Chemical and Isotopic Signatures of Hot Springs From East-Central Sonora State, Mexico: A New Prospection Survey of Promissory Low-To-Medium Temperature Geothermal Systems

Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Almirudis et al.

REVISTA MEXICANA DE CIENCIAS GEOLÓGICAS v. 35, núm. 2, 2018, p. 116-141

Chemical and isotopic signatures of hot springs from


east-central Sonora State, Mexico: a new prospection survey of
promissory low-to-medium temperature geothermal systems

Erika Almirudis1, Edgar R. Santoyo-Gutierrez2,*, Mirna Guevara2,


Francisco Paz-Moreno3, and Enrique Portugal4
1
Posgrado en Ingeniería (Energía), Instituto de Energías Renovables, Universidad Nacional Autónoma de México,
Priv. Xochicalco s/n, Col. Centro, 62580, Temixco, Morelos, Mexico.
2
Coordinación de Geoenergía, Instituto de Energías Renovables, Universidad Nacional Autónoma de México,
Priv. Xochicalco s/n, Col. Centro, 62580, Temixco, Morelos, Mexico.
3
Departamento de Geología, Universidad de Sonora,
Blvd. Luis Encinas s/n, Hermosillo, Son., 83000, Mexico.
4
Gerencia de Geotermia, Instituto Nacional de Electricidad y Energías Limpias,
Reforma 113 Col. Palmira, Cuernavaca, Morelos, 62490, Mexico.
* [email protected]

ABSTRACT than the mean temperature inferred from the geothermometers, it


was suggested as an optimistic maximum reservoir temperature of the
A promissory low-to-medium temperature geothermal system Sonora geothermal system.
located in Sonora (Mexico) has been studied. In the present work, a Using 150 °C and 200 °C as rounded-off reservoir temperatures
detailed geochemical survey was carried out to understand the hydro- (or min-max estimates), geochemical equilibria modelling based on
geochemical signatures of hot spring waters. A field work campaign fluid-mineral stability diagrams was carried out. An equilibrium pro-
was conducted for collecting water samples from twelve hot springs cess among local hydrothermal waters and albite-potassium feldespar
placed in four major zones (NW, NE, C, and S). The collected samples and muscovite-prehnite-laumontite mineral assemblages was found.
were analysed by chemical and isotopic methods for determining their These minerals were proposed as representative mineral assemblages of
chemical (major and trace elements) and isotopic (18O/16O and D/H) low-grade metamorphism, which seems to indicate that the geothermal
compositions. Using geochemometric analyses of the fluid composi- fluid equilibria were probably reached within the intermediate to acidic
tion and fractionation, depletion and enrichment processes exhibited volcanic rocks from the Tarahumara Formation.
by major and trace elements were analysed.
Hydrogeochemical classification was used to indicate the presence Key words: exploration; geochemometrics; low-to-medium enthalpy;
of sodium-sulphate (Na-SO4) waters in the North (NW and NE) and fluid geochemistry; rare-earth elements; geothermal energy; Mexico.
South hydrothermal zones; whereas calcium-magnesium-bicarbonate
(Ca-Mg-HCO3) waters were identified for the Central zone. Some hot
spring waters located in the NE zone were also typified as sodium- RESUMEN
bicarbonate (Na-HCO3). In relation to the isotopic signatures of 18O/16O
and D/H, four water samples from NE and C zones lie near to the global Se estudió geoquímicamente un sistema geotérmico promisorio de
meteoric water line; whereas the remaining eight samples showed a baja-media temperatura localizado en el Estado de Sonora (México).
shift for both oxygen and deuterium isotopes. A mixing line with a En el presente trabajo, se llevó a cabo un estudio geoquímico para en-
small shift of δ18O was identified and used as a proxy to discriminate tender las firmas hidrogeoquímicas de los manantiales termales. Con
waters with different isotopic signatures. estos propósitos, se llevó a cabo una campaña de trabajo de campo para
After applying a geochemometric outliers detection/rejection and recolectar muestras de agua de doce manantiales termales localizados
an iterative ANOVA statistical test, the mean temperature inferred en cuatro zonas principales (NW, NE, C y S). Las muestras recolectadas
from the most reliable solute geothermometers was 149±40 °C, which se analizaron mediante métodos químicos e isotópicos para determinar
suggests to be considered as the minimum value of the reservoir sus principales rasgos químicos (elementos mayores y traza) e isotópicos
temperature. As most of the hot spring waters fall outside of the full (18O/16O y D/H). Usando las firmas químicas y un análisis geoquimiomé-
equilibrium curve, the original reservoir conditions were corrected by trico de la composición del fluido y de los procesos de fraccionamiento se
using a mixing conductive model, which predicted a deep equilibrium examinaron los procesos de empobrecimiento y enriquecimiento exhibidos
temperature of 210±11 °C. As this temperature is considerably higher por los elementos mayores y traza.

Almirudis, E., Santoyo, E., Guevara, M., Paz-Moreno, F., Portugal, E., 2018, Chemical and isotopic signatures of hot springs from east-central Sonora State (Mexico):
A new prospection survey of promissory low-to-medium temperature geothermal systems: Revista Mexicana de Ciencias Geológicas, v. 35, núm. 2, p. 116-141.

116 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

La clasificación hidrogeoquímica indicó la presencia de aguas de tipo Ledesma and Canet, 2004; Gutiérrez-Negrín et al., 2015; Morales-
sulfatadas-sódicas (Na-SO4) en las zonas hidrotermales del Norte (NW y Arredondo et al., 2016). The primary use of these resources has been
NE) y Sur; mientras que aguas bicarbonatadas-cálcico-magnésicas (Ca- in balneology, notwithstanding the large amount of resources available
Mg-HCO3) se identificaron en la zona Central. Algunas aguas termales for direct uses, which is estimated in 40,589 MWth (Romo-Jones et
ubicadas en la zona NE se caracterizaron como bicarbonatadas sódicas al., 2017). A complete inventory of these geothermal resources was
(Na-HCO3). Con relación a las firmas isotópicas de 18O/16O y D/H, cuatro reported by Iglesias (2003).
muestras de agua de las zonas hidrotermales NE y C se ubicaron cerca From this inventory, the presence of promissory hydrothermal
de la Línea Meteórica Mundial; mientras que las ocho muestras restantes systems in the northwestern part of Mexico has been highlighted by
mostraron un desplazamiento tanto para los isótopos de oxígeno como Hiriart et al. (2011). These geothermal systems have been delimited
para deuterio. Se identificó una línea de mezclado con un pequeño des- from geological, geophysical, and geochemical prospecting studies.
plazamiento de δ18O, el cual se utilizó como una guía para discriminar Prol-Ledesma and Juárez (1986) quantified, in a preliminary geo-
aguas con una firma isotópica diferente. physical study carried out in the same region of Mexico, heat fluxes of
A partir de un análisis geoquimiométrico basado en un algoritmo 100 mW/m2, which were based on reservoir temperatures inferred
de detección/eliminación de valores discordantes y una prueba iterativa from SiO2 geothermometry. Based on direct measurements of heat flux
de ANOVA, la temperatura media del yacimiento predicha por los performed in exploration wells, Prol-Ledesma (1991) measured more
geotermómetros de solutos fue de 149±40 °C, la cual se consideró accurate heat fluxes up to 191 mW/m2 for the same regions.
como la temperatura mínima del yacimiento. Como la mayoría de las Other promissory zones with a geothermal potential have also
aguas termales están fuera del equilibrio, las condiciones originales been identified in Sonora (Mexico) by Verdugo-Mariscal (1983), Prol-
del yacimiento se corrigieron utilizando un modelo de mezcla, que Ledesma (1991), Torres et al. (1993), Barragán et al. (2001), Iglesias et
estima una temperatura de equilibrio profundo de 210±11 °C. Como al. (2011), and Almirudis et al. (2015): see Figure 1b. Verdugo-Mariscal
esta temperatura resultó considerablemente más alta que el valor (1983) reported the presence of geothermal zones that coincide with
promedio calculado por los geotermómetros, se refirió a ésta como una NE-SW morphological structures possibly related with the spreading
estimación optimista o como la temperatura máxima del reservorio centres of the Gulf of California. This author suggested the existence
geotérmico. of a hydrothermal system characterized by two types of SO4-Cl waters:
Utilizando 150 °C y 200 °C como los valores mínimos y máximos (i) medium-enthalpy fluids with estimated geothermometric tem-
de la temperatura del yacimiento, se desarrolló un modelo de equilibrio peratures up to 189±21 °C, which interact with rocks of lithological
basado en diagramas de estabilidad fluido-mineral. Se encontró un units from the Paleozoic to the Cretaceous and Tertiary volcanics; and
proceso de equilibrio entre las aguas termales locales y los sistemas mi- (ii) low-to-medium enthalpy fluids with estimated geothermometric
nerales de albita-feldespato potásico y muscovita-prehnita-laumontita. temperatures up to 143±5 °C, which are interacting with rocks from
Estos minerales se proponen como un conjunto mineral representativo the continental Tertiary, which were sealed by younger clay rocks.
del metamorfismo de bajo grado, que parecen indicar que se alcanzó el Prol-Ledesma (1991) carried out a study for evaluating thermalized
equilibrio de los fluidos geotérmicos con las rocas volcánicas intermedias water aquifers of Guaymas (Sonora). Using the chemical composition
a ácidas de la Formación Tarahumara. of water samples collected from wells, the fluids were classified as HCO3
and Na-Cl waters with a variable concentration of SO4. By applying
Palabras clave: exploración; geoquimiometría, baja-mediana entalpía, some solute geothermometers, deep equilibrium temperatures up to
geoquímica de fluidos, elementos de las tierras raras, energía geotérmica; 115 °C were predicted, which were in good agreement with direct
México. measurements of geothermal gradients.
Torres et al. (1993) reported the existence of thermal springs
of Na-HCO3 waters, which are associated with NNW–SSE regional
INTRODUCTION structures in eastern Sonora. Using a comprehensive database, Iglesias
et al. (2011) reported the existence up to 154 hydrothermal springs
A large portion of the Mexican territory is privileged with the pres- located in Sonora with surface temperatures ranging from 28 to 88 °C.
ence of volcanic and tectonic activity, which has led to the formation Barragan et al. (2001) reported a geochemical survey for studying
of convective hydrothermal systems (Arango-Galván et al., 2015). An Na-Cl fluids with temperatures up to 192 °C, emerging from artesian
effective installed capacity of ~920.4 MWe is currently produced from and deep wells drilled in the Riito zone (located in the eastern side of
high-temperature geothermal resources, which contribute with nearly the Colorado river in the NW part of Sonora). These fluids were typi-
2.1% of the electricity annual production (Bertani, 2016; SENER, 2017). fied as high-saline sedimentary waters, which are isotopically similar
High-temperature geothermal fields of Cerro Prieto in Baja California, to those waters found in the Mexicali valley, and which are poorly
Los Azufres in Michoacán, Los Humeros in Puebla, Las Tres Vírgenes affected by meteoric waters. This hydrothermal system was related to
in Baja California, and the Domo of San Pedro in Nayarit are cur- adjacent fault systems of Cerro Prieto.
rently exploited for electricity production with an effective installed Finally, Almirudis et al. (2015) carried out an exploration campaign
capacity of 570 MWe, 225 MWe, 80.4 MWe, 10 MWe, and 35 MWe, in the central and eastern zones of Sonora, and briefly reported the
respectively (SENER, 2017). The Cerritos Colorados geothermal probable origin of the geothermal fluids found in some hot springs.
field in Jalisco, with an estimated potential of 75 MWe, has also been Reservoir temperatures and water-rock interaction processes were
identified as a promissory zone (Flores-Armenta, 2013; Pandarinath roughly analysed for developing a preliminary geochemical model. A
and Domínguez, 2015). An updating map of geothermal resources of detailed study on these zones was proposed for a better characteriza-
Mexico that includes the geothermal fields under exploitation, and tion of these low-to-medium geothermal resources, from where the
the most promissory zones under exploration are shown in Figure 1a. present study is derived.
Low-to-medium temperature geothermal systems, t<200 °C (also All these previous geochemical surveys, along with the Quaternary
referred as low-to-medium enthalpy systems) in several regions of and recent volcanism (Paz-Moreno et al., 2003; García-Abdeslem and
Mexico have not been exploited yet for electricity generation, although Calmus, 2015), the relative geographical nearness to Cerro Prieto and
a total installed capacity of ~156 MWth has been quantified (Prol- Las Tres Vírgenes geothermal fields, and the active tectonics of the

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 117


Almirudis et al.

c)

b)

a)

Figure 1. a) General map showing the current geothermal fields under exploitation and promissory hydrothermal sites of Mexico; b) Projected simplified map of
Sonora state; c) Study area and location of the sampled thermal springs in four zones: Huasabas (HSB), Bacadehuachi (BCD), Cumpas (CMP), Granados (GRN),
Arivechi (ARV), San Marcial (SMR), Tecoripa (TCP), Tonichi (TCH), Divisaderos (DVS), Tonibabi (TNB), Matape (MTP), and Aconchi (ACH).

Gulf of California, seem to identify the Sonora State as a region of GEOLOGICAL SETTING
Mexico with a promissory geothermal potential (1b). The geological,
geophysical and geochemical features enable the hydrothermal systems The geology of southwestern North America was dominated by a
of Sonora to be studied in more detail for a better evaluation of their subduction period since at least the mid-Cretaceous until the middle
geothermal reserves. Miocene: from southern Mexico up to Canada, the Farallon-Kula
The goal of the present study is to report a more comprehensive oceanic plates were subducted to the East beneath the North American
geochemical model for the evaluation of the low-to-medium tempera- continental plate (Atwater, 1989). As a result of this subduction process,
ture geothermal systems located in the east-central part of Sonora. The between 90 and 40 Ma, a large volume of magmatic intrusive and
new geochemical model is based on a fluid geochemistry database of volcanic rocks were emplaced in the Sonora State. The plutonic rocks
shallow hot springs (containing the composition of major and trace produced from this intrusive event are known as the Sonora Laramide
elements, as well as the isotopic signatures of oxygen and deuterium), Batholith (Damon et al., 1983), whereas the volcanic rocks from the
including the pseudo-equilibrium state conditions found in the system same geological event are referred as the Tarahumara Formation
as a result of the fluid-rock interaction that governed the formation of (Wilson and Rocha, 1949; McDowell et al., 2001) or the Lower Volcanic
these hydrothermal zones. Complex (McDowell and Keizer, 1977; Ferrari et al., 2005, 2007).
The study also aims to describe the major hydrochemistry sig- These rocks are widely distributed in the eastern and central Sonora
natures of hot springs in terms of fluid classification, the geochemo- State.
metric determination of the most probable reservoir temperatures, Towards 50 and 42 Ma, a reorientation of the Pacific plates occurred
the geothermal fluid sources, and the local fluid mixing processes. together with a decrease in the convergence velocity between the
Meteorological data available for these geothermal sites were also Farallon-Kula and North American plates (Stock and Molnar, 1988).
analysed to evaluate the possible influence of meteoric water recharge, This reorientation triggered a progressive change in the magmatism of
and the variability of the most common climatic parameters, including Mexico and southwestern United States, inducing the emplacement of
the surface and ambient temperatures. enormous volumes of ignimbrites during the late Eocene-Oligocene.

