You are on page 1of 14

/ Journal of Volcanology and Geothermal Research 346 (2017) 81–94

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

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

Gas chemistry of Icelandic thermal fluids


Andri Stefánsson
Institute of Earth Sciences, University of Iceland, Sturlugata 7, 101 Reykjavík, Iceland

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

Article history: The chemistry of gases in thermal fluids from Iceland was studied in order to evaluate the sources and processes
Received 14 November 2016 affecting volatile concentrations in volcanic geothermal systems at divergent plate boundaries. The fluids includ-
Received in revised form 15 March 2017 ed vapor fumaroles and two-phase well discharges with temperatures of ~100–340 °C. The vapor was dominated
Accepted 2 April 2017
by H2O accounting for 62–100 mol% and generally for N 99 mol%, with CO2, H2S and H2 being the dominant gases
Available online 7 April 2017
followed by N2, CH4, and Ar. Overall mineral-gas and gas-gas equilibria were not observed for the major gases,
Keywords:
including CO2, H2S, H2 and CH4 within the geothermal reservoirs. Instead the system proved to be controlled
Hydrothermal by source(s) and their ratios and various metastable equilibria along a fluid-rock reaction progress with gas con-
Fluid centrations controlled by such metastable equilibria varying at particular temperatures as a functional extent of
Equilibrium reaction. The concentrations of H2S and H2 closely reflect mineral-fluid metastable equilibria, whereas CO2 con-
centrations are controlled by the input of magma gas corresponding to N0.1 to b 5% mass input. With fluid ascent
to the surface, boiling and condensation may occur, further changing the gas concentrations and hence surface
fumaroles may not reflect the reservoir fluid characteristics but rather secondary processes.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction These equilibria have been based on the concentration of a single reac-
tive gas, like CO2, H2S or H2 (Arnórsson and Gunnlaugsson, 1985), two
The chemistry of gases in thermal fluids has been extensively stud- or more reactive gases like CO2-H2, H2S-H2 and CH4-CO2-CO
ied for over four decades, for the purpose of tracing the source(s) and (Arnórsson and Gunnlaugsson, 1985; Chiodini and Cioni, 1989) or a
reactions of gases as well as for exploration and utilization of geother- ratio of reactive and inert gas like CO2/Ar, H2/Ar and CO2/N2
mal resources (e.g., D'Amore and Nuti, 1977; Giggenbach, 1980, 1987, (e.g., Arnórsson, 1987; Giggenbach, 1991), the last introduced to ac-
1993; D'Amore and Panichi, 1980; Arnórsson and Gunnlaugsson, count for possible gas condensation in the fluid conduit in geothermal
1985; Taran, 1986, 2011; Arnórsson et al., 1990; Chiodini and Cioni, systems.
1989; Chiodini and Marini, 1998; Stefánsson and Arnórsson, 2002; Gases in thermal fluids in Icelandic geothermal systems have been a
Scott et al., 2014). The major gases in thermal fluids include H2O, H2, subject of numerous studies, both their isotopes and chemical composi-
CO2, CH4, H2S and N2 as well as, sometimes, NH3, CO, He and Ar tions (e.g., Árnason, 1977; Ármannsson et al., 1982; Arnórsson and
(e.g., Giggenbach, 1975; Arnórsson et al., 2007). They are considered Gunnlaugsson, 1985; Sano et al., 1985; Arnórsson, 1987; Arnórsson
to be derived from the atmosphere, mantle, and crust and often reflect et al., 1987, 2007; Marty et al., 1991; Poreda et al., 1992; Hilton et al.,
processes occurring within geothermal systems, such as fluid-rock and 1998; Stefánsson and Arnórsson, 2002; Füri et al., 2010; Stefánsson
fluid-fluid interaction and thermal decomposition of organic matter et al., 2011; Kaasalainen and Stefánsson, 2012; Barry et al., 2014; Scott
(e.g., Giggenbach, 1992; Taran et al., 1995; Chiodini and Marini, 1998; et al., 2014; Pope et al., 2015). Based on these studies it is evident that
Hilton et al., 2002; Stefánsson et al., 2017a). the vapors of Icelandic geothermal systems are dominated by water,
It has been suggested that local equilibria between geothermal min- in most cases N97%. From the δD and δ18O relationships, the source of
erals and gaseous species prevail in some geothermal systems the water is considered to be meteoric and/or seawater (e.g.,
(e.g., Giggenbach, 1980; Chiodini and Marini, 1998). Local mineral-gas Sveinbjörnsdóttir et al., 1986; Pope et al., 2015). The non-reactive vola-
and gas-gas equilibria are not always observed, however, as they are tiles in the source vapors, like Ar and N2, are considered to be dominant-
thought to be influenced by the flux of gases and/or slow reaction kinet- ly atmospheric, whereas recent data on δ15N isotopes have suggested a
ics, particularly for oxidation and reduction reactions (e.g., Stefánsson minor mantle contribution in some cases (Óskarsson et al., 2015). A
and Arnórsson, 2002; Lowenstern et al., 2015). The attainment of local deep-seated source for He has also been suggested (Poreda et al.,
equilibria involving gases constitutes the basis for gas geothermometry. 1992; Hilton et al., 1998; Füri et al., 2010). With respect to reactive ele-
ments like CO2, H2S, CH4, H2 and Cl, melt degassing and/or rock leaching
have been suggested as a source of these elements (e.g., Ármannsson
E-mail address: as@hi.is. et al., 1982; Arnórsson and Barnes, 1983; Poreda et al., 1992;

http://dx.doi.org/10.1016/j.jvolgeores.2017.04.002
0377-0273/© 2017 Elsevier B.V. All rights reserved.
82 A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94

