You are on page 1of 7

Journal of Volcanology and Geothermal Research 305 (2015) 56–62

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

journal homepage: www.elsevier.com/locate/jvolgeores

Short communication

Thermal history of the Acoculco geothermal system, eastern Mexico:


Insights from numerical modeling and radiocarbon dating
Carles Canet a,⁎, Frederic Trillaud b, Rosa María Prol-Ledesma a, Galia González-Hernández a, Berenice Peláez a,
Berenice Hernández-Cruz a, María M. Sánchez-Córdova c
a
Instituto de Geofísica, Universidad Nacional Autónoma de México, Del. Coyoacán, 04150 México D.F., Mexico
b
Instituto de Ingeniería, Universidad Nacional Autónoma de México, Del. Coyoacán, 04150 México D.F., Mexico
c
Posgrado en Ciencias de la Tierra, Universidad Nacional Autónoma de México, Del. Coyoacán, 04150 México D.F., Mexico

a r t i c l e i n f o a b s t r a c t

Article history: Acoculco is a geothermal prospective area hosted by a volcanic caldera complex in the eastern Trans-Mexican
Received 10 July 2015 Volcanic Belt. Surface manifestations are scarce and consist of gas discharges (CO2-rich) and acid-sulfate springs
Accepted 14 September 2015 of low temperature, whereas hydrothermal explosive activity is profusely manifested by meter-scale craters and
Available online 25 September 2015
mounds of hydrothermal debris and breccias. Silicic alteration extends for several square kilometers around the
zone with gas manifestations and explosive features, affecting surficial volcanic rocks, primarily tuffs and
Keywords:
Heat flow regime
breccias. In the subsurface, an argillic alteration zone (ammonium illite) extends down to a depth of ∼600 m,
Water-rock interaction and underneath it a propylitic zone (epidote–calcite–chlorite) occurs down to ∼1000 m.
1D finite difference analysis Thermal logs from an exploratory borehole (EAC-1, drilled in 1995 down to 1810 m) showed a conductive heat
Sealed geothermal reservoir transfer regime under high geothermal gradient (∼140 °C/1000 m). In contrast, the thermal profile established
Geothermal exploration from temperatures of homogenization of fluid inclusions—measured on core samples from the same drill
Hydrothermal explosions hole—suggests that convection occurred in the past through the upper ~1400 m of the geothermal system. A
drop in permeability due to the precipitation of alteration minerals would have triggered the cessation of the
convective heat transfer regime to give place to a conductive one. With the purpose of determining when the
transition of heat transfer regime occurred, we developed a 1D model that simulates the time-depth distribution
of temperature. According to our numerical simulations, this transition happened ca. 7000 years ago; this date is
very recent compared to the lifespan of the geothermal system. In addition, radiocarbon chronology indicates
that the hydrothermal explosive activity postdates the end of the convective heat transfer regime, having
dated at least three explosive events, at 4867–5295, 1049–1417 and 543–709 y cal. BP. Therefore, hydrothermal
explosions arise from the self-sealing of the Acoculco geothermal system, involving a natural hazard that could
affect future geothermal-power infrastructure.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction changes allow to reconstruct the thermal history of geothermal systems


(e.g. Dalrymple et al., 1999), which in turn can be used to ascertain at
Geothermal fields are complex and inherently dynamic systems that what evolutionary stage a particular system is, thus determining the
can evolve over short periods of time (cf. Barbier, 1997). On the one potentiality and type of a current geothermal play (e.g. Stimac et al.,
hand they can undergo physical changes as a consequence of exploita- 2001). Therefore the knowledge of the thermal evolution of geothermal
tion operations, e.g. changes in fluid flow and temperature distribution. systems offers a useful approach that can assist in exploring and evalu-
These changes take place on a time scale of years and have direct effects ating potential geothermal resources.
on the efficiency and sustainability of geothermal power exploitation, Numerical models such as those by Norton and Knight (1977) and
explaining thereby why they are subject of predictive modeling studies by Norton and Hulen (2001) proved that cooling of a magmatic source
(e.g. Rybach and Mongillo, 2006; Axelsson, 2010; Lei and Zhu, 2013). On by pure conductive heat transfer is a process that can take millions of
the other hand there are changes that geothermal systems underwent years. However hydrothermal circulation is able to generate enough
prior to exploitation, i.e. in relation to their geological evolution and convective heat fluxes to shorten drastically the cooling time. Rock
on a time scale from thousands to millions of years (e.g. Norton and permeability is the key factor determining the dominating mechanism
Knight, 1977). The understanding and timing of these ‘geological’ of heat transfer in a geothermal system. Indeed, if the threshold value
of 0.001 millidarcy (10− 8 m2) is exceeded, convective heat flow
⁎ Corresponding author. Tel.: +52 55 56224133; fax: +52 55 55502486. prevails (Norton and Knight, 1977). Permeability of rocks overlying a
E-mail address: ccanet@geofisica.unam.mx (C. Canet). magmatic heat source may change over time similarly to the heat flow

http://dx.doi.org/10.1016/j.jvolgeores.2015.09.019
0377-0273/© 2015 Elsevier B.V. All rights reserved.
C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62 57

