You are on page 1of 15

Journal of Volcanology and Geothermal Research 192 (2010) 1–15

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / j v o l g e o r e s

Fluid geochemistry of hydrothermal systems in the Arica-Parinacota, Tarapacá and


Antofagasta regions (northern Chile)
F. Tassi a,b,⁎, F. Aguilera c,d, T. Darrah e,f, O. Vaselli a,b, B. Capaccioni g, R.J. Poreda e, A. Delgado Huertas h
a
Department of Earth Sciences, University of Florence, Via G. La Pira 4, 50121, Florence, Italy
b
CNR-IGG Institute of Geosciences and Earth Resources, Via G. La Pira 4, 50121, Florence, Italy
c
Departamento de Geologia, Universidad de Atacama, Copayapu 485, Copiapó, Chile
d
Programa de Doctorado en Ciencias mención Geologia, Universidad Católica del Norte, Av. Angamos 0610, Antofagasta, Chile
e
Department of Earth and Environmental Sciences, University of Rochester, 227 Hutchinson Hall, Rochester, NY 14627, USA
f
Environmental Earth and Ocean Sciences, University of Massachusetts Boston, 100 Morrissey Blvd., Boston, MA 02125, USA
g
Department of Earth and Geological-Environmental Sciences, University of Bologna, Piazza Porta San Donato, 1, 40126, Bologna, Italy
h
Estacion Experimental de Zaidin (CSIC), Prof. Albareda 1, 18008, Granada, Spain

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

Article history: We investigate the chemical and isotopic composition of water and gas thermal discharges from six
Received 25 August 2009 hydrothermal systems in the Tarapacà and Antofagasta regions (northern Chile): Surire, Puchuldiza-Tuja,
Accepted 9 February 2010 Pampa Lirima, Pampa Apacheta, El Tatio and Torta de Tocorpuri, to determine the chemical–physical
Available online 13 February 2010
conditions at the fluid source. The chemical facies of the thermal discharges vary from Na+–Cl− (El Tatio,
Puchuldiza-Tuja and part of Surire where SO2− +
4 -rich waters also occur) to Na (Ca
2+
)–Cl−(SO2−
4 ) (Pampa
Keywords: −
hydrothermal system
Lirima), Ca2+–SO2− 4 (HCO3 ) (Torta de Tocorpuri) and Ca
2+
–SO2−
4 (Pampa Apacheta). The gas seeps are
fluid geochemistry characterized by the dominance of CO2, H2S and CH4 with N2/Ar ratios that occasionally are up to 450.
geothermal prospection Significant amounts of SO2 and HCl are present at Pampa Apacheta. Water and gas geothermometry suggests
geothermometry that El Tatio and Puchuldiza-Tuja fluid reservoirs have relatively high reservoir equilibrium temperature (up
northern Chile to 270 °C). Gases from Pampa Apacheta apparently equilibrated at unusually high temperature for
hydrothermal fluids (up to 350 °C), being likely related to an active magmatic system, as testified by the
presence of highly acidic gas compounds such as SO2 and HCl. On the contrary, low equilibrium temperatures
were calculated for the Surire fluids (b 200 °C), suggesting interactions of deep-originated fluids with
meteoric water at shallow depth. Shallow, low temperature aquifers mask any signal of deep fluids in the
Pampa Lirima and Torta de Tocorcupi thermal discharges. Despite the variations in temperatures, He isotopic
compositions and redox conditions of the fluid reservoirs consistently indicate a significant contribution
from magmatic degassing. Although geophysical studies are necessary to better constrain the geothermal
potential of northern Chile, the geothermometric estimations carried out in the present study indicate that El
Tatio, Pampa Apacheta, Surire and Puchuldiza-Tuja can be regarded as the most promising areas and deserve
more detailed geological investigations.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction Guallatiri, Olca, Ollagüe, Putana, Alítar and Lastarria volcanoes


currently exhibit fumarolic degassing and no historical description
The Andean Central Volcanic Zone (ACVZ) in northern Chile of eruptive events is available (Casertano, 1963; deSilva and Francis,
includes 44 active or potentially active volcanic systems (deSilva and 1991; Matthews et al., 1997). The anomalous geothermal gradient
Francis, 1991). This intense volcanism of northern Chile is related to related to the presence of a convergent plate boundary over this large
the subduction process thrusting the oceanic Nazca Plate beneath the region produces diffuse hydrothermal activity not necessarily asso-
South America Plate (deSilva and Francis, 1991; Stern, 2004). Phreatic ciated with the volcanic structures (Lahsen, 1976; Hauser, 1997).
to phreato-magmatic events were recently recorded at Lascar, Preliminary investigations (e.g., Mahon, 1970; Hochstein, 1971;
Guallatiri, Isluga, Irruputuncu and San Pedro volcanoes (BGVN, Cusicanqui, 1979; ELC, 1980; Lahsen, 1988) indicate that areas such
1997; Gardeweg and Medina, 1994; Céspedes et al., 2004). Tacora, as El Tatio, Puchuldiza and Surire, are characterized by the presence of
extended hydrothermal systems, although the local economic condi-
tions have not allowed a systematic geothermal development.
⁎ Corresponding author. Department of Earth Sciences, University of Florence, Via G.
This study presents and discusses the compositional data of fluid
La Pira 4, 50121, Florence, Italy. Tel.: +39 055 2757477; fax: +39 055 284571. discharges collected, from October 2005 through May 2008, from N to S,
E-mail address: franco.tassi@unifi.it (F. Tassi). at Surire, Puchuldiza-Tuja, Pampa Lirima, Pampa Apacheta, El Tatio

0377-0273/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jvolgeores.2010.02.006
2 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

and Torta de Tocorpuri hydrothermal systems (Fig. 1). The aim is to 133 thermal discharges occurred in an area of about 15 km2 (Trujillo,
investigate the chemical–physical conditions acting on the fluid 1972). Nevertheless, the use of water during boron mining activity in
reservoirs to provide useful information for the evaluation of their the northern sector of the local evaporite deposits (Surire Salar)
geothermal potential. provoked a decrease of the water table. Consequently, many thermal
springs dried out (Cusicanqui, 1979). Presently, most of the bubbling
2. Regional setting pools and thermal springs of this system are located along the
southern border of the salar (Fig. 2a).
The main hydrothermal systems of northern Chile are located
within NS-, NW-trending grabens (Lahsen, 1969; Trujillo, 1972; 2.2. Puchuldiza-Tuja
Lahsen, 1976; JICA, 1979; Marinovic and Lahsen, 1984) in the western
side of the Pliocene–Holocene Central Andean Volcanic Zone (CAVZ) The Puchuldiza-Tuja hydrothermal system is located at an altitude
that runs parallel to the Andean sedimentary chain (e.g., Springer and of about 4100–4200 m a.s.l., 27 km SW of the active Isluga volcano
Förster, 1998; Riller et al., 2001; Matteini et al., 2002a,b; Petrinovic et characterized by permanent fumarolic activity (Céspedes et al., 2004).
al., 2006; Acocella et al., 2007; Tibaldi et al., 2008). This area has a Two distinct zones have bubbling and mud pools, thermal springs, and
relatively homogeneous geological setting, consisting of Lower geysers: (1) Puchuldiza that covers an area of approximately 1 km2,
Miocene–Pleistocene ignimbrite deposits and andesitic–rhyolitic and (2) Tuja (0.15 km2), at 6 km NW of Puchuldiza. The fluid dis-
volcanic products overlying Middle Cretaceous–Upper Miocene charges within these areas are likely controlled by the Churicollo,
volcano-sedimentary formations (Francis and Rundle, 1976; Lahsen, Puchuldiza and Tuja faults (Fig. 2b). Several thermal springs with low
1976; JICA, 1979; Marinovic and Lahsen, 1984; Montgomery and gas emission surround the main emission areas (Letelier, 1981).
Rosko, 1996; Polanco and Gardeweg, 2000; Ahumada and Mercado,
2008). The latter likely host the main hydrothermal reservoirs that are 2.3. Pampa Lirima
dominantly consisting of andesitic lava and pyroclastic flows,
conglomerates, breccias, sandstones, siltstones, limestones, marls The Pampa Lirima hydrothermal system occurs at an altitude of
and evaporites (Marinovic and Lahsen, 1984; García et al., 2004). 4000 m a.s.l., 25 km SW of the Sillajguay volcanic chain. Thermal
Evaporitic surficial deposits, locally named “salares” and composed of springs and pools with low-flux gas bubbling exist along the western
borates (mostly ulexite), subordinate sulfates, carbonates and side of the Pampa Lirima graben (Fig. 2c). Sinter deposits mark the
chlorides, are locally present, i.e. in the Surire area (Chong et al., whole emission area.
2000). Hydrothermal activity commonly produces deposits charac-
terized by peculiar mineralogical association of teruggite, nobleite, 2.4. Pampa Apacheta
ulexite, opal-A and traces of illite–smectite (Rodgers et al., 2002).
For each of the studied systems a brief geographical setting and Pampa Apacheta is a flat-floored valley within the 3 × 5 km wide
descriptions of the geothermal manifestations at surface is reported. Pabelloncito graben, located 105 km NE of Calama City and 55 km NW
of the El Tatio hydrothermal system (Fig. 1). A 180 m deep well (PAE-1)
2.1. Surire drilled by the Chilean National Mining Company (CODELCO) in 1998
produced steam measured at 88 °C (Urzúa et al., 2002). Fluid discharges
The Surire hydrothermal system is located 25 km NW of the active emitting superheated steam (up to 118 °C; Urzúa et al., 2002) with high
Isluga volcano at an altitude of 4000–4300 m a.s.l. In 1972, as many as flow rates are along the eastern flank of the 5150 m high Apacheta
volcano (Fig. 2d).

2.5. El Tatio

El Tatio is located 100 km E of the town of Calama (Fig. 1) at an


altitude of 4300 m a.s.l. Dozens of thermal springs, fumaroles, geysers
and boiling and mud pools occur in four main areas (Fig. 2e): (1)
“Central” (2.13 km2 wide) located in the central portion of the El Tatio
graben; (2) “West” (0.19 km2), at the western border of the El Tatio
graben; (3) “Corfo” (0.1 km2), located at the western flank of the El
Tatio volcanoes, 4 km SE of the Central zone; and (4) “Geyser Blanco”
(0.06 km2), located 1.5 km SW of the Central one. Extensive zones
covered by sinter deposits and polychrome algae mark the hydro-
thermal terrains. Hydrogeological models (Cusicanqui et al., 1975;
Giggenbach, 1978) indicate that meteoric waters infiltrate in recharge
areas some 15 km E from the field. The main hydrothermal reservoir is
confined within the permeable Puripicar Formation and the Salado
Member. An important secondary aquifer occurs in the Tucle Dacite
subunit that is capped by the impermeable Tatio Ignimbrite subunit
(Cusicanqui et al., 1975).

2.6. Torta de Tocorpuri

The Torta de Tocorpuri hydrothermal system is located at the


eastern border of the El Tatio volcanic chain, 100 km E of the city of
Calama (Fig. 1) and 20 km NE of the Putana volcano. Thermal
discharges, consisting of small bubbling pools with relatively low
Fig. 1. A schematic map of northern Chile with the location of the Puchulzida-Tuja, Surire, flow occur at an altitude of approximately 5000 m a.s.l. in a restricted
Pampa Lirima, Pampa Apacheta, El Tatio and Torta de Tocorpuri hydrothermal systems. zone N of the Pleistocene–Holocene dome Torta de Tocorpuri (Fig. 2f).
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 3

Fig. 2. a–f. Schematic geologic maps of a) Surire, b) Puchulzida-Tuja, c) Pampa Lirima, d) Pampa Apacheta, e) El Tatio and f) Torta de Tocorpuri areas and location of the sampling sites.
4 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