118 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

The resulting rocks constitute the Sierra Madre Occidental (SMO), the Northwest (NW) zone which is characterized by the shallow hot
which are known as the Upper Volcanic Complex, and are recognized springs from Aconchi (ACH) and Cumpas (CMP); the Northeast
as one of the largest ignimbrite provinces in the world (McDowell and (NE) zone which is characterized by the shallow hot springs from
Keizer, 1977; Cochemé, 1985; Ferrari et al., 2007). During Oligocene- Tonibabi (TNB), Huasabas (HSB), Divisaderos (DVS), Granados
Miocene, the North American plate and the Eastern Pacific ridge (GRN), and Bacadehuachi (BCD); the Central (C) zone which is
came into contact, giving end to the subduction period. A sinking of characterized by the shallow hot springs from Matape (MTP) and
the lithosphere was produced and a thermal flux along Sonora and Arivechi (ARV); and the South zone (S) which is characterized by
California resulted in low angle fault movements (Stock and Molnar, the shallow hot springs from San Marcial (SMR), Tecoripa (TCP) and
1988). During the middle Cenozoic a distension period of large Tonichi (TCH).
scale was produced in southern Sonora, with a northwest-southeast The NE zone contains a Lower Cretaceous detritus-carbonated
orientation (Gans, 1997). The volcanic plateau was fragmented to the sedimentary sequence (Lawton et al., 2004), which is cut by Upper
east and west of the SMO, resulting the current morphology of the Basin Cretaceous volcanic rocks, and correlated with the Tarahumara
and Range (Stewart, 1978), which is also present in the Sonora and Formation. Both sedimentary and volcanic sequences are intruded
Chihuahua States in Mexico. This morphology of parallel mountains by a series of plutonic rocks (Roldán-Quintana, 1994; Almirudis,
and valleys, delimited by normal faults, has different degrees of 2010) (Figure 2). Tertiary rocks correlated with the Upper Volcanic
sedimentary accumulation in their basins. The most ancient sediments, Complex cover the whole series of rocks (Cochemé, 1985). The basins
locally known as the Baucarit Formation (King, 1939; Cochemé et al., are filled with alluvium and sediments from the Baucarit Formation,
1988), present basalt intercalations dated at 23–19 Ma (Demant et al., and, particularly the Moctezuma valley, with basalts of the Quaternary
1989; Paz-Moreno, 1993; McDowell et al., 1997). Moctezuma volcanic field. The NW, C, and S zones present similar
A peralkaline acidic ignimbritic volcanism in Central Sonora, dated lithologies to those in the NE zone, although lacking Quaternary vol-
at 12.5 Ma, emphasizes an important change in the source of magmas. canism. Furthermore, some of their outcrops correspond to ancient
It corresponds to the latest stage of continental extension prior to the rocks composed mainly by carbonated sedimentary rocks (limestones
marine invasion and the development of spreading centers in the Gulf and dolomites) and sandstones (see Figures 2 and 3).
of California (Vidal-Solano et al., 2005). This ignimbritic episode is
related to a large-scale extension and lithosphere thinning that induced Climatological data
the formation of an intra-continental rift. This ‘proto-Gulf ’ occurred The climate of the study area can be categorized as semiarid. The
before trans-tensional deformation and spreading operated in the Gulf highest annual rainfall generally occurs on July, and it ranges from
of California as a consequence of new Pacific-North America plate 114 to 175 mm, whereas for the dry season, which generally occurs on
boundary motions (Oskin and Stock, 2003). May, the rainfall is lower than 3 mm. The lowest and highest annual
As result of the establishment of a transitional regime in the evaporation values are 84 mm and 331 mm, respectively. The air
Gulf of California, between a divergent limit in the central and temperature over this area varies from a maximum value of 42 °C in
southern areas and a transforming limit to the north, the present the summer time (on June) up to a minimum temperature of 2 °C in
tectonic configuration was established with the formation of the winter (mainly on January). Table 1 summarizes the average annual
first pull-apart basins approximately 4 Ma (Martín-Barajas, 2000). climatological measurements collected from six meteorological stations
Nowadays the limit between the North American and the Pacific located in the study area (BCD, HSB, DVS, MTP, ARV, and TCH).
plates is governed by the rift system of the Gulf of California-San These data were compiled from the Eric III Climatological Database
Andreas fault. v.1.1 (IMTA, 2013).
Quaternary volcanism of the Sonora state is roughly represented
by different rock outcroppings (Paz-Moreno, 1993; Lynch et al., 1993).
One of these outcrops is the Moctezuma basaltic field, which is placed METHODOLOGY
within the northeast zone of the geothermal study area of the present
research. This volcanic field is characterized by the transition from The methodology used in this prospection study is based on a
fissure type tholeiitic basaltic volcanism to alkaline scoria cones, novel and fundamental diagram of water-rock interaction (WRI) and
with intermittent activity between 1.7 and 0.5 Ma (Paz-Moreno et al., chemical partitioning processes associated with geothermal environ-
2003). More recently, a basalt flow from this Quaternary volcanism, ments, which was originally described by Libbey and Williams-Jones
that impounds a Bison-bearing fossil deposit at Térapa, Sonora, was (2016). A simplified diagram showing some tracing paths of these
dated to 440–130 ka using whole rock 40Ar/39Ar (Bright et al., 2010). processes (applied to the present study) is shown in Figure 4. For de-
A simplified geological map of the study area is available in the map scribing the fluid fractionation, the WRI paths used are shown with
collection of the Servicio Geológico Mexicano (Geological Survey of double dot connection lines, and frameworks without filled patterns
Mexico https://www.sgm.gob.mx/GeoInfoMexGobMx/). A detailed (‘A-A1-A2-A3-A4-A5-A7’).
map of the geological setting of the northeast zone of Sonora is Chemical and isotopic signatures were defined as key geochemical
schematically shown in Figure 2. parameters for the fluid fractionation study. The chemical mobility
of major and trace elements was therefore studied by using chemical
Promissory geothermal zones of the Sonora State analyses of hot spring waters from the four hydrothermal zones of
The study area containing the promissory geothermal zones is Sonora (NW, NE, C, and S). Reactive (Na, K, Ca, Mg, SO4, SiO2, F,
located approximately 50 km in straight line to the east of Hermosillo among others) and conservative (Cl, Br, the ‘Rare Alkalis’: Li, Rb, Cs,
city (the capital of the Sonora State), in the SMO mountain portion of among other) elements were used for finding out some geochemical
east-central Sonora (see Figures 1b and 1c). Schematic geological cross signatures of the fluids produced in the reservoir (Giggenbach, 1991).
sections of the study area are shown in Figure 3. In this area, twelve The use of bulk-rock components as geochemical tracers of fluid paths
geothermal localities were identified with a hydrothermal activity in contributes to a comprehensive understanding of element mobility
association with hot spring waters (see Figure 1c). These localities and reactivity in these geothermal systems (Libbey and Williams-
have been geochemically grouped in four major hydrothermal zones: Jones, 2016).

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 119


Almirudis et al.

Figure 2. A simplified geological map of the north-eastern zone of the study area.

SW NW NW SE
Yaqui-Tonichi
river
1200 m 1000 m
DVS GRN
BCD
TCH
TNB HSB TCP
800 m 200 m

Figure 3. Schematic geological cross sections of the northeastern and southern zones of the study area (same legend as in Figure 2).

120 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

Table 1. Annual climatological measurements collected from the Eric III National Database v1.1 (IMTA, 2003).

Station Longitude Latitude Altitude Average record time Min Temp Intervala Max Temp Intervala Rainfall Intervala Evaporation Intervala
name W N (m a.s.l.) (days) (oC) (oC) (mm) (mm)

BCD -109.283° 29.767° 706 1129±67 2±3 (Jan) 21±4 (Jan) 6±12 (May) 84±21 (Jan)
21±2 (Jul) 38±3 (Jun) 114±52 (Jul) 317±68 (Jul)
12±5 (Oct) 31±5 (Oct) 27±32 (Oct) 153±35 (Oct)
HSB -109.300° 29.903° 570 198±21 4±3 (Jan) 22±4 (Jan) 4±7 (May) 102±32 (Jan)
22±2 (Jul) 41±4 (Jun) 132±61 (Aug) 331±33 (Jun)
13±4 (Oct) 324 (Oct) 41±40 (Oct) 147±8 (Oct)
TNB -109.567° 28.583° 217 1538±58 8±3 (Jan) 27±5 (Jan) 5±10 (Apr) N.R.
24±2 (Jul) 42±4 (Jun) 175±79 (Jul) N.R.
18±4 (Oct) 364 (Oct) 31±38 (Oct) N.R.
DVS -109.033° 29.600° 680 279±35 5±3 (Jan) 21±3 (Jan) 5±8 (May) N.R.
22±3 (Jul) 38±3 (Jun) 146±58 (Jul) N.R.
15±4 (Oct) 315 (Oct) 30±33 (Oct) N.R.
MTP -109.500° 29.100° 718 453±33 8±3 (Jan) 23±5 (Jan) 3±6 (May) 112±24 (Jan)
19±5 (Jul) 37±3 (Jun) 168±80 (Jul) 308±54 (Jun)
15±4 (Oct) 31±5 (Oct) 23±26 (Oct) 158±44 (Oct)
ARV -109.183° 28.933° 566 853±45 8±3 (Jan) 24±4 (Jan) 5±10 (May) 151±0. (Sep)
19±5 (Jul) 40±3 (Jun) 160±61 (Jul) 293±0. (Jul)
13±4 (Oct) 34±5 (Oct) 26±33 (Oct) 158±0. (Oct)

a
The parameter measurement interval considers records from the lowest to the highest values. The climatological data of 2003 are reported as a reference for the
time of the sample collection campaign.

Water Rock Interac�on Processes


Path A Path B

Fluid frac�ona�on Rock frac�ona�on

Chemical element mobility B1 Compa�bility B2


A1 A2

Immobile elements Mobile elements Compa�ble Incompa�ble

Mass transfer
Reac�vity
Chemo-stra�graphy A3 A4

Reac�ve Conserva�ve

Vola�lity Vola�lity
A5 A6 A7 A8

Non-vola�le Vola�le Non-vola�le Vola�le


elements elements elements elements

Hydrogeochemistry & Isotopy Fluid origin (source & mixing) Hydrogeochemistry & Isotopy Fluid origin (source & mixing)
Fluid origin(source & mixing) Soil-gas vectoring Fluid origin (source & mixing) Soil-gas vectoring
Solute geothermometry Gas geothermometry Solute geothermometry Gas geothermometry
Fluid-Mineral Equilibrium Mineral assemblages Fluid-Mineral Equilibrium Mineral assemblages
Reservoir proper�es Reservoir proper�es Reservoir proper�es Reservoir proper�es
and others and others and others and others

Figure 4. Schematic diagram showing the water-rock interaction (WRI) and chemical partitioning processes associated with geothermal environments (modified
after Libbey and Williams-Jones, 2016).

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 121


Almirudis et al.

Fluid sampling and chemical analyses area. Small gas bubbling sites were intermittently found in a few zones
For the fieldwork, a geochemometric programme was used for the with very low discharge flow rates, which made difficult their collection.
collection and analysis of representative water samples from the identi- Polyethylene bottles of 500, 100, and 50 mL volume (previously
fied hydrothermal zones. Major physicochemical parameters were in situ washed with HCl and rinsed three times with deionized water) were
measured. A HACH portable multimeter (equipped with temperature used for collecting water samples to be used in the chemical analysis of
and electrical conductivity sensors), and a HANNA portable potenti- major (anions and cations, and other compounds) and trace (metals)
ometer were used for measuring temperature, and pH of the hot spring elements, as well as for the isotopic analyses (18O/16O, and D/H). For
waters. All the pH measurements were recorded with the temperature the analysis of anions, hot spring waters were collected without adding
compensation option after calibration with standardized buffer solu- acid, whereas for the determination of cations, the water samples were
tions of pH 4.0, 7.0 and 10.0. Measurement errors of pH (accuracy and preserved by adding ultrapure HNO3 to pH < 2. All these water samples
precision) were typically better than ±0.1 pH units (±1.4%); whereas for were previously filtered through a 0.2 μm membrane filter.
the temperature and electrical conductivity measurements, the errors Chemical analyses (major and trace elements) were carried out at
were better than ±1.1 °C (±2.6%) and ± 33 µS/cm (±3.0%), respectively. Actlabs Laboratories (Canada), whereas the isotopic analyses were con-
Ten replicates of all these measurements were systematically recorded ducted at the Isotope Laboratory of the University of Arizona (USA).
in the hot spring discharges. Total alkalinity (HCO3– and CO32–) was Major cations and trace elements were determined by Inductively
determined in situ, by triplicate, after sampling by using a titration Coupled Plasma-Mass Spectrometry and Inductively Coupled Plasma-
method with a micro-burette according to the standardized analytical Optical Emission Spectrometry, whereas the anions were analysed by
methodology proposed by Nicholson (1993). Gas samples were not col- ion chromatography. The stable isotopes of 18O/16O and D/H were
lected because steam vents or fumaroles were not observed in the study analysed by gas source Isotope-Ratio Mass Spectrometry. Table 2

Table 2. Precision, accuracy and detection limits for the elemental chemical analyses.

Chemical Precision Accuracy LOD   Chemical Precision Accuracy LOD


Elements (%) (%) Elements (%) (%)
Major elements     Trace elements
Cations       Ru ND ND 0.01
Li+ 9 5.9 0.001 Pd ND ND 0.01
Na+ 5 9.5 0.005 Ag ND ND 0.2
K+ 1 10 0.03 Cd 8.3 -2.1 0.01
Mg2+ 2.2 1.6 0.002 In ND ND 0.001
Ca2+ 2.3 5.9 0.7 Sn ND ND 0.1
Anions  Sb 5 -2.4 0.01
F- 8.3 -1.6 0.01 Te ND ND 0.1
Cl- 1.3 0.4 0.03 I ND ND 1
NO2- 2.6 3.6 0.01 Cs 0.6 15 0.001
NO3- 0.9 -0.7 0.01 Ba 0.8 5 0.1
SO42- 1 1.6 0.03 Hf ND ND 0.001
PO43- 2.1 5.5 0.02 Ta ND ND 0.001
HCO3- 3 3 1 W ND ND 0.02
Re ND ND 0.001
Trace elements Os ND ND 0.002
Be ND ND 0.0001 Pt ND ND 0.3
Al 0.5 10.3 2 Hg ND ND 0.2
Si 1.6 10.5 200 Tl 2.6 -12.7 0.001
Sc ND ND 1 Pb ND -11.5 0.01
Ti 4.6 -14.3 0.1 Bi ND -12.1 0.3
V 10 4.5 0.1 Th 5.7 7.3 0.001
Cr 3 0.5 U 5.7 -12.7 0.001
Mn 0.7 0.2 0.1        
Fe 1.1 2 10 REE      
Co 0.1 11.7 0.005 La 1.1 5.2 0.001
Ni 0.1 -3.9 0.3 Ce 0.7 -20 0.001
Cu 1.5 -6.5 0.2 Pr 0.8 6.6 0.001
Zn 0.9 -14.6 0.5 Nd 3.1 -13 0.001
Ga ND ND 0.01 Sm 1.4 4.4 0.001
Ge ND ND 0.01 Eu 5.8 -7.6 0.001
As 8.3 -5.2 0.03 Gd 2.4 -6.3 0.001
Se ND ND 0.2 Tb 0.1 -7.9 0.001
Br 0.4 -0.1 3 Dy 3.4 -8.2 0.001
Rb 0.3 4.3 0.005 Ho 4.7 -2.7 0.001
Sr 1.8 3.4 0.04 Er 0.9 -2.5 0.001
Y 0.9 11.2 0.003 Tm 5.2 -4.1 0.001
Zr ND ND 0.01 Yb 2 -2.6 0.001
Nb ND ND 0.005 Lu 0.1 -5.4 0.001
Mo 8.3 0.1

LOD: Limit of detection (estimated according to the statistical methodology suggested by IUPAC (Long and
Winefordner, 1983); i.e., 3s. LOD concentration units: Major elements (mg/kg); trace elements (µg/kg).

