each stage are provided in Table D.1.

Table D.1
Simulation statistics.
Simulation Description of the stage Simulation duration (in Time steps, N Simulation time
stages years) (min)

1 initial background model steady-state - 138


2 emplacement of LH magma reservoir 18 286 159 1527
3 eruption and cooling of LH until LP emplacement began 61 696.5 297 1774
4 emplacement of LP magma reservoir 33 304 153 970
5 eruption and cooling of LP until Group 1 magma pockets emplacement began 60099.5 297 1711
6 emplacement of Group 1 magma pockets 901 36 452
7 eruption and cooling of Group 1 magma pockets until emplacement of Group 2 99.5 98 504
began
8 emplacement of Group 2 magma pockets 901 36 330
9 eruption and cooling of Group 2 magma pockets until emplacement of Group 3 2 199.5 98 647
began
10 emplacement of Group 3 magma pockets 1 001 20 202
11 eruption and cooling of Group 3 magma pockets until emplacement of Group 4 99.5 98 555
began
12 emplacement of Group 4 magma pockets 901 36 349
13 eruption of group 4 magma pockets and cooling until present day 2 799.5 98 650

19
P. Deb et al. Geothermics 96 (2021) 102162

References Carrasco-Núñez, GerardoG., Branney, M.J., 2005. Progressive assembly of a massive


layer of ignimbrite with a normal-to-reverse compositional zoning: the Zaragoza
ignimbrite of central Mexico. Bull. Volcanol. 68 (1), 3–20. https://doi.org/10.1007/
Aguilar, A.A., Deb, P., Izquierdo, G., 2019. Conceptual analysis of geothermal
s00445-005-0416-8.
neighboring zones characterization with contrasting behavior: applied to a Mexican
Carrasco-Núñez, GerardoG., McCurry, M., Branney, M.J., Norry, M., Willcox, C., 2012.
geothermal field. Int. J. Hydrol. 3 (3), 175–183. https://doi.org/10.15406/
Complex magma mixing, mingling, and withdrawal associated with an intra-Plinian
ijh.2019.03.00178.
ignimbrite eruption at a large silicic caldera volcano: Los Humeros of central Mexico.
Andaverde, J., Verma, S.P., Santoyo, E., 2005. Uncertainty estimates of static formation
Bull. Geol. Soc. Am. 124 (11–12), 1793–1809. https://doi.org/10.1130/B30501.1.
temperatures in boreholes and evaluation of regression models. Geophys. J. Int. 160
Cashman, K.V., Giordano, G., 2014. Calderas and magma reservoirs. J. Volcanol.
(3), 1112–1122. https://doi.org/10.1111/j.1365-246X.2005.02543.x.
Geotherm. Res. 288, 28–45. https://doi.org/10.1016/j.jvolgeores.2014.09.007.
Annen, C., 2009. From plutons to magma chambers: thermal constraints on the
Cashman, K.V., Sparks, R.S.J., Blundy, J.D., 2017. Vertically extensive and unstable
accumulation of eruptible silicic magma in the upper crust. Earth Planet. Sci. Lett.
magmatic systems: a unified view of igneous processes. Science 355, 6331. https://
284 (3–4), 409–416. https://doi.org/10.1016/j.epsl.2009.05.006.
doi.org/10.1126/science.aag3055.
Annen, CatherineC., Blundy, J.D., Sparks, R.S.J., 2006. The genesis of intermediate and
Castillo-Roman, J., Verma, S.P., Andaverde, J., 1991. Modelacion de temperaturas bajo
silicic magmas in deep crustal hot zones. J. Petrol. 47 (3), 505–539. https://doi.org/
la Caldera de Los Humeros, Puebla, Mexico, en termicos de profundidad de la
10.1093/petrology/egi084.
camara magmatica. Geofisica Int. 30 (3), 149–172.
Arellano, V.M., García, A., Barragán, R.M., Izquierdo, G., Aragón, A., Nieva, D., 2003. An
Castro, J.M., Dingwell, D.B., Nichols, A.R., Gardner, J.E., 2005. New insights on the
updated conceptual model of the Los Humeros geothermal reservoir (Mexico).
origin of flow bands in obsidian. Spec. Pap.-Geol. Soc. Am. 396, 55. https://doi.org/
J. Volcanol. Geotherm. Res. 124 (1–2), 67–88. https://doi.org/10.1016/S0377-0273
10.1130/0-8137-2396-5.55.
(03)00045-3.
Cathles, L.M., 1977. An analysis of the cooling of intrusives by ground-water convection
Arellano, V.M., Barragán, R.M., Ramírez, M., López, S., Paredes, A., Aragon, A.,
which includes boiling. Econ. Geol. 72 (5), 804–826. https://doi.org/10.2113/
Tovar, R., 2015. The Response to Exploitation of the Los Humeros (México)
gsecongeo.72.5.804.
Geothermal Reservoir. In: Horne, R., Boyd, T. (Eds.), Proceedings World Geothermal
Cathles, L.M., Erendi, A.H.J., Barrie, T., 1997. How long can a hydrothermal system be
Congress 2015, 20 Apr 2015 - 24 Apr 2015. Melbourne, Australia.
sustained by a single intrusive event? Econ. Geol. 92, 766–771. https://doi.org/
Ascencio, F., García, A., Rivera, J., Arellano, V., 1994. Estimation of undisturbed
10.2113/gsecongeo.92.7-8.766.
formation temperatures under spherical-radial heat flow conditions. Geothermics 23
Cavazos, J.A., Carrasco-Núñez, G., 2020. Anatomy of the Xáltipan ignimbrite at Los
(4), 317–326. https://doi.org/10.1016/0375-6505(94)90027-2.