3. Geothermal exploration in the CAVZ and (2) silicone/tygon tubes connected to a plastic funnel, to sample
gases from fumarolic vents (at the Pampa Apacheta and El Tatio
Initial geothermal exploration in the CAVZ took place in late 1960s hydrothermal systems) and bubbling pools (at the Surire, Puchuldiza-
in response to increasing Chilean energy demands. At El Tatio a pre- Tuja, Pampa Lirima and Torta de Tocorpuri hydrothermal systems),
feasibility investigation, funded in 1967 by the Corporation for the respectively. Gas samples were collected until the head-space pres-
Promotion of Development and the United Nations Development sure equilibrated with the discharging vent. Water vapor and the
Program (CORFO/UNDP), in 1968–1980 was followed by geological main acidic gas species (CO2, SO2 and HCl) dissolved in the alkaline
(Lahsen, 1969; Trujillo, 1971), geophysical (McDonald, 1969, 1974; solution whereas H2S formed insoluble CdS. Low-solubility gas
Hochstein, 1971; ELC, 1980) and geochemical (Mahon, 1974; Lahsen, species (N2, O2, CO, H2, He, Ar, Ne, CH4 and light hydrocarbons)
1976; Cusicanqui, 1978a,b) surveys. Six 600 m deep exploration wells, were concentrated in the sampling flask head-space (Giggenbach and
drilled between 1969 and 1971, encountered temperatures up to Gougel, 1989; Montegrossi et al., 2001; Vaselli et al., 2006). Residual
250 °C. In 1973 and 1974, production wells in the SE sector of El Tatio inorganic gases and CH4 was analyzed by using a Shimadzu 15 A gas
were projected to produce approximately 30 MW (Huttrer, 1996). chromatographic system equipped with a 9 m, 5 A molecular sieve
The full potential exploitation of this geothermal field was estimated column and thermal conductivity detector (TCD). Light hydrocarbons
to be as high as 100–400 MW (Lahsen, 1988; Huttrer, 1996). were determined by a Shimadzu 14 A gas-chromatograph equipped
At Puchuldiza-Tuja, geological (Healy, 1969), geochemical (Mahon, with a 10 m long stainless steel column (ϕ = 2 mm) packed with
1970, 1974) and geophysical (Risk, 1970; Hochstein, 1971) studies Chromosorb PAW 80/100 mesh coated with 23% SP 1700 and a flame
were performed by CORFO/UNDP (from 1968 to 1974) and by Japan ionization detector (FID). The solid precipitate, separated from the
International Cooperation Agency (JICA) (from 1978 to 1980) to alkaline solution by centrifugation at 4000 rpm for 30 min, was oxidized
evaluate geothermal potential. Temperatures up to 166 °C were mea- by H2O2 to determine H2S as SO2− 4 by ion-chromatography (Methrom
sured in exploration wells at the depth of 900 m (JICA, 1979, 1981). 761). The alkaline solution was used for determining the concentrations
In the Surire zone geological and geochemical (Trujillo, 1972; of: (1) CO2, analyzed as CO2− 3 by acidimetric titration with 0.1 N HCl
Cusicanqui, 1979) investigations were carried out by CORFO between and (2) HCl as Cl− by ion-chromatography, SO2 as SO24, after oxidation
1972 and 1979. Reservoir temperatures up to 230 °C were estimated with H2O2, by ion-chromatography. Analytical error is b5% for the main
by geothermometric calculations based on the water chemistry of the gas components and b10% for minor and trace gas compounds.
thermal discharges (Cusicanqui, 1979).
Geothermal exploration in the CAVZ was abandoned in 1982 4.3. Isotopic analysis of He (3He/4He ratio), C (13C/12C ratio of CO2 and
because of both the remote location of the hydrothermal systems CH4) and H (2Η/1Η ratio of CH4)
and economic factors. A Chilean energy shortage which occurred in
the late nineties caused a renewal of interest in alternative energy The 13C/12C ratios of CO2 (expressed as δ13C–CO2 ‰ V-PDB) were
resources. In response, after almost three decades, private and measured by using 2 mL of the soda solution after the addition of
governmental companies (e.g., CORFO, Geotermica del Norte S.A., ∼ 5 mL of anhydrous phosphoric acid for the exsolution of CO2.
Geotermica del Tatio, and Empresa Nacional de Geotermia) have Isotopic equilibration was achieved in a thermal bath at the tem-
planned to conduct a new phase of geothermal exploration in the perature of 25 ± 0.1 °C for at least 8 h. The exsolved CO2 was then
systems investigated in 1969–1982 as well as in other areas of extracted and purified by using liquid N2 and N2-trichloroethylene
northern Chile, i.e. Pampa Lirima, Torta de Tocorpuri and Pampa cryogenic traps. The 13C/12C ratio was analyzed with a Finningan Delta
Apacheta, showing presence of thermal fluid discharges. S mass spectrometer. Internal (Carrara and San Vincenzo marbles)
and international (NBS18 and NBS19) standards were used to esti-
4. Methods mate external precision. The analytical error and the reproducibility
are ±0.05‰ and ±0.1‰, respectively.
4.1. Chemical and isotopic (δ18O and δD) analysis of water samples The 13C/12C and 2H/1H ratios of CH4 (expressed as δ13C–CH4 ‰
V-PDB and δD-CH4 ‰ V-SMOV, respectively) were analyzed by isotope
Temperature, pH and HCO3− concentrations (acidimetric titration ratio mass spectrometry (Varian MAT 250) following the procedure
with 0.01 N HCl) were determined in the field. Water samples were described by Schoell (1980). Analytical precision is ±0.15‰.
analyzed for major cations (Na+, K+, Ca2+, Mg2+ and Li+) and anions The 3He/4He ratios (expressed as R/Ra ratios, where R is the
(Cl−, SO2− −
4 , NO3 , Br

and F−) by atomic absorption spectropho- 3
He/4He measured ratio and Ra is the 3He/4He ratio in the air:
tometry (AAS: Perkin-Elmer AAnalyst 100) and ion-chromatography 1.39 10−6; Mamyrin and Tolstikhin, 1984) were determined using a
(IC: Metrohm 761), respectively. Boron, NH+ 4 and SiO 2 were VG 5400 Rare Gass Mass Spectrometer with a Faraday cup (resolution
determined by molecular spectrophotometry (MS: Hach 2010). The of 200) and a Johnston electron multiplier (resolution of 600) following
analytical error for AAS, IC and MS was ≤5%. the procedure of Poreda and Farley (1992). Assuming that Ne in
The 18O/16O and 2H/1H isotopic ratios (expressed as δ18O and δD ‰ magmatic and crustal gases is negligible, and that nearly all Ne in
V-SMOW, respectively) were determined using a Finningan Delta Plus geothermal gas samples is atmospheric, R was corrected for the pres-
XL mass spectrometer according to standard protocols. Oxygen ence of atmospheric He using the He/Ne ratio of the sample.
isotopes were analyzed by using the CO2–H2O equilibration method Analytical error for R/Ra determination is ≤0.3% using Rochester,
proposed by Epstein and Mayeda (1953). The hydrogen isotopic NY air and a secondary standard of Yellowstone Park gas (MM =
measurements were carried out on H2 obtained after the reaction of 16.5 × air) (Craig et al., 1978).
10 µL of water with metallic zinc at 500 °C (Coleman et al., 1982). The
experimental error was + 0.1‰ and + 1‰ for δ18O and δ2H values, 5. Results
respectively, using EEZ-3 and EEZ-4 as internal standards that were
previously calibrated vs. V-SMOW and SLAP reference standards. 5.1. Chemical and isotopic (δ18O and δD) composition of waters

4.2. Sampling and chemical analysis of gas samples Table 1 reports the temperature, pH, and the chemical composition
of the hot and cold water discharges from Surire (S), Puchuldiza-Tuja
Pre-evacuated 60 mL glass flasks, equipped with a Teflon stopcock, (PT), Pampa Lirima (PL), Pampa Apacheta (PA), El Tatio (ET) and Torta
filled with 20 mL of a 5N NaOH and 0.15 M Cd(OH)2 suspension were de Tocorpuri (TT) hydrothermal systems. Water temperatures range
used on a line with (1) a titanium tube and double-wall glass dewars from 7.4 (TT11) to 87.9 °C (ET2), while pH values range between
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 5

Table 1
Outlet temperatures (in °C), pH, chemical composition and δ18O and δD (‰ V-SMOW) values of the thermal discharges from the Surire (S), Puchuldiza-Tuja (PT), Pampa Lirima (PL),
Pampa Apacheta (PA), El Tatio (ET) and Torta de Tocorpuri (TT) hydrothermal systems, northern Chile. Concentrations of solutes are in mg/L; n.a.: not analyzed.

Sample Cord N Cord E Altitude T pH HCO3− F– Cl− Br− NO3− SO2−


4 Na+ K+ Ca2+ Mg2+ NH+
4 Li+ B SiO2 δD δ18O TSiO2
(°C)

PT01 7854527 504729 4226 87.7 5.61 228 3.5 2700 5.50 0.01 135 1600 197 34 2.5 1.61 13 71 146 −95.18 −10.24 160
PT02 7854532 504695 4228 87.2 6.16 186 3.3 2675 5.60 0.01 140 1575 197 30 1.5 1.48 13 78 140 −96.52 −10.23 157
PT03 7854027 504614 4232 77.6 6.31 198 3.8 2850 6.20 0.40 145 1706 219 19 0.5 1.29 14 85 134 −92.93 −9.00 154
PT04 7853817 504403 4226 51.3 5.99 271 3.3 2710 3.90 1.20 125 1700 214 47 0.5 2.00 13 81 140 −90.01 −9.11 157
PT05 7853875 503619 4174 79.6 6.36 250 3.3 2900 5.30 0.05 130 1812 158 45 0.5 1.22 15 77 136 −95.22 −9.00 155
PT06 7853773 504020 4181 77.3 6.22 253 4.0 2650 5.10 0.30 130 1585 177 40 0.5 1.35 13 76 139 −93.82 −8.72 157
PT07 7853045 504252 4206 86.9 6.06 328 2.7 1684 1.82 1.55 71 1090 102 57 1.2 1.13 9 80 140 −100.60 −11.42 157
PT08 7853045 504252 4230 84.0 7.03 210 5.5 2405 3.78 1.27 103 1694 188 22 0.2 1.48 12 80 150 −96.81 −8.04 161
PT09 7856345 499763 4107 74.9 6.54 84 1.4 2277 2.09 0.24 71 1300 92 70 3 2.69 9 503 140 −98.78 −10.37 157
PT10 7856654 499497 4100 81.8 6.85 167 4.3 3323 4.38 0.44 89 2170 207 58 0.2 2.78 15 1020 138 −93.80 −8.64 156
PT11 7856603 499404 4112 73.5 5.16 22 5.5 2496 2.47 5.60 219 1660 144 69 0.2 6.62 12 572 139 −87.76 −5.36 157
PT12 7856415 499626 4097 87.3 1.76 0 3.2 2444 3.23 1.50 119 1731 147 64 0.4 2.61 12 658 135 −98.56 −7.89 155
S01 7908676 502282 4097 85.0 6.97 277 2.3 2093 2.33 0.47 240 1337 156 100 9.5 0.77 9 303 101 −113.15 −13.22 138
S02 7908676 502282 4283 83.3 6.72 260 2.8 2193 1.95 0.44 245 1519 146 112 12 0.54 9 341 100 −112.51 −13.38 137
S03 7909071 501559 4287 80.8 6.61 278 2.7 1698 1.71 0.95 177 1120 138 82 11 0.48 7 289 95 −116.58 −13.96 134
S04 7909071 501559 4287 78.4 6.50 259 1.9 1604 1.37 0.32 171 1065 127 87 11 0.36 8 288 99 −113.16 −13.81 137
S05 7908818 500096 4291 43.2 6.45 177 2.3 1719 1.89 0.56 887 1180 198 207 45 0.84 6 248 100 −109.53 −10.78 137
S06 7908818 500096 4291 50.3 6.15 156 2.3 1659 1.95 10.10 846 1201 178 213 45 0.90 7 236 102 −111.42 −11.68 138
S07 7908818 500096 4291 49.8 6.29 167 1.8 1661 1.73 1.88 901 1205 183 210 46 0.44 7 235 107 −110.87 −12.48 141
S08 7908818 500096 4291 55.7 6.41 172 2.1 1692 1.51 0.35 880 1219 184 209 44 1.17 7 246 106 −108.34 −12.54 141
S09 7908818 500096 4291 58.6 6.38 168 2.5 1686 1.66 1.15 881 1201 194 233 50 0.99 6 238 102 −106.57 −12.79 138
S10 7908818 500096 4291 42.2 6.41 174 2.5 1713 1.83 0.55 879 1162 182 220 51 1.07 6 219 95 −108.60 −10.53 134
PL01 7804917 510031 4022 42.0 6.57 180 4.0 237 0.42 0.08 259 255 29 46 4.0 0.51 2.6 45 79 −94.81 −12.57 125
PL02 7804883 509909 4016 44.0 6.75 149 4.3 249 0.40 0.10 267 265 30 41 2.2 0.76 2.6 50 72 −98.11 −13.64 120
PL03 7804773 509857 4012 54.3 6.68 123 4.6 261 0.46 0.16 264 268 33 50 3.2 0.61 0.8 55 76 −102.11 −13.15 123
PL04 7804726 509847 4012 46.3 6.83 143 4.3 297 0.49 0.13 279 286 31 37 2.7 0.63 2.8 52 79 −97.84 −13.70 125
PL05 7804942 509780 4016 72.8 7.05 136 4.7 267 0.79 0.28 247 276 32 37 1.5 0.48 2.8 35 79 −95.52 −13.33 125
PL06 7804904 509782 4020 55.7 7.30 126 4.8 274 0.48 0.28 256 284 33 39 1.7 0.56 2.8 52 70 −94.29 −12.27 118
PL07 7804892 509782 4013 56.3 6.74 137 1.4 267 0.29 0.04 274 285 28 38 1.8 0.44 2.8 53 75 −93.02 −12.12 122
PL08 7804837 509804 4011 76.8 7.38 144 4.8 270 0.47 0.75 260 285 33 38 2.0 0.49 2.8 47 80 −102.21 −13.24 125
PL09 7804875 509762 4016 56.2 7.03 162 4.9 273 0.57 0.20 273 286 31 40 2.0 0.55 2.8 53 76 −92.59 −12.79 123
PL10 7804919 509789 4014 66.9 7.16 142 4.3 266 0.48 0.11 262 280 37 41 2.0 0.56 2.9 54 79 n.a. n.a. 125
PL11 7804982 509770 4016 73.7 6.87 134 4.7 261 0.37 1.02 249 270 31 37 2.5 0.59 2.7 51 85 −95.73 −13.39 128
PL12 7804840 509688 4017 51.0 6.70 127 4.4 266 0.49 0.18 266 276 33 38 2.2 0.41 2.9 56 82 n.a. n.a. 127
PA01 7583644 586520 5214 83.3 3.11 0 0.5 1.8 0.01 4.35 265 9.2 12 41 11.5 31 0.1 0.1 56 −42.34 −2.39 108
PA02 7583644 586520 5214 83.2 2.85 0 0.5 1.4 0.01 14.50 205 8.8 6.9 45 6.2 24 0.1 1.7 59 −37.49 −9.84 110
PA03 7583644 586520 5214 84.3 3.57 0 1.3 0.9 0.01 5.20 3751 35 22 995 207 48 0.1 1.2 51 −37.68 −4.99 103
ET01 7530599 602643 4290 87.6 5.35 26 0.8 5100 6.2 3.3 30 2650 395 177 1.0 2.80 30 113 n.a. −98.78 −10.37
ET02 7529326 601945 4290 87.9 6.48 89 1.3 4650 6.0 5.1 40 2575 152 202 4.5 0.39 26 105 n.a. −98.56 −7.89
ET03 7529277 601850 4290 82.8 5.31 32 1.9 6500 14.6 4.6 61 3390 182 267 2.1 1.22 36 139 188 −63.29 −3.12 176
ET04 7526562 604308 4514 84.8 5.84 43 8.5 5400 20.0 0.80 170 3240 335 204 1.5 1.18 37 138 186 −70.73 −5.80 176
ET05 7529876 600439 4240 85.1 5.84 43 8.5 5400 20.0 0.8 170 3240 335 204 1.5 1.18 37 138 180 −70.73 −5.80 173
ET06 7529528 600416 4230 87.8 5.75 64 7.0 5750 5.5 0.5 42.5 3230 336 215 0.7 1.15 39 134 192 −68.19 −5.41 178
TT01 7522497 607487 5000 11.4 7.20 215 0.1 2.0 0.02 0.31 584 28 7.1 172 65 0.61 0.1 0.1 52 n.a. n.a. 104
TT02 7522586 607178 5000 19.6 3.97 0 3.7 1.4 0.01 5.17 690 135 6.2 111 28 1.42 0.1 0.4 51 n.a. n.a. 103
TT03 7522508 509857 5011 27.6 6.25 157 0.2 1.1 0.01 0.76 306 22 4.0 108 27 0.55 0.1 0.3 50 −60.41 −9.36 102
TT04 7522174 607164 4990 13.8 6.47 758 0.3 0.8 0.17 0.19 421 58 7.1 276 43 0.43 0.1 1.3 48 n.a. n.a. 100
TT05 7522174 607164 4990 10.8 6.13 320 0.3 1.8 0.01 2.47 1142 100 9.0 262 109 0.55 0.1 0.5 56 n.a. n.a. 108
TT06 7522174 607164 4990 23.9 6.32 676 0.2 1.9 0.01 0.05 1068 122 2.4 298 127 0.44 0.1 0.4 51 −57.29 −8.85 103
TT07 7522174 607164 4990 21.9 6.44 660 0.2 1.9 0.01 0.07 1336 130 13.0 355 148 0.76 0.1 0.3 48 −58.31 −8.84 100
TT08 7522174 607164 4990 16.8 6.40 561 0.2 1.5 0.01 0.09 792 63 6.0 272 105 0.39 0.1 0.6 42 n.a. n.a. 94
TT09 7522174 607164 4990 16.4 6.51 795 0.5 2.8 0.01 0.11 846 72 8.8 318 112 0.66 0.1 0.3 45 n.a. n.a. 97
TT10 7522174 607164 4990 8.7 6.61 606 0.5 2.4 0.01 0.44 606 55 6.7 249 95 0.51 0.1 0.6 44 n.a. n.a. 96
TT11 7521686 607345 4970 7.4 5.90 28 0.1 0.7 0.01 0.20 92 17 4.8 25 6 0.44 0.1 0.7 39 −58.98 −8.22 91