122 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

summarizes the precision and accuracy errors (in %) reported for the for obtaining low and high temperatures of the system, keeping in mind
chemical analyses, including the limits of detection (LOD’s), for each that these tools should be applied with caution.
one of the elements. Accuracy was evaluated by analysing water certi- To evaluate the actual equilibrium state that exhibit the fluids col-
fied standards as external testers. Ionic charge balances were calculated lected in the hydrothermal zones of Sonora, the well-known Na-K-Mg
among the major cations (Li+, Na+, K+, Ca2+, and Mg2+) and anions (F−, ternary diagram proposed by Giggenbach (1988) was used. The equi-
Cl−, NO3–, SO42–, PO43–, and HCO3–) for checking the quality of the librium isothermal curves were plotted from 75 °C to 350 °C by using
chemical analyses of hot spring waters. the equations of the Na-K and K-Mg geothermometers (TNaKVS97
and TKMgG88, respectively). The hot spring waters sampled in the
Hydrogeochemistry geothermal system were also plotted in the same ternary diagram.
For ascending hot spring waters that lack equilibrium conditions
Fluid classification (i.e., temperatures and chemical compositions falling out of the full
Major ion concentrations for all the hot spring samples were recal- equilibrium curve of the Giggenbach’s ternary diagram), the influence
culated to 100%, and plotted in standardized geochemical diagrams of of conductive cooling processes was examined by using a well-known
Schoeller, Piper, and the ternary diagram of anion variation (Cl-SO4- mixing geochemical model (i.e., a plot between the silica and fluid
HCO3) for the fluid classification. Kriging interpolation maps were used enthalpy) to evaluate the mixing degree with colder groundwaters
to describe the distribution patterns of physicochemical parameters in assuming slow fluid flow rates (usually <0.5 L/s) and lengthy flow
the studied zones. The geochemistry of trace elements in hot spring paths (Nicholson, 1993). Fournier (1977) stated that, whether or not
waters was also studied by analyzing their compositions, the major equilibrium is achieved after mixing, the equilibrium temperature of
distribution patterns and ternary diagrams (Li-Rb-Cs), as well as some hot spring waters may not be directly estimated from either a simple
chemical mobility and reactivity signatures. The Li-Rb-Cs triangular solubility relationship or solute geothermometer unless the mixing
plot was used to delineate common processes or sources that may be process is analysed. The application of mixing models to hot spring
responsible for the composition of the hot spring waters. Alkali metals and cold waters by using silica composition and temperatures was per-
(Li, Rb and Cs) were studied because it is expected that they behave formed for inferring the original fluid composition and temperature at
conservatively within a geothermal system (Ellis and Mahon 1967). reservoir conditions. In these models, the initial silica composition of
Lithium is a key conservative element because when it is partitioned hot spring waters is inferred from the curve of quartz solubility (with
into the geothermal fluid, it remains in the liquid phase as there is no an approximated uncertainty of ±5%), assuming that no further solu-
other major sink for this element in the system (Kipng’ok and Kanda, tion or deposition of silica occurs before or after mixing.
2012). Less reactive elements are usually added at depth, and their
concentration depends on the host rock composition surrounding Fluid-mineral equilibrium
the reservoir. Uptake as impurities by secondary minerals like zeolite, For this geochemical study, it was important to understand how
quartz and illite occasionally affects their concentration in geothermal the hot spring waters chemically interact with host rock-minerals (i.e.,
waters (Reyes et al., 2002). In this way, these elements may depict fluid fractionation and chemical mobility processes), as well as their
differences among waters, mainly transported by shallow processes. transient transport towards the surface. The local equilibrium state
between the hydrothermal waters and the surrounding mineral phases
Fluid isotopy was thermodynamically investigated by using a comparison between
Natural waters from meteoric, oceanic, geothermal or magmatic the water composition and the solubility of specific minerals at the
origins typically exhibit systematic differences in D/H and 18O/16O most probable reservoir temperature conditions (Torres-Alvarado,
compositions (Taylor, 1979). According to this isotopic behaviour, the 2002). The prediction of the distribution of chemical species in the
D/H and 18O/16O signatures were used as tracer indicators for defining fluid, and the saturation indexes of minerals were then determined
the most probable origin of the fluids collected in the hydrothermal by using the software “The Geochemist’s Workbench” (GWB v. 8.0:
area under study. The isotopic data were reported as δD and δ18O, React and Act2 programs). React is a computer program for modeling
where δ represents the variations of the samples with respect to the equilibrium states and geochemical processes in systems that contain
Standard Mean Ocean Water (SMOW). For the present study, the equa- an aqueous fluid; whereas Act2 is a program that calculates and plots
tion δD=8δ18O+10, originally proposed by Craig (1961) as the Global activity-activity diagrams showing the stability of minerals and the
Meteoric Water Line (GMWL), was used as a world isotopic reference. predominance of aqueous species in geochemical systems.
React program calculates the equilibrium distribution of aqueous
Solute geothermometry and mixing models species in a fluid, a fluid’s saturation state with respect to minerals.
Reservoir temperatures were estimated by solute geothermometry The program may also trace the geochemical evolution of a system as
using the SolGeo software, which was programmed by Verma et al. it undergoes reversible or irreversible reaction in an open or closed
(2008). The following solute geothermometers were used in this study: system, either at a given temperature or for an interval of temperatures.
(i) Na-K geothermometers: TNaKF79 (Fournier, 1979); TNaKVS97 GWB software uses a thermodynamic database created by the Lawrence
(Verma and Santoyo, 1997); and TNaKDSR08 (Díaz-González et al., Livermore National Laboratory (Delany and Lundeen, 1990). As part of
2008); (ii) Na-K-Ca geothermometer: TNaKCaFT73 (Fournier and this modelling, a set of non-linear governing equations describing the
Truesdell, 1973); (iii) Na-Li geothermometer: TNaLiVS97 (Verma and equilibrium state of a geochemical system is generally proposed. The
Santoyo, 1997); and (iv) SiO2 geothermometer: TSiO2VS97 (Verma resulting non-linear equation system is solved for multi-component
and Santoyo, 1997). equilibrium by means of the Newton-Raphson method. A full descrip-
Table 3 summarizes the geothermometer equations and their ap- tion of the mathematical formulation is reported by Bethke (2008).
plicability constraints. The Na-K geothermometers were selected for For plotting the ionic activity diagrams of the hot springs, the fol-
the prediction of the equilibrium temperatures because these tools lowing assumptions were considered: (i) an aqueous solution should be
generally provide more consistent and reliable temperature estimates continuously present in the system; (ii) the aluminium (Al3+) content in
than other solute geothermometers (Verma et al., 2008). The K-Mg, the minerals must be present as a solid phase; (iii) the silica concentra-
Na-K-Ca, Na-Li, and SiO2 geothermometric equations have been used tion must be determined by the solubility curve of quartz; and (iv) the

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 123


Almirudis et al.

Table 3. Geothermometer equations applied in this work and applicability constraints.

Geothermometer Geothermometer Equation to obtain temperature Reference


abbreviation (in °C)

1217(±93.9)
TNaKF79 Na/K t= - 273.15 (a) Fournier, 1979
æ Na ö
log ç ÷ + 1.483( ±0.2076)
è K ø
1289(±76)
TNaKVS97 Na/K t= - 273.15 Verma and Santoyo, 1997
æ Na ö
log ç ÷ + 0.615( ±0.179)
è K ø

876.3(±26.26)
t= - 273.15
TNaKDSR08 Na/K æ Na ö (b) Díaz-González et al., 2008
log ç ÷ + 0.8775(±0.0508)
è K ø

4410
t= - 273.15
TKMgG88 K/Mg æ K2 ö Giggenbach, 1988
14.0 - log ç ÷
è Mg ø

ì ü
ï ï
TNaKCaFT73 Na-K-Ca ï 1647 ï Fournier and Truesdell, 1973
t =í + 2.47 ý - 273.15 (c)
ï log æ Nam ö + b log æ Cam ö + 2.06
0.5
ï
ï ç ÷ ç ÷ ï
î è Km ø è Nam ø þ

1049(±44)
t= - 273.15 (d)
TNaLiVS97 Na/Li æ Nam ö Verma and Santoyo, 1997
log ç ÷ + 0.44( ±0.10)
è Lim ø

1267(±35)
t= - 273.15 (e)
æ Na ö
log ç m ÷ + 0.07(±0.10)
è Lim ø
TSiO2VS97 SiO2 t= −{44.119(±0.438)} + {0.24469(±0.00573)}S− Verma and Santoyo, 1997
{1.7414×10−4 (±1.365×10−5)}S2 + {79.305(±0.427)}log S (f)

t= [140.82(±0.00)] + [0.23517(±0.00179)]S (g)

(a) Applicable for temperatures >150 °C; certainly not recommended for temperatures <100 °C when this geothermometer might give
anomalously high temperatures.
(b) The database for proposing both geothermometric equations was for temperatures between 30 and 350 ºC.
(c) Concentrations are in mol/kg (molal) β=4/3 for t < 100 ºC, β=1/3 for t > 100 ºC and β=1/3 for t < 100 ºC and log(Ca0.5/Na) < 0; Mg
correction is recommended, see Fournier (1991) or D’Amore and Arnórsson (2000) for details.
(d) For Cl < 0.3 mol/kg.
(e) For Cl > 0.3 mol/kg.
(f) Applicable for 20–210 ºC, S is the concentration of SiO2 in mg/kg.
(g) Applicable for 210–333 ºC.

temperature and the corresponding pressure of the system are constant. the gas-phase composition, which generally result in large uncertain-
The ionic activities in solution corresponding to the water samples at ties for the calculation of the total molalities and ion activities; (ii) the
pseudo- or equilibrium conditions were obtained as activity-activity effect of degassing on the aqueous species, which in combination with
diagrams. The results inferred from this fluid-equilibria modelling pH have significant effects for mineral super- or under-saturation;
were examined with caution because the steam/gas phase composi- (iii) the effect of boiling may also produce an irregular dispersion of
tion for the fluid samples was not available. The lack of gas-phase log(Q/K) curves due to the combined effects of aqueous component
composition limited the use of the multi-component geothermometry concentration caused by water loss, and the pH change due to CO2
methodology [i.e., log(Q/K) curves depending on temperature] for loss; (iv) the effect of dilution also produces a shift and dispersion of
inferring the equilibrium temperatures with accuracy. Regarding this the mineral log(Q/K) curves causing a displacement in most curve
missing information, Reed and Spycher (1984) and Pang and Reed intersections (with log Q/K = 0) with implications on the calculation of
(1998) identified several uncertainties that may affect the calculation of low equilibrium temperatures; and (v) the uncertainty of the aluminium
equilibrium temperatures, among which are the following sources: (i) analyses (or analytical errors) which may increase the propagated errors
the large pH differences expected between the fluid with and without by up to 35 oC in the estimation of reservoir equilibrium temperature.

124 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

RESULTS AND DISCUSSION spring waters with shallow temperatures greater than 40 °C are mostly
found in the NW, NE and S hydrothermal zones, whereas the lower
Fluid sampling temperatures were mostly observed at the C zone, and the anomalous
From the fluid sampling campaign, it was observed that the ACH, saline hot spring DVS (NE); (ii) a pH transition from 6.8 (neutral) to
TNB, TCP, and TCH hot spring waters are geologically correlated with 9.0 (strongly alkaline) was observed for the fluids from the C to NE
the Laramide granitoids (Figure 3), and probably associated to NW- geothermal zone, with the exception of the anomalous DVS sample
SE oriented regional faults at the basin limits. The CMP and DVS hot which exhibits a pH of 6.0 (slightly acidic); and (iii) the highest ionic
spring waters rise towards the surface probably through fractures and concentrations (given by the high electrical conductivity values) were
secondary faults and interact with ignimbrite outcrops characteristic located at the NE zone, whereas the lowest values were mostly found at
of the Sierra Madre Occidental (Figure 3). The BCD, GRN, and SMR the C zone, which agree with the temperature and pH patterns observed
thermal spring waters ascend to the surface directly on alluvium, probably as a product of the WRI processes.
whereas the HSB hot springs ascend on ancient alluvium that has been
correlated with the Baucarit Formation (Figure 3). The MTP hot spring Chemical analyses of water samples
water upflow to the surface occurs perhaps through a more than 10 The chemical analyses for the hot spring waters sampled in the
km long regional fracture present in limestones and covered with sedi- geothermal system of Sonora under study are compiled in Table 5. The
ments, whereas the waters of hot spring ARV rise towards the surface content of major elements (anions and cations) and SiO2 (in mg/kg)
to interact with limestones from older Formations. and the isotopic analyses of δ18O and δD (in ‰) have been reported
in the first section of Table 5. The next section was used to report the
Physicochemical measurements abundance of trace elements in µg/kg, whereas the last section was used
The physicochemical parameters measured in situ in the hot to present the concentration of the rare earth elements (REE) in µg/kg.
springs are summarized in Table 4. The highest shallow temperatures, A statistical descriptive plot showing the main concentration patterns
which ranged from 41.5±7.5 °C to 63.1±0.6 °C, were measured for hot of major and trace elements in all hot spring samples is presented in
spring waters that emerge mostly on granite formations: TCP (S), GRN Figure 6 (as box-plots with some statistical parameters given by the
(NE), TNB (NE), ACH (NW), and TCH (S). Intermediate shallow median, percentile, and min-max values), whereas the REE average
temperatures between 31.4±0.2 °C and 40.0±0.2 °C were measured for concentration patterns for the four hydrothermal zones (normalized
thermal springs emerging on sedimentary formations: SMR (S), BCD with respect to chondrite) are plotted in Figure 7.
(NE), HSB (NE), and ARV (C). The lowest temperatures were meas- The statistical descriptive plot (Schoeller type) clearly indicates that
ured in waters coming from the hot springs DVS (29.3±0.2 °C), MTP the hot spring waters are mainly depleted in the major elements F–,
(30.9±2.9 °C), and SMR (31.4±0.2 °C), which emerge on ignimbrite, Cl–, K+ and Li+, and enriched in HCO3–, SO42–, Na+ and Ca2+ (Figure 6).
limestone, and alluvium rocks, respectively. Regarding the trace element contents, depletions were mainly observed
In the case of the pH measurements, the values varied from 6.0±0.1 for Cs, Sr, Sb, V, Co and Tl, whereas enrichments were mostly given
(for the DVS hot spring) to 9.0±0.1 (for the HSB), whereas those related for Rb, Ba, As, Al, Mn and Fe. A high chemical mobility from the host
with the electrical conductivity show a larger variability with values rock to the fluid (fractionation) was mainly observed in Mn, Fe, Sr, As,
ranging from 426±12 μS/cm (for the BCD hot spring) to 3670±110 Rb, Ba, including the light REE (La and Ce). Intermediate and heavy
μS/cm (for the DVS). Total dissolved solids (TDS) were also estimated REE were immobile elements that were not partitioned into the fluid
for all the hot springs sampled, and they ranged between 343±10 ppm due to the alkaline pH conditions (pH>6) that dominated most of the
(for the TCP hot spring waters, S) and 4519±136 ppm (for the DVS, WRI and water mixing processes observed in the hydrothermal system.
NE): see Table 4. An interesting geochemical signature of the NE hot spring waters
Shallow temperature, pH, and electrical conductivity measure- is the zig-zag pattern observed for the REE, with positive and nega-
ments were plotted in distribution maps using the Surfer software tive anomalies of Eu and Gd, respectively. This REE pattern is roughly
(Figure 5). These distribution maps show the spatial distribution of observed for the NW and S hot spring waters, whereas for the C zone
these parameters and the following geochemical signatures: (i) the hot appears almost as a flat pattern (see Figure 7).

Table 4. Physicochemical parameters measured in-situ in waters collected from hot springs located in the promissory geothermal system of east-central Sonora
State (Mexico).