regime. Changes of permeability in geothermal systems obey primarily overall thickness of ∼300 m. Simultaneous with caldera formation ba-
fluid-rock interactions. Alteration minerals are the rock-preserved saltic andesite was erupted outside the caldera structure; subsequently
evidence of these interactions and their formation involves the convec- (1.0–0.24 Ma) basalts were extruded, covering partially the caldera
tion of huge masses of high-temperature hydrothermal fluids (e.g. products (López-Hernández et al., 2009).
Moore and Gunderson, 1995). However, the very process of crystalliza-
tion of alteration minerals may result in the occlusion of permeability, 2.2. The Acoculco geothermal system
preventing the convection of hydrothermal fluids and causing the self-
sealing of the geothermal system (Facca and Tonani, 1967; Muffler Surface manifestations in Acoculco consist of gas discharges
et al., 1971; Norton and Hulen, 2001). and bubbling, acid-sulfate springs of low temperature (Fig. 2A), concen-
Acoculco is a geothermal prospective area in the state of Puebla, trated at two zones separated from each other by about 1750 m. In
Mexico, in which to date there has been no progress toward commercial such manifestations CO2 is the predominant gas and the 3He/4He ratio
exploitation. Thermal logs of an exploratory borehole drilled in 1995 by (R/Rair) attains up to 8.5; this value is among the highest reported for
the Comisión Federal de Electricidad—the Mexican state-owned the whole TMVB and suggests an active magmatic source (Polak et al.,
electricity supplier—in Acoculco (well EAC-1) show a conductive heat 1982). Besides that, there are meter-sized craters with associated
transfer regime under high geothermal gradient (López-Hernández mounds of sand- to gravel-sized debris of altered volcanic rock that
et al., 2009); that is why this area has been regarded as a candidate bear witness to recent, explosive hydrothermal activity (Fig. 2B and
site for future hot-dry-rock development and has been virtually ruled C). These geothermal landforms are found only at the northernmost
out for a conventional one (Pulido et al., 2011). According to Lopez- zone with gas emission (referred to as Alcaparrosa) and fossilize
Hernandez and Castillo-Hernandez (1997), Acoculco may be a case of charcoal-rich paleosols (Fig. 2C). Beneath the debris deposits there is a
self-sealed geothermal field, for which fracture filling by precipitation ~ 15 cm thick horizon of fine grained (average: 5.3 ϕ = 25 μm),
of alteration minerals produced a general decrease in rock permeability, unconsolidated particles of alteration minerals, likely deposited as a
resulting in a ∼ 500 m-thick lithocap that disabled the convective hydrothermal explosion ash fall. In addition, a land-slide scarp exposing
regime. In contrast, fluid inclusion analyses done on core samples pervasively altered volcanic rocks that occurs further up-slope from the
from the same drill hole gave temperatures of homogenization (TH) craters might have been triggered by violent hydrothermal explosions.
systematically above—at equivalent depths—those of thermal log Perhaps the most striking surface feature of Acoculco is hydrother-
(López-Hernández et al., 2009), but consistent with the temperatures mal alteration, which extends for an area of several square kilometers
indicated by the alteration mineralogy (Canet et al., 2015), describing by the center of the caldera complex (Fig. 1), affecting tuffs, breccias
a tortuous geothermal profile with conspicuous temperature peaks. and, to a lesser degree, lavas (Canet et al., 2015). Silicic alteration is
Therefore, fluid inclusions and alteration minerals in Acoculco, taken the most widespread type of alteration that undergone the surficial
altogether, bear witness to a convective geothermal regime that might rocks. This alteration occurs as a pervasive replacement of the pyroclas-
have occurred under a higher geothermal gradient (Canet et al., 2015). tic deposits by an assemblage of opal, tridimite and anatase, developed
To understand the thermal evolution of Acoculco and therefore under temperature conditions below ~ 150 °C and near-neutral pH
to evaluate its geothermal potential, it is important to determine (Canet et al., 2015). Advanced argillic alteration has a much more
when the transition of heat transfer regime—from convective to restricted distribution in Acoculco, occurring principally nearby the ac-
conductive—occurred. To that end we developed a 1D model simulating tive gas manifestations. This type of alteration is expressed by secondary
the time-depth distribution of temperature. The output curves are mineral assemblages rich in kaolinite and sulfates (alunite and
compared with the experimental data provided by fluid inclusions and ammoniojarosite), and it could be the consequence of a steam-heated
by the borehole thermal log. The time estimates obtained by such an overprint in the geothermal system. Advanced argillic overprints devel-
approach are then compared with radiocarbon dates of hydrothermal op above the water table associated with the oxidation of H2S in steam
explosion features, which were reported by Canet et al. (2015) and condensate, under low pH (2–3) and at near 100 °C (e.g. White and
represent the most recent manifestation of hydrothermal fluids Hedenquist, 1995).
reaching the ground surface. Concerning the subsurface rocks, two major zones of alteration can
be distinguished according to Canet et al. (2010): (a) a shallow zone
2. Study area with ammonium illite, and (b) a deeper zone with an alteration assem-
blage of epidote–calcite–chlorite. The shallow zone extends to a depth
2.1. Geological setting: the Acoculco caldera of 500–600 m below ground surface (m bgs), and consists of pervasive
ammonium-argillic alteration of the ignimbrites and lavas, indicating
The Acoculco geothermal system is hosted by a volcanic caldera temperatures above 200 °C. The deeper zone was recognized down to
complex that extends over the Puebla–Hidalgo state boundary, at the at least ∼1000 m bgs. It shows a typical propylitic assemblage of alter-
eastern portion of the Trans-Mexican Volcanic Belt (TMVB) (Fig. 1). ation suggesting temperatures of ∼240 °C. Underneath, metamorphism
This complex is ∼ 18 km in diameter and contains up to ∼ 900 m of of limestones produced marbles as well as skarns with wollastonite,
Pliocene to Pleistocene calcalkaline volcanic rocks overlying a Mesozoic garnet and diopside (López-Hernández et al., 2009); this calc-silicate
sedimentary basement (López-Hernández et al., 2009). The basement paragenesis suggests temperatures above ~350 °C (e.g. Einaudi, 1982).
rocks include, from base to top: (a) Jurassic sandstones with interbed- The geothermal interest of Acoculco, in line with the above consider-
ded pelites, biopelites and hydrocarbon-rich limestones and dolomites, ations, led the Comisión Federal de Electricidad to drill two exploratory
conformably overlain by (b) Cretaceous calcarenites and reef lime- wells: EAC-1 in 1995 and EAC-2 in 2008, attaining depths of 1810
stones, with minor pelites and siltstones (Viniegra-Osorio, 1965; and 1900 m bgs, respectively (cf. Viggiano-Guerra et al., 2011).
Morales and Garduño, 1984). A granitic intrusion of presumed Late The two wells were drilled very close to each other, being located
Cretaceous age metamorphosed the sedimentary section beneath the both in the southernmost zone with active gas emission, known as Los
caldera complex (Campos-Enríquez et al., 2003). Azufres (Fig. 1). According to the review by López-Hernández et al.
The Pliocene to Pleistocene volcanic rocks are, for the most part, (2009), the bottom-hole temperature of EAC-1, measured after well
products of the caldera, resulting from two main periods of volcanic ac- completion, attained 307 °C. This implies a geothermal gradient of
tivity (López-Hernández et al., 2009). The older episode (3.0–2.6 Ma) ~140 °C/1000 m, which is about three times the average for the TMVB
produced up to ∼600 m of dacitic to rhyodacitic lavas and pyroclastic (Ziagos et al., 1985). On the other hand, the downhole temperature
deposits. The earlier episode occurred at 1.7–1.26 Ma producing rhyolit- profile indicates a pure conductive heat transfer regime (Fig. 3; López-
ic domes, ignimbrite deposits and minor dacite lava flows, with an Hernández et al., 2009).
58 C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62