1.76 (PT12) and 7.30 (PL06). Waters from each hydrothermal system The PA and TT waters show the lowest concentrations of both Li+ and
have a similar composition: S, PT and ET waters are dominated by Na+ SiO2, which vary from 0.1 to 1,7 and from 39 to 89 mg/L, respectively.
and Cl−, whereas those of PL have relatively high (up to 274 mg/L) SO2−4 Conversely, SiO2 concentrations of the S, PT and ET waters are
concentrations. In comparison, PA and TT waters have Ca2+–SO2− 4 and relatively high, ranging between 230 and 315 mg/L, whereas those

Ca2+–SO2−4 (HCO3 ) compositions, respectively. measured at PL are b136 mg/L.
Value of the Total Dissolved Solids (TDS) are relatively high for the Analytical data for δ18O and δD vary between −13.96 and −2.39
Na+–Cl− waters (ranging from 4003 to 10,886 mg/L), whereas those and −116.58 and −37.49‰ V-SMOV, respectively. Each hydrother-
from PL, PA and TT are b2700 mg/L, with the only exception of the mal system shows a well-defined isotopic signature with δ18O and δD
PA03 sample (TDS = 5128 mg/L). Boron concentrations have a wide values varying in relatively narrow ranges (Table 2).
range, from extremely low at the PA and TT systems (between 0.1 and
1.7 mg/L), to between 35 and 56 mg/L at PL, and up to 1020, 341 and 5.2. Chemical composition of gases
138 mg/L at PT, S and ET, respectively. The ET waters have the highest
Li+ concentrations (from 26 to 39 mg/L), followed by PT (from 9 to The chemical composition of the dry gas fraction and gas/H2Ovap
15 mg/L), S (from 6 to 9 mg/L) and PL (from 0.8 to 2.9 mg/L) waters. ratios of the thermal discharges from the six investigated
6 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

Table 2
Chemical composition (dry gas fraction) and gas/H2Ovap of gases from the thermal discharges of the Surire (S), Puchuldiza-Tuja (PT), Pampa Lirima (PL), Pampa Apacheta (PA), El
Tatio (ET) and Torta de Tocorpuri (TT) hydrothermal systems, northern Chile. Concentrations of gas compounds are in µmol/mol; n.d.: not detected.

Sample CO2 HCl SO2 H2S N2 CH4 Ar O2 Ne H2 He CO C2H6 C3H8 C3H6 i- n- i- C6H6 C7H8 etil- m-p- Gas/
C4H10 C4H10 C4H8 C6H6 C8H10 H2Ovap

PT01 971034 n.d. n.d. 80 25264 1931 473 56 0.25 1153 4.3 n.d. 2.164 0.101 0.00009 0.651 0.045 0.001 0.027 0.001 0.0002 0.0002 4.0
PT02 968433 n.d. n.d. 61 27853 1801 573 34 0.34 1237 3.9 n.d. 2.195 0.118 0.00016 0.799 0.045 0.002 0.026 0.001 0.0002 0.0002 3.5
PT03 974587 n.d. n.d. 218 22044 1459 515 244 0.29 925 4.0 n.d. 1.468 0.114 0.00012 1.043 0.387 0.004 0.250 0.006 0.0011 0.0017 5.8
PT04 957857 n.d. n.d. 117 35907 3970 820 191 0.45 1132 1.9 n.d. 2.113 0.175 0.00019 1.146 0.717 0.006 0.388 0.009 0.0027 0.0029 6.2
PT05 964014 n.d. n.d. 152 31227 2531 772 76 0.45 1215 8.7 n.d. 1.598 0.139 0.00014 0.259 0.402 0.002 0.294 0.006 0.0019 0.0016 13.1
PT06 959913 n.d. n.d. 131 35154 2493 685 163 0.39 1440 14 n.d. 2.759 0.193 0.00020 0.901 0.678 0.007 0.434 0.012 0.0029 0.0037 15.7
PT07 966658 n.d. n.d. 105 29407 2096 588 101 0.30 1027 12 n.d. 2.097 0.104 0.00017 1.011 0.755 0.004 0.515 0.013 0.0016 0.0031 17.8
PT08 968715 n.d. n.d. 50 27868 1690 530 96 0.28 1036 11 n.d. 1.685 0.121 0.00023 0.082 0.845 0.001 0.565 0.020 0.0012 0.0019 11.8
PT09 980174 n.d. n.d. 532 16416 1034 54 27 0.03 1761 0.58 n.d. 0.931 0.010 0.00005 0.464 0.013 0.002 0.162 0.013 0.0003 0.0006 20.4
PT10 975403 n.d. n.d. 547 20984 1174 70 119 0.05 1701 0.65 n.d. 1.043 0.006 0.00003 0.681 0.007 0.002 0.104 0.005 0.0003 0.0001 17.9
PT11 973066 n.d. n.d. 426 23564 914 85 31 0.06 1910 0.75 n.d. 1.176 0.009 0.00004 0.788 0.005 0.002 0.131 0.012 0.0003 0.0006 22.3
PT12 977239 n.d. n.d. 470 17865 2335 103 88 0.06 1898 0.46 n.d. 0.497 0.008 0.00003 0.643 0.002 0.002 0.235 0.018 0.0002 0.0010 11.9
S01 984205 n.d. n.d. 34 14256 1039 237 66 0.13 160 0.65 n.d. 0.703 0.027 0.00002 0.994 0.002 0.003 0.197 0.014 0.0001 0.0007 19.6
S02 974813 n.d. n.d. 13 22773 1389 569 237 0.34 204 0.75 n.d. 0.733 0.027 0.00004 0.752 0.001 0.002 0.165 0.012 0.0002 0.0004 19.4
S03 977910 n.d. n.d. 28 19738 1535 467 167 0.24 154 0.97 n.d. 0.349 0.016 0.00003 0.539 0.001 0.001 0.240 0.028 0.0004 0.0008 15.6
S04 976885 n.d. n.d. 118 20915 1082 479 347 0.25 172 0.65 n.d. 0.934 0.035 0.00002 0.332 0.003 0.001 0.104 0.007 0.0001 0.0001 19.1
S05 975607 n.d. n.d. 15 21741 1784 509 168 0.26 174 0.42 n.d. 1.200 0.094 0.00002 0.335 0.007 0.001 0.178 0.010 0.0001 0.0001 12.7
S06 984442 n.d. n.d. 66 12813 2198 301 58 0.15 119 0.51 n.d. 1.439 0.132 0.00001 0.637 0.013 0.002 0.142 0.012 0.0001 0.0007 18.5
S07 983948 n.d. n.d. 63 14695 945 200 32 0.11 114 0.60 n.d. 0.803 0.071 0.00003 0.760 0.006 0.002 0.147 0.011 0.0001 0.0001 12.9
S08 985010 n.d. n.d. 133 12564 1981 166 27 0.10 118 0.65 n.d. 0.706 0.044 0.00004 0.565 0.004 0.002 0.163 0.014 0.0002 0.0007 17.6
S09 977326 n.d. n.d. 35 19724 2222 433 63 0.23 195 0.76 n.d. 0.931 0.066 0.00002 0.502 0.006 0.001 0.118 0.010 0.0002 0.0001 19.1
S10 981641 n.d. n.d. 39 16206 1530 342 118 0.19 121 0.55 n.d. 1.183 0.104 0.00010 0.860 0.009 0.002 0.157 0.013 0.0001 0.0008 16.2
PL05 979248 n.d. n.d. n.d. 18055 46 412 2193 0.23 0.31 45 n.d. 0.300 0.032 n.d. n.d. 0.001 n.d. 0.003 n.d. n.d. n.d. 15.8
PL06 976109 n.d. n.d. n.d. 20116 75 452 3184 0.25 0.52 62 n.d. 0.291 0.023 n.d. n.d. 0.001 n.d. 0.008 n.d. n.d. n.d. 41.5
PL07 977958 n.d. n.d. n.d. 18671 63 437 2820 0.25 0.77 49 n.d. 0.554 0.028 n.d. n.d. 0.001 n.d. 0.003 n.d. n.d. n.d. 22.2
PL08 975826 n.d. n.d. n.d. 20059 48 486 3533 0.27 1.00 46 n.d. 0.498 0.033 n.d. n.d. 0.001 n.d. 0.004 n.d. n.d. n.d. 41.5
PL09 977569 n.d. n.d. n.d. 18890 40 451 3000 0.26 1.23 48 n.d. 0.368 0.032 n.d. n.d. 0.001 n.d. 0.003 n.d. n.d. n.d. 22.2
PL10 977366 n.d. n.d. n.d. 18592 38 465 3484 0.26 1.02 54 n.d. 0.485 0.037 n.d. n.d. 0.001 n.d. 0.003 n.d. n.d. n.d. 22.3
PL11 975753 n.d. n.d. n.d. 20901 62 503 2731 0.30 0.78 48 n.d. 0.404 0.043 n.d. n.d. 0.002 n.d. 0.003 n.d. n.d. n.d. 11.4
PA01 979399 563 126 7529 10103 4.5 26 77 0.01 2171 6.2 n.d. 0.792 0.020 0.00526 0.010 0.013 0.210 0.042 0.023 n.r. n.r. 0.01
PA02 980795 568 110 5996 10168 4.0 23 119 0.01 2214 5.8 n.d. 0.838 0.021 0.00536 0.011 0.011 0.159 0.032 0.014 n.r. n.r. 0.01
PA03 960914 607 146 7987 25926 9.6 66 244 0.04 4097 8.4 n.d. 1.863 0.045 0.01245 0.026 0.024 0.411 0.088 0.037 n.r. n.r. 0.01
ET01 989862 n.d. n.d. 3285 6148 116 42 32 0.02 514 0.71 n.d. 1.863 0.154 0.0025 0.019 0.047 0.013 0.673 0.036 0.0016 0.0191 0.11
ET02 989703 n.d. n.d. 3669 5878 77 47 25 0.03 596 1.9 n.d. 1.413 0.778 0.0160 0.141 0.285 0.086 1.009 0.131 0.0108 0.0203 0.12
ET03 993054 n.d. n.d. 2020 4083 197 41 59 0.02 544 1.6 n.d. 0.990 0.139 0.0013 0.014 0.010 0.015 0.425 0.017 0.0011 0.0141 0.14
ET04 992534 n.d. n.d. 1181 5397 219 55 57 0.03 555 1.5 n.d. 1.051 0.139 0.0021 0.016 0.012 0.016 0.520 0.036 0.0015 0.0154 0.11
ET05 993202 n.d. n.d. 725 5072 416 50 46 0.03 487 1.8 n.d. 0.926 0.105 0.0018 0.011 0.012 0.013 0.494 0.019 0.0012 0.0072 0.16
TT06 978114 n.d. n.d. n.d. 19312 80 466 1878 0.25 8.02 69 n.d. 0.340 0.024 n.d. n.d. 0.001 n.d. 0.002 n.d. n.d. n.d. 22.2
TT07 979689 n.d. n.d. n.d. 17913 84 434 1714 0.24 8.36 84 n.d. 0.517 0.020 n.d. n.d. 0.001 n.d. 0.002 n.d. n.d. n.d. 41.3
TT08 977298 n.d. n.d. n.d. 19880 65 519 2083 0.28 6.49 89 n.d. 0.705 0.015 n.d. n.d. 0.001 n.d. 0.003 n.d. n.d. n.d. 22.3
TT09 981019 n.d. n.d. n.d. 16997 59 436 1346 0.24 5.93 83 n.d. 0.415 0.021 n.d. n.d. 0.001 n.d. 0.002 n.d. n.d. n.d. 23.6