Location Site Latitude N Longitude W Elevation Date Tsup pH EC TDS


(m a.s.l.) (°C) (μS/cm) (ppm)

NW ACH 29°50’36.31 110°16’38.32 664 02/10/2011 50.8 ± 4.1 7.6 ± 0.1 728 ± 41 555 ± 17
NW CMP 30°07’33.61 109°52’38.48 896 24/09/2011 45.5 ± 0.2 7.5 ± 0.1 767 ± 7 676 ± 20
NE BCD 29°48’57.73 109°09’38.53 706 18/09/2011 34.8 ± 0.9 8.5 ± 0.1 426 ± 12 418 ± 13
NE HSB 29°56’37.11 109°18’38.97 570 18/09/2011 35.9 ± 0.4 9.0 ± 0.1 858 ± 3 719 ± 22
NE GRN 29°51’17.64 109°17’09.64 611 25/09/2011 46.1 ± 1.2 6.8 ± 0.1 1777 ± 32 1569 ± 47
NE TNB 29°50’22.78 109°33’46.49 791 01/10/2011 49.8 ± 1.0 7.2 ± 0.1 1400 ± 44 1202 ± 36
NE DVS 29°42’12.51 109°25’54.28 834 01/10/2011 29.3 ± 0.5 6.0 ± 0.1 3670 ± 110 4519 ± 136
C MTP 29°16’44.59 109°52’02.28 762 02/10/2011 30.9 ± 2.9 6.8 ± 0.1 444 ± 25 483 ± 14
C ARV 28°56’40.02 109°15’17.53 609 28/09/2011 40.0 ± 0.2 6.9 ± 0.1 546 ± 2 555 ± 17
S TCP 28°36’43.49 109°55’08.89 406 29/09/2011 41.5 ± 0.6 8.4 ± 0.1 525 ± 9 343 ± 10
S SMR 28°34’30.79 110°15’21.68 333 29/09/2011 31.4 ± 0.2 7.1 ± 0.1 826 ± 14 638 ± 19
S TCH 28°37’58.54 109°29’41.88 365 29/09/2011 63.1 ± 0.6 7.6 ± 0.1 1428 ± 11 1093 ± 33

Tsup: surface temperature; EC: electrical conductivity; TDS: total dissolved solids.

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 125


Almirudis et al.

As seen in Table 5, boron (a highly mobile and conservative ele-


ment) was not measured in the hot spring waters because it is generally
present in high-temperature geothermal systems (Aggarwal et al. 2000;
Gurav et al., 2016). According to the local WRI processes (i.e., fluid
fractionation or chemical mobility) that occur in the hydrothermal
system of Sonora, the non-volatile elements characterized as non-
reactive or conservative compounds were mainly associated to the
contents of Cl, Li, Rb, and Cs, whereas the reactive or geoindicator
elements were typified by the concentrations of major cations (Na, K,
Ca, Mg), and SiO2.
A multivariate statistical analysis of the linear relationships for all
the physicochemical parameters, major and trace elements was applied
by using a correlation matrix (Table 6a and 6b). Correlations among
these variables provided valuable information of the major hydro-
geochemical processes that controlled part of the chemical signatures
found in the hot spring waters. Applying the methodology suggested
by Bevington and Robinson (2003) for a small data sample (n=12),
correlation coefficients of r>0.823 provide an acceptable correlation
a) for a confidence level of 99.9%, whereas r values ranging from 0.5 to
0.7 shows a moderate correlation among the parameters. From this
analysis, the physicochemical parameters (EC and TDS) resulted in
high correlations (r>0.82) for both the cation (Na+, K+, Li+ and Ca2+)
and anion (SO42– and HCO3–) contents (see Table 6a). From the cor-
relation matrix, it was found that TDS and EC of the hot spring waters
were mainly controlled by Li+-HCO3– and Na+-SO42– ions, respectively.
On the other hand, the cation concentrations (Na+, K+, Li+, and
Ca ) were highly correlated with most of the major elements found
2+

in the hot spring waters, as well as with the trace elements Mn, Fe,
Ge, Rb, Y, Zr, and Nb. In relation to the anion compositions, SO42–
resulted in a good correlation with Be, Ti, Ge, and Rb; whereas the
HCO3– concentrations were well correlated with Mn, Fe, Y, Zr, and
Nb. Among trace elements, Mn, Fe, Y, and Nb showed the highest
correlation coefficients.

Hydrogeochemistry (fluid classification)


Major components of the hot spring waters were plotted in a Piper
plot (Figure 8). As seen from this diagram, the geothermal fluids pro-
duced in the hot springs of the Central zone (MTP and ARV) were
classified as calcium and magnesium-bicarbonate (Ca-Mg-HCO3)
b) waters, compositions that are in concordance with the sedimentary
formation from where they emerge after WRI and mixing processes.
The waters produced in the NW (ACH and CMP), two NE (GRN and
TNB), and S (TCP, SMR and TCH) hot springs were mostly grouped as
sodium-sulphate (Na-SO4) waters, whereas those fluids coming from
three hot springs from the NE hydrothermal zones (BCD, HSB and
DVS) were mainly characterized as sodium-bicarbonate (Na-HCO3)
waters. As seen in Figure 8, the hot spring waters of the Central zone
(MTP and ARV) have entirely different chemistry compared to the
other springs, which indicates that these waters probably come from
a totally different fluid source or reservoir. The major-ion data of the
remaining hot spring waters very well matches with more systematic
chemical signatures of a geothermal system of low-to-medium tem-
perature, which mostly dominates along the NW, NE and S zones.
An additional geochemical classification of fluids is also repre-
sented in the Cl-HCO3-SO4 ternary diagram proposed by Giggenbach
(1988). According to this diagram (Figure 9), the hot spring waters
produced at the Central (MTP and ARV), South (TCP and SMR) and
c) NE (BCD and DVS) zones may be roughly classified as peripheral
waters (i.e., bicarbonate and diluted chloride waters with a low concen-
Figure 5. Distribution of physicochemical parameters measured in the hot tration of sulphates) with a tendency toward volcanic waters; whereas
spring waters of the Sonora geothermal system: (a) temperature, (b) pH, and the remaining hot spring waters fall on the plot section of “steam
(c) electrical conductivity. heated waters” due to the relative high proportion of sulphates found

126 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

Table 5. Chemical composition of geothermal waters collected from hot springs of central-eastern Sonora (major elements in mg/kg; trace elements in μg/kg).

Sample ACH CMP BCD HSB GRN TNB DVS MTP ARV TCP SMR TCH
Zone NW NW NE NE NE NE NE C C S S S

Major Cations
Na+ 154.0 156.0 110.0 210.0 400.0 289.0 785.0 5.8 17.3 100.0 160.0 260.0
K+ 5.09 3.24 0.77 0.45 14.30 11.20 38.00 4.23 5.11 2.36 4.83 9.31
Li+ 0.470 0.410 0.144 0.110 1.070 0.660 3.100 0.020 0.020 0.172 0.260 0.530
Ca2+ 19.4 40.0 0.9 1.1 51.9 71.3 273.0 75.3 66.2 4.8 38.2 83.8
Mg2+ 0.159 0.646 0.044 0.012 1.210 0.620 58.400 19.700 32.300 0.226 3.380 0.240
Major Anions
F− 9.56 6.58 1.32 2.45 6.61 7.24 < 0.01 < 0.01 < 0.01 3.73 4.44 6.74
Cl− 44.9 12.4 3.5 21.5 87.5 41.4 32.4 3.3 6.4 49.8 158.0 148.0
SO42− 180.0 279.0 12.1 228.0 581.0 630.0 773.0 12.2 90.3 58.0 106.0 546.0
CO32− < 0.02 < 0.02 36 61 < 0.02 < 0.02 < 0.02 < 0.02 < 0.02 11 < 0.02 < 0.02
HCO3− 141.3 177.9 253.2 193.2 426.1 151.5 2559.4 362.0 337.6 112.9 162.7 38.6
SiO2 77.01 56.91 49.42 45.14 71.45 92.20 61.18 22.89 31.23 55.41 42.57 70.38
II*(%) -0.23 2.42 8.41 7.36 3.03 1.74 4.39 4.46 5.15 3.51 1.51 2.27
Stable Isotopes
δ18O ‰ -8.9 -7.4 -9.5 -8.2 -8.8 -9.6 -8.8 -8.0 -6.8 -7.2 -7.4 -8.5
δD ‰ -63 -57 -67 -61 -67 -69 -64 -55 -50 -52 -53 -58
Trace Elements
Be 0.3 0.4 0.1 0.1 0.7 0.6 0.7 0.1 0.1 0.1 0.1 0.2
Al 9 7 17 21 6 7 <2 3 3 44 4 10
Sc 6 4 4 3 6 7 5 2 2 4 3 6
Ti 4.7 3.3 2.6 2.4 4.6 5.7 5.7 2.0 2.2 3.4 2.8 4.5
V 0.2 0.1 42.6 40.0 3.7 0.1 0.1 0.6 1.8 31.9 0.1 0.1
Cr 0.5 0.5 4.2 13.3 0.5 0.5 2.5 0.5 0.5 0.5 0.5 0.5
Mn 30.9 81.6 0.4 0.2 3.0 205.0 568.0 2.1 0.2 15.3 54.5 18.9
Fe < 10 < 10 < 10 < 10 < 10 30 240 < 10 < 10 < 10 20 < 10
Co < 0.005 0.019 0.218 0.007 < 0.005 < 0.005 0.031 < 0.005 0.010 < 0.005 < 0.005 < 0.005
Ni 0.7 0.7 10.8 0.3 4.5 0.8 0.5 1.2 1.2 2.4 1.1 8.4
Cu 1.1 2.6 15.4 1.8 7.9 3.0 2.3 1.5 3.1 5.3 3.7 8.3
Zn 32.2 5.1 4.2 15.6 13.0 6.3 6.2 56.9 12.8 3.6 7.3 6.6
Ga 1.59 0.09 2.28 2.90 0.11 0.62 0.01 0.01 0.01 1.51 0.09 1.57
Ge 5.05 5.14 1.85 0.71 10.30 10.40 12.80 0.18 0.20 3.80 3.60 5.74
As 25.8 169.0 105.0 40.1 138.0 21.3 2.4 10.9 14.3 61.9 67.0 42.2
Se 0.2 0.3 0.4 1.8 0.7 < 0.2 0.5 0.4 0.9 0.3 0.9 0.8
Rb 35.80 25.40 4.02 2.07 89.50 103.00 123.00 13.20 10.90 9.95 29.40 75.30
Sr > 200 > 200 4 4 > 200 > 200 > 200 > 200 > 200 115 > 200 > 200
Y 0.019 0.040 0.006 0.003 0.019 0.030 0.136 0.031 0.015 0.014 0.006 0.014
Zr 0.09 0.12 0.04 0.14 0.10 0.11 0.33 0.01 0.02 0.04 0.03 0.06
Nb 0.009 0.006 < 0.005 < 0.005 0.006 0.011 0.055 < 0.005 < 0.005 0.006 0.005 0.007
Mo 119.0 9.8 1.9 12.5 19.4 41.8 3.6 1.0 1.0 94.0 38.5 21.1
Ru < 0.01 < 0.01 < 0.01 < 0.01 0.01 0.01 0.04 < 0.01 < 0.01 < 0.01 < 0.01 0.01
Pd < 0.01 0.01 0.01 < 0.01 0.01 0.01 0.06 0.01 0.01 < 0.01 0.02 < 0.01
Ag < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2
Cd < 0.01 0.04 0.01 < 0.01 0.01 < 0.01 0.03 0.04 0.02 < 0.01 < 0.01 < 0.01
Sn 0.2 < 0.1 0.1 0.1 0.1 0.1 < 0.1 < 0.1 < 0.1 0.1 < 0.1 0.1
Sb 0.28 1.11 0.19 0.27 4.74 0.32 0.01 0.57 0.14 0.09 1.26 0.22
Cs 12.50 28.80 4.73 1.16 111.00 74.80 35.00 2.67 3.16 1.44 12.90 16.40
Ba 21.2 34.1 0.9 0.7 17.2 31.5 16.2 > 400 81.2 10.3 64.4 37.5
Hf 0.002 0.004 0.001 0.003 0.004 0.002 0.001 < 0.001 0.001 0.001 0.001 0.002
Ta 0.004 0.003 < 0.001 0.001 0.003 0.003 0.003 < 0.001 < 0.001 0.001 0.002 0.003
W > 20.0 > 20.0 5.23 1.02 > 20.0 > 20.0 0.78 0.33 0.17 > 20.0 > 20.0 > 20.0
Re 0.003 0.001 < 0.001 0.002 0.003 < 0.001 < 0.001 0.006 0.006 0.005 < 0.001 < 0.001
Os < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002 < 0.002
Pt 0.3 0.3 0.5 0.3 0.3 0.4 0.3 0.8 0.4 0.3 0.3 < 0.3
Hg < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2 < 0.2
Tl 0.062 0.025 < 0.001 < 0.001 0.246 0.128 0.030 0.180 0.113 < 0.001 0.025 0.114
Pb 0.69 0.34 0.27 0.23 0.13 0.38 0.03 0.74 1.80 0.21 0.23 0.14
Bi < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3 < 0.3
Th 0.003 0.006 < 0.001 < 0.001 0.005 0.009 0.007 < 0.001 < 0.001 < 0.001 < 0.001 0.002
U 4.070 0.073 6.200 5.440 1.200 0.021 4.840 0.710 1.250 34.100 0.003 0.030
continues

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 127


Almirudis et al.

Table 5 (cont.). Chemical composition of geothermal waters collected from hot springs of central-eastern Sonora (major elements in mg/kg; trace elements in μg/kg).

Sample ACH CMP BCD HSB GRN TNB DVS MTP ARV TCP SMR TCH
Zone NW NW NE NE NE NE NE C C S S S
Rare Earth Elements (REE)
La 0.027 0.284 0.044 0.016 0.146 0.038 0.010 0.050 0.057 0.056 0.244 0.035
Ce 0.017 0.047 0.016 0.007 0.016 0.020 0.005 0.009 0.009 0.044 0.008 0.014
Pr 0.002 0.005 0.002 0.002 0.003 0.003 < 0.001 0.001 0.002 0.004 0.001 0.002
Nd 0.005 0.018 0.007 0.002 0.008 0.010 0.005 0.004 0.005 0.017 0.003 0.004
Sm 0.003 0.006 0.003 0.003 0.003 0.004 0.003 0.003 0.003 0.004 0.001 0.002
Eu < 0.001 0.001 < 0.001 0.002 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001
Gd < 0.001 0.004 0.002 0.002 0.002 0.002 0.003 0.003 0.002 0.003 < 0.001 0.002
Tb < 0.001 < 0.001 < 0.001 0.003 < 0.001 < 0.001 0.001 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001
Dy 0.002 0.003 0.001 0.004 0.001 0.003 0.007 0.002 0.001 0.002 < 0.001 0.001
Ho < 0.001 0.001 < 0.001 0.003 < 0.001 0.001 0.002 < 0.001 < 0.001 < 0.001 < 0.001 0.001
Er 0.001 0.003 < 0.001 0.004 0.001 0.002 0.006 0.001 < 0.001 0.001 < 0.001 0.001
Tm < 0.001 < 0.001 < 0.001 0.005 < 0.001 0.001 < 0.001 < 0.001 < 0.001 < 0.001 < 0.001 0.001
Yb 0.002 0.002 0.001 0.008 0.001 0.002 0.005 0.002 0.001 0.001 < 0.001 0.002
Lu 0.001 < 0.001 < 0.001 0.008 < 0.001 0.002 0.001 < 0.001 < 0.001 < 0.001 < 0.001 0.002
* II: Ionic Imbalance. Obtained Br- values were < 0.03 mg/kg, except for ACH (0.57 mg/kg), TCP (0.72 mg/kg), SMR (2.24 mg/kg), and TCH (2.54 mg/kg). Obtained NO3– values were
< 0.01 mg/kg, except for BCD (0.27 mg/kg), HSB (1.13 mg/kg), MTP (0.70 mg/kg), ARV (0.63 mg/kg), and TCP (0.03 mg/kg).

in the samples (NW: ACH and CMP; NE: HSB, GRN and TNB; and are usually typified by a high concentration of sulphates, which are
the anomalous TCH sample of the S zone, which seems to be the most commonly discharged from hot springs nearby to the vent of active
representative water sample by its higher surface temperature, and typi- (degassing) volcanoes. However, these authors also describe that
fied as a mix chloride-sulphate or volcanic water). Some water samples slightly acid geothermal fluids (with pH>5) may be found close to
(e.g. GRN, TNB, and HSB) exhibited anomalous alkaline pH between lateral outflows some distance away from the main degassing vent.
6.8 and 9, which could not be strictly considered as characteristic of Under these conditions, “alkaline sulphate waters” might be discharged
“steam heated waters”. The classification of these samples is difficult to with a diluted ionic composition.
understand because it would be expected to have acid waters (pH<3) Furthermore, reactions among dissolved carbon dioxide and host
instead of slightly acid sulphate waters with pH >6. To support these rocks tend to form HCO3–, the concentration of which is probably
anomalous geochemical signatures, some recent geochemical findings affected by permeability and lateral flow. As a result of this process,
reported in the literature on the possible origin of these geothermal hot springs that are fed directly from the reservoir tend to exhibit the
fluids are discussed. lowest HCO3– concentrations (which is actually characterized by the
Smith et al. (2010) pointed out that in magmatic–hydrothermal TCH water sample: 38.6 mg/kg and the highest surface temperature).
and associated geothermal systems, acid magmatic fluids (with pH<3) This enables the HCO3–/SO42– ratio to be used as an indicator of flow

Alkaline Radioactive
5000000 Anions Alkali metals earth metals Metalloids Transition metal and metals elements

500000

50000 Samples collected from


hot springs in Sonora
Concentration [µg/L]

5000 (September-October, 2011)

500

50

0.5

0.05
Median
25%-75% percentile
0.005 Min-Max

0.0005
Cl F Na Li Rb Mg Ba Sb Ti Mn Co Cu Ga Pb U
HCO3 SO 4 K Cs Ca Sr As Al V Fe Ni Zn Tl Th
Figure 6. Statistical descriptive plot showing the main compositional features observed for major and trace elements measured in the hot spring samples of the
Sonora geothermal system.