Fig. 1. Geological map of the Acoculco volcanic complex, eastern Mexico, with the location of the exploratory well EAC-1 (19°55′16.30″ N/98°08′34.85″ W) and of the two main zones with
manifestations of geothermal activity (i.e. bubbling, acid-sulfate warm springs); the surface distribution of hydrothermal alteration minerals is also shown (after López-Hernández et al.,
2009; Canet et al., 2015). TMVB, Trans-Mexican Volcanic Belt.

3. Methods integration of all the boundary conditions under the main assumption
of a pure conduction heat transfer, to solve the following heat balance
3.1. Modeling constraints, experimental data and computational methods equation:

To assess the evolution of geothermal activity in Acoculco, the time ∂T ∂T2


¼ Dth 2 ð1Þ
and 1D spatial (vertical) temperature distribution has been numerically ∂t ∂z
computed using the Finite Difference Method (Fig. 3). This has been
done from the surface down to 1800 m bgs, corresponding to the where T is the temperature depending on time (t) and depth (z), and Dth
depth interval over which experimental data are available from the is the thermal diffusivity, which also depends on z and is defined as
exploratory well EAC-1. Two sets of experimental data are used for follows:
modeling the variation with time of geotherms i.e. depth-temperature
profiles: (a) TH of fluid inclusions, and (b) borehole temperature k
Dth ¼ ð2Þ
log values. For modeling purposes, the former were chosen to ρc
plot the initial profile (time = 0), while the latter represent the
current geotherm—or final profile—for the same location (19°55′ where k is the thermal conductivity, c the specific heat capacity and
16.30″N/98°08′34.85″W). The microthermometric analysis of fluid in- ρ the mass density. The model used k as a numerical parameter,
clusions was done by López-Hernández et al. (2009) in calcite and whereas c and ρ are experimental values (given in Table 1). Thus,
quartz crystals from core samples recovered from the exploratory drill depth-temperature profiles were obtained at different times.
hole. With regard to the borehole temperature log, the considered The proposed model assumes an underground layered structure
data were those measured after well completion. such that each layer is isotropic and has constant material properties
The two experimental curves are extrapolated by the model down to (summarized in Table 1). Such stratified model is consistent with the
4000 m bgs, which is the depth suggested by López-Hernández et al. geological characteristics revealed by the exploratory well EAC-1,
(2009) for the magmatic heat source. This extrapolation allows the which intersected ~ 1100 m of the meta-sedimentary and plutonic
C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62 59