Humeros Volcanic Complex; the largest eruption of the Trans-Mexican Volcanic Belt.
Bachmann, O., Bergantz, G., 2008. The magma reservoirs that feed supereruptions.
J. Volcanol. Geotherm. Res. 392, 106755 https://doi.org/10.1016/j.
Elements 4 (1), 17–21. https://doi.org/10.2113/GSELEMENTS.4.1.17.
jvolgeores.2019.106755.
Bär, K., & Weydt, L.M., 2019. Final Report on rock and fluid samples and their physical
Cavazos-Álvarez, J.A., Carrasco-Núñez, G., Dávila-Harris, P., Peña, D., Jáquez, A.,
properties in the Acoculco and Los Humeros regions Comprehensive report on the
Arteaga, D., 2020. Facies variations and permeability of ignimbrites in active
rock and fluid samples and their physical properties in the Acoculco and Los
geothermal systems; case study of the Xáltipan ignimbrite at Los Humeros Volcanic
Humeros regions. H2020 project GEMex Scientific Deliverable D6.1. GEMex Report.
Complex. J. South Amer. Earth Sci. 104, 102810 https://doi.org/10.1016/j.
https://data.d4science.net/zVLc. Last accessed –09 January 2021.
jsames.2020.102810.
Benediktsdóttir, A., Hersir, G.P., Vilhjalmsson, A.M., et al., 2019. Report on resistivity
Cedillo-Rodriguez, F., 2000. Hydrogeologic model of the geothermal reservoir from Los
modelling and comparison with other SHGS. H2020 project GEMex Scientific
Humeros, Puebla, Mexico. Geotherm. Resour. Council Trans. 24, 381–409.
Deliverable D5.2. GEMex Report.
Chiozzi, P., Barkaoui, A., Rimi, A., Verdoya, M., Zarhloule, Y., 2017. A review of surface
Bonini, M., Maestrelli, D., Corti, G., et al., 2021. Modelling intra-caldera resurgence
heat-flow data of the northern Middle Atlas (Morocco). J. Geodyn. 112, 58–71.
settings: laboratory experiments with application to the Los Humeros Volcanic
https://doi.org/10.1016/j.jog.2017.10.003 (November).
Complex (Mexico). J. Geophys. Res. 126, e2020JB020438 https://doi.org/10.1029/
Claiborne, L.L., Miller, C.F., Flanagan, D.M., Clynne, M.A., Wooden, J.L., 2015. Zircon
2020JB020438.
reveals protracted magma storage and recycling beneath Mount St. Helens. Geology
Brown, S.J.A., Burt, R.M., Cole, J.W., Krippner, S.J.P., Price, R.C., Cartwright, I., 1998.
38 (11), 1011–1014. https://doi.org/10.1130/G31285.1.
Plutonic lithics in ignimbrites of Taupo Volcanic Zone, NewZealand; sources and
Clauser, C., 1984. A climatic correction on temperature gradients using surface
conditions of crystallisation. Chem. Geol. 148 (1–2), 21–41. https://doi.org/
temperature series of various periods. Tectonophysics 103, 33–46. https://doi.org/
10.1016/S0009-2541(98)00026-6.
10.1016/0040-1951(84)90072-6.
Calcagno, P., Evanno, G., Trumpy, E., C. Gutiérrez-Negrín, L., MacIás, J. L., Carrasco-
Clauser, C., 2003. Numerical Simulation of Reactive Flow in Hot Aquifers. Springer,
Núñez, G., & Liotta, D. (2018). Preliminary 3-D geological models of Los Humeros
Berlin-Heidelberg https://doi.org/10.1007/978-3-642-55684-5.
and Acoculco geothermal fields (Mexico)-H2020 GEMex Project. Adv. Geosci., 45,
Clauser, C., 2020a. Radiogenic heat production of rocks. In: Gupta, H. (Ed.),
321–333. https://doi.org/10.5194/adgeo-45-321-2018.
Encyclopedia of Solid Earth Geophysics. Encyclopedia of Earth Sciences Series.
Calcagno, P., Trumpy, E., Carlos Gutiérrez-Negrín, L., et al., 2020. Report on the
Springer, Cham. https://doi.org/10.1007/978-3-030-10475-7_74-1.
integrated geomodels of Los Humeros and Acoculco. H2020 project GEMex Scientific
Clauser, C., 2020b. Thermal storage and transport properties of rocks, I: heat capacity
Deliverable D3.1. 1. https://data.d4science.net/wT4V.
and latent heat. In: Gupta, H. (Ed.), Encyclopedia of Solid Earth Geophysics.
Calcagno, P., Trumpy, E., Carlos Gutiérrez-Negrín, L., Norini, G., Macías, J.L., Carrasco-
Encyclopedia of Earth Sciences Series. Springer, Cham. https://doi.org/10.1007/
Núñez, G., Liotta, D., Garduño-Monroy, V.H., Páll Hersir, G., Vaessen, L., Evanno, G.,
978-3-030-10475-7_238-2.