Stefánsson and Arnórsson, 2002; Barry et al., 2014; Stefánsson et al., reservoir fluid to study their chemical characteristics. Based on such
2015, 2016a; Stefánsson and Barnes, 2016). Based on thermodynamics, well data, the primary or reservoir fluids have typically close to a neutral
it has been suggested that the major gas composition, i.e. CO2, H2S and pH value, with Na, Cl, S, C and Si being the dominant elements. The
H2, is controlled close to equilibrium with secondary minerals. In con- chemical composition of secondary fluids is much more variable. Pro-
trast, overall gas-gas equilibria are not always observed (Arnórsson cesses like depressurization boiling, condensation and mixing may
and Gunnlaugsson, 1985; Stefánsson and Arnórsson, 2002). Moreover, alter their composition from the reservoir to the surface. The boiled liq-
the ratio of magma gas to meteoric and/or seawater is also poorly uid that is typically discharged by hot springs is usually slightly enriched
known, as well as the effects of progressive fluid-rock interaction on in the non-volatile element concentrations and depleted in volatile con-
the volatile concentrations. centration, like CO2 and H2S, relative to the primary fluids. As the fluids
In this study, the available data on the vapor composition of have lost their main acids their pH is typically alkaline. In contrast, the
Icelandic thermal fluids were summarized, including well discharges vapor phase formed upon boiling that is typically discharged at the sur-
and fumaroles, of data reported in scientific papers, reports and unpub- face by fumaroles is enriched in the volatiles, including CO2, H2S, H2,
lished data. From these data, the reservoir volatile concentrations and CH4, N2 and Ar relative to the boiled water and primary fluids. The
mole fractions were used and compared with equilibrium mineral and vapor may also mix and condense with non-thermal water at shallow
gas reactions, as well as reaction modeling involving gas-fluid-rock in- depths and surface steam-heated water, typically observed as mudpots
teraction, in order to quantify the source(s) of the gases in the vapor and bubbling springs. This means that secondary fluids may have un-
and the effects of mineral-gas and gas-gas reactions on their chemistry. dergone various chemical and physical changes from the reservoir to
the surface, making their usage for assessment of reservoir chemical
2. Geothermal activity in Iceland characteristics more difficult compared with well fluids.

Geothermal activity in Iceland has been divided into high- and low- 3. Methods
temperature systems (Bödvarsson, 1961; Fridleifsson, 1979). The high-
temperature geothermal activity is commonly associated with active 3.1. Collection and analysis of well and fumarole discharges
volcanic complexes located along the rift, whereas low-temperature
systems are associated with off-rift activity on the North American The dataset used in the present study comes from various sources,
and Eurasian plates (Fig. 1). spanning samples collected over six decades (Líndal and
Icelandic thermal fluids are of meteoric and seawater origin or a Hermannsson, 1951; Ármannsson and Hauksson, 1980; Gíslason et al.,
mixture of them, with temperatures between ~ 10 to ~ 450 °C, pH of 1984; Ármannsson et al., 1985; Arnórsson and Gunnlaugsson, 1985;
b2 to N10 and Cl concentrations of b1 to N20,000 ppm (Kaasalainen Benjamínsson, 1985; Benjamínsson et al., 1985; Arnórsson, 1987;
and Stefánsson, 2012; Ármannsson, 2015; Kaasalainen et al., 2015; Arnórsson et al., 1987, 1996; Ólafsson, 1991; Poreda et al., 1992;
Stefánsson et al., 2016b). They have been divided into two groups: pri- Bjarnason, 1996, 2002; Bjarnason and Ólafsson, 2000; Gudmundsson
mary and secondary type fluids (Arnórsson et al., 2007). Primary fluids, and Arnórsson, 2002, 2005; Hjartarson and Ólafsson, 2005a, 2005b;
sometimes referred to as reservoir fluids, are those reaching the deepest Giroud, 2008; Benjamínsson and Hauksson, 2010, 2011; Stefánsson
level within the geothermal system. With ascent to the surface, they can et al., 2011, 2015; Kristinsson et al., 2013a, 2013b; Scott et al., 2014;
undergo chemical and physical changes leading to the formation of sec- Kaasalainen et al., 2015; Stefánsson and Barnes, 2016). Various methods
ondary geothermal fluids. were applied in collecting and analyzing the samples. However, in the
Deep wells, of up to N2500 m, have been drilled into many of the past ~40 years, a similar method has been applied in sample collection
geothermal systems in Iceland, thus allowing direct access to the and analysis for both fumarole gases and single and two-phase well