128 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
0.001

0.0001 NE
HSB
BCD
GRN
DVS
TNB
1E-005
n=5

0.001

0.0001
NW
? ? CMP
ACH

n=2
1E-005

0.001
Abundance/Chondrite

0.0001

? C
? ? ARV
1E-005 ? MTP

n=2

0.001

0.0001
?
? S
SMR
TCP
1E-005 TCH

n=3

0.001

0.0001

All
zones
1E-005

n=12

La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu
Figure 7. Rare Earth Element (REE) patterns normalized to chondrite values for the hot spring waters of the NE, NW, C and S zones. All the symbols correspond
to the mean value for each zone each zone; bars indicate minimum and maximum values.

direction. The flow of a fluid away from the upflow yields the suitable Additionally, and to support the detection of relative high propor-
conditions for rock-water reaction, and therefore for an increase of tions of sulphates in six water samples from the NW, NE and S zones
HCO3–. This, combined with the loss of H2S by rock-water reactions (see Figure 9), three precipitated salt samples were collected in situ at
with increased lateral flow, leads to an increase in the HCO3–/SO42– ra- the CMP, TCH and TNB hot springs, and analysed by X-Ray Diffraction
tios away from the upflow zone with values ranging from 0.07 (TCH) (XRD) for determining the major mineral compositions. The results
to 2 (TCP) for hot springs located at low altitudes of ~400 m a.s.l. (S of these XRD analyses show that the minerals that precipitated from
zone), and from 0.25 to 11 for hot springs situated at an interval of the CMP waters were mainly characterized as quartz (SiO2), thenar-
altitudes between 600 and 900 m a.s.l. (NW, NE, and C zones). dite (Na2SO4), and calcite (CaCO3), whereas those for TCH and TNB

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 129


Almirudis et al.

Table 6a. Correlation matrix of the physicochemical parameters and the major ion content in water samples from hot springs in central-eastern Sonora (the best
correlations were considered for r>0.823, n=12, and α=0.001; according to Bevington and Robinson, 2002).

  Tsup pH EC TDS Na+ K+ Li+ Ca2+ Mg2+ F– Cl– SO42– HCO3– SiO2

Tsup 1
pH 0.1394 1
EC -0.0702 -0.5959 1
TDS -0.2129 -0.6138 0.9822 1
Na+ -0.0661 -0.4781 0.9826 0.9555 1
K+ -0.1432 -0.7160 0.9743 0.9825 0.9298 1
Li+ -0.1609 -0.6080 0.9751 0.9862 0.9630 0.9782 1
Ca2+ -0.2125 -0.7647 0.8888 0.9296 0.8001 0.9445 0.8932 1
Mg2+ -0.4920 -0.6833 0.6345 0.7447 0.5223 0.7480 0.6825 0.8634 1
F– 0.7683 0.1240 -0.0758 -0.2197 -0.0005 -0.1615 -0.1091 -0.3301 -0.6555 1
Cl– 0.3691 -0.1229 0.1628 0.0301 0.1672 0.0849 0.0502 0.0055 -0.2811 0.4418 1
SO42– 0.3390 -0.4811 0.8687 0.7777 0.8649 0.7943 0.7785 0.6761 0.2948 0.2784 0.2654 1
HCO3– -0.4551 -0.5937 0.8682 0.9448 0.8290 0.9148 0.9242 0.9125 0.8717 -0.4523 -0.1709 0.5370 1
SiO2 0.6667 -0.0311 0.3925 0.2657 0.4652 0.3117 0.3540 0.0977 -0.2971 0.7749 0.2784 0.6837 0.0074 1

Table 6b. Summarized correlation matrix of the major and trace elements in water samples from hot springs in central-eastern Sonora (the best correlations were
considered for r>0.823, n=12, and α=0.001; according to Bevington and Robinson, 2002).

  Na+ K+ Li+ Ca2+ SO42- HCO3− Be Ti Mn Fe Ge Rb Y Zr Nb


Na +
1
K+ 0.9298 1
Li+ 0.9630 0.9782 1
Ca2+ 0.8001 0.9445 0.8932 1
SO42- 0.8649 0.7943 0.7785 0.6761 1
HCO3− 0.8290 0.9148 0.9242 0.9125 0.5370 1
Be 0.7941 0.7532 0.7662 0.5827 0.8649 0.5426 1
Ti 0.7530 0.7008 0.7106 0.5317 0.8620 0.4255 0.8287 1
Mn 0.8537 0.9061 0.9166 0.8962 0.6793 0.9014 0.6514 0.6427 1
Fe 0.8591 0.9246 0.9379 0.9211 0.5930 0.9787 0.5491 0.5196 0.9595 1
Ge 0.8587 0.8040 0.8179 0.6241 0.9028 0.5590 0.9354 0.9327 0.7200 0.6176 1
Rb 0.8446 0.8446 0.8038 0.7270 0.9518 0.5696 0.8770 0.9048 0.7164 0.6266 0.9442 1
Y 0.8039 0.9064 0.9167 0.9314 0.6050 0.9445 0.6310 0.5365 0.9431 0.9515 0.6408 0.6380 1
Zr 0.9102 0.8308 0.9022 0.7520 0.7311 0.8434 0.6992 0.6259 0.8800 0.8738 0.7019 0.6506 0.8559 1
Nb 0.8705 0.9332 0.9500 0.9196 0.6253 0.9691 0.5840 0.5805 0.9621 0.9946 0.6521 0.6573 0.9584 0.8901 1

hot springs were typified as gypsum (CaSO4.2H2O), and a mixture of all the hydrothermal zones (ranging from 4 to 123 µg/kg for the NE;
thenardite (Na2SO4) and halite (NaCl), respectively (see Figure 10). from 12.50 to 28.80 µg/kg for the NW; from 3.16 to 16.4 µg/kg for the
Taking into account these XRD analyses, the composition of these S; and from 2.67 to 3.16 for the C). This pattern is shown in the ternary
hot spring waters (as a product of the interaction with or dissolution diagram of Li-Rb-Cs (Figure 11). The plot shows chemical signatures
of these precipitated minerals) may be in agreement with the unusual related to a group of hot spring samples falling between the basalt and
concentrations of sulphates reported for these water samples (see Figure rhyolite rock boundaries, which are well characterized by K/Rb and K/
9). All these classification signatures are mostly in agreement with the Cs ratios ranging from 109 to 192, and from 112 to 567, respectively.
grouping inferences obtained from the Piper diagram. According to Goguel (1983), it is assumed that these waters may interact
with illite at temperatures between 190 and 210 oC.
Ternary diagram of Li-Rb-Cs On the other hand, the enrichment of Rb in presence of low
The lithium concentrations in the hot spring waters varied between concentrations of Li seems to indicate that these hot waters reached
0.66 and 3.1 mg/kg for the NE zone, whereas lower concentrations were temperatures around 210 °C at the reservoir conditions before mix-
found for the NW and S hydrothermal zones which ranged from 0.41 ing with shallow cold waters during their ascent towards the surface.
to 0.47 mg/kg, and from 0.17 to 0.53 mg/kg, respectively. Hot spring As a result of these chemical signatures, the study of the ternary
waters that emerge from the C zone show the lowest lithium concen- diagram of Li-Rb-Cs may also be used as a secondary proxy indicator
trations (0.020 mg/kg). On the other hand, Rb concentrations were for inferring the theoretical reservoir temperatures of the Sonora hot
mostly the highest for the NE zone (up to 123 µg/kg), whereas for the spring waters.
NW and S zones, intermediate concentrations ranging from 25.40 to
35.80 µg/kg, and from 9.95 to 75.3 µg/kg, respectively, were found. The Isotopic analyses
lowest concentrations of Rb were actually measured for the C zone. The water isotopic composition (D/H and 18O/16O) is considered
In relation to Cs, a systematic composition pattern was identified for as a conservative signature, and it is a good geochemical indicator for

130 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

l)

Ca
(C
80 80

lciu
ide

m
r
hlo

(C
60 60

+C

a)
+
4)

Ma
O
(S

gn
40 40

es
te
lfa

ium
Su

(M
20 TCH
20

g)
ARV
TNB

MTP

SMR
CMP
GRN
ACH

100 0 DVS HSB 0 100


TCP
So

80 20 20 80
diu

3)
TNB

CO
m

TCH
(N

(H

3
60 40 40 CMP
60
N
a) a

ate

SO 4
Mg

GRN
+

HSB

on
Po

ACH
K

arb
BCD
tas

40 ARV 60 Bic 60 40
siu

MTP
m

DVS TCP
(K

20 80 80 ARV SMR 20
)

DVS

TNB GRN BCD BCD


0 SMR
TCH TCP 100 100 MTP 0
CMP ACH HSB
100 80 60 40 20 0 0 20 40 60 80 100
Ca Cl
Figure 8. Piper diagram for the hot spring waters showing their geochemical classification.

Cl
inferring water origin, mixing and evaporation processes, as well as 0 100
a proxy for identifying the intensity degree of water rock interaction
when the rock permeability is low (Bahati et al., 2005). Results of the
isotopic analyses (δ18O and δD) for the hot spring samples collected
20 80
Ma

in Sonora are reported in Table 5. The isotopic composition of these


tur

waters shows an approximated dispersion of 2.7 units (from -6.9 to


ew
ate

-9.6 with an accuracy of ±0.08 ‰) for 18O/16O, and 19 units (from -50
r
s

to -69 with an accuracy of 0.9 ‰) for D/H. 40 60


All these data points are shown in a δ18O – δD plot (Figure 12),
where the Global Meteoric Water Line (GMWL) is also represented.
s
ter
wa

Isotopic data from waters of three nearby rivers Ures (UR), Aconchi
60 40
ic

(AR) and Sonora (SR) are also plotted as reference of the local isotopic
lcan

composition of cold water in the study area. As seen in Figure 12, the
Vo

SMR

isotopic compositions of some hot spring waters lie near to the GMWL,
Pe
rip

particularly those corresponding to the hot springs TCH and MTP 80 20


he

TCP
ra

from the S and C hydrothermal zones, respectively. The remaining TCH


lw

samples (TNB, BCD, ACH, GRN, DVS, HSB, CMP, SMR, TCP, and
ate

Steam heated waters ACH


rs

ARV) show a clear shift for both δ18O and δD, which is described by GRN
ARV MTP 0
100 TNB HSB
the dashed mixing line (Figure 12). This signature is generally the CMP
DVS BCD

result of mixing with isotopically heavier water. The hot spring waters
0 20 40 60 80 100
are probably a mixture of waters similar to the most depleted local SO4 HCO3
groundwaters with a composition that results from the intersection
of the mixing line with the GMWL, and an isotopically heavier water Figure 9. Ternary diagram showing the variation of major anions (Cl–, SO42– and
not identified in the study zone. HCO3–) in the hot spring waters.

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 131


Almirudis et al.

3000
Q - SiO2
T - Na2SO4
2500 C - CaCO3

2000
Intensity (counts)

1500

1000

500

a) 0
10 20 30 40 50 60 70 80
Dos-Theta (degrees)

140 CaSO4.2H2O

120

100
Intensity (counts)

80

60

40

20
b)
x10^3
10 20 30 40 50 60 70 80
Dos-Theta (degrees)

T - Na2SO4
80 H - NaCl

70

60
Intensity (counts)

50

40

30

20

10
c) 0
14 20 30 40 50
Dos-Theta (degrees)

Figure 10. XRD diffractograms of precipitated salt samples collected at the (a) CMP, (b) TCH and (c) TNB hot springs.

Fluid geothermometry The Na/K ratios ranged between 20.7 and 48.1 for the spring waters
Pseudo-equilibrium temperature estimates were obtained by the of the NW, NE (except BCD: 143 and HSB: 467), and S zones, which
application of seven solute geothermometers (Na/K: 3, K-Mg, Na-K- provided reliable temperature approaches for the reservoir. These esti-
Ca, Na-Li, and SiO2). The results of these calculations are reported mates must be considered as lower limits of the reservoir temperatures
in Table 7. because the chemical signatures of the hot spring waters show evidence
of mixing with shallow groundwater, as well as the interaction with
Na/K geothermometer surface precipitated minerals. In this context, hot spring waters that
A good agreement among the temperatures calculated by three are subject to conductive cooling, lateral flow or near surface reactions
different versions of the Na/K geothermometer was roughly obtained. tend to exhibit high Na/K ratios (between 12.3 and 60) for temperatures

132 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

Li of equilibrium conditions between solutes and alteration minerals


0 100
(which is evidenced by the presence of incompletely equilibrated
spring waters, probably due to mixing or dilution processes); see
the classical Na-K-Mg geothermometer diagram (Figure 13). In this
20 HSB 80
triangular plot, a fast equilibrating process is represented by the K/Mg
DVS
TCP geothermometer, whereas the slow equilibrating process is given by
BCD the Na/K geothermometer (Giggenbach 1988).
40 ACH 60 By excluding the extreme temperatures inferred from ARV and
Outflow
MTP (the highest temperature estimates), and from HSB, and BCD
TCH
SMR CMP thermal springs (the lowest temperature estimates) (see Table 7), and
60
GRN
40 after applying a geochemometric analysis based on outliers detection/
TNB rejection under the assumption of a normal distribution, the mean
flow
reservoir temperature predicted with the TNaKVS97 equation for the
lt