Galván, C.A., 2021. Updating the 3D Geomodels of Los Humeros and Acoculco
Cole, J.W., Milner, D.M., Spinks, K.D., 2005. Calderas and caldera structures: a review.
Geothermal Systems (Mexico)-H2020 GEMex Project (accepted). In: Proceedings
Earth Sci. Rev. 69 (1–2), 1–26. https://doi.org/10.1016/j.earscirev.2004.06.004.
World Geothermal Congress 2020, 21 May 2021 - 26 May 2021. Reykjavik, Iceland.
Cornejo, N., Schill, E., Perez, M., Carrillo-Lopez, J., 2019. Gravity Modelling. H2020
Campos-Enríquez, J.O., Duran, M.F., 1986. Determinación preliminar del campo de
project GEMex Scientific Deliverable D5.6 https://cordis.europa.eu/project/id/
temperaturas en Los Humeros, Puebla. Geotermia, Revista Mexicana de Geoenergía
727550/results.
2, 141–152.
Costa, F. (2008). Residence times of silicic magmas associated with calderas. In J.
Campos-Enríquez, J.O., Garduño-Monroy, V.H., 1987. The shallow structure of Los
Gottsmann & J. Marti (Eds.), Caldera Volcanism Analysis, Modelling and Response
Humeros and Las Derrumbadas geothermal fields, Mexico. In: Geothermics, 16,
Developments in Volcanology (Vol. 10, pp. 1-55). Elsevier, Amsterdam.
pp. 539–554. https://doi.org/10.1016/0375-6505(87)90038-1.
Coleman, D.S., Gray, W., Glazner, A.F., 2004. Rethinking the emplacement and evolution
Caricchi, L., Blundy, J., 2015. The temporal evolution of chemical and physical
of zoned plutons: geochronologic evidence for incremental assembly of the
properties of magmatic systems. Geol. Soc. Spec. Publ. 422 (1), 1–15. https://doi.
Tuolumne Intrusive Suite, California. Geology 32 (5), 433–436. https://doi.org/
org/10.1144/SP422.11.
10.1130/G20220.1.
Caricchi, L., Simpson, G., Schaltegger, U., 2014. Zircons reveal magma fluxes in the
Créon, L., Levresse, G., Carrasco-Nuñez, G., Remusat, L., 2018. Evidence of a shallow
Earth’s crust. Nature 511, 457–461. https://doi.org/10.1038/nature13532 htttps://
magma reservoir below Los Humeros volcanic complex: Insights from the
doi.org/10.1038/nature13532.
geochemistry of silicate melt inclusions. J. South Amer. Earth Sci. 88, 446–458.
Carrasco-Núñez, G. Hernández, J., De León, L., Dávila, P., Norini, G., Bernal, J.P., Jicha,
https://doi.org/10.1016/j.jsames.2018.09.017.
B., Navarro, M., & López, P. (2017a). Geologic Map of Los Humeros volcanic
Dahren, B., Troll, V.R., Andersson, U.B., Chadwick, J.P., Gardner, M.F., Jaxybulatov, K.,
complex and geothermal field, eastern Trans-Mexican Volcanic Belt /Mapa geológico
Koulakov, I., 2012. Magma plumbing beneath Anak Krakatau volcano, Indonesia:
del complejo volcánico Los Humeros y campo geotérmico, sector oriental del
Evidence for multiple magma storage regions. Contrib. Mineral. Petrol. 163 (4),
Cinturón Volcánico Trans-Mexicano. In Terra Digitalis. 10.22201/igg.
631–651. https://doi.org/10.1007/s00410-011-0690-8.
terradigitalis.2017.2.24.76.
Davidzon, M.I., 2012. Newton’s law of cooling and its interpretation. Int. J. of Heat and
Carrasco-Núñez, G., Bernal, J.P., Dávila, P., Jicha, B., Giordano, G., Hernández, J., 2018.
Mass Transfer 55, 5397–5402. https://doi.org/10.1016/j.
Reappraisal of Los Humeros Volcanic Complex by New U/Th Zircon and 40Ar/39Ar
ijheatmasstransfer.2012.03.035.
Dating: Implications for Greater Geothermal Potential. Geochem. Geophys. Geosyst.
Davies, J.H., 2013. Global map of solid Earth surface heat flow. Geochem. Geophys.
19 (1), 132–149. https://doi.org/10.1002/2017GC007044.
Geosyst. 14 (10) https://doi.org/10.1002/ggge.20271.
Carrasco-Núñez, G., Gómez-Tuena, A., Lozano, V.L., 1997. Geologic map of Cerro Grande
Dávila-Harris, P., Carrasco-Núñez, G., 2014. An unusual syn-eruptive bimodal eruption:
volcano and surrounding area, central Mexico. Geological Society of America.
the Holocene Cuicuiltic Member at Los Humeros caldera, Mexico. J. Volcanol.
Geological Society of America, Boulder, Colo.
Geotherm. Res. 271, 24–42. https://doi.org/10.1016/j.jvolgeores.2013.11.020.
Carrasco-Núñez, G, López-Martínez, M., Hernández, J., Vargas, V., 2017b. Subsurface
Deb, P., Knapp, D., Marquart, G., Clauser, C., 2019a. Thermal model of the Los Humeros
stratigraphy and its correlation with the surficial geology at Los Humeros
super-hot geothermal system, Mexico. H2020 project GEMex Scientific Report-
geothermal field, eastern Trans-Mexican Volcanic Belt. Geothermics 67, 1–17.