Fig. 1. Geology of Iceland and sampling locations. REY: Reykjanes, SVA: Svartsengi, KRÝ: Krýsuvík, HEL: Hellisheidi, HVER: Hveragerdi, NES: Nesjavellir, GEY: Geysir; SLL: Southern Low-
lands, TORF: Torfajökull, KERL: Kerlingarfjöll, HVERV: Hveravellir, VON: Vonarskard, KVER: Kverkfjöll, ASK: Askja, NÁM: Námafjall, KRA: Krafla and THEI: Theistareykir.
A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94 83

discharges. These have been described in detail by Arnórsson et al. one untreated and used for F, Cl and S2O3 determination and another
(2006) and are only briefly summarized here. to which 2% Zn acetate solution was added to precipitate dissolved sul-
For fumaroles, the common gases analyzed include H2O, CO2, H2S, fide as zinc sulfide, leaving dissolved SO4 in solution for analysis. All
H2, CH4, N2 and Ar, whereas He, CO, NH3 and light hydrocarbons anion analyses were carried out using ion chromatography with the ex-
(C2+) other than CH4 have not commonly been determined. Samples ception of CO2 in samples that were analyzed using the modified alka-
for SO2 determination have rarely been collected, because of the low linity titration previously mentioned (Stefánsson et al., 2007).
sampling temperatures of fumaroles (~100 °C) and the non-magmatic Dissolved sulfide [ΣS-II = HS− and H2S(aq)] were titrated in situ using
reservoir temperatures (b 350 °C) of geothermal systems in Iceland. In the method previously described (Arnórsson et al., 2006). The pH was
fact SO2 is rarely observed in the thermal fluids except during and analyzed in situ and in-line within a few seconds of sampling at
soon after volcanic eruptions (e.g., Ármannsson et al., 1982). ~20 °C using a flow-through pH cell.
In this study, the vapor, in each case, was collected by placing a plas- A typical sampling setup for fumarole gas collection and two-phase
tic funnel over or inserting a titanium tube into the vapor outlet and well discharges is shown in Fig. 2.
connecting a gas bottle using a siliconee tube. The gas bottle ranged in
volume from 50 to 250 mL was evacuated prior to sampling and 3.2. Calculation of reservoir fluid composition for well discharges
contained ~ 10 mL of 50% KOH per 100 mL. The condensable gases,
CO2 and H2S, were then analyzed using modified alkalinity titration The WATCH program (Bjarnason, 2010) was used to calculate
(Stefánsson et al., 2007) and Hg-acetate precipitation titration with the composition of geothermal reservoir fluids, aqueous speciation
dithizone as an indicator, respectively (Arnórsson et al., 2006). The and gas concentrations from data on two-phase well discharges. In the
non-condensable gases (H2, CH4, N2 and Ar) were analyzed using gas case of vapor only fumarole discharges, the analytical data were directly
chromatography and the H2O determined by gravimetry and mass used. These calculations of reservoir fluid composition from the
balance. chemical composition of vapor and liquid well discharges involve
For two-phase well discharges, the liquid and vapor phase were sep- solving together equations describing conservation of mass and
arated at the wellhead using a Webre separator. Vapor samples were enthalpy (Arnórsson et al., 2007). The conservation of mass may be
collected and analyzed in a similar manner as for fumarole vapor, as de- expressed by,
scribed previously. The liquid phase samples were cooled down using a  
stainless steel spiral connected to the Webre separator and then filtered mif ;t ¼ md;t d;v d;v
i ¼ X mi þ 1−X
d;v
md;lq
i ð1Þ
through a 0.2 μm cellulose acetate filter into polypropylene (for major
anion and cation analyses) or amber glass bottles (for CO2 analyses).
and the conservation of enthalpy by,
Samples for major cation analysis were acidified with 0.5 mL concen-
trated HNO3 (Suprapur®, Merck) per 100 mL sample and determined  
f ;t d;t d;v d;lq
h ¼h ¼ X d;v h þ 1−X d;v h ð2Þ
using ICP-OES. Two samples for major anion analysis were collected,

Fig. 2. Sampling methods for two-phase well discharges and fumaroles. (A–B) The fumarole samples were collected by placing a funnel over or inserting a titanium tube into the vapor
stream, connecting the outflow by silicone tube to a gas bottle. (C–D) For the two-phase well discharges the liquid and vapor phases were separated using a Webre separator and the
two phases sampled, the vapor phase collected in the same manner as for fumaroles. For further details see Arnórsson et al. (2006).
84 A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94

where mi denotes the molal concentration of the i-th component, X is pressures of gas species including H2O,
the vapor fraction, h is the enthalpy and f, t, d, v and lq are the fluid,
total, discharge, vapor and liquid, respectively. The vapor fraction can
be calculated from, P total ¼ ∑ P i ð5Þ
i