hydrothermal systems are reported in Table 2. The gas/H2Ovap ratios lack of CO in the head-space of the sampling flasks (Giggenbach and
of the steam released from the S, PT, PL and TT bubbling pools, Matsuo, 1991; Arnósson et al., 2006), since chemical analyses were
strongly depending on water condensation at the surface, are carried out about 1 month after the gas collection. Argon concentra-
relatively high (from 3.5 to 41.5). Conversely, the fumarolic tions are low in the PA and ET gases (b66 µmol/mol) and vary
discharges at PA and ET, although the outlet temperatures are between 54 and 820 µmol/mol in the S, PT, PL and TT gases. Oxygen
b88 °C (Table 1), have very low gas/H2Ovap ratios (b0.16). The dry concentrations vary over a wide range, high in the PL and TT gases
gas fraction of all the gas samples is largely dominated by CO2 (from 1346 to 3533 µmol/mol) and significantly lower (b244 µmol/
(ranging from 957,857 to 993,202 µmol/mol) followed by N2 (up to mol) in the other hydrothermal systems. Helium is up to 89 µmol/mol.
35,888 µmol/mol). Methane concentrations are relatively high in the Several C2–C8 alkanes, alkenes and aromatics hydrocarbons were
PT and S gases (up to 3970 and 2222 µmol/mol, respectively), and detected, although they constitute a small fraction of the total gas in
do not exceed 416, 84, 75 and 9.6 µmol/mol at the ET, TT, PL and PA each sample (b3 µmol/mol).
hydrothermal systems, respectively. PT, ET, and PA gases have high H2
concentrations (up to 4097 µmol/mol) when compared with those of
the S (b204 µmol/mol), TT (b8.4 µmol/mol) and PL (b1.23 µmol/mol) 5.3. Isotopic composition of gases
gases. Water-soluble acidic gas species, such as HCl and SO2, are
practically absent (b0.01 µmol/mol) in all the gas samples, with the The values of R/Ra, δ13C–CO2, δ13C–CH4 and δD–CH4 measured in
only exception of the PA gases (up to 607 and 146 µmol/mol, selected gas samples are listed in Table 3. The R/Ra values of the S, PT
respectively). Hydrogen sulphide is present at relatively high and ET gases are in a narrow range (from 2.67 to 3.07), whereas those
concentrations in the PA and ET gases (up to 7987 and 3669 µmol/ of the PL, PA and TT gases are significantly lower (between 1.09 and
mol, respectively), whereas at the PT and S systems it is b547 µmol/ 1.31). The PT, S, PA and ET gases are characterized by similar δ13C–CO2
mol. The PL and TT gases are H2S-free. Carbon monoxide is below the (from −8.95 to −3.64‰ V-PDB), δ13C–CH4 (from −33.12 to −22.54‰
detection limit (b1 µmol/mol) possibly because this gas readily V-PDB) and δD–CH4 (from −111.53 to −91.17‰ V-SMOW) values.
dissolves in shallow aquifers forming HCOOH (Shock, 1993). Conversely, the PL and TT hydrothermal systems have a carbon
However, time-dependent reactions in NaOH could also explain the isotopic signature of both CO2 and CH4 significantly more negative
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 7

Table 3 occur at the margin of the ET hydrothermal system, an area with


δ13C in CO2 and CH4 (‰ V-PDB), δD in CH4 (‰ V-SMOW), R/Ra values and CO2/3He and relatively low input of geothermal fluids. The PA water pools are
CH4/3He ratios of gases from the thermal discharges of the Surire (S), Puchuldiza-Tuja
(PT), Pampa Lirima (PL), Pampa Apacheta (PA), El Tatio (ET) and Torta de Tocorpuri
characterized by a Ca2+–SO2− 4 composition and show very low pH
(TT) hydrothermal systems, northern Chile; n.a.: not analyzed. %L, %S and %M are the (b3.57). Steam-heated acidic-sulfate waters are produced by shallow
limestone, organic sediments and mantle end-members, as defined by Sano and Marty aquifers heated by inputs of deep vapor phases rich in S-bearing

(1995). compounds. The TT fluid discharges show a Ca2+–SO2− 4 (HCO3 )

Sample R/ δ13C– δ13C– δD– CO2/3He CH4/3He % L %S %M composition, low outlet temperatures (b28 °C), and relatively high
Ra CΟ2 CΗ4 CΗ4 (109) (106) Mg2+ concentrations (Fig. 3b). The chemical features are typical of
PT01 n.a. −5.84 n.a. n.a. n.a. a shallow aquifer that has absorbed relatively low amounts of CO2–
PT02 2.67 −6.42 −29.51 −111.1 66 124 76.8 20.9 2.3 H2S-rich fluids without significant heating. The PL waters, charac-
PT03 n.a. −5.10 n.a. n.a. n.a. terized by a Na+–Cl−(SO2− 4 ) composition and relatively low TDS
PT04 2.77 −7.88 −30.18 −93.2 130 55 72.9 25.9 1.2 values (b1250 mg/L), plot at intermediate position between the
PT05 n.a. n.a. n.a. n.a. n.a.
mature and immature water fields (Fig. 3a and b), indicating a
PT06 n.a. n.a. n.a. n.a. n.a.
PT07 n.a. −7.64 n.a. n.a. n.a. significant hydrothermal fluid contribution partially mitigated by
PT08 n.a. −3.95 n.a. n.a. n.a. presence of solute from secondary gas–water–rock interactions.
PT09 n.a. −3.98 n.a. n.a. n.a. As commonly observed in hydrothermal systems and active vol-
PT10 n.a. n.a. n.a. n.a. n.a.
canoes (e.g., Arehart et al., 2003), mature (or partially mature) waters,
PT11 3.07 −3.64 −24.74 −100.5 304 285 87.5 12.0 0.5
PT12 n.a. n.a. n.a. n.a. n.a. i.e. those from the ET, PT and S hydrothermal systems, are enriched in
S01 2.73 −4.72 −33.12 −102.5 399 421 84.0 15.6 0.4 Li+ (Fig. 4), an element released during water–rock interactions and
S02 n.a. n.a. n.a. n.a. n.a. intimately dependent on the temperature (e.g. Fouillac and Michard,
S03 n.a. n.a. n.a. n.a. n.a. 1981). Boron and Cl− do not readily enter secondary mineral struc-
S04 n.a. n.a. n.a. n.a. n.a.
tures during alteration of igneous rocks and sediments at hydrother-
S05 2.79 −4.53 −31.44 −98.9 605 1107 84.7 15.0 0.2
S06 n.a. n.a. n.a. n.a. n.a. mal conditions, and thus they can be considered as conservative
S07 n.a. n.a. n.a. n.a. n.a. chemical species (Seyfried et al., 1984; Palmer et al., 1987; Spivack
S08 n.a. n.a. n.a. n.a. n.a. and Edmond, 1987).
S09 n.a. n.a. n.a. n.a. n.a.
Volatile-bearing natural waters are characterized by B/Cl− ratios
S10 n.a. n.a. n.a. n.a. n.a.
PL05 n.a. n.a. n.a. n.a. n.a.
varying from that of seawater (b0.001) up to 1, the latter being related
PL06 1.09 −14.98 −66.54 −288.1 10 0.8 38.3 46.6 15.1 to geothermal waters affected by magmatic degassing (e.g., Arnósson
PL07 n.a. n.a. n.a. n.a. n.a. and Andresdottir, 1995; Aggarwal et al., 2000). As shown in Fig. 5, the
PL08 n.a. n.a. n.a. n.a. n.a. mature waters of the Chilean geothermal systems fall along two
PL09 n.a. n.a. n.a. n.a. n.a.
different trends when plotting B vs. Cl−: (1) trend A (ET and PT01–
PL10 n.a. n.a. n.a. n.a. n.a.
PL11 n.a. n.a. n.a. n.a. n.a. PT08 waters), showing B/Cl− ratios b0.03; and (2) trend B (S and
PA01 1.81 −4.76 n.a. n.a. 63 0.3 82.3 15.3 2.4 PT09–PT12), characterized by higher B/Cl− ratios (between 0.13 and
PA02 n.a. n.a. n.a. n.a. n.a. 0.30). This could imply that the B-enriched S and PT09–PT12
PA03 n.a. n.a. n.a. n.a. n.a.
discharges, the latter being clustered in the Tuja area (Fig. 2a), are
ET01 2.94 −7.87 −28.06 −98.3 343 40 73.4 26.2 0.4
ET02 n.a. n.a. −24.73 −97.4 n.a.
characterized by a higher fraction of hydrothermal fluids than that of
ET03 2.79 −8.95 −28.63 −91.2 164 32 69.4 29.6 0.9 the ET and PT01–PT08 fluid discharges. Alternatively, relatively high
ET04 n.a. n.a. −26.28 −95.1 n.a. B/Cl− ratios may be ascribed to contribution of waters circulating
ET05 n.a. n.a. −22.54 −99.9 n.a. within salar deposits, like those present in correspondence of the S
TT06 1.31 −11.19 −63.11 −297.4 8 0.6 47.9 33.2 18.9
discharges (Fig. 2a).
TT07 n.a. n.a. n.a. n.a. n.a.
TT08 n.a. −12.06 n.a. n.a. n.a. In geothermal prospecting, NH+ 4 is a useful geochemical indicator
TT09 n.a. n.a. n.a. n.a. n.a. because it can be transported in the gaseous phase (e.g., Tonani, 1970;
Martini et al., 1984). Accordingly, the high NH+ 4 concentrations in the
PA waters may relate to the high gas flux of these emissions (Aguilera,
(b−11.19 and b−63.11‰ V-PDB, respectively), as well as that of H2 of 2008) that prevented NH+ 4 oxidation during the fluid ascent.
CH4 (δD–CH4 b−288.11‰ V-SMOW). In the δD vs. δ18O diagram (Fig. 6), the Na+–Cl− waters plot off of
the Local Meteoric Water Line (LMWL) proposed by Aravena (1995),
6. Discussion Chaffaut et al. (1998), Aravena et al. (1999) and Gonfiantini et al.
(2001) for the same area. The isotopic signature of these waters
6.1. Water geochemistry may be produced by a combination of different processes including:
(1) evaporation, (2) fluid–rock interaction, and (3) mixing between
The Cl−–HCO3−–SO2− 4 and Na+–SO2−4 –Mg
2+
triangular diagrams meteoric and “andesitic” water, the latter defined by Taran et al.
(both in meq/L) (Fig. 3a and b) indicate that the ET, PT and some of (1989) and Giggenbach (1992). To quantify the effects of evaporation
the S (S01–S04) waters have a Na+–Cl− composition and relatively at shallow depth on δD and δ18O values of steaming pools at ET,
high TDS (Total Dissolved Solids) values (up to 10,829 mg/L). These Cortecci et al. (2005) calculated the slope (S) of the straight line
characteristics are a common compositional feature of thermal springs depicted on a δD vs. δ18O diagram by local meteoric water heated by
sourced from geothermal reservoirs as Cl− is primarily derived from steam released from a hydrothermal reservoir at 270 °C, using the
magmatic HCl degassing, and secondarily can be leached from rock equation proposed by Giggenbach and Stewart (1982), as follows:
(e.g., Ellis and Mahon, 1977), Na+ is released by intense leaching
of host rocks, and Mg2+ is depleted by incorporation into clay min- 18 18
S ¼ ðδDs −δDg þ ε2H Þ=ðδ Os −δ Og þ ε18O Þ ð1Þ
erals (e.g., Bischoff and Seyfried, 1978; Thornton and Seyfried, 1987;
Giggenbach, 1988, 1991). The S05–S10 samples (Table 1) show a In Eq. (1) the subscripts s and g refer to the steam and ground-
significant SO2− 4 -enrichment (Fig. 3a) that may result from either H2S water, respectively, while ε is the kinetic isotope factor, which is close
dissolution or leaching of evaporitic rocks, the latter being exten- to 50‰ for hydrogen and 16‰ for oxygen. As shown in Fig. 6, the ET
sively present in the area (Salas, 1975). Previous studies (Giggen- samples do not fit within the steam-heated water line. In agreement
bach, 1978; Cortecci et al., 2005) reported that SO2−
4 -rich waters also with previous authors (Giggenbach, 1978; Cortecci et al., 2005), the
8 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

Fig. 3. Cl−–HCO3−–SO2− + 2−
4 (a) and Na –SO4 –Mg
2+
(b) ternary diagrams (in meq/L) for the Surire (open down triangle), Puchulzida-Tuja (open up triangle), Pampa Lirima (open
diamond), Pampa Apacheta (closed circle), El Tatio (closed square) and Torta de Tocorpuri (open circle) waters.