Up
sa

hot springs CMP, GRN, SMR, TCP, TCH, DVS, TNB, and ACH was
Ba

80 20
MTP ARV
142±16 °C (see Figure 14).
yol
ite For the outlier detection/rejection, the univariate analysis applied
Rh
100 0 the Dixon, Grubbs, skewness and kurtosis statistical tests at the 99%
Rock dissolution
10.Cs
confidence level by using the computer code UDASYS of Verma et al.
4 Rb
.
0 20 40 60 80 100 (2013). A distribution map showing the most attractive geothermal
Figure 11. Ternary diagram showing relative Li, Rb, and Cs contents in the hot
potential zones (NW and NE) based on these reservoir temperature
spring waters. Outflow and upflow are schematically shown. Dotted area com- approaches is shown in Figure 15a. According to this map, lower
prises the group of hot spring samples falling between the basalt and rhyolite reservoir temperatures are mainly distributed in the S hydrothermal
rock boundaries, which limit the waters thay may have interacted with illite at zone (SMR, TCP and TCH).
temperatures between 190 and 210 °C (Goguel, 1983).
Na-Li geothermometer
Na/Li ratios were calculated after converting the ion concentra-
ranging from 100 °C to 200 °C (Nicholson, 1993; O’Brien, 2010). It is tions into mol/kg units. Na/Li ratios for all the hot spring waters
well-known that the prediction capability of this geothermometer at ranged from 76.5 to 230.7. Reservoir temperatures were estimated
temperatures below to 200 °C (i.e., with Na/K ratios ranging from 13 using the corresponding TNaLiVS97 equations by considering their
to 27) provide greater uncertainties than those ratios (ranging between molality constraints (chlorides) indicated in Table 3. After applying
3 and 13) corresponding to high temperatures (between 200 °C and the appropriate TNaLiVS97 equations, estimates of reservoir tempera-
340 °C). This limitation of the Na/K geothermometer is due to the tures for nearly all the hydrothermal zones under evaluation ranged
long WRI times needed for achieving equilibrium between Na and K from 101±21 °C to 178±27 °C. An anomalous lower temperature of
in controlled laboratory experiments at low-to-medium temperatures 55±17 °C was estimated for the HSB hot spring. For identifying the
(<200 °C; Pérez-Zarate et al., 2015). origin of this anomalous temperature, the Na/Li ratios of all the water
The TNaKVS97 geothermometer systematically provided higher samples were analysed. The highest value of this ratio was obtained
temperature estimates, whereas lower temperature values were com- for the HSB hot spring (Na/Li=576.3). High Na/Li ratios may suggest
puted with the TNaKDSR08 geothermometer. Anomalous higher either a preferential leaching of Na or a removal of some alteration
temperatures of 328±62 °C and 462±87 °C were also calculated for the minerals of Li (e.g., Li-chlorite, Fouillac and Michard, 1981).
ARV and MTP hot spring, respectively (values that were excluded from Following the same methodology used with the TNaKVS97 geo-
the Table 7). These unrealistic temperature estimates were expected thermometer, the geochemometric analysis based on outliers detec-
because such samples exhibited anomalous chemical and isotopic tion/rejection was again applied. From this statistical analysis, it was
signatures (probably caused by the high content of Mg), in comparison found that the anomalous lower temperature estimated for the HSB
with the remaining water samples. On the other hand, temperatures
corresponding to the HSB and BCD (which are alkaline together with
TCP) hot springs could not be determined with confidence due to the
-20
complex chemical and isotopic signatures that revealed some mixing in e
or diluting effects caused by the local hydrogeochemical processes rL
at e
W
c UR SR
recorded by these samples (see Figures 8–12). or i
eteAR
ᵟ D (‰, SMOW)

To analyse the anomalous temperature estimates, the Na/K -40


al
M
ob
concentration ratios of the geothermal fluids were computed. The Gl
TCP Hydrothermal zones
highest values of the Na/K ratios were obtained for the HSB and BCD TCH
MTP
SMR
ARV NE
NW
(466.7 and 142.9, respectively), whereas the lowest values correspond -60 ACHDVS CMP C
to the MTP and ARV hot springs (1.4 and 3.4, respectively). High BCD GRN HSB S
Na/K ratios may suggest either a preferential leaching of Na or River waters
TNB
Mixing line
removal of some of the K in the form of secondary alteration products
-80
(Giggenbach, 1988), which could be confirmed in future prospection
studies by collecting fresh and altered rock samples for analysing -12 -10 -8 -6 -4 -2

ᵟ O (‰, SMOW)
18

the presence of alteration minerals (e.g., muscovite, microcline).


Low Na/K ratios were associated to a prediction of unrealistic high Figure 12. Plot of δD versus δ18O of all hot spring water samples collected in the
temperatures. These geochemical anomalies may be attributed either four hydrothermal zones (NW, NE, C and S). AR, UR and SR are water samples
to ion exchange processes among clay minerals or simply to the lack collected in nearby rivers of Sonora.

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 133


Almirudis et al.

Table 7. Deep reservoir temperatures estimated from solute geothermometers (°C).

Zone Sample Tsup TNa/K1 TNa/K2 TNa/K3 TK−Mg4 TNa−K−Ca5 TNa/Li6 TSiO27

NW ACH 50.8 ± 4.1 137.5 ± 42.8 143.2 ± 34.4 98.4 ± 8.3 101.0 200.1 157.6 ± 25.4 123.3 ± 1.0
NW CMP 45.5 ± 0.2 111.3 ± 38.9 117.7 ± 31.4 69.1 ± 7.4 71.7 321.3 146.4 ± 24.4 108.4 ± 0.9
NE BCD 34.8 ± 3.9 N.A. 68.8 ± 25.9 N.A. 69.5 83.1 101.1 ± 20.7 101.9 ± 0.9
NE HSB 35.9 ± 0.4 N.A. 27.7 ± 21.8 N.A. 72.1 52.4 54.6 ± 17.2 97.8 ± 0.9
NE GRN 46.1 ± 1.2 142.3 ± 43.5 147.9 ± 35.0 103.9 ± 8.5 101.5 136.4 147.7 ± 24.5 119.5 ± 1.0
NE TNB 49.8 ± 1.0 147.3 ± 44.3 152.7 ± 35.6 109.7 ± 8.6 104.0 229.1 136.4 ± 23.6 132.8 ± 1.1
NE DVS 29.3 ± 0.5 161.8 ± 46.6 166.8 ± 37.4 126.5 ± 9.2 76.6 147.8 178.3 ± 27.2 111.9 ± 1.0
C MTP 30.9 ± 2.9 N.A. N.A. N.A. 40.9 21.7 144.9 ± 24.3 69.2 ± 0.7
C ARV 40.0 ± 0.2 N.A. N.A. N.A. 39.8 187.3 104.5 ± 20.9 81.9 ± 0.8
S TCP 41.5 ± 7.2 118.2 ± 39.9 124.4 ± 32.2 76.7 ± 7.6 76.6 116.2 117.6 ± 22.0 107.2 ± 0.9
S SMR 31.4 ± 0.2 132.1 ± 42.0 138.0 ± 33.8 92.3 ± 8.1 61.9 118.5 115.0 ± 21.8 95.2 ± 0.9
S TCH 63.1 ± 0.6 142.4 ± 43.5 148.0 ± 35.0 104.0 ± 8.5 112.3 126.8 128.7 ± 22.9 118.8 ± 1.0
1
Fournier (1979); 2,6,7Verma and Santoyo (1997); 3Díaz-González et al. (2008); 4Giggenbach (1988); 5Fournier and Truesdell (1973). N.A.  the geothermometer
equations were not applied because the estimates provided unrealistic temperature values.

hot spring belongs to the same statistical sample of the TNaLiVS97 Anomalous higher temperatures of 321 °C, 229 °C, and 200 °C
temperature estimates (i.e., it was not an outlier value). Therefore, the were calculated for the CMP, TNB, and ACH hot springs, respectively.
mean reservoir temperature inferred from the TNaLiVS97 indicated On the other hand, lower temperatures of 22 °C, 52 °C and 83 °C for
values of 134±24 °C (Figure 14). A distribution map showing the the MTP, HSB and BCD hot spring waters were respectively estimated
obtained TNaLiVS97 reservoir temperatures, and the most attractive mainly due to the low Ca concentrations observed in these samples.
geothermal potential zones is shown in Figure 15b. Similarly to the In order to analyse these temperature estimates, the √Ca/Na ratios of
contour map of TNaKVS97 temperatures, lower reservoir temperatures the hot spring waters were analysed. Highest values of √Ca/Na ratios
were associated to the S hydrothermal zone (i.e., SMR, TCP and TCH). were measured for the ARV and MTP hot springs (53.96 and 170.77,
respectively), whereas the lowest values were determined in HSB and
Na-K-Ca geothermometer BCD (0.57 and 0.99, respectively). High √Ca/Na ratios may suggest
Before applying the Na-K-Ca (TNaKCaFT73) geothermometer, either a preferential leaching of Ca (e.g. in carbonated rock formations)
√Ca/Na ratios were first computed. Such ratios ranged between 2.07 or a removal of some of the Na in the form of secondary alteration
and 4.65, which provided acceptable reservoir temperatures for the products (e.g., NaSO4). Low √Ca/Na ratios may be also due to either
hydrothermal zones under evaluation. These temperature approaches Ca precipitation (e.g., travertine) or to the lack of equilibrium between
varied between 116 °C and 148 °C. solutes and alteration minerals observed in these zones.

Na/1000
0 100

20 80

40 Fully equilibrated HSB 60


waters
140°
180°
100°
220°
80°
60 40
260°
Partially equilibrated
or mixed waters BCD
300° TCH
ACH
80 TNB GRN 20

TCP
CMP
Immature waters
DVS
SMR
100 0
MTP ARV
K/100 √Mg
0 20 40 60 80 100

Figure 13. Ternary equilibrium diagram showing the relative Na, K and Mg contents for all the hot spring samples collected in the four hydrothermal zones of the
Sonora geothermal system (based on Giggenbach, 1988).

134 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

400 are statistically the same; Ha2: not all temperature estimates among
the geothermometers are statistically the same. For α=0.05 and the
obtained p-value=8.5×10-4, Ho2 is rejected;
145.1±81.2 °C
(n=12)
2) Cycle # 2: Statistical hypotheses (after removing all the tem-
perature estimates predicted by the TNaKDSR08 geothermometer
300 because these data were identified as the lowest estimates according
to the outlier detection algorithm) — According to the calculated F
Temperature estimates (°C)

values and following the same statistical methodology, it is concluded


that Ho1 cannot be rejected; whereas the Ho2 is rejected; and,
3) Cycle # 3: Statistical hypotheses (after removing all the tem-
200
142.3±15.7 °C
134.4±23.7 °C perature estimates predicted by the TSIO2VS97 geothermometer
(n=11)
136.6±16.1 °C
(n=8)
(n=8) with the same statistical argument) — According to the calculated F
105.6±17.9 °C values and following the same statistical methodology, it is concluded
97.6±18.3 °C
(n=8)
(n=12) that both hypotheses Ho1 and Ho2 cannot be rejected, and therefore,
the mean reservoir temperature may be estimated as 149±40 °C using
100 the temperature estimates of the ACH, TCP, TNB, TCH, CMP, DVS,
GRN, and SMR hot springs (see Table 8). This conclusion highlights
the importance of applying suitable geothermometric tools at each
geothermal site.

0 Mixing geochemical model


TNaKF79 TNaKVS97 TNaKDSR08 TNaKCaFT73 TNaLiVS97 TSiO2VS97 Taking into consideration the water mixing processes evidenced
Geothermometers from the classification diagrams (Figures 8, 9, 11) and the D/H and
Figure 14. Results of the statistical outlier detection/rejection analysis applied
18
O/16O plot (Figure 12), a mixing geochemical model based on a silica
for all the solute geothermometers, and mean reservoir temperatures estimated and fluid enthalpy plot was applied for inferring the water composition
by each solute geothermometer. and temperature at the original reservoir conditions. The silica content
of ascending hot spring waters (which fall outside of the full equilibrium
curve: Figure 13) and their corresponding fluid enthalpy were plotted
After applying the geochemometric analysis (based on outliers in Figure 16. The influence of a conductive cooling was quantitatively
detection/rejection under the assumption of a normal distribution), a examined for determining the mixing degree with colder waters.
mean reservoir temperature from the TNaKCaFT73 was estimated as After applying the model, it was inferred that the original reser-
145±81 °C (Figure 14). A distribution map showing the geothermal voir conditions indicate a deep reservoir temperature of 210±11 °C,
potential zones (NW and NE) from the reservoir temperatures pre- together with a silica content of 300±27 mg/kg, and a fluid enthalpy
dicted by the TNaKCaFT73 is presented in Figure 15c, which also agrees of 950±85 kJ/kg (see the blue point C in Figure 16). This reservoir
with the temperature distribution field indicated by the TNaKVS97. temperature is higher than those estimates predicted from the solute
geothermometers (Table 7), and it may be considered as the maximum
SiO2 geothermometer reservoir temperature of the Sonora geothermal system.
The use of the TSiO2VS97 geothermometer for the determination According to the Li-Rb-Cs diagram (Figure 11), this reservoir
of reservoir temperatures predicted mean values of 106±18 °C, which temperature of 210±11 °C is in good agreement with the temperature
were systematically lower than those reservoir temperatures estimated that controls the illite mineral interaction, and the temperature interval
by the Na/K, Na/Li, and Na-K-Ca geothermometers. inferred from the K/Rb and K/Cs ratios. All the geochemical signatures
observed at deep reservoir conditions reflect the degree of mixing of
Estimation of the mean reservoir temperature the waters during their ascent towards the surface.
A two-way analysis of variance (ANOVA), with one sample by
group at α=0.05 (or 5%), was applied to the reservoir temperature Fluid-mineral equilibria
estimates (excluding the rejected values from the outliers analysis, For evaluating the pseudo-equilibrium state that actually exhibit
Table 7). The ANOVA test was iteratively applied for evaluating both most of the waters sampled in the twelve hot springs, the Na-K-Mg
the statistical differences among temperature estimates for hot springs diagram was again examined (Figure 13). This diagram associates the
and solute geothermometers under the following two strict cases of hydrothermal fluids with their most probable reservoir temperatures.
analysis: (i) if the reservoir temperature estimates of the hot spring As seen in Figure 13, most of the hot spring waters fall in the region
waters were statistically the same among hot springs (Ho1); and (ii) if of mixed waters or partial equilibrium (e.g., ACH and CMP from the
there are statistical significant differences among the predictions made NW zone; BCD, GRN, TNB, and DVS from the NE zone; and TCP,
by the solute geothermometers (Ho2). When both strict conditions were SMR and TCH from the S zone). Some exceptions were given by water
statistically fulfilled, the estimation of the mean reservoir temperature samples falling outside or in the limits of the full equilibrium line
was then calculated. Table 8 summarises the results obtained from the (HSB and BCD from the NE zone), and two remaining water samples
iterative ANOVA analysis. After three cycles of this analysis, the fol- located in the section of immature waters (MTP and ARV from the C
lowing results were obtained: zone). This geochemical shift was probably caused by the loss of some
1) Cycle # 1: Statistical hypotheses — Ho1: all temperature estimates ions (e.g., K) by these two water samples. Another possible cause may
among the hot springs are statistically the same; whereas Ha1: not all be related to the total dissolved carbonate concentration, which in
temperature estimates among the hot springs are statistically the same; the specific case of the HSB and BCD water samples comprises both
For α=0.05 and the obtained p-value=0.225, Ho1 cannot be rejected (see HCO3– and CO32– species). The total carbonate concentration may also
Table 8). Ho2: all temperature estimates among the geothermometers be influenced by the CO2 partial pressure in the deep geothermal fluid

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 135


Almirudis et al.

a) b)

c)
c)

c) d)
Figure 15. Distribution map showing the most attractive geothermal zones (NW and NE) based on the most reliable reservoir temperature approaches (inferred
from the solute geothermometers): a) TNaKVS97; b) TNaLiVS97; and c) TNaKCaFT73. The distribution of the surface temperatures measured in the four hy-
drothermal zones is also plotted as reference (d).