Fig. 2. Surface manifestations of hydrothermal activity in the Acoculco geothermal system (Alcaparrosa zone). (A) Bubbling pool of warm, acid-sulfate water; (B) crater formed as a result
of a hydrothermal explosion on a landslide deposit; (C) hummocky deposit of sand- to gravel-sized debris of hydrothermally altered volcanic rocks; the top-right inset shows a close-up
view of a charcoal-rich paleosol horizon found at the basis of such deposit.

basement and the entire caldera sequence (~900 m). Thermal proper- there is no heat source (or sink) besides that of magmatic nature that
ties of rocks were taken from literature (Eppelbaum et al., 2014 and is supposed to be at depths greater than 4000 m bgs.
references therein), but volumetric mass density was measured in Even though the boundary and initial conditions, under the above
core and surface samples through the water displacement method assumptions, are sufficient to solve the Eq. (1), by themselves they do
(Table 1). Besides, the model assumes fixed temperature boundary not allow simulating the variation of geotherms between the initial
conditions of 15 °C at the surface and of 750 °C at a depth of 4000 m and the final thermal profiles. To get the best results that fit with the
bgs; the former value corresponds to the annual average temperature experimental depth-temperature curves, an appropriate set of material
of the region (Peláez Pavón, 2015), whereas the latter is the presumed properties for each layer (k, c′ and ρ), compatible with the reported
temperature of the heat source considered as a silicic magma (e.g. lithologies, must be chosen. Given that k is directly related to the ther-
Scaillet et al., 1998). An additional assumption of the model is that mal gradient, we considered it as an optimal parameter choice among

Fig. 3. Temperature–depth profiles for the exploratory well EAC-1, Acoculco geothermal system. Profiles shown on the right correspond to experimental data (compiled by López-
Hernández et al., 2009): temperatures of homogenization (TH) of fluid inclusions and log temperatures measured after well completion. To the left is shown the modeled variation of
geotherm curves with time, considering that the initial profile is that given by fluid inclusions and the final one corresponds to the temperature log. The elapsed time between the initial
and final profiles is, according to the numerical simulation, of 7000 years.
60 C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62

Table 1
Thermal proprieties, measured mass density and mineral composition of selected samples of hydrothermally altered volcanic rocks from the Acoculco geothermal zone.

# Sample Depth Volcanic rock type/description of alterationa Mineralsb Density Thermal Heat capacityc
(m) (kg/m3) conductivityc (J/(kg·K))
(W/(m·K))

1 AC-24b 0 Breccia with pervasive silicic alteration Op, Q, Ana (Ill) 2124 1.65 840
2 EAC-2 100 Ignimbrite with pervasive argillic (ammonium) alteration Q, Ill-NH4 (Kao, Py) 2341 1.65 840
3 EAC-1 N-2 325 Welded lapilli tuff, slightly altered (only the mafic minerals) Pl, Hb, Aug, Bio (Chl, Sme, Op, Cal) 2295 1.65 840
4 EAC-1 N-3 602 Dacite lava, with argillic alteration Q, Pl, FK, Ill (Py) 2290 1.87 840
5 EAC-1 N-4 825 Coarse ash tuff with propylitic alteration Pl, Q, Cal, Chl (Kao, Py) 2698 1.65 840
6 EAC-2P 1520 Marble, unaltered Cal (Q) 2359 2.37 890
7 EAC-1 N-6 1815 Granite, slightly altered Pl, FK, Q, Hb, Bio (Chl) 2761 2.68 950
a,b
From Canet et al. (2010, 2015).
b
Obtained by XRD and SWIR reflectance spectrometry. In order of abundance; minerals in trace amounts are given in parentheses. Key: Ana, anatase; Aug, augite; Bio, biotite; Cal, calcite;
Chl, chlorite; FK, K-feldspar; Hb, hornblende; Ill, illite (Ill-NH4, ammonium illite); Kao, kaolinite; Op, opal; Pl, plagioclase; Py, pyrite; Q, quartz; Sme, smectite s.l.
c
From Eppelbaum et al. (2014).