Deliverable D6.3. Zenodo Repository. https://doi.org/10.5281/zenodo.3719193.
https://doi.org/10.1016/j.geothermics.2017.01.001.

20
P. Deb et al. Geothermics 96 (2021) 102162

Deb, P., Knapp, D., Marquart, G., Montegrossi, G., 2019b. Report on the calibrated IEA (2018). 2017 Annual Report IEA Geothermal. July. Internatuonal Energy Agency
reservoir model for the super-hot reservoir at Los Humeros and its calibration against (IEA), Paris. http://iea-gia.org/publications-2/annual-reports/Last accessed –09
available field data. H2020 project GEMex Scientific Report- Deliverable D6.6. January 2021.
Zenodo Repository. https://doi.org/10.5281/zenodo.3719193. Isaia, R., Vitale, S., Marturano, A., Aiello, G., Barra, D., Ciarcia, S., Iannuzzi, E.,
Del Rio, L., L., 1982. Busqueda de zonas con tectonismo extremo, Teoria Y una Tramparulo, F.D., 2019. High-resolution geological investigations to reconstruct the
Aplicacion a la Geotermia: Caldera Los Humeros. Geofisica Internacional 21 (3), long-term ground movements in the last 15 kyr at Campi Flegrei caldera (southern
265–294. Italy). J. Volcanol. Geotherm. Res. 385, 143–158. https://doi.org/10.1016/j.
DePaolo, D.J., 1981. Trace element and isotopic effects of combined wallrock jvolgeores.2019.07.012.
assimilation and fractional crystallization. Earth Planet. Sci. Lett. 53 (2), 189–202. Karlstrom, L., Dufek, J., Manga, M., 2010. Magma chamber stability in arc and
https://doi.org/10.1016/0012-821X(81)90153-9. continental crust. J. Volcanol. Geotherm. Res. 190 (3–4), 249–270. https://doi.org/
Di Renzo, V., Wohletz, K., Civetta, L., Moretti, R., Orsi, G., Gasparini, P., 2016. The 10.1016/j.jvolgeores.2009.10.003.
thermal regime of the Campi Flegrei magmatic system reconstructed through 3D Keller, J., Rath, V., Bruckmann, J., Mottaghy, D., Clauser, C., Wolf, A., Seidler, R.,
numerical simulations. J. Volcanol. Geotherm. Res. 328, 210–221. https://doi.org/ Bücker, H.M., Klitzsch, N., 2020. SHEMAT-Suite: An open-source code for simulating
10.1016/j.jvolgeores.2016.11.004. flow, heat and species transport in porous media. SoftwareX 12. https://doi.org/
Di Vito, M.A., Acocella, V., Aiello, G., Barra, D., Battaglia, M., Carandente, A., Del 10.1016/j.softx.2020.100533.
Gaudio, C., de Vita, S., Ricciardi, G.P., Ricco, C., Scandone, R., Terrasi, F., 2016. Kutasov, I.M., Eppelbaum, L.V., 2005. Determination of formation temperature from
Magma transfer at Campi Flegrei caldera (Italy) before the 1538 AD eruption. Sci. bottom-hole temperature logs—a generalized Horner method. J. Geophys. Eng. 2
Rep. 6 (1), 32245. https://doi.org/10.1038/srep32245. (2), 90–96. https://doi.org/10.1088/1742-2132/2/2/002.
Domra Kana, J., Djongyang, N., Raïdandi, D., Nouck, P.N., Dadjé, A., 2015. A review of Lee, Y., Deming, D., 1998. Evaluation of thermal conductivity temperature corrections
geophysical methods for geothermal exploration. Renew. Sustain. Energy Rev. 44, applied in terrestrial heat flow studies. J. Geophys. Res. 103, 2447–2454. https://
87–95. https://doi.org/10.1016/j.rser.2014.12.026. doi.org/10.1029/97JB03104.
Dowdle, W.L., Cobb, W.M., 1975. Static Formation temperature From Well Logs - an Lee, C.T.A., Lee, T.C., Wu, C.T., 2014. Modeling the compositional evolution of
Empirical Method. J. Petroleum Technol. 27, 1326–1330. https://doi.org/10.2118/ recharging, evacuating, and fractionating (REFC) magma chambers: Implications for
5036-PA. differentiation of arc magmas. Geochim. Cosmochim. Acta 143, 8–22. https://doi.
Edmonds, M., Cashman, K.V., Holness, M., Jackson, M., 2019. Architecture and dynamics org/10.1016/j.gca.2013.08.009.
of magma reservoirs. Philos. Trans. R. Soc., A 377 (2139). https://doi.org/10.1098/ Lee, B., Unsworth, M., Árnason, K, Cordell, D., 2019. Imaging the magmatic system
rsta.2018.0298. beneath the Krafla geothermal field, Iceland: A new 3-D electrical resistivity model
Ellis, B.S., Barry, T., Branney, M.J., Wolff, J.A., Bindeman, I., Wilson, R., Bonnichsen, B., from inversion of magnetotelluric data. Geophys. J. Int. 220 (1), 541–567. https://
2010. Petrologic constraints on the development of a large-volume, high doi.org/10.1093/gji/ggz427.
temperature, silicic magma system: the Twin Falls eruptive centre, central Snake Lejeune, A., Richet, P., 1995. Rheology of crystal-bearing silicate melts: an experimental
River Plain. Lithos 120 (3–4), 475–489. https://doi.org/10.1016/j. study at high viscosity. J. Geophys. Res. 100, 4215–4229. https://doi.org/10.1029/
lithos.2010.09.008. 94JB02985.