(PCO2) and the pH (Nicholson, 1993). The loss of CO2 during boiling with some evaporated water sources (Figure 11).
processes normally raises the pH of the fluid by consuming protons From the thermal point of view, most of the water samples were
through the involved reactions (i.e., the water usually becomes more distributed in the partial equilibrium region where the reservoir tem-
alkaline, as in the case of HSB and BCD samples which exhibited the peratures vary from 75 °C to 175 °C, most of them falling close to the
highest pH values of 9.0 and 8.5, respectively; see Table 5). 150 °C isotherm, which seems to be the lower limit of the reservoir
The lack of equilibrium with the surrounding host rocks is also temperatures of the geothermal system.
confirmed by the mixing model (see Figure 16), and explained either Using 150 °C and 200 °C as rounded-off reservoir temperatures
as a result of a mixing between hydrothermal waters and colder shallow (or min-max values) and the Geochemist’s Workbench software,
fluids during the path towards the surface, or simply due to meteoric geochemical equilibrium modelling based on fluid-mineral stability
waters coming from different sources or river waters (Figure 11; e.g., diagrams was carried out. The dominant species and mineral phases
MTP and ARV). The origin of the HSB hot spring water (located above that are stable over this temperature interval (i.e., mineral stability
the full equilibrium line in Figure 13) is probably related to mixing diagrams of the primary and secondary minerals for these reactions)

136 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

Table 8. Two factor variance analysis with one sample by group applied to the equilibrium temperature estimates (α=0.05).

Sample TNaKF79 TNaKVS97 TNaKDSR08 TNaKCaFT73 TNaLiVS97 TSiO2VS97

ACH 137.5 143.2 98.4 200.1 157.6 123.3


CMP 111.3 117.7 69.1 321.3 146.4 108.4
GRN 142.3 147.9 103.9 136.4 147.7 119.5
TNB 147.3 152.7 109.7 229.1 136.4 132.8
DVS 161.8 166.8 126.5 147.8 178.3 111.9
TCP 118.2 124.4 76.7 116.2 117.6 107.2
SMR 132.1 138 92.3 118.5 115 95.2
TCH 142.4 148 104 126.8 128.7 118.8
Iterative Cycle # 1
Variation origin Square addition Degrees of freedom Square means F Probability F critical value
Rows 10242.63 7 1463.23 1.43 0.22536 2.28
Columns 27740.86 5 5548.17 5.42 0.00085 2.49
Error 35853.67 35 1024.39
Total 73837.16 47
Iterative Cycle # 2
Variation origin Square addition Degrees of freedom Square means F Probability F critical value
Rows 10673.48 7 1524.78 1.29 0.29097 2.35
Columns 14693.01 4 3673.25 3.1 0.03092 2.71
Error 33082.75 28 1181.52
Total 58449.25 39
Iterative Cycle # 3
Variation origin Square addition Degrees of freedom Square means F Probability F critical value
Rows 11504.36 7 1643.48 1.1 0.39 2.48
Columns 7306.84 3 2435.61 1.63 0.21 3.07
Error 31326.45 21 1491.73
Total 50137.66 31

Iterative Cycle # 1: Statistical hypotheses  Ho1: all temperature estimates among the hot springs are statistically the same, i.e., come from the same
statistical sample; whereas Ha1: not all temperature estimates among the hot springs are statistically the same; For α=0.05 and the obtained p-value=0.225,
Ho1 cannot be rejected. It is therefore concluded that all the temperature estimates among the hot springs are statistically the same. Ho2: all temperature
estimates among the geothermometers are statistically the same; Ha2: not all temperature estimates among the geothermometers are statistically the same.
For α=0.05 and the obtained p-value=8.5E-04, Ho2 is rejected. It is also concluded that not all the temperature estimates among the geothermometers
are statistically the same.
Iterative Cycle # 2: Statistical hypotheses (after removing TNaKDSR08 temperature estimates)  According to the F calculated values and following the
same statistical methodology, it is concluded that Ho1 cannot be rejected; whereas the Ho2 is rejected.
Iterative Cycle # 3: Statistical hypotheses (after removing TSIO2VS97 temperature estimates)  According to the F calculated values and following the same
statistical methodology, it is concluded that Ho1 and Ho2 cannot be rejected, and therefore, the mean reservoir temperature may be estimated as 149±40 oC.

were investigated. The following water-rock (mineral) systems were interaction shows much better equilibrium conditions at temperatures
found as the major chemical and mineral signatures underlying the close to 200 °C than those data corresponding to 150 °C (e.g., Henley
geothermal system under study (see the mineral stability diagrams in and Ellis, 1983; Hedenquist et al., 2000). These interactions were mainly
Figures 17a and 17b): observed in the hot spring waters of the NW (ACH and CMP), NE
1) A Na2O-K2O-Al2O3-SiO2-H2O system, which exhibited ionic (BCD, GRN and TNB), and S (SMR and TCH) zones.
exchange between Na and K that seems to be controlled either by the 2) A CaO-K2O-Al2O3-SiO2-H2O system, where the ionic exchange
albite/muscovite or the albite/K-feldspar mineral interactions under between Ca and K is described by the prehnite-laumontite-muscovite
metastable equilibrium conditions. This system reflects the hydro- mineral assemblage (see Figure 17). These WRI processes seem to be
thermal alteration of primary plagioclase minerals in agreement with associated with the hot spring waters from the NE (DVS and HSB),
the following reaction: and S (TCP) zones. Laumontite is a Ca-rich zeolite typically observed
in some geothermal fields at temperatures between 170 °C and 250 °C
Na-Plag + K+ + H+ ↔ K-mica + Na+ (Reyes, 1990; Torres-Rodríguez, 2000). It would be expected that hot
spring waters in equilibrium with laumontite indicate the existence
The discharge fluid composition tends to be placed along an exten- of a reservoir with a temperature interval between 120 °C and 210 °C
sion from either the albite/K-feldspar limit through the muscovite field, (Reyes, 1990).
or through phases chemically similar like sericite and illite (Hedenquist, Most of these minerals belong to mineral assemblages of low
1991). This behaviour can be interpreted as the result of the primary grade metamorphism (Liou et al., 1985). This WRI signature seems to
dissolution of feldspars in the reservoir under metastable equilibrium indicate that, according to the regional geology of the study area, the
conditions between albite and K-feldspar (Torres-Alvarado, 2002). As hot spring waters are reaching the chemical equilibria at the contact
seen in Figure 17, the albite and muscovite (K-mica as illite or sericite) zones between the plutonic rocks (from the Laramide Batholith) and

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 137


Almirudis et al.

800
CONCLUSIONS
Hydrothermal zones
NE
700 NW The presence of hot springs, Quaternary volcanism, and tectonic
C activity at the east-central Sonora state (Mexico) was reported in this
S
prospection study. A field work campaign for the collection of rep-
600 Quartz solubility curve
resentative samples from twelve hot springs was carried out. From

e
Point A (Model)

v
chemical and isotopic analyses of these samples, a study of the major

Cur
Point C (Model)
500 chemical and isotopic signatures of fluids produced in this low-to-

ility
Mixing line: SiO2 = 0.2632 * HF + 14.812 (n=8)
medium temperature geothermal system was inferred.
SiO2 (mg/kg)

ub
Sol
400
Hydrothermal waters with the highest surface temperatures (up to

rtz
63.1 °C) were mainly associated to hot springs emanating on granite

Qua
Point C rocks (in the basin and range limits marked by regional faults), whereas
300 (900.0 kJ/kg, 252 mg/kg)
= 210±11 oC the lowest surface temperatures (up to 29.3 °C) were mostly observed
for hot springs that emanate on recent sediments (alluvium) within
the valleys. Based on the chemical signatures of the hot spring waters
200
line
ing (major and trace elements), depletion (F–, Cl–, K+, and Li+ and Cs, Sr,
Mix
TNB
Sb, V, Co and Tl) and enrichment (in HCO3–, SO42–, Na+ and Ca2+ and
100
DVS
GRN
BCD HSB TCH
Rb, Ba, As, Al, Mn and Fe) in fluid compositions were identified. A high
TCP ACH
CUMBEN
SMR
MTP
ARV CMP
chemical mobility from the host rock to the fluid was also observed in
0 Mn, Fe, Sr, As, Rb, Ba, and the light REE (La and Ce).
0 100 200 300 400 500 600 700 800 900 1000 1100 1200 The interpretation of hydrogeochemistry classification diagrams
Fluid enthalpy (kJ/kg) mostly indicated the presence of sodium-sulphate (Na-SO4) waters in
the North (NW: ACH and CMP; and NE: GRN and TNB) and South
Figure 16. SiO2-enthalpy mixing model and mixing line obtained for the hot
(TCP, SMP and TCH) hydrothermal zones; whereas calcium and
spring waters of the Sonora geothermal system.
magnesium-bicarbonate (Ca-Mg-HCO3) waters were found for the
Central zone (MTP and ARV). Two hot spring waters located in the NE
(BCD, HSB and DVS) zone were characterized as sodium-bicarbonate
the tectonic basins (i.e., at regional fault zones under the influence of (Na-HCO3) type waters. The signatures of stable isotopes (δ18O and
those plutonic rocks and low-grade metamorphism rocks). The litho- δD) indicate a meteoric origin for four hot spring waters (NE: BCD
logical units that represent the former metamorphic features are the and TNB; S: TCH; and C: MTP), which lie near to the global meteoric
intermediate to acidic volcanic rocks from the Tarahumara Formation. water line; whereas a small shift for both isotopes was observed in the
The interpretation of the most probable pseudo- or equilibrium rest of the samples (NW: ACH and CMP; NE: HSB, GRN and DVS;
conditions among the hot spring waters and the mineral assemblages C: ARV; and S: TCP and SMR). This small δ18O shift typified a mixing
in the Na2O-K2O-Al2O3-SiO2-H2O and CaO-K2O-Al2O3-SiO2-H2O line which was used as a proxy to discriminate waters with a different
systems seem to be in a good agreement with the expected reservoir isotopic signature.
temperatures observed in the hydrothermal zones, and the theoretical After applying a geochemometric analysis (based on outlier de-
thermal stability of hydrothermal minerals by Henley and Ellis (1983) tection/rejection and ANOVA), the mean reservoir temperature was
and Hedenquist et al. (2000): Figure 18. estimated as 149±40 °C. By considering the mixing process among hot

a) b)

Figure 17. Fluid-mineral stability diagrams obtained for the (a) Na2O-K2O-Al2O3-SiO2-H2O and (b) CaO-K2O-Al2O3-SiO2-H2O systems.

138 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

a) b)
Figure 18. Thermal stability of various hydrothermal minerals (modified after Hedenquist et al., 2000 and Henley and Ellis, 1983). (a) Mineral assemblages associated
to pH and temperature; (b) Mineral assemblages associated only to temperature. Dashed orange lines indicate the pH and estimated temperature values prevailing
in the hot spring waters reported this study (purple and red dashed lines correspond to minimum and maximum estimates respectively).

spring and shallow colder waters, the temperature of the reservoir was infrastructure support for a sabbatical leave program carried out on
corrected by applying a conductive cooling model, from which a deep 2016-2017. The authors want to thank to the anonymous reviewers and
reservoir temperature of 210±11 °C was inferred. Both temperatures the RMCG Editor in Chief (A. Nieto) for their constructive comments
149±40 °C and 210±11 °C are considered with confidence as the that enabled the manuscript to be improved. Finally, the authors want
minimum (realistic) and maximum (optimistic) reservoir temperatures to dedicate this work in tribute to Dr. Ignacio S. Torres-Alvarado
of the Sonora geothermal system. Using some approaches to these (1964-2012).
reservoir temperatures (150 °C and 200 °C), as the most probable
equilibrium temperatures (min-max), it was found that the hot spring REFERENCES
waters are in equilibrium with mineral assemblages of the Na2O-K2O-
Al2O3-SiO2-H2O (NW: ACH-CMP; NE: BCD-GRN-TNB; and S: SMR Aggarwal, J.K., Palmer, M.R., Bullen, T.D., Arnórsson, S., Ragnarsdóttir, K.V.,
and TCH) and CaO-K2O-Al2O3-SiO2-H2O (NE: DVS-HSB; and S: 2000, The boron isotope systematics of Icelandic geothermal waters: 1.
TCP) systems. These minerals are proposed as representative mineral Meteoric water charged systems: Geochimica et Cosmochimica Acta,
assemblages of low-grade metamorphism. 64, 579-585.
Almirudis, E., 2010, Petrogénesis y Geocronología 40Ar-39A del Plutonismo
The integrated geochemical model here proposed enabled the
Laramídico del área Sobai Satechi, Sonora, México: Hermosillo, México,
evolution of the Sonora hydrothermal reservoir to be described, where Universidad de Sonora, Tesis de Licenciatura, 106 pp.
a deep fluid circulation characterised by low-to-medium temperatures Almirudis, E., Guevara, M., Santoyo, E., Torres-Alvarado, I.S., Paz-Moreno, F.,
(t<200 °C) dominates in the system. Based on this geochemical model, 2015, Geothermal energy potential of a promissory area in the Central
it is concluded that the hot spring waters from the geothermal system and Eastern zones of Sonora, Mexico: A preliminary geochemical study,
of east-central Sonora have the potential to be used not only for heat in Proceedings of the World Geothermal Congress 2015: Melbourne,
direct use, but also for electricity production through plants of binary Australia, 19-25 April 2015, p. 9.
cycle, which are suitable to be safely operated with fluid temperatures Arango-Galván, C., Prol-Ledesma, R.M., Torres-Vera, M.A., 2015, Geothermal
prospects in the Baja California Peninsula: Geothermics, 55, 39-57.
between 100 °C and 180 °C and with very low environmental issues.
Atwater, T., 1989, Tectonic Maps of the Northeast Pacific in Winterer E.L.,
Hussong D.M., Decker R.W. (eds.), The Eastern Pacific Ocean and Hawaii:
Geological Society of America, Boulder, Colorado, The Geology of North
ACKNOWLEDGEMENTS America, 15-20.
Bahati, G., Pang, Z., Ármannsson, H., Isabirye, E.M., Kato, V., 2005, Hydrology
The authors want to thank to DGAPA–PAPIIT-UNAM project and reservoir characteristics of three geothermal systems in western
(IT101014) and Project P09 CeMIE-Geo (Project 207032 CONACyT- Uganda: Geothermics, 34, 568-591.
SENER) for partial financial support. We also thank Veda Lome Barragán, R., Birkle P., Portugal E., Arellano V., Alvarez H., 2001, Geochemical
survey of medium temperature geothermal resources from the Baja
for drawing some of the figures. The first author acknowledges the
California peninsula and Sonora, México: Journal of Volcanology and
Engineering (Energy) Graduate Programme of UNAM and CONACyT Geothermal Research, 110, 101-119.
for scholarship support. The second author also wants to thank to Bertani, R., 2016, Geothermal power generation in the world 2010-2014 update
CONACyT and CIICAp-UAEM for their generous financial and report: Geothermics, 60, 31-43.

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 139


Almirudis et al.