the material properties. A heuristic procedure, seeking to get the best dissolved in dead spectrophotometric-grade benzene) in 3-mL Teflon®
accuracy in the reproduction of the two experimental profiles, allowed vials in order to reduce the background signal. Analyses were performed
us to establish a set of k values suitable for the stratified model. These in a Quantulus 1220 ultra-low-level liquid scintillation spectrometer.
values, available in Table 2, are within the range found in the literature Each sample was analyzed for 2500 min distributed in 50 cycles, alter-
for the different rock types intersected by the EAC-1 exploratory well, as nating sample vials with Oxalic acid II (SRM 4990C) standard and back-
shown in Table 1. ground vials. The counting window was set to optimize the figure
In order to solve the diffusion Eq. (1) and to produce the simulated of merit with a 14C counting efficiency higher than 65% and the
geotherms depicted in Fig. 3, a program was written using the Python background less than 0.2 CPM/g C. The conventional radiocarbon date
scripting language (Python Software Foundation, n.d). The code is was calibrated using the Calib 7.0 software (Stuiver and Reimer, 1993)
available in a supplementary file and it may be freely used and/or with the calibration curve IntCal_13 (Reimer et al., 2013).
modified.
4. Results
3.2. Radiocarbon measurements
4.1. Thermal profiles: comparison and model results
Four soil samples and one charcoal sample were dated at the
Laboratorio Universitario de Radiocarbono, Universidad Nacional
Two distinct geotherms, representing different ages, arise from ex-
Autónoma de México (UNAM) by liquid scintillation spectrometry,
perimental data (Fig. 3). The initial or former profile is that defined by
following standard procedures reported in detail in Beramendi-Orosco
fluid-inclusion TH values; in this case the variation of temperature
et al. (2006). Sample descriptions and provenance are provided in
with depth is not linear, showing two conspicuous maxima of 178°
Table 3. Prior to benzene synthesis, soil samples were dried at 50 °C
and 282 °C at 400 and 1200 m bgs, respectively. Contrastingly, the
and sieved to pass mesh 10 (2 mm), before inspection under the micro-
final or current profile, which was obtained from the borehole thermal
scope in order to remove rootlets and other macro-contaminants. After
log, shows an almost constant, negative slope indicating a geothermal
this, samples were washed with distilled water at 40 °C for 24 h follow-
gradient of ~ 140 °C/1000 m. The offset between the two curves is
ed by acid wash with 1 M hydrochloric acid (HCl) at 50 °C for 24 h, then
more pronounced at the depths of the TH maxima (400 and 1200 m
neutralized by washing with distilled water and dried at 50 °C. The
bgs) but becomes negligible from 1400 m bgs and below.
charcoal sample was dried and inspected under the microscope in
The simulated geotherms produced by the 1D model for times among
order to remove rootlets and other contaminants prior to the Acid-
1000 and 9000 years vary progressively between the two experimental
Alkali-Acid (AAA) pretreatment. The AAA method consists in treating
profiles, being the curve of 7000 years which has the best fit to the
samples with 1 M HCl at 50 °C for 24 h, 0.1 M sodium hydroxide
final profile (Fig. 3). Thence, the geothermal gradient progressively de-
(NaOH) at 50 °C for 4 h, and 1 M HCl at 50 °C for 24 h. Samples were neu-
creases (down to ~90 °C/1000 m not considering the extrapolated seg-
tralized by washing with distilled water and dried at 50 °C.
ment) with increasing time until reaching the steady state regime,
Prior to analysis, all samples were transformed to benzene (1.5 mL)
under a pure conduction heat transfer, at 250,000 y (after the initial
in a vacuum synthesis line and mixed with 0.5 mL of scintillation
profile); the slope of these curves, although always negative, varies
cocktail (2,5-diphenyloxazole + 1,4-Bis(5-phenyl-2-oxazolyl) benzene
with depth in response to the underground layered structure (Fig. 3).

4.2. Hydrothermal explosions: radiocarbon dating


Table 2
Numerical parameters used to reproduce the geothermal profiles as a function of time
depicted in Fig. 3.
The age of different deposits and landforms that is related to hydro-
thermal explosions was constrained by radiocarbon dating of paleosol
Depth (m) k (W/m-°C) c (J/kg-K) ρ (kg/m3) Dth (10−7 · m2/s) horizons (Table 3). The oldest radiocarbon age (conventional date) is
0 to 50 1.84 840 2120 10.33 4440 ± 80 BP and corresponds to an ash-fall horizon likely deposited
50 to 250 1.41 840 2340 7.17 by a hydrothermal eruption. Moreover, two deposits of sand-sized
250 to 400 1.36 840 2300 7.04
hydrothermal debris, separated each other by 250 m on the footslope,
400 to 600 1.44 840 2300 7.45
600 to 800 1.46 840 2290 7.59 gave ages of 1230 ± 65 and 1380 ± 80 BP; the latter was confirmed
800 to 1000 1.68 840 2700 7.41 with the 14C analysis of a charcoal fragment separated from the same
1000 to 1500 1.84 890 2700 7.66 horizon (Table 3). The youngest hydrothermal event in the zone,
1500 to 1900 1.66 890 2360 7.90 dated to 680 ± 65 BP, produced conical mounds of gravel-sized hydro-
N1900 2.10 950 2760 8.01
thermal debris preserved on the toeslope (Fig. 2C).
C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62 61

Table 3
Radiocarbon ages of paleosol horizons fossilized by hydrothermal explosions in the Acoculco geothermal zone. A close−up photograph of the paleosol UNAM-1421 is provided in Fig. 2C
(top-right inset).
14
Sample Location Description Total C C age Calibrated range (2σ)