Feng, W., Zhu, Y., 2018. Decoding magma storage and pre-eruptive processes in the Lesher, C.E., Spera, F.J., 2015. Chapter 5 - thermodynamic and transport properties of
plumbing system beneath early Carboniferous arc volcanoes of southwestern silicate melts and magma. In: Sigurdsson, E. (Ed.), The Encyclopedia of Volcanoes.
Tianshan, Northwest China. Lithos 322, 362–375. https://doi.org/10.1016/j. Academic Press, pp. 113–141. https://doi.org/10.1016/B978-0-12-385938-
lithos.2018.09.030. 9.00005-5.
Ferriz, H., Mahood, G.A., 1984. Eruption rates and compositional trends at Los Humeros Lipman, P.W., 2007. Incremental assembly and prolonged consolidation of Cordilleran
volcanic center, Puebla, Mexico. J. Geophys. Res. 89 (B10), 8511–8524. https://doi. magma chambers: evidence from the Southern Rocky Mountain volcanic field.
org/10.1029/JB089iB10p08511 https://doi.org/10.1029/JB089iB10p08511. Geosphere 3 (1), 42–70. https://doi.org/10.1130/GES00061.1.
Ferriz, H., Mahood, G., 1987. Strong Compositional Zonation in a Silicic Magmatic Liu, L., Lowell, R.P., 2011. Modeling heat transfer from a convecting, crystallizing,
System: Los Humeros, Mexican Neovolcanic Belt. J. Petrol. 28 (1), 171–209. https:// replenished silicic magma chamber at an oceanic spreading center. Geochem.
doi.org/10.1093/petrology/28.1.171. Geophys. Geosyst. 12 (9) https://doi.org/10.1029/2011GC003612.
Fitz-Díaz, E., Lawton, T.F., Juárez-Arriaga, E., Chávez-Cabello, G., 2018. The Cretaceous- López-Hernández, A., 1995. Estudio Regional Volcánico y Estructural del Campo
Paleogene Mexican orogen: structure, basin development, magmatism and tectonics. Geotérmico de Los Humeros, Puebla, México. Geotermia, Revista Mexicana de
Earth-Sci. Rev. 183, 56–84. https://doi.org/10.1016/j.earscirev.2017.03.002 Geoenergía 11 (1), 17–36.
https://doi.org/10.1016/j.earscirev.2017.03.002. Lorenzo-Pulido, C.D., 2008. Borehole Geophysics and Geology of Well H43, Los Humeros
Fjeldskaar, W., Helset, H.M., Johansen, H., Grunnaleite, I., Horstad, I., 2008. Thermal Geothermal Field, 23. United Nations University, Puebla, Mexico, pp. 387–425.
modelling of magmatic intrusions in the Gjallar Ridge, Norwegian Sea: implications Geothermal Training Programme Report.
for vitrinite reflectance and hydrocarbon maturation. Basin Res. 20 (1), 143–159. Lucci, F., Carrasco-Núñez, G., Rossetti, F., Theye, T., Charles White, J., Urbani, S.,
https://doi.org/10.1111/j.1365-2117.2007.00347.x. Azizi, H., Asahara, Y., Giordano, G., 2020. Anatomy of the magmatic plumbing
Galushkin, Y.I., 1997. Thermal effects of igneous intrusions on maturity of organic system of Los Humeros Caldera (Mexico): Implications for geothermal systems. Solid
matter: A possible mechanism of intrusion. Org. Geochem. 26 (11–12), 645–658. Earth 11 (1), 125–159. https://doi.org/10.5194/se-11-125-2020.
https://doi.org/10.1016/S0146-6380(97)00030-2. Lucci, F., Rossetti, F., Becchio, R., Theye, T., Gerdes, A., Opitz, J., Giordano, G., 2018.
García-Gutiérrez, A., Arellano, V., Barragán, R.M., Espinosa-Paredes, G., 2002. Initial Magmatic Mn-rich garnets in volcanic settings: Age and longevity of the magmatic
temperature field in the Los Humeros geothermal reservoir. Geofisica Internacional plumbing system of the Miocene Ramadas volcanism (NW Argentina). Lithos 322,
41 (3). 238–249. https://doi.org/10.1016/j.lithos.2018.10.016.
Giordano, G., Lucci, F., Calzolari, G., 2019. Report on the volcanological conceptual Magaritz, M., Hofmann, A.W., 1978. Diffusion of Sr, Ba and Na in obsidian. Geochim.
models of Los Humeros and Acoculco. H2020 project GEMex Scientific Deliverable Cosmochim. Acta 42 (6), 595–605. https://doi.org/10.1016/0016-7037(78)90004-
D3.2. https://data.d4science.net/tem2. Last accessed –09- January 2021. 2.