d;t d;lq
h −h
X d;v ¼ d
ð3Þ Vapor in geothermal systems is usually dominated by H2O, and has a
L relatively low gas partial pressure. Under these conditions it is reason-
able to assume the fugacity coefficients to be unity, i.e. Γi = 1 and that
where Ld is the latent heat of vaporization. The values for hd,t can fi = Pi and that,
either be obtained from the measured liquid to vapor fraction of
discharge fluids, here referred to as measured enthalpy (meas. h) or
evaluated from the reservoir temperature assuming a liquid only reser- xvapor
i ¼ P i =P total ¼ nvapor
i =nvapor
total
ð6Þ
voir, i.e. h f,t = hd,t = hTreslq where hTreslq is the enthalpy of liquid at the res-
ervoir temperature. The values for hd,t, hd,v and Ld can be obtained from where nvapor is the number of moles in the vapor phase including H2O.
i
the properties of liquid and vapor water (steam tables). In a two-phase liquid and vapor system (lq + v), the equilibrium sol-
Here, the two approaches were used to calculate the reservoir fluid ubility constant (Kh,i) of gas in the liquid phase is given by,
composition and vapor fraction at sampling and in the reservoir, i.e.
model 1 assumes a liquid only reservoir and hf , t = hd , t = hTreslq and
model 2 takes the measured discharge vapor fraction to calculate the K h;i ¼ ai = f i ¼ mi γi =P i Γ i ð7Þ
reservoir vapor fraction when hf,t = hd,t N hTreslq.
The reservoir temperatures for two-phase well discharges were cal-
where ai, mi and γi are the activity, molal concentration and activity co-
culated assuming quartz equilibrium (Gunnarsson and Arnórsson,
efficient of the i-th aqueous species, respectively.
2000). With respect to fumaroles, the reservoir temperature was calcu-
For vapor, the values of Pi, Ptotal, xi, mi and γi for the liquid and vapor
lated based on gas geothermometry involving CO2, H2S and H2
phases were calculated with the aid of the WATCH program (Bjarnason,
(Arnórsson and Gunnlaugsson, 1985). It should be pointed out that
2010). For these calculations the thermodynamic database of Arnórsson
equilibrium my not have been attained with respect to individual nor
et al. (1982) was used with updates on gas solubilities (Fernandez-Prini
gas ratios, i.e. the gas geothermometry temperatures need to be consid-
et al., 2003). The reaction equilibrium constants involving minerals and
ered with care.
gases were calculated with the aid of the Supcrt92 program and the
slop07.dat database (Johnson et al., 1992; http://geopig.asu.edu/tools).
3.3. Thermodynamic relations
The mineral-gas and gas-gas equilibrium constants used in the study
are summarized in Table 1. The reactions involve commonly observed
The equilibrium constant (Ki) of reactions involving gases can be de-
secondary minerals in Icelandic geothermal systems at N200 °C
scribed by the general equation,
including quartz, calcite, pyrite, pyrrhotite, magnetite, epidote,
clinozoesite, chlorite, prehnite, grossular, wollastonite and anhydrite
ν
K ¼ ∏ f i i ¼ ∏ xνi i Γ νi i P total ð4Þ (e.g., Kristmannsdóttir, 1978; Lonker et al., 1993). For the calculations
i i
the activities of minerals was assumed to be unity except for prehnite,
clinozoesite and epidote where the activity was taken to be equal to
where fi, xi and Γi are the gas fugacities, mole fractions and fugacity co- the mole fraction of the mineral in the solid solution (X = a) commonly
efficients for the i-th gas species, respectively, and vi are the reaction observed for these minerals of Xpre = 0.5, Xczo = 0.3 and Xepi = 0.7, re-
stoichiometric coefficients, positive for products and negative for reac- spectively (Sveinbjörnsdóttir, 1992; Lonker et al., 1993; Stefánsson and
tants. The total fluid pressure, Ptotal, is defined as the sum of all partial Arnórsson, 2002)

Table 1
Chemical reactions and equilibrium constants (logKi). Calculated based on Supcrt92, slop07.dat.a,b

Reactions Temperature (°C)

25 100 150 200 250 300 350

Gas-gas reactionsc
CO2(g) + 4H2(g) = CH4(g) + 2H2O(g) 19.73 14.04 11.23 8.98 7.14 5.59 4.29
pyr + H2(g) = pyrr + H2S(g) −4.54 −2.76 −1.91 −1.22 −0.65 −0.17 0.23
3pyr + 2H2(g) + 4H2O(g) = mt + 6H2S(g) −30.99 −22.93 −18.98 −15.87 −13.35 −11.26 −9.49

Gas-mineral reactions
czo + cc + 3/2qtz + H2O(g) = 3/2pre + CO2(g) −0.89 −0.95 −0.93 −0.90 −0.88 −0.86 −0.83
2/5czo + cc + 3/5qtz = 3/5gro + 1/5H2O(g) + CO2(aq) −6.91 −3.74 −2.27 −1.12 −0.19 0.58 1.22
2/3gro + 1/3pyr + 1/3pyrr + 2/3qtz + 4/3H2O(g) = 2/3epi + 2/3wo + H2S(g) −5.34 −5.23 −4.99 −4.73 −4.48 −4.21 −3.92
2gro + 1/2pyr + 1/2mt + 2qtz + 2H2O(g) = 2epi + 2wo + H2S(g) −2.28 −2.69 −2.77 −2.83 −2.88 −2.92 −2.96
1/4pyr + 1/2pyrr + H2O(g) = 1/4mt + H2S(aq) −5.40 −4.35 −3.79 −3.36 −3.01 −2.73 −2.49
6pyr + 1.5pre + 10H2O(g) = 2mt + 1.5qtz + czo + anh + 11H2S(g) −84.03 −63.66 −53.63 −45.72 −39.31 −33.99 −29.49
4/3pyrr + 2/3pre + 2/3H2O(g) = 2/3epi + 2/3pyr + H2(g) −1.74 −1.57 −1.49 −1.46 −1.46 −1.45 −1.46
2/3gro + 4/3pyrr + 2/3qtz + 4/3H2O(g) = 2/3epi + 2/3wo + 2/3pyr + H2(g) 0.18 −0.99 −1.49 −1.93 −2.28 −2.59 −2.85
6gro + 2mt + 6qtz + 4H2O(g) = 6epi + 6wo + H2(g) 8.67 3.41 1.17 −0.57 −1.96 −3.12 −4.12
3/2pyrr + H2O(g) = 3/4pyr + 1/4mt + H2(g) −0.94 −1.59 −1.87 −2.14 −2.37 −2.56 −2.72
1.5pyr + 4.5pre + 8H2O(g) = 0.5mt + 4.5qtz + 3czo + 3anh + 11H2(g) −81.65 −64.83 −56.50 −49.90 −44.52 −40.05 −36.27
a
geopig.asu.edu/tools.
b
pyr: pyrite; pyrr: pyrrhotite; mt: magnetite; czo: clinozoesite; cc: calcite; qtz: quartz; pre: prehnite; gro: grossular; epi: epidote; wo: wollastonite; anh: anhydrite.
c
apre = 0.5; aczo: 0.3 and aepi:0.7.
A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94 85