% (BP ± 1σ) cal. BC/AD cal. BP Pr. %

UNAM-1419 19°56′17.15″N Paleosol underneath a deposit of sand-sized hydrothermal debris 2.2 1380 ± 80 AD 533–778 1172–1417 93.4
98°08′35.23″O
UNAM-1418 19°56′17.15″N Charcoal fragment from the paleosol horizon UNAM-1419 – 1370 ± 70 AD 540–778 1172–1410 97.9
98°08′35.23″O
UNAM-1420 19°56′19.04″N Paleosol underneath a deposit of sand-sized hydrothermal debris 1.1 1230 ± 65 AD 663–901 1049–1287 94.3
98°08′27.20″O
UNAM-1421 19°56′15.79″N Paleosol underneath a deposit of gravel-sized hydrothermal debris 5.4 680 ± 65 AD 1241–1407 543–709 98.2
98°08′34.44″O
UNAM-1422 19°56′17.24″N Paleosol underneath a silt-sized horizon of hydrothermal ash 4.3 4440 ± 80 BC 3346–2918 4867–5295 100.0
98°08′25.28″O

5. Discussion exceeds the strength of the overlying rock violently disrupting them;
an effective mechanism that can lead to a hydrothermal explosion is a
The shape of the geotherm established from fluid inclusions (‘initial sudden release in pressure, which in turn can be triggered by an earth-
profile’ in Fig. 3), with marked TH maxima at 400 and 1200 m bgs that quake or by hydrothermal fracturing (Christiansen et al., 2007).
may be taken as indicators for thermal paleoaquifers, suggests that In Acoculco, recent hydrothermal explosive activity is manifested in
convection has occurred in the past through the upper ~1400 m of the small craters and mounds of hydrothermal debris and breccias, formed
geothermal system of Acoculco. Furthermore, hydrothermal alteration in a zone with active gas emission where the surface volcanic rocks are
is much more intense and pervasive in the same (upper) portion of pervasively altered (Fig. 2). The incidence of hydrothermal explosions in
the system (Canet et al., 2015), which can also be interpreted as Acoculco is consistent with two defining characteristics of the present-
evidence of convection (e.g. Norton and Hulen, 2001). Contrastingly, day geothermal field: (a) an active magmatic heat source that holds
at depths greater than 1400 m bgs, the TH geotherm becomes roughly high geothermal gradient; and (b) a very low permeability in the
linear and overlaps with the current geotherm (i.e. which is given by volcanic rocks of the caldera sequence due to hydrothermal alteration.
the borehole thermal log; plotted as ‘final profile’ in Fig. 3), indicating The radiocarbon dating of paleosols revealed that the hydrothermal
that: (a) at such depths conduction is the only operating heat transfer explosive activity in Acoculco postdates the end of the convective heat
mechanism, both for the present and the past, and (b) fixing tempera- transfer regime, which happened, according to our 1D numerical
ture boundary conditions at depth is a reasonable assumption for the modeling, ca. 7000 years ago. At least three events of this activity are
1D model. recorded in the studied paleosols, dated to 4867–5295, 1049–1417
According to our numerical simulation, the elapsed time between and 543–709 y cal. BP (Table 3). Accordingly, our general chronology
the initial and final profiles is of 7000 y; put another way, the transition of events, constructed from the results of numerical simulations plus
14
of heat transfer regime from convective to conductive in Acoculco likely C ages, indicates that hydrothermal explosions are consequence of
happened ca. 7000 years ago. This time lapse is too short to allow signif- the self-sealing of the geothermal system, which is consistent with the
icant changes in the deep, magmatic heat source (cf. Norton and Knight, models of Muffler et al. (1971). Furthermore, we can deduce from this
1977), but it could reasonably account for rapid changes in the upper chronology that (a) the change in heat transfer regime and, thus, in
portion of the geothermal system. In the case of Acoculco, a sudden the way to dissipate the excess thermal energy to the surface, occurred
drop in permeability due to the precipitation of alteration minerals rapidly and recently compared to the lifespan of the geothermal system;
would have triggered the cessation of the convective heat transfer and (b) the period of recurrence for hydrothermal explosions could be
regime to give place to a conductive one (López-Hernández et al., within hundreds to a few thousands of years, implying a natural hazard
2009; Canet et al., 2015). A noticeable effect of these rapid changes that should not be dismissed for any future geothermal-power
is a decrease in the local geothermal gradient, manifested by an development.
offset between the two experimental geotherms (Fig. 3). This effect
only affects the rocks of the upper portion of the geothermal system 6. Conclusions
(~ 0–1400 m bgs), which began to cool when stopped the interaction
with hydrothermal fluids (ca. 7000 years ago). Without convection, Convection in Acoculco has occurred in the past through the upper
the cooling rate of the magmatic heat source may be decreased by an ~1400 m of the geothermal system. A drop in permeability due to the
order of magnitude (Norton and Knight, 1977). precipitation of alteration minerals triggered the cessation of convective
Hydrothermal explosions occur in areas with modern geothermal heat transfer to give place to a conductive regime. According to 1D
activity and produce a variety of features such as craters of variable numerical simulations, this transition of heat transfer regime happened
diameter (from several to hundreds of meters), buttes, and deposits of ca. 7000 years ago; this date is very recent in comparison with the
debris and breccias of hydrothermally altered rocks (Muffler et al., lifespan of the geothermal system.
1971 and references there in). They take place when hot water (over Hydrothermal explosive activity is manifested in Acoculco by meter-
~250 °C) that is contained in the rocks of the upper portion of a geother- scale craters and mounds of hydrothermal debris and breccias. These
mal system boil and flash to steam, disrupting the confining rocks and features occur in a zone with thermal gas manifestations where the
ejecting rock fragments, mud, water and steam (Muffler et al., 1971). surface volcanic rocks show pervasive alteration. Hydrothermal
However, in conventional geothermal systems heat is efficiently explosions in Acoculco are consequence of the self-sealing of the
dissipated by hot-springs and/or geysers, thereby avoiding the geothermal system: upon cessation of convection, excess thermal
occurrence of major hydrothermal explosions. It is expected, therefore, energy accumulates triggering sudden events of flash boiling that
that these explosions occur in geothermal systems with very low violently disrupt the confining rocks.
rock permeability (Muffler et al., 1971). Under this scenario, where Radiocarbon chronology confirms that the hydrothermal explosive
convection does not occur, energy brought from a deep heat source activity in Acoculco postdates the end of the convective heat transfer
accumulates in the upper portion of the system until fluid pressure regime, having dated at least three explosive events, at 4867–5295,
62 C. Canet et al. / Journal of Volcanology and Geothermal Research 305 (2015) 56–62