Gómez-Tuena, A., Carrasco-Núñez, G., 2000. Cerro Grande volcano: the evolution of a Magee, C., Stevenson, C.T.E., Ebmeier, S.K., et al., 2018. Magma plumbing systems: a
Miocene stratocone in the early Trans-Mexican Volcanic Belt. Tectonophysics 318 geophysical perspective. J. Petrol. 59 (6), 1217–1251. https://doi.org/10.1093/
(1), 249–280. https://doi.org/10.1016/S0040-1951(99)00314-5. petrology/egy064.
Gudmundsson, A. (2008). Chapter 8- magma-chamber geometry, fluid transport, local Marianelli, P., Sbrana, A., Proto, M., 2006. Magma chamber of the Campi Flegrei
stresses and rock behaviour during collapse caldera formation. In J. Gottsmann & J. supervolcano at the time of eruption of the Campanian Ignimbrite. Geology 34 (11),
Marti (Eds.), Caldera Volcanism Analysis, Modelling and Response Developments in 937–940. https://doi.org/10.1130/G22807A.1.
Volcanology (Vol. 10, pp. 314-346). Elsevier, Amsterdam. https://doi.org/10.1016/ Martı ́nez-Serrano, R.G., 2002. Chemical variations in hydrothermal minerals of the
S1871-644X(07)00008-3. LosHumeros geothermal system. Mexico. Geothermics 31 (5), 579–612. https://doi.
Goutorbe, B., Lucazeau, F., Bonneville, A., 2007. Comparison of several BHT correction org/10.1016/S0375-6505(02)00015-9.
methods: a case study on an Australian data set. Geophys. J. Int. 170 (2), 913–922. Michel, J., Baumgartner, L., Putlitz, B., Schaltegger, U., Ovtcharova, M., 2008.
https://doi.org/10.1111/j.1365-246X.2007.03403.x https://doi.org/10.1111/ Incremental growth of the Patagonian Torres del Paine laccolith over 90 k.y.
j.1365-246X.2007.03403.x. Geology 36 (6), 459–462. https://doi.org/10.1130/G24546A.1.
Gualda, G.A., Ghiorso, M.S., 2013. The Bishop Tuff giant magma body: an alternative to ... Moghadam, H.S., Li, Q.L., Griffin, W.L., Stern, R.J., Ishizuka, O., Henry, H.,
the Standard Model. Contrib. Mineral. Petrol. 166, 755–775. https://doi.org/ Ghorbani, G., 2020. Repeated magmatic buildup and deep “hot zones” in continental
10.1007/s00410-013-0901-6 https://doi.org/10.1007/s00410-013-0901-6. evolution: the Cadomian crust of Iran Earth Planet. Sci. Lett. 531, 115989 https://
Hamada, M., Laporte, D., Cluzel, N., Koga, K.T., Kawamoto, T., 2010. Simulating bubble doi.org/10.1016/j.epsl.2019.115989.
number density of rhyolitic pumices from Plinian eruptions: constraints from fast Nabelek, P.I., Hofmeister, A.M., Whittington, A.G., 2012. The influence of temperature-
decompression experiments. Bull. Volcanol. 72 (6), 735–746. https://doi.org/ dependent thermal diffusivity on the conductive cooling rates of plutons and
10.1007/s00445-010-0353-z. temperature-time paths in contact aureoles. Earth Planet. Sci. Lett. 317, 157–164.
Hersir, G.P., Benediktsdóttir, A., Árnason, K., et al., 2020. Report on the integrated https://doi.org/10.1016/j.epsl.2011.11.009.
geophysical model of Los Humeros and Acoculco. H2020 project GEMex Scientific Norini, G., Carrasco-Núñez, G., Corbo-Camargo, F., Lermo, J., Rojas, J.H., Castro, C.,
Deliverable D5.10. https://data.d4science.net/qNfH 10. Last Accessed- 09 June Bonini, M., Montanari, D., Corti, G., Moratti, G., Piccardi, L., Chavez, G., Zuluaga, M.
2021. C., Ramirez, M., Cedillo, F., 2019. The structural architecture of the Los Humeros
Hofmann, A.W., 2012. Magma chambers on a slow burner. Nature 491 (7426), 677–678. volcanic complex and geothermal field. J. Volcanol. Geotherm. Res. 381, 312–329.
https://doi.org/10.1038/491677a. https://doi.org/10.1016/j.jvolgeores.2019.06.010.
Horner, D.R., 1951. Pressure Build-up in Wells. In: Brill, E.J. (Ed.), Proceedings 3rd Norini, G., Groppelli, G., Sulpizio, R., Carrasco-Núñez, G., Dávila-Harris, P., Pellicioli, C.,
World Petroleum Congress 28 May-6 June. The Hague, the Netherlands. Zucca, F., De Franco, R., 2015. Structural analysis and thermal remote sensing of the

21
P. Deb et al. Geothermics 96 (2021) 102162

Los Humeros Volcanic Complex: Implications for volcano structure and geothermal Sword, 2020. Kuster downhol gauges –K10 Geothermal. http://swordtek.com/yahoo_si
exploration. J. Volcanol. Geotherm. Res. 301, 221–237. https://doi.org/10.1016/j. te_admin/assets/docs/K10_Geothermal.314195737.pdf. Last accessed –09 January
jvolgeores.2015.05.014. 2020.