3.4. Geochemical simulation of gas-fluid-rock interaction i.e. Geysir, Torfajökull, Kerlingafjöll and Askja, whereas the others are
dominated by basaltic rocks.
To gain further insight into gas-fluid-rock interaction and the effects The well and fumarole vapor was dominated by H2O, accounting for
of gas (acid) supply source and reactions of C, S and H to the system, re- 92.2 to 99.9 mol% and 62.3–100 mol%, respectively (Fig. 3). Carbon diox-
action path simulations were conducted using the PHREEQC program ide, H2S and H2 were the dominant gases with concentration ranges
(Parkhurst and Appelo, 1999) and the llnl.dat database. The calculations (median in parenthesis) in well and fumarole vapor discharges of
involve mixing non-thermal water with volcanic gas of different ratios 0.01–37.8 (0.37) mol%, 7·10− 5-0.30 (0.06) mol% and 7·10−5-0.68
(0.1–10%), and reacting the liquid solution with basaltic rock in steps, (0.05) mol% followed by N2, CH4, and Ar with concentrations, typically
allowing secondary minerals to precipitate when saturated. The calcula- b0.03 mol%, b0.002 and b 0.001 mol%, respectively.
tions were conducted at 275 °C and water vapor saturation pressure Considerable differences in gas concentrations were observed be-
(Psat). The secondary minerals included in the calculations were those tween different geothermal systems. Elevated CO2 concentrations
commonly associated with alteration of basalts at N200 °C and included were observed in some samples from Krafla, Askja, Krýsuvík and
quartz, calcite, pyrite, pyrrhotite, magnetite, epidote, wollastonite, gros- Torfajökull, whereas vapor in Námafjall was high in H2 and CH4 relative
sular, clionzoestie, chlorite, prehnite, wairakite, albite, microcline and to fluids from most other geothermal systems. The concentrations of
anhydrite (e.g., Kristmannsdóttir, 1978; Lonker et al., 1993). The results H2S and H2 were generally lower for geothermal systems with a seawa-
of the reaction path simulations were then compared with the observed ter source fluid than for systems with meteoric source fluids. Moreover,
fluid composition in order to evaluate the role of volcanic gas input and fluids having low temperatures such as in the Southern Lowlands and at
progressive mineral-fluid reactions in the gas chemistry. The input con- Geysir and Hveravellir had lower gas concentrations in general com-
ditions for the modeling, i.e. non-thermal water, basalt and volcanic gas pared to geothermal systems having higher reservoir temperatures.
composition, are listed in Table 2. The basalt is the average Icelandic ba- These observations suggest that multiple factors may affect gas concen-
salt composition accounting for N80% of the rock type in Iceland trations in thermal fluids in Iceland, including the sources of gases, rock
(Kaasalainen and Stefánsson, 2012), the non-thermal water is repre- and melt types and reactions within the geothermal systems and upon
sented by typical surface water in Iceland and has been previously fluid ascent to surface.
used for modeling (Kaasalainen and Stefánsson, 2012), and the volcanic
gas is based on direct measurements of magma gases in Iceland 4.2. Reservoir fluids
(Sigvaldason and Elísson, 1968; Gíslason et al., 2015; Gauthier et al.,
2016; Stefánsson et al., 2017b). The reservoir fluid composition was calculated based on the liquid
and vapor composition of the two-phase well discharges. For these cal-
4. Gas concentrations of thermal fluids in Iceland culations, two approaches were used. Firstly, the geothermal reservoir
was assumed to be liquid only and the total fluid enthalpy calculated
4.1. Fumaroles and well fluid discharges based on the temperature and the liquid enthalpy of the reservoir.
This model is referred to as model 1 and calculated enthalpy
The dataset compiled in the present study includes well fluid dis- (i.e., model 1, calc. h). Secondly, the vapor fraction of the well discharge
charges (n = 275) and fumarole vapor (n = 587) from over 15 geother- and reservoir were calculated based on the measured enthalpy of the
mal systems in Iceland (Fig. 1). The geothermal systems include fluid discharge and the reservoir temperature. This model is referred
Reykjanes, Svartsengi, Krýsuvík, Hengill including Hellisheidi, to as model 2 and measured enthalpy (i.e., model 2, meas. h). The results
Nesjavellir and Hveragerdi, Geysir, Southern Lowlands, Torfajökull, of the mole fractions of CO2, H2S and H2 as a function of reservoir tem-
Kerlingafjöll, Hveravellir, Vonarskard, Askja, Kverkfjöll, Námafjall, Krafla perature have been plotted in Fig. 4.
and Theisareykir. All systems are located within the active volcanic zone The resulting reservoir vapor composition calculated based on well
of Iceland with the exception of the systems on the Southern Lowlands. liquid and vapor discharges was highly variable depending on the ap-
Representative fluid composition for these geothermal systems are proach used to calculate the reservoir fluid composition. Assuming
given in Table 3 (the full ThermoFluid dataset, Stefánsson et al., 2016c, only liquid in the geothermal reservoir (model 1, calc. h), the mole frac-
can be obtained from Ríkey Kjartansdóttir, rikey@hi.is). The source tions of gases for dilute fluids (b 500 ppm Cl), including CO2, H2S and H2,
fluid for most of the geothermal systems is considered to be meteoric were generally higher compared to when assuming two-phase liquid
water, whereas at Reykjanes and Svartsengi the source of the fluids is and vapor reservoirs. It should be noted, however, that in the former
both seawater and seawater-meteoric water (Arnórsson, 1995). Four case the mole fraction of gases in the vapor phase was purely a thermo-
geothermal systems are associated with major silicic rock formations, dynamic convention as there was no vapor phase present.