ET waters may derive from mixing between andesitic water and 6.2. Gas geochemistry
meteoric precipitation, the latter having δD and δ18O values of −107
and −14.6‰ V-SMOV, respectively. Mahon and Cusicanqui (1980) Hydrothermal fluids in convergent plate boundaries are common-
suggested a similar explanation for the PT waters. Waters from the S ly related to three possible source components, including: (1) a
system may have suffered the same process, although they seem to be meteoric component, represented by atmospheric gases released
related to a different meteoric end-member (Fig. 6), with δD and δ18O from air saturated groundwater (ASW); (2) a deep component from
values of −137 and −18.8‰ V-SMOV, respectively, and corre- magma degassing, rich in water vapor, CO2, sulphur gas compounds
sponding to that of hydrothermal systems in Peru (Scandiffio et al., and N2; and (3) a crustal component, showing relatively high
1992; Barragan et al., 1999; Steinmüller, 2001). The TT and PL water concentrations of Ar and He, from radiogenic decay, and hydrocarbons
samples are shifted along LMWL likely by evaporation at relatively originating from thermal breakdown of organic matter buried in
low temperature, indicating that these waters have a meteoric origin. sedimentary formations. In hydrothermal reservoir H2 and CO are
The PA thermal springs appear to have experienced strong evapora- produced under reducing conditions by chemical reaction between
tion. The isotopic compositions of steam condensates from the same circulating fluids and embedding rocks (e.g., Giggenbach, 1980, 1991;
fluid discharges reported by Urzúa et al. (2002) are significantly more Chiodini and Marini, 1998).
negative than those analyzed in the present study, supporting this Contributions to the Chilean gases from the fluid source regions
hypothesis. The negative δ18O-shift shown by PA02 sample (Fig. 6) described above can be examined in the ternary diagrams of Fig. 7 a–c.
was likely caused by isotopic exchange between meteoric water and ET, PA and part of the PT samples, i.e. those of the Tuja area (Fig. 2a),
CO2 (Negrel et al., 1999). This process has also likely affected the PA01 display a significant N2-excess (Fig. 7a) indicating relevant gas
and PA03 waters, masking potential contributions from an andesitic contribution originating from sediments overlying the subducted
fluid source. slab. This process can produce N2/Ar ratios up to 2000 (Matsuo et al.,
1978; Giggenbach, 1997; Snyder et al., 2003). These evidences,
coupled with the relatively high concentrations of H2S (Fig. 7b) and H2

Fig. 4. Li+ vs. Cl− binary diagram (in mg/L). Symbols as in Fig. 3. Fig. 5. B vs. Cl− binary diagram (in mg/L). Symbols as in Fig. 3.
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 9

et al., 1994, 2001) within the liquid-dominated environment over-


lying the degassing magma may explain the lack of strong acidic gas
compounds at the surface. In comparison, the strong flux character-
izing the ascending fluids feeding the PA fumaroles did not allow a
complete dissolution of the highly soluble magmatic gases (Table 2). A
similar mechanism was proposed for explaining the presence of these
compounds in the low temperature gas discharges collected in 2002
from a submarine fumarolic field 3 km offshore of Panarea Island,
Italy (Capaccioni et al., 2007). The PAE-1 exploration well drilled by
CODELCO in the valley 4.5 km E of the PA fumaroles, discharging
sulphur-depleted gas species (Urzúa et al., 2002), is located at the
periphery of this system. The PL and TT samples, and at a minor extent
the S and PT ones, show significant enrichments in He and Ar (Fig. 7a,c).
This evidence, coupled with the relatively high O2 concentrations
(Table 2), indicates a contribution from both radioactive isotope decay
in crustal sediments and ASW (Air Saturated Water: Giggenbach, 1993).
It is worth noting that the CH4 concentrations of the S and the PT
discharges of the Puchuldiza area are up to two orders of magnitude
higher than in the other gases (Fig. 7b; Table 2). Such high CH4
abundances could be ascribed to bacteria-driven maturation of organics
Fig. 6. δ2H vs. δ18O binary diagram. The “andesitic water” field (Taran et al., 1989; transported by crustal fluids. Nevertheless, other genetic processes
Giggenbach, 1992) and the Local Meteoric Water Line (LMWL: δ2H = 8.15δ18O + 15.3; should be taken into consideration. At hydrothermal conditions thermal
Chaffaut et al. (1998)) are also shown. Symbols as in Fig. 3. decomposition processes are indeed able to produce thermogenic CH4
(e.g., Schoell, 1980, 1988; Whiticar, 1999). Moreover, a laboratory
(Fig. 7c), likely imply the occurrence of significant contribution from a experiment has shown that CH4 formation can proceed through
deep magmatic-related source for these gas discharges, although the abiogenic processes, including (1) reduction of graphite (Holloway,
typical magmatic gases, i.e. SO2 and HCl, were not detected at the 1984), (2) thermal decomposition of siderite (McCollom, 2003) and (3)
surface (Table 2). Occurrence of “scrubbing” processes (Symonds reduction of gaseous or dissolved carbon oxides (Berndt et al., 1996;

Fig. 7. Ar–N2–He (a), CH4–Ar–H2S (b), He–CO2–H2 and (c) ternary diagrams. ASW: Air Saturated Water (Giggenbach, 1993). Symbols as in Fig. 3.
10 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

Horita and Berndt, 1999; Foustoukos and Seyfried, 2004). Although the Poreda, 1993; Patino et al., 2000; Snyder et al., 2001). The mixing
natural occurrence of pure inorganic CH4 production is still a matter of model of Sano and Marty (1995) was used to calculate the relative
debate (e.g., Sugisaki and Mimura, 1994; Kenney, 1995; Abrajano et al., proportions of carbon sources, i.e. MORB (M), organic sediment (S)
1988; Szatmari, 1989; Giggenbach, 1997; Sherwood-Lollar et al., 2002; and limestone (L) including slab components, contributing to CO2 in a
Taran et al., 2002; Fiebig et al., 2004, 2007), by coupling the carbon and subduction zone, according to the following equations:
hydrogen isotopes a microbial, thermogenic and abiogenic origin of
geothermal CH4 can be discriminated (e.g., Welhan and Craig, 1979; MþSþL¼1 ð2Þ
Schoell, 1980, 1988; Whiticar, 1999; Bréas et al., 2001). The relative
abundance of the C1–C4 alkanes can provide further indication about the 13 13 13 13
ðδ C−CO2 Þobs ¼ ðδ C−CO2 ÞMORB M þ ðδ C−CO2 ÞLim L þ ðδ C−CO2 ÞSed S
CH4 genesis (Oremland et al., 1987; Whiticar and Suess, 1990; Kiyosu
ð3Þ
et al., 1992; Darling, 1998). Gases from medium-to-high temperature
hydrothermal reservoir are characterized by CH4/(C2H6 + C3H8) and
(C2H6 + C3H8)/(isoC4H10 + normalC4H10) ratios ranging between 1 and 12 3 12 3 12 3 12 3
1=ð C= HeÞobs ¼ M=ð C= HeÞMORB þ L=ð C= HeÞLim þ S=ð C= HeÞSed
100. As shown in Fig. 8a and b and in Table 2, CH4 isotope and light ð4Þ
alkane compositions of the PT, S and ET gases are compatible with a
prevalent thermogenic source, i.e. genetic processes occurring within where subscripts obs, MORB, Lim and Sed refer to the sample, MORB,
the hydrothermal reservoirs. Minor contribution of abiogenic methane, limestone including slab carbonate and sediment, respectively.
likely related to the Sabatier process involving CO2 and H2 at high Calculations were performed adopting the following values: (δ13C–
temperature and pressure, cannot be excluded. Conversely, the origin of CO2)MORB = −6.5‰, (δ13C–CO2)Lim = 0‰, (δ13C–CO2)Sed = −30‰,
CH4 in PL and TT gases seems to be related to bacterial activity in fresh (12C/3He)MORB = 1.5·109, (12C/3He)Lim = 1·1013, and (12C/3He)Sed =
water (Fig. 8a) and it is independent of the thermogenic processes likely 1·1013.
producing the C2–C3 alkanes (Fig. 8b). The results (Table 3) indicate that L–S–M proportions for the PT, S,
The isotopic signatures of CO2 and He provide further evidence of PA and ET gases are similar and dominated by an L component, whose
mantle gas contribution to the PT, S and ET hydrothermal systems. The percentage (up to 87.5%) is higher than the average of arc volcanoes
δ13C–CO2 values (from −8.95 to −3.64‰ V-PDB) indeed overlap the (L = 74.7%) (Sano and Marty, 1995) and very low (b2.4%) carbon
range typical of mantle CO2 (Rollinson, 1993; Hoefs, 1997; Ohmoto contribution from MORB. In comparison, the PL and TT gases are fed
and Goldhaber, 1997) and the R/Ra values (from 2.67 to 3.07) are by a relatively high fraction of MORB carbon (up to 18.9%). These
consistent with those commonly characterizing gases from arc results conflict with the chemical and isotopic features of the gas and
volcanoes, which typically vary in a range (between 5 and 8) water phases of these fluid discharges, which indicate a shallow fluid
depending on the degree of crustal contamination affecting the source. One possible explanation is that the relatively low CO2/3He
mantle source (Welhan et al., 1988; Poreda and Craig, 1989; Hilton et ratios of these gases (8 − 10 × 109; Table 3), which strongly affect the
al., 1993; Giggenbach and Poreda, 1993; Hulston and Lupton, 1996). calculation of the L–S–M proportions, are produced by CO2 dissolution
Similar considerations can also be applied for CO2 and He of the PA into shallow thick aquifers, a process that has no significant influence
gases, although the relatively low R/Ra value (Table 2) suggests a on He.
higher crustal contribution. As already evidenced by water and gas The CH4/3He ratio, a further diagnostic parameter to investigate
chemistry, the PL and TT systems are completely distinct with respect the origin of gases discharged in hydrothermal and volcanic
to the rest of the studied areas, being characterized by the highest environments (e.g., Poreda et al., 1988; Giggenbach et al., 1993;
fraction of crustal He and presence of 13C-depleted CO2 of likely Giggenbach, 1995), is between 32 × 106 and 110 × 107 in the PT, S and
organic origin. ET gas discharges. These values are intermediate between those
The CO2/3He ratios of volcanic gases from arc setting are up to two measured in sediment-free mid-ocean ridge environment (between
orders of magnitude higher than those of MORB and OIB gases (Des 1 × 105 and 1 × 106; Snyder et al., 2003) and those of thermogenic
Marais and Moore, 1984; Marty and Jambon, 1987) due to the gases (up to 1×1012; Poreda et al., 1988), supporting the hypothesis of
addition of CO2 from carbonate-rich sediments subducted with the a mixing between abiogenic and thermogenic CH4 consistent with the
down-going slab (e.g., Marty and Giggenbach, 1990; Giggenbach and isotopic signature of carbon and hydrogen of this compound (Fig. 8a).

Fig. 8. δ13C-CH4 vs. δD-CH4 (a) and CH4/(C2H6+ C3H8) vs. δ13C-CH4 (b) binary diagrams. Symbols as in Fig. 3.
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 11

6.3. Geothermometry strongly affected by interactions with shallow aquifers (Shock, 1993).
Therefore, in the present study geothermometric evaluations are re-
6.3.1. Water geothermometry stricted to the gas equilibria in the CO2–CH4–H2 system.
An ionic solute geothermometer in the Na+–K+–Mg2+–Ca2+ The following reaction provides the equilibrium constrains for the
system (Giggenbach, 1988) is graphically represented in the 10 K+/ CH4–CO2 pair:
(Na+ + 10 K+) vs. 10 Mg2+/(10 Mg2+ + Ca2+) diagram (Fig. 9),
where the curves of the water composition expected for equilibrium CH4 þ 2H2 O⇔CO2 þ 4H2 ð5Þ
with an average crustal rock (Giggenbach, 1984), and that produced Assuming that log fH2O = 4.9−1820/T and RH = log(H2/H2O)
via isochemical dissolution of crustal rocks are reported. (Giggenbach, 1987), the dependence of the log(XCO2/XCH4) values
Most ET and PT waters seem to attain a full equilibrium at about (where XCO2 and XCH4 are the mole fractions of CO2 and CH4,
220–250 °C, a range of temperatures consistent with that measured in respectively) on temperature and RH in steam produced by boiling of a
a 600 m deep exploratory well at ET (253 °C: Lahsen and Trujillo, liquid phase can be expressed by the following equation:
1976). The S and PL water samples follow a common trend suggesting
a transition from rock dissolution to water–rock equilibrium at logðXCH4 =XCO2 ÞV ¼ 4RH þ 5181=TðKÞ ð6Þ
temperatures similar to those of the ET and PT systems. Conversely,
the PA and TT samples are scattered in an area far from the rock–water whereas in the liquid phase is, as follows:
equilibrium curve. Equilibrium temperatures of the silica (quartz)
logðXCH4 =XCO2 ÞL ¼ 4RH þ 5181=T þ logðBCO2 Þ−logðBCH4 Þ ð7Þ
geothermometer (Fournier and Potter, 1982) are significantly lower
than those indicated by the composition of the main cations, ranging BCO2 and BCH4 are the vapor/liquid distribution coefficients of CO2
between ∼175 °C (ET waters) and ∼100 °C (TT waters) (Table 1). The and CH4, respectively.
SiO2 equilibrium temperatures (≤200 °C) of fluids from the explora- Assuming that Ar is fixed by the equilibrium between the
tion wells drilled in the ET area (Giggenbach, 1978) suggest that the atmosphere and air saturated water (ASW), the dependence of H2
disagreement between the silica and cation geothermometers can on RH in equilibrated vapor is given by (Giggenbach, 1991):
only partially be related to SiO2 precipitation at decreasing tem-
perature along the fluid ascending pathways. As shown by the logðXH2 =XAr ÞV ¼ RH þ 6:52 ð8Þ
distribution of water samples in Fig. 10, lower SiO2 equilibrium where Ar* values were calculated, as follows:
temperatures (TSiO2) correspond to lower Cl− concentrations, sug-
gesting that the hydrothermal fluids are diluted by relatively cold, 
Ar ¼ Ar−ðO2 =22Þ ð9Þ
shallow aquifers. The effect of this process is particularly evident for
the PL and TT fluids that are depleted of gas soluble compounds such The O2 /22 values correspond to the amounts of Ar from
as H2S (Table 2). Accordingly, the TT waters have relatively high SO2− atmospheric contamination, considering that O2 is absent in pristine
4
concentrations (up to 1336 mg/L; Table 1), likely produced by H2S hydrothermal fluids.
dissolution. The relation of log(XH2/XAr*) vs. RH in the saturated liquid phase is
described by the following equation:
6.3.2. Gas geothermometry
logðXH2 =XAr ÞL ¼ RH −logðBH2 Þ þ 6:52 ð10Þ
Gas geothermometry in the H2O–CO2–H2–CO–CH4 system can be
usefully applied to hydrothermal fluids (Chiodini and Marini, 1998). The BCO2, BCH4 and BH2 values at different temperatures are
However, the H2Ovap concentrations in bubbling pools, which are the calculated using the fourth-order polynomial equations reported by
prevalent fluid discharges in the investigated hydrothermal systems, Sepulveda et al. (2007), which are based on the linear expression of
are essentially controlled by water boiling at the surface to very BCO2, BCH4 and BH2 proposed by Giggenbach (1980).
shallow depth. Moreover, CO concentrations in uprising gases are According to Eqs. (6)–(8) and (10), thermodynamic conditions
of the CO2–CH4–H2 equilibria in the various hydrothermal systems can
be shown by the log(XH2/XAr*) vs. log(XCH4/XCO2) diagram (Giggenbach,