O’Neill, H.S.C., Jenner, F.E., 2012. The global pattern of trace-element distributions in Suter, M., 1987. Structural traverse across the Sierra Madre Oriental fold-thrust belt in
ocean floor basalts. Nature 491 (7426), 698–704. https://doi.org/10.1038/ east-central Mexico. Geol. Soc. Am. Bull. 98, 249–264. https://doi.org/10.1130/
nature11678. 0016-7606(1987)98<249:STATSM>2.0.CO;2.
Orsi, G., De Vita, S., Di Vito, M.A., Isaia, R., 2005. The Campi Flegrei nested caldera Tarasewicz, J., White, R.S., Woods, A.W., Brandsdóttir, B., Gudmundsson, M.T., 2012.
(Italy): a restless, resurgent structure in a densely populated area. In: Balmuth, M.S., Magma mobilization by downward-propagating decompression of the Eyjafjallajkull
Chester, D.K., Johnston, P.A. (Eds.), Cultural Responses to the Volcanic Landscape: volcanic plumbing system. Geophys. Res. Lett. 39 (19), 1–5. https://doi.org/
The Mediterranean and Beyond. Archaeological Institute of America, Boston, 10.1029/2012GL053518.
pp. 71–90. Torres-Rodriguez, M.A., 1995. Characterization of the Reservoir of the Los Humeros,
Pasquale, V, Verdoya, M., & Chiozzi, P. (2014). Heat flow and geothermal resources in México, Geothermal Field. In: Proceedings of the World Geothermal Congress 1995,
northern Italy. Renew. Sustain. Energy Rev., 36, 277–285. https://doi.org/10.1016/ Florence, Italy, May 18-31, 3, pp. 1561–1567.
j.rser.2014.04.075. Urbani, S., Giordano, G., Lucci, F., Rossetti, F., Acocella, V., Carrasco-Núñez, G., 2020.
Pasquale, VincenzoV., Verdoya, M., Chiozzi, P., 2017. Heat conduction and Estimating the depth and evolution of intrusions at resurgent calderas: Los Humeros
thermophysical parameters. In: Pasquale, V., Verdoya, M., Chiozzi, P. (Eds.), (Mexico). Solid Earth 11 (2), 527–545. https://doi.org/10.5194/se-11-527-2020.
Geothermics Heat Flow in the Lithosphere. SpringerBriefs in Earth Sciences. Urbani, S., Giordano, G., Lucci, F., Rossetti, F., Acocella, V., Carrasco-Núñez, G., 2021.
Springer, Cham, pp. 17–51. https://doi.org/10.1007/978-3-319-52084-1_2, 2nd ed. Structural studies in active caldera geothermal systems. Reply to Comment on
Probe, 2020. Kuster geothermal and ultra-high temperature products. Probe (Adelaide). “Estimating the depth and evolution of intrusions at resurgent calderas: Los Humeros
https://www.probe1.com/product-category/kuster Last accessed –09 January 2021. (Mexico) by Norini and Groppelli (2020). Solid Earth Discussions 1–21. https://doi.
Prol Ledesma, R.M., Gonzalez-Moran, T., 1982. Modelo Preliminar del Regimen Termico org/10.5194/se-2020-218.
Conductivo en la Caldera de Los Humeros, Puebla. Geofisica Internacional 21 (3), Ushioda, M., Takahashi, E., Hamada, M., Suzuki, T., Niihori, K., 2018. Evolution of
295–3047. Magma Plumbing System in Miyakejima Volcano: constraints From Melting
Prol-Ledesma, R.M., 1988. Reporte de los estudios petrograficos y de inclusiones fluidas Experiments. J. Geophys. Res. 123 (10), 8615–8636. https://doi.org/10.1029/
en nucleos de pozos de exploracion en el campo geotermico de Los Humeros, 86. 2018JB015910.
Comunicaciones Tecnicas, Instituto de Geofısica, UNAM, Puebla, Mexico, pp. 1–75. Verma, M.P., Verma, S.P., & Sanvicente, H. (1990). Temperature field simulation with
Prol-Ledesma, R.M., 1998. Pre-and post-exploitation variations in hydrothermal activity stratification model of magma chamber under los humeros caldera, Puebla, Mexico.
in Los Humeros geothermal field, Mexico. J. Volcanol. Geoth. Res. 83, 313–333. Geothermics, 19(2), 187–197. https://doi.org/10.1016/0375-6505(90)90015-4.
https://doi.org/10.1016/S0377-0273(98)00024-9. Verma, S.P., 1984. Alkali and alkaline earth element geochemistry of Los Humeros
Pyle M., D., 1992. The volume and residence times of magma beneath active volcanoes Caldera, Puebla, Mexico. J. Volcanol. Geotherm. Res. 20 (1–2), 21–40. https://doi.
determined by decay-series disequilibria methods. Earth Planet. Sci. Lett. 112, org/10.1016/0377-0273(84)90064-7.
61–73. https://doi.org/10.1016/0012-821X(92)90007-I. Verma, S.P., 1985b. On the magma chamber characteristics as inferred from surface
Roche, O., Druitt, T.H., 2001. Onset of caldera collapse during ignimbrite eruptions. geology and geochemistry: examples from Mexican geothermal areas. Phys. Earth
Earth Planet. Sci. Lett. 191 (3–4), 191–202. https://doi.org/10.1016/S0012-821X Planet. Inter. 41, 207–214. https://doi.org/10.1016/0031-9201(85)90035-4.