Table 2 5. Mineral-gas and gas-gas equilibria


Chemical compostion of basalt, non-thermal water and volcanic gas used in the geochem-
ical model calculations.
5.1. Mineral-gas equilibria
Basalta,b Non-thermal water a Volcanic gas c

SiO2 wt% 50.0 pH 7.45 H2O mol% 91.6 Mineral-gas equilibria have been suggested as a major factor con-
Al2O3 wt% 14.9 SiO2 ppm 11.5 CO2 mol% 4.95 trolling CO2, H2S and H2 concentrations in thermal fluids in Iceland
TiO2 wt% 1.52 Na ppm 7.79 SO2 mol% 3.22 (Arnórsson and Gunnlaugsson, 1985; Stefánsson and Arnórsson,
FeO wt% 10.8 K ppm 0.47 HCl mol% 0.23 2002), based on the composition of reservoir fluid composition and
MnO wt% 0.19 Ca ppm 3.89 HF mol% 0.05
mineral-gas equilibrium values. However, based on the comparison of
MgO wt% 7.86 Mg ppm 1.41
CaO wt% 12.3 CO2 ppm 19.3 the observed and calculated mole fractions (Fig. 4) it cannot be conclud-
Na2O wt% 2.14 Cl ppm 3.37 ed whether mineral-gas reactions controls CO2, H2S and H2 concentra-
K2O wt% 0.2 F ppm 0.14 tions in vapors in Icelandic geothermal systems; the scatter of data
S ppm 300 SO4 ppm 4.8
and equilibrium values is simply too great. Moreover, the uncertainties
Cl ppm 220
CO2 ppm 145b about the calculated reservoir fluid composition exceed in some cases
a
the range of equilibrium values calculated for the proposed mineral-
Kaasalainen and Stefánsson (2012) and references therein.
b
Barry et al. (2014).
gas reactions as well as the changes on boiling.
c
Based on Sigvaldason and Elísson (1968), Gíslason et al. (2015), Gauthier et al. (2016) Looking at the gas concentrations for a given geothermal system
and Stefánsson et al. (2017b). sometimes reveals a small range, for example at Hengill, whereas in
86
Table 3
Selected discharge well and fumarole gas compostion.a,b

Sample # Location/well # Type c Tsampling model#1 (calc. h) d,e


model #2 (meas. h) d,e
Liquid phase (ppm) Vapor phase (μmol/mol)

Tres Xsampling Xres Tres Xsampling Xres Cl CO2 H2S H2O CO2 H2S H2 CH4 N2 Ar

Askja
1983–0001 f 311 971,089 26,537 1787 474 0.75 110 2.02
1983–0013 f 312 971,948 26,204 908 580 1.05 350 8.12
1983–0014 f 311 971,544 26,820 1012 471 149 2.77

Geysir
1982–0005 f 226 997,628 2335 17.1 0.027 18.6 0.34
1982–0006 f 248 995,077 2615 5.07 8.18 2248 47.2
1982–0007 f 256 997,232 2615 23.9 1.23 1.18 122 4.80
1982–0008 f 246 996,036 3434 6.74 0.326 513 10.7
1982–0097 f 263 996,645 1602 7.59 4.12 5.64 1703 33.5
1982–0098 f 259 996,448 1788 6.51 2.28 0.81 1706 48.5