1049–1417 and 543–709 y cal. BP. Hence it follows that the period of Eppelbaum, L., Kutasov, I., Pilchin, A., 2014. Thermal properties of rocks and density of
fluids. In: Eppelbaum, L., Kutasov, I., Pilchin, A. (Eds.), Applied Geothermics. Springer,
recurrence for hydrothermal explosions could be within hundreds to a Berlin Heidelberg, pp. 99–149 http://dx.doi.org/10.1007/978-3-642-34023-9_2.
few thousands of years, implying a natural hazard that could affect Facca, G., Tonani, F., 1967. The self-sealing geothermal field. Bull. Volcanol. 30, 271–273.
future geothermal-power infrastructure. Lei, H., Zhu, J., 2013. Numerical modeling of exploitation and reinjection of the Guantao
geothermal reservoir in Tanggu District, Tianjin, China. Geothermics 48, 60–68.
Lopez-Hernandez, A., Castillo-Hernandez, D., 1997. Exploratory drilling at Acoculco,
Acknowledgments Puebla, Mexico: a hydrothermal system with only nonthermal manifestations.
Geotherm. Resour. Counc. Trans. 21, 429–433.
López-Hernández, A., García-Estrada, G., Aguirre-Díaz, G., González-Partida, E., Palma-
Funding was provided by the project 151453 (Fondo Mixto Guzmán, H., Quijano-León, J.L., 2009. Hydrothermal activity in the Tulancingo–
Conacyt — Gobierno del Estado de Hidalgo). This research was framed Acoculco Caldera Complex, central Mexico: exploratory studies. Geothermics 38,
by the Red de Experimentos en Laboratorios Subterráneos (DGECI, 279–293.
Moore, J.N., Gunderson, R.P., 1995. Fluid inclusion and isotopic systematics of an evolving
UNAM), led by Juan Carlos d'Olivo (ICN, UNAM). Lucy Mora and Kumiko
magmatic-hydrothermal system. Geochim. Cosmochim. Acta 59, 3887–3907.
Shimada are thanked for the total carbon analyses of soil samples, and Morales, G.J., Garduño, M.V.H., 1984. Estudio tectónico-estructural en el prospecto
Faustino Juárez for the rock density measurements. The authors would Huauchinango, Pue. Internal report. Instituto Mexicano del Petróleo, Mexico.
like to thank Guillermo López Flores and Sergio Salinas for their valuable Muffler, L.J.P., White, D.E., Truesdell, A.H., 1971. Hydrothermal explosion craters in
Yellowstone National Park. Bull. Geol. Soc. Am. 82, 723–740.
assistance during fieldwork. Norton, D.L., Hulen, J.B., 2001. Preliminary numerical analysis of the magma-
hydrothermal history of the Geysers geothermal system, California, USA.
Appendix A. Supplementary data Geothermics 30, 211–234.
Norton, D., Knight, J., 1977. Transport phenomena in hydrothermal systems: cooling
plutons. Am. J. Sci. 277, 937–981.
Supplementary data to this article can be found online at http://dx. Peláez Pavón, L.B., 2015. Anális físico-geográfico de la Caldera de Acoculco, Puebla.
doi.org/10.1016/j.jvolgeores.2015.09.019. Unpublished B. Thesis in Geography, Facultad de Filosofía y Letras, Universidad
Nacional Autónoma de México, 101 pp.
Polak, B.G., Prasalov, E.M., Kononov, V.I., Verkovsky, A.B., González, A., Templos, L.A.,
References Espíndola, J.M., Arellano, J.M., Mañón, A., 1982. Isotopic composition and concentra-
tion of inert gases in Mexican hydrothermal systems (genetic and applied aspects).
Axelsson, G., 2010. Sustainable geothermal utilization - case histories; definitions; Geofis. Int. 21, 193–227.
research issues and modelling. Geothermics 39, 283–289. Pulido, C.L., Armenta, M.F., Silva, G.R., 2011. Caracterización de un yacimiento de roca seca
Beramendi-Orosco, L.E., González-Hernández, G., Urrutia-Fucugauchi, J., Morton-Bermea, caliente en la zona geotérmica de Acoculco, Pue. Geotermia 24, 59–69.
O., 2006. The radiocarbon laboratory at the National Autonomous University of Python Software Foundation, d. Python Language Reference, version 3.4 Available at