Bethke, C.M., 2008, Geochemical and Biogeochemical Reaction Modeling: concentrations in the geothermal springs along the west coast of
Second Edition. Cambridge University Press, New York, 16 pp. Maharashtra, India: Arabian Journal of Geosciences, 9(1), 44, 15 pp,
Bevington, P.R., Robinson, D.K., 2003, Data Reduction and Error Analysis for https://doi.org/10.1007/s12517-015-2139-2.
the Physical Sciences: Third Edition, McGraw-Hill Higher Education, Gutiérrez-Negrín, L.C.A., Maya-González, R., Quijano-León, J.L., 2015, Present
Boston, 338 pp. Situation and Perspectives of Geothermal in Mexico, in Proceedings World
Bright, J., Kaufman, D.S., Forman, S.L., McIntosh, W.C., Mead, J.I., Baez, A., Geothermal Congress 2015: Melbourne, Australia, April 19-24, 2015.
2010, Comparative dating of a Bison-bearing late-Pleistocene deposit, Hedenquist, J., 1991, Boiling and dilution in the shallow portion of the Waiotapu
Térapa, Sonora, Mexico: Quaternary Geochronology, 5, 631-643. geothermal system, New Zealand: Geochimica et Cosmochimica Acta,
Cochemé, J., 1985, Le magmatisme dans le Nord-Ouest du Mexique. Cartographie 53, 2235-2257.
de la région de Yécora-Maicoba-Mulatos. Illustration magmatique de la fin Hedenquist, J.W., Arribas, A., Gonzalez-Urien, E., 2000, Exploration for
d’un régime en subduction et du passage à un régime distensif : Marseille, epithermal gold deposits: Reviews in Economic Geology 13(2), 45-77.
France, Université Aix-Marseille III, These d’Etat, 209 pp. Henley, R.W., Ellis, A.J., 1983, Geothermal systems ancient and modern: a
Cochemé, J., Demant, A., Aguirre, L., Hermitte, D., 1988, Présence de heulandite geochemical review: Earth-Science Reviews, 19(1), 1-50.
dans les remplissages sédimentaires liés au ’basin and range’ (Formation Hiriart, G., Gutiérrez-Negrín, L., Quijano-León, J., Ornelas-Celis, A., Espíndola,
Báucarit) du nord de la Sierra Madre Occidental (Mexique) : Comptes S., Hernández, I., 2011, Evaluación de la Energía Geotérmica en México:
Rendus de l’Académie des Sciences Paris, 307, 643-649. México, D.F. Informe para el Banco Interamericano de Desarrollo y
Craig, H., 1961, Isotopic variations in meteoric waters: Science, 133, 1702-1703. la Comisión Reguladora de Energía (CRE), http://www.cre.gob.mx/
Damon, P.E., Shafiqullah, M., Roldán-Quintana, J., Cochemé J.J., 1983, El documento/2026.pdf; 164 pp.
batolito larámide (90-40) de Sonora in Proceedings of XV Convención Iglesias, E., 2003, First assessment of Mexican low- to medium- temperature
Nacional de Ingeniería en Minas, Metalurgia y Geología de México, 63-95. geothermal reserves: Energy Sources, 25, 161-173.
D’Amore, F., Arnórsson, S., 2000, Geothermometry in Arnórsson, S. (ed.), Isotopic Iglesias, E., Torres, R., Martínez-Estrella, I., Reyes-Picasso, N., 2011, Resumen de
and Chemical Techniques in Geothermal Exploration, Development and la evaluación 2010 de los recursos geotérmicos mexicanos de temperatura
Use: Sampling Methods, Data Handling, Interpretation: Vienna, Austria, intermedia a baja: Geotermia, Revista Mexicana de Geoenergía, 24, 39-48.
International Atomic Energy Agency, 152-199. IMTA (Instituto Mexicano de Tecnología del Agua), 2013, Eric III
Delany, J., Lundeen, S., 1990, The LLNL Thermochemical Database: Lawrence Versión 3.2 - Extractor rápido de información climatológica:
Livermore National Laboratory, 150 pp. available at <https://www.imta.gob.mx/productos/software/
Demant, A., Cochemé, J., Delpretti, P., Piguet, P., 1989, Geology and petrology of eric-iii-version-3-2-extractor-rapido-de-informacion-climatolo-detail>.
the Tertiary volcanics of the northwest Sierra Madre Occidental, Mexico: INEGI (Instituto Nacional de Estadística y Geografía), 2017, Fallas fracturas,
Bulletin de la Societé Géologique de France, 8, 737-748. in Datos Vectoriales escala 1:1000000,<http://www.inegi.org.mx/geo/
Díaz-González, L., Santoyo, E., Reyes-Reyes, J., 2008, Tres nuevos geotermómetros contenidos/recnat/geologia/infoescala.aspx>, downloaded december 1,
mejorados de Na/K usando herramientas computacionales y 2017.
geoquimiométricas: aplicación a la predicción de temperaturas de sistemas King, R., 1939, Geological reconnaissance in Northern Sierra Madre Occidental
geotérmicos: Revista Mexicana de Ciencias Geológicas, 25(3), 465-482. of México: Geological Society of America, 50, 1625-1722.
Ellis, A. J., Mahon, W.A.J., 1967, Natural hydrothermal systems and experimental Kipng’ok, J., Kanda, I., 2012, Introduction to geochemical mapping, in Short
hot water/rock interactions (Part II): Geochimica et Cosmochimica Acta, Course VII on Exploration for Geothermal Resources: Lake Bogoria and
31, 519-538. Lake Naivasha, Kenya, Oct. 27-Nov. 18, UNU-GTP, GDC and KenGen,
Ferrari, L., Valencia-Moreno, M., Bryan, S., 2005, Magmatismo y tectónica en 13 pp.
la Sierra Madre Occidental y su relación con la evolución de la margen Lawton, T.F., González-León, C.M., Lucas, S.G., Scott, R.W., 2004, Stratigraphy
occidental de Norteamérica: Boletín de la Sociedad Geológica Mexicana, and sedimentology of the upper Aptian-upper Albian Mural Limestone
57(3), 343-378. (Bisbee Group) in northern Sonora, Mexico: Cretaceous Research, 25, 43-60.
Ferrari, L., Valencia-Moreno, M., Bryan, S., 2007, Magmatism and tectonics Libbey, R.B., Williams-Jones, A.E., 2016, Lithogeochemical approaches in
of the Sierra Madre Occidental and its relation with the evolution of the geothermal system characterization: An application to the Reykjanes
western margin of North America: Geological Society of America, Special geothermal field, Iceland: Geothermics, 64, 61-80.
Paper 422, 1-39. Liou, J.G., Maruyama, S., Cho, M., 1985, Phase equilibria and mineral
Flores-Armenta, M., 2013, Evolución de la geotermia en el servicio público parageneses of metabasites in low-grade metamorphism: Mineralogical
mexicano, in Foro Internacional sobre Energía Geotérmica: México, D.F., Magazine, 49, 321-333.
Secretaría de Energía, 58-63. Long, G.L., Winefordner, J.D. 1983, Limit of detection A closer look at the
Fouillac, C., Michard, G., 1981, Sodium/lithium ratio in water applied to IUPAC definition: Analytical Chemistry, 55(7), 712A-724A.
geothermometry of geothermal reservoirs: Geothermics, 10(1), 55-70. Lynch, D.J., Musselman, T.E., Gutmann, J.T. & Patchett, P.J., 1993, Isotopic
Fournier, R.O., 1977, Chemical geothermometers and mixing models for Evidence for the origin of Cenozoic volcanic rocks in the Pinacate volcanic
geothermal systems: Geothermics, 5, 41-50. field, northwestern Mexico: Lithos, 29, 295-302.
Fournier, R.O., 1979, A revised equation for the Na/K geothermometer: Martín-Barajas, A., 2000, Volcanismo y extensión en la Provincia Extensional del
Geothermal Resources Council, 3, 221-224. Golfo de California: Boletín de la Sociedad Geológica Mexicana, 53, 72-83.
Fournier, R., Truesdell, A., 1973, An empirical Na-K-Ca geothermometer for McDowell, F., Keizer, R., 1977, Timing of mid-Tertiary volcanism in the
natural waters: Geochimica et Cosmochimica Acta, 37, 1255-1275. Sierra Madre Occidental between Durango city and Mazatlán, México:
Gans, P., 1997, Large-magnitude Oligo-Miocene extension in southern Sonora: Geological Society of America Bulletin, 88, 1479-1487.
Implications for the tectonic evolution of Northwest Mexico: Tectonics, McDowell, F., Roldán-Quintana, J., Amaya-Martínez, R., 1997, Interrelationship
16, 388-408. of sedimentary and volcanic deposits associated with Tertiary extension in
García-Abdeslem, J., Calmus, T., 2015, A 3D model of crustal magnetization at Sonora, Mexico: Geological Society of America Bulletin, 109, 1349-1360.
the Pinacate Volcanic Field, NW Sonora, Mexico: Journal of Volcanology McDowell, F., Roldán-Quintana, J., Connelly, J., 2001, Duration of late
and Geothermal Research, 301, 29-37. Cretaceous-Early Tertiary magmatism in East-Central Sonora, México:
Giggenbach, W., 1988, Geothermal solute equilibria. derivation of Na-K-Mg- Geological Society of America Bulletin, 113, 521-531.
Ca geoindicators: Geochimica et Cosmochimica Acta, 52, 2749-2765. Morales-Arredondo, I., Rodríguez, R., Armienta, A., Villanueva-Estrada,
Giggenbach, W.F., 1991, Chemical Techniques in Geothermal Exploration, in R.E., 2016, A low temperature geothermal system in central Mexico:
D’Amore F. (ed.), Application of Geochemistry in Geothermal Reservoir Hydrogeochemistry and potential heat source: Geochemical Journal,
Development UNITAR, 119-144. 50(3), 211-225.
Goguel, R.L., 1983, The rare alkalies in hydrothermal alteration at Wairakei Nicholson, K., 1993, Geothermal Fluids: Chemistry and Exploration Techniques:
and Broadlands, geothermal fields, New Zealand: Geochimica et Springer-Verlag, Berlin Heidelberg, 263 pp.
Cosmochimica Acta, 47, 429-437. O'Brien, J.M., 2010, Hydrogeochemical Characteristics of the Ngatamariki
Gurav, T., Singh, H.K., Chandrasekharam, D., 2016, Major and trace element Geothermal Field and a Comparison with the Orakei Korako Thermal

140 RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397


Low-to-medium temperature geothermal systems in Sonora

Area, Taupo Volcanic Zone, New Zealand: New Zealand, University of Smith, D.J., Jenkin, G.R.T., Naden, J., Boyce, A.J., Petterson, M.G., Toba, T.,
Canterbury, M.Sc. Thesis, 168 pp. Darling, W.G., Taylor, H., Millar, I.L., 2010, Anomalous alkaline sulphate
Oskin, M.E., Stock, J., 2003, Pacific-North America plate motion and opening fluids produced in a magmatic hydrothermal system—Savo, Solomon
of the Upper Delfín basin, northern Gulf of California, Mexico: Geological Islands: Chemical Geology, 275(1-2), 35-49.
Society of America Bulletin, 115, 1173-1190. Stock, J., Molnar, P., 1988, Uncertainties and implications of the Late Cretaceous
Pandarinath, K., Domínguez, H., 2015, Evaluation of the solute geothermometry and Tertiary position of North America relative to the Farallon, Kula and
of thermal springs and drilled wells of La Primavera (Cerritos Colorados) Pacific Plates: Tectonics, 7, 1339-1384.
geothermal field, Mexico. A geochemometrics approach: Journal of South Stewart, J., 1978, Rift systems in the western United States, in Ramberg I.B.,
American Earth Sciences, 62, 109-124. Newmann E.R. (eds.), Tectonics and Geophysics of Continental Rifts:
Pang, Z.H., Reed, M., 1998, Theoretical chemical thermometry on geothermal Reidel Public Company, 89-109.
waters: problems and methods: Geochimica et Cosmochimica Acta 62, Taylor, H., 1979, Oxygen and hydrogen isotope relationships in hydrothermal
1083-1091. mineral deposits, in Geochemistry of hydrothermal ore deposits: New
Paz-Moreno, F.A., 1993, Le volcanisme mio-plio-quaternaire de l’Etat du York, Wiley-Interscience, 236-277.
Sonora (nord- ouest du Mexique): Évolution spatiale et chronologique; Torres, V., Arellano, V., Barragán, R., González, E., Herrera, J., Santoyo, E.,
implications pétrogénétiques : Marseille, France, Université Aix-Marseille, Venegas, S., 1993, Geotermia en México: México, Programa Universitario
Ph.D. Thesis, 212 pp. de Energía, Coordinación de la Investigación Científica (CIC-UNAM),
Paz-Moreno, F.A., Demant, A., Cochemé, J.J., Dostal, J., Montigny, R., 2003, The 161 pp.
Quaternary Moctezuma volcanic field: a tholeiitic to alkali basaltic episode Torres-Alvarado, I.S., 2002, Chemical Equilibrium in Hydrothermal Systems:
in the central Sonoran Basin and Range Province, Mexico: Geological The Case of Los Azufres Geothermal Field, Mexico: International Geology
Society of America, Special Paper 374, 439-455. Review, 44, 639-652.
Pérez-Zárate, D., Santoyo, E., Guevara, M., Torres-Alvarado, I.S., Peiffer, L., Torres-Rodríguez, V., 2000, Geothermal Chart of Mexico scale 1:2000000, in
Martínez-Frías, J., 2015, Geochemometric modeling and geothermal Proceedings World Geothermal Congress 2000: Japan, 1867-1870.
experiments of Water/Rock Interaction for the study of alkali-feldspars Verdugo-Mariscal, 1983, Geotermia en Sonora: Hermosillo, Sonora, Universidad
dissolution: Applied Thermal Engineering 75, 1244-1261. de Sonora, Tesis de Licenciatura, 101 pp.
Prol-Ledesma, R., 1991, Terrestrial heat flow in Mexico, in Cermák V., Rybach, Verma, S.P., Santoyo, E., 1997, New improved equations for Na/K, Na/Li and
L., Exploration of the Deep Continental Crust: Springer-Verlag, Berlin SiO2 geothermometers by outlier detection and rejection: Journal of
Heidelberg, 475-485. Volcanology and Geothermal Research, 79, 9-23.
Prol-Ledesma, R., Juárez, G., 1986, Geothermal map of Mexico: Journal of Verma, S.P., Pandarinath, K., Santoyo, E., 2008, SolGeo: A new computer
Volcanology and Geothermal Research, 28, 351-362. program for solute geothermometers and its applications to Mexican
Prol-Ledesma, R., Canet, C., 2004, Vent fluid chemistry in Bahía Concepción geothermal fields: Geothermics, 37, 597-621.
coastal submarine hydrothermal system, Baja California Sur, Mexico: Verma, S.P., Cruz-Huicochea, R., Díaz-González, L., 2013, Univariate data
Journal of Volcanology and Geothermal Research, 137, 311-328. analysis system: deciphering mean composition of island and continental
Reed, M., Spycher, N., 1984, Calculation of pH and mineral equilibria in arc magmas, and influence of the underlying crust: International Geology
hydrothermal waters with application to geothermometry and studies of Review, 55, 1922-1940.
boiling and dilution: Geochimica et Cosmochimica Acta 48, 1479-1492. Vidal-Solano, J., Paz-Moreno, F.A., Iriondo, A., Demant, A., Cochemé, J.-J., 2005,
Reyes, A.G., 1990, Petrology of Philippine geothermal systems and the Middle Miocene peralkaline ignimbrites in the Hermosillo region (Sonora,
application of alteration mineralogy to their assessment: Journal of Mexico): Geodynamic implications: Comptes Rendus Geoscience, 337,
Volcanology and Geothermal Research, 43(1), 279-309. 1421-1430.
Reyes, A.G., Trompetter, W.J., Britten, K., Searle, J., 2002, Mineral deposits in Wilson, F.I., Rocha S.V., 1949, Coal Deposits of the Santa Clara District Near
the Rotokawa geothermal pipelines, New Zealand: Journal of Volcanology Tonichi, Sonora, Mexico: 962A, U.S. Geological Survey Bulletin, 80 pp.
and Geothermal Research, 119, 215-239.
Roldán-Quintana, J., 1994, Geología del Sur de la Sierra Oposura, Moctezuma,
Estado de Sonora, México: Revista Mexicana de Ciencias Geológicas,
11(1), 1-10.
Romo-Jones, J.M., Gutiérrez-Negrin, L.C., Flores-Armenta, M., Del Valle, J.L,
García, A., 2017, 2016 México Country Report: IEA Geothermal, <http://
iea-gia.org/wp-content/uploads/2014/12/IEA-Geothermal-2016-Mexico-
Country-Report.pdf>, 10 pp.
SENER (Secretaría de Energía), 2017, Mapa de Ruta Tecnológica en Geotermia: Manuscript received: july 21, 2016
Reporte SENER, <https://www.gob.mx/cms/uploads/attachment/ Corrected manuscript received: march 10, 2018
file/279714/MRTGEO_SENER_V_20_Oct_Rev_2-OPT.pdf>, 64 pp. Manuscript accepted: march 11, 2018

RMCG | v. 35 | núm. 2 | www.rmcg.unam.mx | DOI: http://dx.doi.org/10.22201/cgeo.20072902e.2018.2.397 141

You might also like