A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94


Hengill including Hellisheidi, Nesjavellir and Hveragerdi
08–3001 HE-07 w 171 263 0.20 0.00 261 0.31 0.14 199 7.4 76.1 998,729 536 436 234 4.45 59.5 0.97
08–3006 HE-05 w 173 268 0.21 0.00 268 0.22 0.01 77 33.1 37.7 998,242 1460 112 11.0 11.9 160 2.67
08–3007 HE-06 w 176 261 0.18 0.00 259 0.39 0.25 107 34.5 56.9 996,559 2808 348 251 3.91 29.7 0.52
09–5199 HE-17 w 188 260 0.16 0.00 284 0.76 0.70 216 15.4 68.7 997,532 1144 842 458 2.78 21.3 0.33
10–5168 HE-12 w 179 281 0.23 0.00 279 0.48 0.33 193 22.1 73.4 998,127 1027 495 331 4.10 16.6 0.34
10–5169 HE-07 w 179 262 0.18 0.00 260 0.29 0.14 185 16.4 73.6 998,552 671 400 234 3.39 138 1.69
10–5171 HE-42 w 179 292 0.26 0.00 301 0.85 0.81 169 13.6 35.0 998,782 310 546 348 1.19 12.5 0.33
1982–0017 f 294 984,611 13,401 1191 507 56.2 222 11.7
1982–0023 f 284 996,482 2931 52 42.3 474 18.0
1982–0025 f 300 993,097 4704 911 1201 11.2 73.5 2.38
2014–0010 f 293 994,512 4945 389 96.8 2.64 53.0 0.74
2014–0011 f 294 993,626 5849 369 103 3.25 50.2 0.74
2014–0012 f 266 998,574 1310 22.2 6.58 0.39 85.7 1.55
81–3017 HV-7 w 186 235 0.10 0.00 186 43.4 23.4 998,260 1260 109 21.1 2.48 331 16.3
81–3016 HV-6 w 178 232 0.11 0.00 171 28.6 23.4 998,534 1018 48.4 23.0 1.56 370 5.08
79–0136 G-2 w 147 191 0.07 0.00 142 44.2 18.4 999,633 338 16.1 11.6 1.20
78–0148 NJ-05 w 210 260 0.12 0.00 259 0.15 0.03 8 89.9 110.0 989,377 8328 850 977 85.0 382
04–3039 NJ-10 w 197 279 0.19 0.00 279 0.19 0.00 122 25.7 50.2 998,698 752 334 28.9 29.1 156 2.71
04–3048 NJ-11 w 168 281 0.25 0.00 278 0.59 0.46 78 15.1 111.7 997,473 590 898 921 12.3 105 1.98
04–3037 NJ-16 w 198 276 0.18 0.00 276 0.20 0.02 65 37.5 86.1 997,258 891 803 931 17.9 98.5 1.46
04–3050 NJ-20 w 200 274 0.17 0.00 273 0.27 0.13 169 32.9 49.9 998,061 1307 383 222 5.9 20.8 0.22
04–3046 NJ-21 w 200 286 0.21 0.00 290 0.80 0.75 111 51.2 83.8 995,713 2523 924 750 7.49 81.7 1.50

Hveravellir
1982–0101 f 258 998,841 1051 36.1 2.76 1.37 67.0 1.19
1982–0102 f 236 998,168 800 31.8 0.18 0.31 981 18.0
1982–0103 f 248 998,963 831 35.2 0.87 0.70 167 2.13
2004–0019 f 237 999,770 164 13.2 1.05 0.68 51.0 0.81
2004–0020 f 260 999,069 721 117 4.08 2.72 84.2 1.70

Kerlingarfjöll
1951–0002 f 284 997,780 1365 225 623 7.50
1982–0119 f 287 996,065 2880 393 433 1.05 225 3.40
2004–0006 f 294 996,075 2604 544 747 9.42 21.2 0.42
2004–0013 f 288 990,422 8298 806 452 4.10 17.2 0.33
2014–0044 f 287 996,839 2157 635 334 2.55 31.3 0.37
2014–0057 f 274 996,242 3331 215 165 8.65 37.8 0.64

Krafla
04–3010 K-13 w 198 234 0.08 0.00 230 0.56 0.53 32 67.9 72.6 995,840 2851 215 850 2.12 237 4.01
04–3015 K-15 w 195 277 0.19 0.00 275 0.33 0.18 36 16.0 93.2 998,455 1047 227 211 0.99 58 1.30
11-KRA-04 K-32 w 176 246 0.15 0.00 241 0.35 0.24 42 59.8 103.2 998,299 1028 424 228 1.43 19.5 0.41
11-KRA-08 K-24 w 137 213 0.14 0.00 213 0.14 0.00 44 45.7 28.4 999,116 792 23.1 4.37 5.08 57.6 1.32
77–1206 K-09 w 206 287 0.20 0.00 286 0.22 0.04 21 245.0 32.2 986,815 12,840 260 36.6 48.8
81–3036 K-16 w 165 259 0.20 0.00 256 0.43 0.29 65 158.5 53.7 993,278 5723 448 349 11.1 188 3.59
88–3245 K-12 w 143 305 0.35 0.00 304 0.86 0.79 110 162.8 44.2 991,890 6987 468 282 3.31 365 4.39
98–3203 K-20 w 189 284 0.22 0.00 285 0.79 0.73 155 204.6 53.5 982,690 15,924 715 625 8.09 36.3 0.77
98–3204 K-17 w 215 270 0.13 0.00 268 0.51 0.44 17 47.1 98.2 995,600 3331 478 538 14.5 38.7 0.91
1979–0001 f 298 835,694 161,051 1707 1515 32.9
1979–0013 f 287 995,296 3834 369 498 2.73
1983–0019 f 300 959,289 39,139 1056 374 0.48 138 2.21
1985–0024 f 279 950,064 49,606 91.9 102 7.22 123 5.17
1995–0018 f 298 972,880 20,460 45.6 6.11 6534 75.2
1995–0019 f 285 989,072 10,094 354 443 3.07 33.4 0.75
2009–0003 f 299 986,402 12,128 885 504 9.91 69.9 1.85
2011–0002 f 281 996,520 3057 190 182 4.90 42.8 1.89
2011–0005 f 314 914,586 81,134 2541 1224 60.8 445 8.92