(01)00428-9. Verma, Surendra S.P., Gómez-Arias, E., 2013. Three-dimensional temperature field
Roure, F., Alzaga-Ruiz, H., Callot, J.P., Ferket, H., Granjeon, D., Gonzalez-Mercado, G.E., simulation of magma chamber in the Los Humeros geothermal field, Puebla, Mexico.
Guilhaumou, N., Lopez, M., Mougin, P., Ortuno-Arzate, S., Séranne, M., 2009. Long Appl. Therm. Eng. 52 (2), 512–515. https://doi.org/10.1016/j.
lasting interactions between tectonic loading, unroofing, post-rift thermal applthermaleng.2012.12.018.
subsidence and sedimentary transfers along the western margin of the Gulf of Verma, Surendra S.P., Gómez-Arias, E., Andaverde, J., 2011. Thermal sensitivity analysis
Mexico: Some insights from integrated quantitative studies. Tectonophysics 475 (1), of emplacement of the magma chamber in Los Humeros caldera, Puebla, Mexico. Int.
169–189. https://doi.org/10.1016/j.tecto.2009.04.012. Geol. Rev. 53 (8), 905–925. https://doi.org/10.1080/00206810903234296.
Rybach, L., 1976. Radioactive heat production: a physical property determined by the Verma, S.P., & Andaverde, J. (2007). Coupling of thermal and chemical simulations in a
chemistry of rocks. In: Strens, R.G.J. (Ed.), The Physics and Chemistry of Minerals 3-D integrated magma chamber-reservoir model: a new geothermal energy research
and Rocks. Wiley & Sons, pp. 309–318. frontier. In H.I. Ueckermann (Ed.), Geothermal Energy Research Trends (Issue
Schön, J. (2015). Physical Properties of Rocks: Fundamentals and Principles of January 2007, pp. 149–188). Nova Science Publishers, Inc.
Petrophysics (J. Schön (ed.); 1st ed.). Elsevier. Weydt, L.M., Ramírez-Guzmán, Á.A., Pola, A., Lepillier, B., Kummerow, J.,
Sekiguchi, K., 1984. A method for determining terrestrial heat flow in oil basinal areas. Mandrone, G., Comina, C., Deb, P., Norini, G., Gonzalez-Partida, E., Avellán, D.R.,
In: Tectonophysics, 103, pp. 67–79. https://doi.org/10.1016/0040-1951(84)90075- Macías, J.L., Bär, K., Sass, I., 2020. Petrophysical and mechanical rock property
1. database of the Los Humeros and Acoculco geothermal fields (Mexico). Earth Syst.
Selva, J., Acocella, V., Bisson, M., et al., 2019. Multiple natural hazards at volcanic Sci. Data Discuss. 2020, 1–39. https://doi.org/10.5194/essd-13-571-2021.
islands: a review for the Ischia volcano (Italy). J. Appl. Volcanol. 8 (1), 1–43. Wohletz, K., Heiken, G., 1992. Calderas and their Geothermal Systems. K. Wohletz & G.
https://doi.org/10.1186/s13617-019-0086-4. Heiken Volcanology and Geothermal Energy. University of California Press,
Shaw, H.R., 1963. Obsidian-H2O viscosities at 1000 and 2000 bars in the temperature pp. 142–177. http://ark.cdlib.org/ark:/13030/ft6v19p151/.
range 700◦ C to 900◦ C. J. Geophys. Res. 68 (23), 6337–6343. https://doi.org/ Wong-Loya, J.A., Andaverde, J.A., del Rio, J.A., 2015. Improved method for estimating
10.1029/JZ068i023p06337. static formation temperatures in geothermal and petroleum wells. Geothermics 57,
Siebert, L., Simkin, T., Kimberley, P., 2010. Volcanoes of the World, 3rd ed. University of 73–83. https://doi.org/10.1016/j.geothermics.2015.06.002.
California Press, London, England, p. 24. Xiao-Yin, T., Gong-Cheng, Z., Jian-She, L., Shu-Chun, Y., Song, R., Sheng-Biao, H., 2014.
Smith, R.L., & Shaw, H.R. (1975). Assessment of Geothermal Resources of the United Modelling of Thermal Effects of Igneous Intrusions on the Temperature Field and
States - 1975. In D.E. White & D.L. Williams (Eds.), Geological Survey Circular Organic Maturity in the Changchang Sag, Qiongdongnan Basin, South China Sea.
(United States) (Issue 726, pp. 58–83). Chin. J. Geophys. 57, 219–229. https://doi.org/10.1002/cjg2.20098.
Smith, R.L., Shaw, H.R., & Muffler, L.J.P. (1978). Igneous-related geothermal systems: Yañez García, C., García Durán, S., 1982. Exploración de la región geotérmica de Los
assessment of Geothermal Resources of the United States - 1978. In L.J.P. Muffler Humeros–Las Derrumbadas, estados de Puebla y Veracruz: México. Comisión Federal
(Ed.), Geological Survey Circular (United States) (Issue 790). de Electricidad, pp. 1–96. Technical Report.
Stimac, J.A., Goff, F., Wohletz, K., 2001. Thermal modeling of the Clear Lake magmatic-
hydrothermal system, California, USA. Geothermics 30 (2–3), 349–390. https://doi.
org/10.1016/S0375-6505(00)00062-6.

22

You might also like