Fig. 9. 10* Mg2+/(10* Mg2++ Ca2+) vs. 10* K+/(10* K++ Na+) binary diagram. The
expected composition of waters in equilibrium with an average crustal rock as function
of temperature and that was produced through isochemical dissolution of crustal rocks
(Giggenbach, 1988) are reported. Concentrations are in mg/L. Symbols as in Fig. 3. Fig. 10. TSiO2 (°C) vs. Cl−(in mg/L) binary diagram. Symbols as in Fig. 3.
12 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

1993; Giggenbach et al., 1993), where the vapor–liquid equilibrium hydrothermal fluids. Accordingly, the Puchuldiza gases are char-
grid at RH ranging from −3.6 to −3.0 and temperatures in the interval acterized by lower H2S concentrations than the Tuja ones
of 150–374 °C are reported (Fig. 11): (Table 2). Argon depletion in deep fluids through open system
boiling may also be admitted as an alternative explanation for the
(1) The PA gases seems to be produced by boiling of a liquid phase relatively high temperature and more reduced conditions of the
at temperatures up to 350 °C and RH = −3.4, i.e. at conditions Tuja samples. A similar process can also be invoked for the PA
significantly more oxidizing than those of the FeO/FeO1.5 buffer fluids, resulting in a shift of the apparent CO2–CH4–H2 equilibria
(RH = −2.8), the most typical redox buffer in hydrothermal toward higher temperatures and more reducing conditions;
environment (Giggenbach, 1987). Such equilibrium conditions (4) The S gases show temperatures not exceeding 200 °C at RH
suggest that the gas phase of this system has a deep origin, ∼−3.3, testifying that the fluid reservoir is cooler than those of
likely related to an active magmatic body releasing hot, the ET and PT systems. However, mixing with shallow O2-rich
oxidizing fluids, in agreement with the amounts of SO2 and aquifers, also supported by water chemistry (Fig. 10), may
HCl detected at the surface (Table 2). A similar hypothesis was trigger secondary oxidation processes able to consume H2 and
also suggested by Urzúa et al. (2002), although they did not thus, affecting the calculated CO2–CH4–H2 temperatures. The
detect highly acidic gases in the same fumaroles. The R/Ra value H2 concentrations in the S gases are indeed significantly lower
(1.81) characterizing the PA gases is significantly lower than than those of the PT and ET ones (Table 2);
those expected for an active subduction environment (Poreda (5) The PL and TT gases plot far from any reasonable equilibrium
and Craig, 1989). This indicates a dilution of the primary 3He- conditions. Boiling of shallow aquifers is capable of completely
rich magmatic component by 4He produced in the thick crust of masking deep fluid contribution.
this part of the Andes, where mantle contamination processes
by crustal assimilation are likely (Thorpe et al., 1984). Chemical equilibrium regulating dehydrogenation of the C3 alkane
Aggressive alteration of host rocks in a magmatic–hydrother- (propane) to produce its homologous alkene (propene) can also
mal environment may also be invoked to explain enhanced provide useful indications about thermodynamic conditions domi-
release of 4He in the PA system (Christenson et al., 2002); nating hydrothermal reservoirs (e.g., Seewald, 1994; Capaccioni and
(2) The ET gases equilibrate at RH ranging between −3.2 and −3.3 Mangani, 2001; Taran and Giggenbach, 2003). The C3–C3 reaction is,
and temperatures (260–275 °C) slightly higher than those as follows:
indicated by the cation and silica geothermometers (Figs. 9 and
10), but consistent with the temperatures (270 °C) calculated C3 H8 ⇔C3 H6 þ H2 ð11Þ
by Cortecci et al. (2005) on the basis of the Na/K ratios (Verma
and Santoyo, 1997); In the equilibrated vapor the temperature dependence of
(3) The PT gases cluster in two groups: the first one includes the reaction (11) is (Capaccioni et al., 2004):
samples from Tuja (Fig. 2b) that are characterized by
temperatures similar to those of the ET gases attained at RH
logðXC3H6 =XC3H8 ÞV þ logðXH2 Þ ¼ 7:15−6600=T ð12Þ
b−3; the second one is referred to samples from the Puchuldiza
zone (Fig. 2b) that are shifted toward temperatures of 210–
Considering that log fH2 O = 4.9 − 1820/T, Eq. (12) can be
240 °C and RH ∼−3.15. The difference in the equilibrium
expressed, as follows:
temperatures of the two PT groups seems to be caused by the
relatively high Ar concentrations of the Puchuldiza gases
(Table 2). Assuming that Ar mainly derives from ASW, logðXC3H6 =XC3H8 ÞV ¼ 2:25−RH −4780=T ð13Þ
variations of Ar concentrations may depend on the degree of
gas–water interaction at shallow depth, in its turn depending The behavior of C3H8 and C3H6 in response to vapor–liquid phase
on the thickness of the local aquifers and the flux of ascending changes was found to be almost identical (Tassi et al., 2007), therefore
it is reasonable to assume that:

logðXC3H6 =XC3H8 ÞV ¼ logðXC3H6 =XC3H8 ÞL ð14Þ

Using in Eq. (13) the temperatures fixed by the CO2–CH4–H2


geothermometer (Fig. 11), most of the PT, S, PA and ET gases plot
along the curve at RH = −5.0 (Fig. 12), approximately corresponding
to those of the Hematite–Magnetite (HM) redox buffer (Giggenbach,
1987). This may imply that the C3H6 ⇔ C3H8 conversion was
quenched at greater depth with respect to the CO2–CH4–H2 equi-
librium (Fig. 11). In fact, reaction (11) is characterized by a relatively
low kinetics at temperature b300 °C (Capaccioni et al., 1995), pre-
venting readjustment of the C3H6/C3H8 ratios at decreasing temper-
ature during the uprising of thermal fluids, whereas H2 is capable of
reflecting almost instantaneous changes of thermodynamic condi-
tions (Giggenbach, 1987). As suggested for the PA gases to justify the
RH values of the CO2–CH4–H2 equilibria (Fig. 11) and the presence of
acidic compounds (Table 2), the highly oxidizing conditions control-
ling the C3–C3 dehydrogenation (Fig. 12) seem to indicate that
magmatic fluid inputs affect the root of the hydrothermal reservoirs at
the PT, S and ET systems. This agrees with previous studies that have
invoked a magmatic contribution to the ET discharges on the basis of
Fig. 11. A binary diagram of log(H2/Ar*) vs. log(CH4/CO2) binary diagram. Ar* = the presence of a low pH deep-seated brine (Giggenbach, 1978) and
Ar−O 2/22; RH = log(H2/H2O) (Giggenbach, 1987). Symbols as in Fig. 3. the hydrocarbon composition (Tassi et al., 2005).
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 13

hydrothermalism in this region is directly associated with the recent


volcanic activity.

Acknowledgements

This research was partially financed by Dirección General de


Investigación y Postgrado (UCN-Chile) and by D-21050592 and
23080036 CONICYT grant (Government of Chile) and the Laboratories
of Fluid and Rock Geochemistry of the Department of Earth Sciences
and CNR-IGG of Florence (Resp. F. Tassi and O. Vaselli).
A. Aiuppa, G. Chiodini and B. Christenson are warmly thanked for
the comments and suggestions they provided in an early version of
the manuscript.
The authors wish to express their gratitude to Jaime Llanos
(Inorganic Chemical Laboratory—UCN—Chile) for facilities in the
samples preparation and to Eduardo Medina, César Donoso, Karen
Guzman, José Luis Mercado and Benigno Godoy (UCN—Chile) for their
help during the sampling campaigns.

References

Fig. 12. Binary diagram of log(C3H6/C3H8) vs. equilibrium temperature (°C) in the Abrajano, T.A., Sturchio, N.C., Bohlke, J.K., Lyon, G.L., Poreda, R.J., Stevens, C.M., 1988.
CO2–CH4–H2 system. Symbols as in Fig. 3. Methane–hydrogen gas seeps, Zambales ophiolite, Philippines: deep or shallow
origin? Chem. Geol. 7, 211–222.
Acocella, V., Vezzoli, L., Omarini, R., Matteini, M., Mazzuoli, R., 2007. Kinematic
7. Conclusions variations across Eastern Cordillera at 24°S (Central Andes): tectonic and magmatic
implications. Tectonophysics 434, 81–92.
Aggarwal, J.K., Palmer, M.R., Bullen, T.D., Arnosson, S., Ragnarsdottir, K.V., 2000. The
Hydrothermal manifestations associated with Quaternary volca- boron isotope systematic of Icelandic geothermal waters: 1. Meteoric water
nism of the Andean cordillera in northern Chile are mainly distributed charged systems. Geochim. Cosmochim. Acta 64, 579–585.
Aguilera, F., 2008. Origin and evolution of fluid in volcanoes, geothermal fields and
within the six areas investigated in the present paper. Geothermal
thermal discharges of Central Volcanic Zone in northern Chile (17°43′S–25°10′S).
exploration projects focused on assessment of resources for electric Ph.D. thesis, Univ. Catol. Norte, Chile (in Spanish).
power generation carried out in the past decades have provided Ahumada, S., Mercado, J, L., 2008. Origin and geological–structural evolution of the
preliminary evaluations of the geothermal potential at S, PT and ET Pampa Apacheta sector, 2nd Region, Antofagasta. Unpub. undergraduate thesis,
Univ Catol. Norte, Chile (in Spanish).
whereas little information was available for the PL, PA and TT Aravena, R., 1995. Isotope hydrology and geochemistry of Northern Chile ground-
hydrothermal systems. Although a reliable estimation of the geother- waters. Bull. Inst. Fr. Études Andines 24, 497–503.
mal potential requires detailed geophysical prospecting, our results Aravena, R., Suzuki, O., Peña, H., Pollastri, A., Fuenzalida, H., Grilli, A., 1999. Isotopic
composition and origin of the precipitation in northern Chile. Appl. Geochem. 14,
confirm the results of previous studies indicating an extended high 411–422.
temperature (up to 270 °C) liquid-dominated reservoir is present at Arehart, G.B., Coolbaugh, M.F., Poulson, S.R., 2003. Evidence for a magmatic source of
ET. Similar chemical–physical conditions characterize the source heat for the Steamboat Springs geothermal system using trace elements and gas
geochemistry. Geotherm. Res. Transact. 27, 12–15.
region of the thermal fluids emerging in the NW sector of the PT Arnósson, S., Andresdottir, A., 1995. Processes controlling the distribution of boron and
system (Tuja zone). The chemistry of the thermal discharges of the chlorine in natural waters in Iceland. Geochim. Cosmochim. Acta 59, 4125–4146.
Puchuldiza zone, located a few km SE of the Tuja area, indicates lower Arnósson, S., Bjarnason, J.Ö., Giroud, N., Gunnarsson, I., Stefánsson, A., 2006. Sampling
and analysis of geothermal fluids. Geofluids 6, 1–14.
reservoir temperature in the range of 210–220 °C, consistent with the
Barragan, R.M., Arellano, V.M., Birkle, P., Portugal, M., Dìaz, H., 1999. Isotopic
geothermometric calculations of Mahon and Cusicanqui (1980). composition and origin of the precipitation in northern Chile. Appl. Geochem. 14,
Therefore, different fluid sources may feed the PT discharges, one 411–422.
Berndt, M.E., Allen, D.E., Seyfried, W.E., 1996. Reduction of CO2 during serpentinization
(at Puchuldiza) relating to a secondary, steam-heated aquifer and the
of olivine at 300 °C and 500 bar. Geology 24, 351–354.
other (at Tuja) to a deeper and hotter reservoir. This hypothesis Bischoff, L.B., Seyfried, W.E., 1978. Hydrothermal chemistry of seawater from 25 °C to
suggests that the compositional differences between the two main 350 °C. Am. J. Sci. 278, 838–860.
discharging zones of the PT system are regulated by secondary Bréas, O., Guillou, C., Reniero, F., Wada, E., 2001. The global methane cycle: isotopes and
mixing ratios, sources and sinks. Isot. Environ. Health Stud. 37, 257–379.
processes, such as thickness of shallow aquifers and local variations of Capaccioni, B., Mangani, F., 2001. Monitoring of active but quiescent volcanoes using
permeability. Secondary gas–water–rock interactions likely affect the light hydrocarbon distribution in volcanic gases: the results of 4 years of
chemistry of hydrothermal fluids of the S system, where realistic discontinuous monitoring in the Campi Flegrei (Italy). Earth Planet. Sci. Lett. 188,
543–555.
reservoir temperature may surpass temperatures estimated by gas Capaccioni, B., Martini, M., Mangani, F., 1995. Light hydrocarbons in hydrothermal and
geothermometry (b200 °C). The geochemical features of the thermal magmatic fumaroles: hints of catalytic and thermal reactions. Bull. Volcanol. 56,
discharges of the PL and TT areas are typical of fluids evolved within 593–600.
Capaccioni, B., Taran, Y., Tassi, F., Vaselli, O., Mangani, F., Macias, J.L., 2004. Source
shallow aquifers. Deep fluid contributions, indicated by the He conditions and degradation processes of light hydrocarbons in volcanic gases: an
isotopic signature above air, are minor in these systems. Conversely, example from El Chichón volcano (Chiapas State, Mexico). Chem. Geol. 20, 81–96.
the fumarolic fluids of the PA volcano summit are produced by a non- Capaccioni, B., Tassi, F., Vaselli, O., Tedesco, D., Poreda, R.J., 2007. Submarine gas burst at
Panarea Island (southern Italy) on 3 November 2002: a magmatic versus
conventional hydrothermal system, since the reservoir calculated
hydrothermal episode. J. Geophys. Res. 112, B05201. doi:10.1029/2006JB004359.
temperatures were extremely high (N320 °C) with significant Casertano, L., 1963. General characteristics of active Andean volcanoes and a summary
amounts of gas species related to magma degassing. It is worth of their activities during recent centuries. Bull. Seismol. Soc. Am. 53, 1415–1433.
Céspedes, L., Clavero, J., Cayupi, J., 2004. Hazard management at Isluga volcano,
noting that estimated redox conditions at the fluid source of the most
Northern Chile: preliminary results. Final Proc. IAVCEI General Assembly, Pucón,
promising hydrothermal systems of northern Chile, i.e. S, PT, ET and Chile, 11447-a.
PA, are significantly more oxidizing than those typical of most Chaffaut, I., Coudrain-Ribstein, A., Michelot, J.L., Pouyaud, B., 1998. Prècipitations
hydrothermal systems (Giggenbach, 1987). This evidence, coupled d'altitude du Nord-Chile, origine des sources de vapeur et donnèes isotopiques.
Bull. Inst. Fr. Études Andines 27, 367–384.
with the He isotopic signature, indicating the presence of a significant Chiodini, G., Marini, L., 1998. Hydrothermal gas equilibria: the H2O–H2–CO2–CO–CH4
fraction of fluids from a deep (mantle) source, suggests that the system. Geochim. Cosmochim. Acta 62, 2673–2687.
14 F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15