Mexico: first set of samples and new 14C internal reference material with an activity http://www.python.org.
of 80.4 pMC. Radiocarbon 48, 485–491. Reimer, P.J., Bard, E., Bayliss, A., Beck, J.W., Blackwell, P.G., Ramsey, C.B., Buck, C.E., Cheng,
Campos-Enríquez, J.O., Alatriste-Vilchis, D.R., Huizar-Álvarez, R., Marines-Campos, R., H., Edwards, R.L., Friedrich, M., Grootes, P.M., Guilderson, T.P., Haflidason, H., Hajdas,
Alatorre-Zamora, M.A., 2003. Subsurface structure of the Tecocomulco subbasin I., Hatté, C., Heaton, T.J., Hoffmann, D.L., Hogg, A.G., Hughen, K.A., Kaiser, K.F., Kromer,
(northeastern Mexico basin), and its relationship to regional tectonics. Geofis. Int. B., Manning, S.W., Niu, M., Reimer, R.W., Richards, D.A., Scott, E.M., Southon, J.R., Staff,
42, 3–24. R.A., Turney, C.S.M., van der Plicht, J., 2013. IntCal13 and Marine13 radiocarbon age
Canet, C., Arana, L., González-Partida, E., Pi, T., Prol-Ledesma, R.M., Franco, S.I., Villanueva- calibration curves, 0–50,000 years cal BP. Radiocarbon 55, 1869–1887.
Estrada, R.E., Camprubí, A., Ramírez-Silva, G., López-Hernández, A., 2010. A statistics- Rybach, L., Mongillo, M., 2006. Geothermal sustainability—a review with identified
based method for the short-wave infrared spectral analysis of altered rocks: an exam- research needs. Transactions - Geothermal Resources Council 30 IIpp. 1083–1090.
ple from the Acoculco Caldera, Eastern Trans-Mexican Volcanic Belt. J. Geochem. Scaillet, B., Holtz, F., Pichavant, M., 1998. Phase equilibrium constraints on the viscosity of
Explor. 105, 1–10. silicic magmas 1. Volcanic–plutonic comparison. J. Geophys. Res. 103, 27257–27266.
Canet, C., Hernández-Cruz, B., Jiménez-Franco, A., Pi, T., Peláez, B., Villanueva-Estrada, R.E., Stimac, J.A., Goff, F., Wohletz, K., 2001. Thermal modeling of the Clear Lake magmatic-
Alfonso, P., González-Partida, E., Salinas, S., 2015. Combining ammonium mapping hydrothermal system, California, USA. Geothermics 30, 349–390.
and short-wave infrared (SWIR) reflectance spectroscopy to constrain a model Stuiver, M., Reimer, P.J., 1993. Extended 14C data base and revised CALIB 3.0 14C age
of hydrothermal alteration for the Acoculco geothermal zone, Eastern Mexico. calibration program. Radiocarbon 35, 215–230.
Geothermics 53, 154–165. Viggiano-Guerra, J.C., Armenta, M.F., Silva, G.R.R., 2011. Evolución del sistema geotérmico
Christiansen, R.L., Lowenstern, J.B., Smith, R.B., Heasler, H., Morgan, L.A., Nathenson, M., de Acoculco, Pue., México: Un estudio con base en estudios petrográficos del pozo
Mastin, L.G., Muffler, L.J.P., Robinson, J.E., 2007. Preliminary assessment of volcanic EAC-2 y en otras consideraciones. Geotermia 24, 14–24.
and hydrothermal hazards in Yellowstone National Park and vicinity. U.S. Geological Viniegra-Osorio, F., 1965. Geología del Macizo de Teziutlán y la cuenca cenozoica de
Survey Open-File Report, 1071. Veracruz. Bol. Asoc. Mex. Geol. Petrol. 17, 100–135.
Dalrymple, G.B., Grove, M., Lovera, O.M., Harrison, T.M., Hulen, J.B., Lanphere, M.A., 1999. White, N.C., Hedenquist, J.W., 1995. Epithermal gold deposits: styles, characteristics and
Age and thermal history of the Geysers plutonic complex (felsite unit), Geysers exploration. Soc. Econ. Geol. Newsl. 23, 9–13.
geothermal field, California: a 40Ar/39Ar and U-Pb study. Earth Planet. Sci. Lett. 173, Ziagos, J.P., Blackwell, D.D., Mooser, F., 1985. Heat flow and the thermal effects of
285–298. subduction in southern Mexico. J. Geophys. Res. 90, 5410–5420.
Einaudi, M.T., 1982. General features and origin of skarns associated with porphyry
copper plutons. Adv. Geol. Porphyry Copper Deposits: Southwest. North Am.,
pp. 185–209

You might also like