Chong, G., Pueyo, J., Demergasso, C., 2000. Borate deposits of Chile. Rev. Geol. Chile 27, Giggenbach, W.F., Poreda, R., 1993. Helium isotopic and chemical composition of gases
99–119 (in Spanish). from the volcanic–hydrothermal systems in the Philippines. Geothermics 22,
Christenson, B.W., Mroczek, E.K., Kennedy, B.M., van Soest, M.C., Stewart, M.K., Lyon, G., 369–380.
2002. Ohaaki reservoir chemistry: characteristics of an arc-type hydrothermal Giggenbach, W.F., Stewart, M.K., 1982. Processes controlling the isotopic composition of
system in the Taupo Volcanic Zone, New Zealand. J. Volcanol. Geotherm. Res. 115, steam and water discharges from steam vents and steam heated pools in
53–82. geothermal areas. Geothermics 11, 71–80.
Coleman, M.L., Shepherd, T.J., Durham, J.J., Rouse, J.E., Moore, G.R., 1982. Reduction of Giggenbach, W.F., Sano, Y., Wakita, H., 1993. Isotopic composition of helium, and CO2
water with zinc for hydrogen isotope analysis. Anal. Chem. 54, 993–995. and CH4 contents in gases produced along the New Zealand part of a convergent
Cortecci, G., Boschetti, T., Mussi, M., Herrera Lameli, C., Mucchino, C., Barbieri, M., 2005. plate boundary. Geochim. Cosmochim. Acta 57, 3427–3455.
New chemical and original isotopic data on waters from El Tatio geothermal field, Global Volcanism Program, 1997. Irruputuncu. Volcanic Activity Reports. BGVN 22:07.
northern Chile. Geochem. J. 39, 547–571. http://www.volcano.si.edu.
Craig, H., Lupton, J.E., Horibe, Y., 1978. A mantle helium component in Circum-Pacific Gonfiantini, R., Roche, M.A., Olivry, J.C., Fontes, J.C., Zuppi, G.M., 2001. The altitude effect
volcanic gases: Hakone, the Marianas, and Mt. Lassen. In: Alexander, E.C., Ozima, M. on the isotopic composition of tropical rains. Chem. Geol. 181, 147–167.
(Eds.), Terrestrial Rare Gases. Japan Sci. Soc. Press, Tokyo, p. 316. Hauser, A., 1997. Register and characterization of mineral and thermal springs of Chile,
Cusicanqui, H., 1978a. Geochemical monitoring of wells 12 and 13 at El Tatio. Unpubl. 50. Nation. Serv. Geol. and Min, Santiago, Chile, p. 90 (in Spanish).
report, Committee for geothermal energy resources (CORFO), 8 pp. (in Spanish). Healy, J., 1969. Geological study of hydrothermal systems in the Tarapacá and
Cusicanqui, H., 1978b. Control geoquimico de los pozos 7, 10 y 11 de El Tatio y estudio Antofagasta provinces: preliminary report of Tarapacá Province. Unpubl report,
geoquimico de los pozos 2, 3, 4, y 5 del campo geotermal de Puchuldiza. Unpubl Committee for geothermal energy resources (CORFO), 16 pp. (in Spanish).
report, Committee for geothermal energy resources (CORFO), 8 pp. (in Spanish). Hilton, D.R., Hammerschmidt, K., Teufel, S., Friedrichsen, H., 1993. Helium isotope
Cusicanqui, H., 1979. Geochemical study of the Suriri thermal area, Arica province, characteristics of Andean geothermal fluids and lavas. Earth Planet. Sci. Lett. 120,
1st Region Estudio geoquímico del área termal de Suriri–Provincia de Arica–I 265–282.
Región. Unpubl report, Committee for geothermal energy resources (CORFO), 29 Hochstein, M.P., 1971. Geophysical survey of the Puchuldiza geothermal area results up
pp. (in Spanish). to December 1970, Survey for geothermal development in northern Chile. Unpubl.
Cusicanqui, H., Mahon, W.A.J., Ellis, A.J., 1975. The geochemistry of the El Tatio report UN, 22 pp.
geothermal field, Northern Chile. 2nd UN Symposium, Development and Utilization Hoefs, J., 1997. Stable Isotope Geochemistry, 4th ed. Springer, Verlag, Berlin, Germany.
of Geothermal Resources, San Francisco, pp. 703–711. Holloway, J.R., 1984. Graphite–CH4–H2O–CO2 equilibria at low-grade metamorphic
Darling, W.G., 1998. Hydrothermal hydrocarbon gases: I. Genesis and geothermometry. conditions. Geology 12, 455–458.
Appl. Geochem. 13, 815–824. Horita, J., Berndt, M.E., 1999. Abiogenic methane formation and isotopic fractionation
Des Marais, D.J., Moore, J.G., 1984. Carbon and its isotopes in mid-oceanic basaltic under hydrothermal conditions. Science 285, 1055–1057.
glasses. Earth Planet. Sci. Lett. 69, 43–57. Hulston, J.R., Lupton, J.E., 1996. Helium isotope studies of geothermal fields in the Taupo
deSilva, S., Francis, P., 1991. Volcanoes of Central Andes. Springer-Verlag, Berlin, Volcanic Zone, New Zealand. J. Volcanol. Geotherm. Res. 74, 297–321.
Germany. 216 pp. Huttrer, G.W., 1996. The status of world geothermal power production 1990–1994.
ELC-Electroconsult, 1980. Geothermal field of El Tatio: evaluation of the geothermal Geothermics 25, 1–27.
potential. Unpubl report, Committee for geothermal energy resources (CORFO), 13 Japan International Cooperation Agency (JICA), 1979. Geothermal power development
pp. (in Spanish). project in Puchuldiza area. Unpubl. Report, 109 pp.
Ellis, A.J., Mahon, W.A.J., 1977. Chemistry and Geothermal Systems. Academic Press. Japan International Cooperation Agency (JICA), 1981. Report on geothermal power
Epstein, S., Mayeda, T.K., 1953. Variation of the 18O/16O ratio in natural waters. development project in Puchuldiza area. Unpubl. Report, 48 pp.
Geochim. Cosmochim. Acta 4, 213–224. Kenney, J.K., 1995. Comment on mantle hydrocarbons: abiotic or biotic? By R. Sugisaki
Fiebig, J., Chiodini, G., Caliro, S., Rizzo, A., Spangenberg, J., Hunziker, J.C., 2004. Chemical and K. Mimura. Geochim. Cosmochim. Acta 59, 3857–3858.
and isotopic equilibrium between CO2 and CH4 in fumarolic gas discharges: Kiyosu, Y., Asada, N., Yoshida, Y., 1992. Origin of light hydrocarbon gases from the
generation of CH4 in arc magmatic–hydrothermal systems. Geochim. Cosmochim. Matsukawa geothermal area in Japan. Chem. Geol. 94, 321–329.
Acta 68, 2321–2334. Lahsen, A., 1969 Geology of the area between El Tatio and Cerros de Ayquina. Unpubl.
Fiebig, J., Woodland, A.B., Spangenberg, J., Oschmann, W., 2007. Natural evidence for report, Committee for geothermal energy resources (CORFO), 75 pp. (in Spanish).
rapid abiogenic hydrothermal generation of CH4. Geochim. Cosmochim. Acta 71, Lahsen, A., 1976. Geothermal exploration in Northern Chile. A summary. Final Proc. En.
3028–3039. Min. Res. Conf., 12–14 July, Honolulu, Hawaii: A.A.P.G. Memoir, 25, pp. 169–175.
Fouillac, C., Michard, G., 1981. Sodium/ lithium ratio in water applied to geothermo- Lahsen, A., 1988. Chilean geothermal resources and their possible utilization.
metry of geothermal reservoirs. Geothermics 10, 55–70. Geothermics 17 (2/3), 401–410.
Fournier, R.O., Potter, R.W., 1982. A revised and expanded silica (quartz) geotherm- Lahsen, A., Trujillo, P., 1976. El campo geotermico de El Tatio, Chile. Int. Rep., CORFO-
ometer. Geotherm. Res. Counc. Bull. 11, 3–12. ONU, p. 21 (in Spanish).
Foustoukos, D.I., Seyfried, W.E., 2004. Hydrocarbons in hydrothermal vent fluids: the Letelier, M., 1981. Geochemistry of thermal manifestations in Puchuldiza and
role of chromium-bearing catalysts. Science 304, 1002–1004. surrounding areas. Unpubl. report, Committee for geothermal energy resources
Francis, P., Rundle, C., 1976. Rates of production of the main magma types in the Central (CORFO), 60 pp. (in Spanish).
Andes. Geol. Soc. Am. Bull. 87, 474–480. Mahon, W., 1970. Preliminary reports on the geochemistry of the hot springs in the
García, M., Gardeweg, M., Clavero, J., Hérail, G., 2004. Arica map: Tarapacá Region, scale Puchuldiza—Lupe Geothermal Area and the fluids discharged from well N° 1 El
1:250.000. Serv. Nac. Geol. Min. 84. Tatio, survey for geothermal development in northern Chile. Unpubl. report,
Gardeweg, M.C., Medina, E., 1994. The 19–20 April 1993 sub-plinian eruption of Lascar Committee for geothermal energy resources (CORFO), 39 pp.
volcano, northern Chile. Final Proc. 7th Chilean Geological Congress, Concepcion, Mahon, W., 1974. The geochemistry of the El Tatio geothermal System, survey for
Chile, pp. 229–304 (in Spanish). geothermal development in northern Chile. Unpubl report, UN, 109 pp.
Giggenbach, W., 1978. The isotope composition of waters from the El Tatio geothermal Mahon, W.A.J., Cusicanqui, H., 1980. Geochemistry of the Puchuldiza and Tuja hot
field, northern Chile. Geochim. Cosmochim. Acta 42, 979–988. springs, Chile. N.Z. J. Sci. 23, 149–159.
Giggenbach, W.F., 1980. Geothermal gas equilibria. Geochim. Cosmochim. Acta 44, Mamyrin, B.A., Tolstikhin, I.N., 1984. Helium isotopes in nature. In: Fyfe, W.S. (Ed.),
2021–2032. Development in Geochemistry. Elsevier, Amsterdam, p. 274.
Giggenbach, W.F., 1984. Mass transfer in hydrothermal alteration systems. Geochim. Marinovic, N., Lahsen, A., 1984. Calama map, Antofagasta Region. Geologic map of Chile
Cosmochim. Acta 48, 2693–2711. no. 58, scale 1:250, 000. Serv. Nac. Geol. Min. ISSN, 0716-0194.
Giggenbach, W.F., 1987. Redox processes governing the chemistry of fumarolic gas Martini, M., Cellini Legittimo, P., Piccardi, G., Giannini, L., 1984. Low temperature
discharges from White Island, New Zealand. Appl. Geochem. 2, 143–161. manifestations in volcanic areas. Rend. Soc. Ital. Miner. Petrol. 39, 401–405.
Giggenbach, W.F., 1988. Geothermal solute equilibria. Derivation of Na–K–Mg–Ca Marty, B., Giggenbach, W.F., 1990. Major and rare gases at White Island volcano, New
geoindicators. Geochim. Cosmochim. Acta 52, 2749–2765. Zealand: origin and flux of volatiles. Geophys. Res. Lett. 17, 247–250.
Giggenbach, W.F., 1991. Chemical techniques in geothermal exploration. Application of Marty, B., Jambon, A., 1987. C/3He in volatile fluxes from the solid earth—implications
Geochemistry in Geothermal Reservoir Development. UNITAR, New York, pp. 253–273. for carbon geodynamics. Earth Planet. Sci. Lett. 83, 16–26.
Giggenbach, W.F., 1992. Isotopic shifts in waters from geothermal and volcanic systems Matsuo, S., Suzuki, M., Mizutani, Y., 1978. Nitrogen to argon ratio in volcanic gases. Adv.
along convergent plate boundaries and their origin. Earth Planet. Sci. Lett. 113, Earth Planet. Sci. 3, 17–25.
495–510. Matteini, M., Mazzuoli, R., Omarini, R., Cas, R., Maas, R., 2002a. The geochemical
Giggenbach, W.F., 1993. Redox control of gas compositions in Philippines volcanic– variations of the upper Cenozoic volcanism along the Calama–Olocapato–El Toro
hydrothermal systems. Geothermics 22, 575–587. transversal fault system in central Andes (24°S): petrogenetic and geodynamic
Giggenbach, W.F., 1995. Variations in the chemical and isotopic composition of fluids implications. Tectonophysics 345, 211–227.
discharged from the Taupo Volcanic Zone, New Zealand. J. Volcanol. Geotherm. Res. 68, Matteini, M., Mazzuoli, R., Omarini, R., Cas, R., Maas, R., 2002b. Geodynamical evolution of
89–116. the central Andes at 24°S as inferred by magma composition along the Calama–
Giggenbach, W.F., 1997. The origin and evolution of fluids in magmatic–hydrothermal Olocapato–El Toro transversal volcanic belt. J. Volcanol. Geotherm. Res. 118, 225–228.
systems, In: Barnes, H.L. (Ed.), Geochemistry of Hydrothermal Ore Deposits, 3rd Matthews, S.J., Gardeweg, M.C., Sparks, R.S.J., 1997. The 1984 to 1996 cyclic activity of
Edition. Wiley, pp. 737–796. Lascar Volcano, northern Chile: cycles of dome growth, dome subsidence,
Giggenbach, W.F., Gougel, R.L., 1989. Method for the collection and analysis of degassing and explosive eruptions. Bull. Volcanol. 59, 72–82.
geothermal and volcanic water and gas samples. NZ-DSIR Report, CD 2387, 53. McCollom, T.M., 2003. Formation of meteorite hydrocarbons from thermal decompo-
Giggenbach, W.F., Matsuo, S., 1991. Evaluation of results from Second and Third IAVCEI sition of siderite (FeCO3). Geochim. Cosmochim. Acta 67, 311–317.
Workshops on Volcanic Gases, Mt. Usu, Japan, and White Island, New Zealand. McDonald, W.J., 1969. Preliminary geophysical report on El Tatio geothermal field.
Appl. Geochem. 6, 125–141. Unpubl. report, Committee for geothermal energy resources (CORFO), 18 pp.
F. Tassi et al. / Journal of Volcanology and Geothermal Research 192 (2010) 1–15 15

McDonald, W.J., 1974. Geophysical survey of El Tatio geothermal field, Survey for Snyder, G., Poreda, R., Fehn, U., Hunt, A., 2003. Sources of nitrogen and methane in Central
Geothermal Development in Northern Chile. Unpubl. report, Committee for American geothermal settings: noble gas and 129I evidence for crustal and magmatic
geothermal energy resources (CORFO), 26 pp. volatile components. Geochem. Geophys. Geosyst. 4. doi:10.1029/2002GC000363.
Montegrossi, G., Tassi, F., Vaselli, O., Buccianti, A., Garofalo, K., 2001. Sulfur species in Spivack, A.J., Edmond, J.M., 1987. Boron isotopic exchange between seawater and the
volcanic gases. Anal. Chem. 73, 3709–3715. oceanic crust. Geochim. Cosmochim. Acta 51, 1033–1044.
Montgomery, E., Rosko, M., 1996. Groundwater exploration and wellfield development Springer, M.H., Förster, A., 1998. Heat flow density across the Central Andean
in the Pampa Lagunillas and Pampa Lirima areas, Iquique Province, Chile. Rev. Geol. subduction zone. Tectonophysics 291, 123–139.
Chile 23, 135–149. Steinmüller, K., 2001. Modern hot springs in the southern volcanic volcanic Cordillera
Negrel, P., Casanova, J., Azaroual, M., Guerrot, C., Cocherie, A., Fouillac, C., 1999. Isotope of Peru and their relationship to neogene epithermal precious-metal deposits.
geochemistry of mineral spring waters in the Massif Central France. In: Armannsson, J. South Am. Earth Sci. 14, 377–385.
H. (Ed.), Geochemistry of Earth's Surface. Rotterdam, The Netherlands, pp. 531–533. Stern, C., 2004. Active Andean volcanism: its geologic and tectonic setting. Rev. Geol.
Ohmoto, H., Goldhaber, M.B., 1997. Sulfur and carbon isotopes. In: Barnes, H.L. (Ed.), Chile 31, 161–206.
Geochemistry of Hydrothermal Ore Deposits. John Wiley and Sons, pp. 517–611. Sugisaki, R., Mimura, K., 1994. Mantle hydrocarbons: abiotic or biotic? Geochim.
Oremland, R.S., Miller, L.G., Whiticar, M.J., 1987. Sources and flux of natural gases from Cosmochim. Acta 58, 2527–2542.
Mono Lake, California. Geochim. Cosmochim. Acta 51, 2915–2929. Symonds, R.B., Rose, W.I., Bluth, W.I., Gerlach, T.M., 1994. Volcanic–gas studies:
Palmer, M.R., Spivack, A.J., Edmond, J.M., 1987. Temperature and pH controls over methods, results, and applications. In: Carrol, M.R., Holloway, J.R. (Eds.), Volatiles in
isotopic fractionation during adsorption of boron on marine clays. Geochim. Magmas. Washington, D.C. Rev. Mineral., 30, pp. 1–66.
Cosmochim. Acta 51, 2319–2323. Symonds, R.B., Gerlach, T.M., Reed, M.H., 2001. Magmatic gas scrubbing: implications
Patino, L.C., Carr, M.J., Feigenson, M.D., 2000. Local and regional variations in Central for volcano monitoring. J. Volcanol. Geochim. Res. 108, 303–341.
American arc lavas controlled by variations in subducted sediment input. Contrib. Szatmari, P., 1989. Petroleum formation by Fischer–Tropsch synthesis in plate tectonics.
Mineral. Petrol. 138, 265–283. AAPG Bull. 73, 989–998.
Petrinovic, I., Riller, U., Brod, J., Alvarado, G., Arnosio, M., 2006. Bimodal volcanism in a Taran, Y.A., Giggenbach, W.F., 2003. Geochemistry of light hydrocarbons in subduction-
tectonic transfer zone: evidence for tectonically controlled magmatism in the related volcanic and hydrothermal fluids. In: Simmons, S.F., Graham, I.J. (Eds.),
southern Central Andes, NW Argentina. J. Volcanol. Geotherm. Res. 152, 240–252. Volcanic, geothermal, and ore-forming fluids: rulers and witnesses of processes
Polanco, E., Gardeweg, M., 2000. Preliminary study of the volcanic stratigraphy of Upper within the Earth. Soc. Econ. Geol. Spec. Issue. Littleton, Colo, pp. 61–74.
Cenozoic at Pampa Lirima and Cancosa, 1st Region highland, Chile (19°45′–20°00′S Taran, Y.A., Pokrovsky, B.G., Esikov, A.D., 1989. Deuterium and oxygen-18 in fumarolic
and 69°00′–68°30′W). Final proc. 9th Chilean Geol. Congr. Puerto Varas, Chile, steam and amphiboles from some Kamchatka volcanoes: “andesitic waters”. Dokl
pp. 324–328. Akad Nauk SSSR 304, 440–443.
Poreda, R.J., Craig, H., 1989. Helium isotope ratios in circum-Pacific volcanic arcs. Nature Taran, Y., Fisher, T.P., Cienfuegos, E., Morales, P., 2002. Geochemistry of hydrothermal
338, 473–478. fluids from an intraplate ocean island: Everman volcano, Socorro Island, Mexico.
Poreda, R.J., Farley, K.A., 1992. Rare-gases in Samoan xenoliths. Earth Planet. Sci. Lett. Chem. Geol. 188, 51–63.
113, 129–144. Tassi, F., Martinez, C., Vaselli, O., Capaccioni, B., Viramonte, J., 2005. The light hydrocarbons
Poreda, R.J., Jeffrey, A.W.A., Kaplan, L.R., Craig, H., 1988. Magmatic helium in as new geoindicators of equilibrium temperatures and redox conditions of geothermal
subduction-zone natural gases. Chem. Geol. 71, 199–210. fields: evidence from El Tatio (northern Chile). Appl. Geochem. 20, 2049–2062.
Riller, U., Petrinovic, I., Ramelow, J., Strecker, M., Oncken, O., 2001. Late Cenozoic Tassi, F., Vaselli, O., Capaccioni, B., Montegrossi, G., Barahona, F., Caprai, A., 2007.
tectonism, collapse caldera and plateau formation in the central Andes. Earth Scrubbing processes and chemical equilibria controlling the composition of light
Planet. Sci. Lett. 188, 299–311. hydrocarbons in natural gas discharges: an example from the geothermal fields of
Risk, G.F., 1970. Geophysical studies of the Puchuldiza geothermal area during Salvador. Geochem. Geophys. Geosyst. 8, Q05008. doi:10.1029/2006GC001487.
December 1969 and January 1970, Survey for Geothermal Development in Thornton, E.C., Seyfried Jr., W.E., 1987. Reactivity of organic-rich sediment in seawater at
northern Chile. Unpubl report, UN, 23 pp. 350 8C, 500 bars: experimental and theoretical constraints and implications for the
Rodgers, K.A., Greatrex, R., Hyland, M., Simmons, S.F., Browne, P.R.L., 2002. A modern, Guaymas Basin hydrothermal system. Geochim. Cosmochim. Acta 51, 1997–2010.
evaporitic occurrence of teruggite Ca4MgB12As2O28 · 18H2O, and nobleite, CaB6O10 · Thorpe, R.S., Francis, P.W., O'Callaghan, L., 1984. Relative roles of source composition,
4H2O, from the El Tatio geothermal field, Antofagasta Province, Chile. Mineral. Mag. fractional crystallization and crustal contamination in the petrogenesis of Andean
66 (2), 253–259. volcanic rocks. Phil. Trans. R. Soc. Lond. A 310, 675–692.
Rollinson, H., 1993. Using Geochemical Data. Longman, London, UK. pp. 352. Tibaldi, A., Corazzato, C., Rovida, A., 2008. Miocene–Quaternary structural evolution
Salas, R., 1975. Geological study of Salar de Surire, Arica Province, Chile. Report IIG-JAA, of the Uyuni–Atacama region, Andes of Chile and Bolivia. Tectonophysics.
Instituto de Investigaciones Geológicas, Arica, Chile, p. 82 (in Spanish). doi:10.1016/j.tecto.2008.09.011.
Sano, Y., Marty, B., 1995. Origin of carbon in fumarolic gases from island arcs. Chem. Tonani, F., 1970. Geochemical methods of exploration for geothermal energy.
Geol. 119, 265–274. Geothermics 2, 492–515.
Scandiffio, G., Verastegui, D., Portilla, F., 1992. Geochemical report on the Challapalca and Trujillo, P., 1971. Sub-surficial geology at El Tatio, Antofagasta Province. Unpubl. report,
Tutupaca geothermal areas, Peru. Report IAEA-TECDOC 641, Estudios Geotérmicos con Committee for geothermal energy resources (CORFO), 14 pp. (in Spanish).
Técnicas Isotópicas y Geoquímicas en América Latina, Vienna, pp. 345–375. Trujillo, P., 1972. Study of the thermal manifestations of Suriri. Unpubl. report,
Schoell, M., 1980. The hydrogen and carbon isotopic composition of methane from Committee for geothermal energy resources (CORFO), 15 pp. (in Spanish).
natural gases of various origins. Geochim. Cosmochim. Acta 44, 649–661. Urzúa, L., Powell, T., Cumming, W., Dobson, P., 2002. Apacheta, a new geothermal
Schoell, M., 1988. Multiple origins of methane in the Earth. Chem. Geol. 71, 1–10. prospect in northern Chile: Geoth. Res. Coun. Transact. , p. 10.
Seewald, J.S., 1994. Evidence for metastable equilibrium between hydrocarbons under Vaselli, O., Tassi, F., Montegrossi, G., Capaccioni, B., Giannini, L., 2006. Sampling and
hydrothermal conditions. Nature 370, 285–287. analysis of volcanic gases. Acta Volcanol. 18, 65–76.
Sepulveda, F., Lahsen, A., Powell, T., 2007. Gas geochemistry of the Cordón Caulle Verma, M.P., Santoyo, E., 1997. New improved equations for Na/K, Na/Li and SiO2
geothermal system, Southern Chile. Geothermics 36, 389–420. geothermometers by outlier detection and rejection. J. Volcanol. Geotherm. Res. 79,
Seyfried, W.E., Janecky, D.R., Mottl, M.J., 1984. Alteration of the oceanic crust: 9–24.
implications for geochemical cycles of lithium and boron. Geochim. Cosmochim. Welhan, J.A., Craig, H., 1979. Methane and hydrogen in East Pacific Rise hydrothermal
Acta 48, 557–569. fluids. Geophys. Res. Lett. 6, 829–831.
Sherwood-Lollar, B., Westgate, T.D., Ward, J.A., Slater, G.F., Lacrampe-Couloume, G., Welhan, J.A., Poreda, R.J., Rison, W., Craig, H., 1988. Helium isotopes in geothermal and
2002. Abiogenic formation of alkanes in the Earth's crust as a minor source for volcanic gases of the western United States. I. Regional variability and magmatic
global hydrocarbon reservoirs. Nature 416, 522–524. origin. J. Volcanol. Geotherm. Res. 34, 185–199.
Shock, E.L., 1993. Hydrothermal dehydration of aqueous organic compounds. Geochim. Whiticar, M.J., 1999. Carbon and hydrogen isotope systematics of bacterial formation
Cosmochim. Acta 57, 3341–3349. and oxidation of methane. Chem. Geol. 161, 291–314.
Snyder, G., Poreda, R., Hunt, A., Fehn, U., 2001. Regional variations in volatile composition: Whiticar, M.J., Suess, E., 1990. Hydrothermal hydrocarbon gases in the sediments of the
isotopic evidence for carbonate recycling in the Central American volcanic arc. King-George Basin, Bransfield Strait, Antarctica. Appl. Geochem. 5, 135–147.
Geochem. Geophys. Geosyst. 2. doi:10.1029/2001GC000163.

You might also like