Krýsuvík
1950–0001 f 281.0040960385 995,882 3538 294 279 7.53

A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94


1981–0008 f 273 994,442 5009 266 136 4.03 138 4.77
1981–0009 f 263 991,650 7822 250 61.2 7.11 204 5.94
1981–0033 f 266 993,864 5135 712 139 1.60 146 1.74
1984–0021 f 254 995,333 4301 297 15.0 1.23 50.6 2.32
1986–0003 f 285 971,349 20,825 1771 169 5792 93.21
2014–0005 f 274 996,323 3306 124 175 1.39 70.3 1.14

Kverkfjöll
1983–0008 f 277 992,964 5995 322 607 6.00 105
1983–0009 f 277 995,312 3658 378 610 2.91 39.2 0.61
1983–0015 f 277 992,155 6362 436 588 1.58 452 5.89
1983–0016 f 277 993,278 5635 450 593 2.80 40.7 0.98
1983–0017 f 278 994,455 3830 478 611 2.34 617 5.82
1983–0018 f 278 995,344 3509 472 614 4.00 55.9 1.72

Námafjall
79–3052 N-08 w 177 233 0.11 0.00 0.16 0.05 26 88.2 132.6 995,646 1703 787 1730 134
81–3022 N-11 w 214 236 0.05 0.00 0.76 0.75 31 10.9 97.8 994,732 1021 692 1439 2.84 2068 45.5
81–3023 N-12 w 216 241 0.05 0.00 0.77 0.76 35 18.1 136.1 995,489 1072 745 1572 5.12 1099 17.5
98–3215 N-04 w 143 236 0.18 0.00 0.19 0.02 40 11.4 10.6 997,134 559 227 17.1 2038 24.2
1950–0012 f 308 989,854 6837 2279 1007 22.7
1982–0109 f 260 998,257 1008 479 162 17.8 74.0 2.11
1982–0114 f 278 996,565 2013 540 619 3.72 255 4.21
1995–0024 f 317 987,584 9951 938 1486 15.1 25.3 1.17
1995–0025 f 315 993,787 3341 914 1910 12.5 34.8 1.03
1996–0016 f 303 993,544 3788 1077 1543 8.29 39.5 0.78
2012–0011 f 307 996,099 992 1125 1744 9.63 30.7 0.54
2013–0010 f 255 999,436 261 180 57 24.3 39.3 1.10
2013–0013 f 309 996,065 2079 1171 635 23.9 24.8 1.16

Reykjanes
1971–0083 RN-08 w 163 280 0.25 0.00 25,700 996,697 3182 100 22.0
1987–0005 RN-08 w 201 276 0.18 0.00 22,950 54.5 1.1 997,822 2079 79.1 2.27 0.27 17.5
1983–0225 RN-09 w 248 298 0.14 0.00 20,292 110.6 6.4 994,417 5310 188 29.3 0.98 55.6
2001–0461 RN-09 w 225 296 0.19 0.00 22,900 29.8 1.9 997,137 2729 119 3.14 11.6
05–3004 RN-12 w 252 285 0.09 0.00 19,288 103.9 11.5 991,437 8148 268 15.9 1.13 128 1.99
05–3003 RN-15 w 223 277 0.14 0.00 19,251 47.4 3.4 992,556 7094 188 15.8 1.22 143 2.22
1950–0002 f 255 997,641 2284 45.1 30.1
1982–0068 f 277 998,317 1604 19.9 4.79 0.13 48.8 5.83
1982–0070 f 292 985,661 13,388 196 63.3 5.60 648 37.5

Southern Lowlands

(continued on next page)

87
88
Table 3 (continued)

Sample # Location/well # Type c Tsampling model#1 (calc. h) d,e


model #2 (meas. h) d,e
Liquid phase (ppm) Vapor phase (μmol/mol)

Tres Xsampling Xres Tres Xsampling Xres Cl CO2 H2S H2O CO2 H2S H2 CH4 N2 Ar

79–3028 Reykholt, w1 w 121 167 0.07 0.00 78 30.4 1.0 999,729 256 5.68 5.68 4.26
79–3030 Reykjaból, w1 w 143 191 0.08 0.00 42 54.3 4.1 999,017 946 17.1 11.4 8.55
88–3002 Efri-Reykir, w23 w 129 172 0.06 0.00 54 30.1 4.8 998,207 796 26.9 0.72 7.27 945 16.8
88–3204 Núpar w 113 172 0.09 0.00 356 49.7 1.7 997,572 1530 3.61 1.08 9.32 870 14.3
88–3206 Bakki w 102 142 0.04 0.00 669 5.9 0.2 995,235 701 0.090 4.34 4.30 4005 49.5
88–3207 Bakki, w2 w 107 136 0.02 0.00 679 6.0 0.3 996,000 529 1.26 1.45 15.3 3392 60.0
88–3208 Bakki, w1 w 111 143 0.03 0.00 811 5.7 0.2 997,073 589 5.06 1.44 16.8 2277 39.0
88–3211 Kotlaugar, w1 w 102 162 0.08 0.00 41 50.8 3.4 994,630 1027 0.18 4.88 5.44 4279 53.5
88–3237 Klausturhólar w 113 193 0.13 0.00 120 235.0 3.5 996,388 3123 25.0 0.36 5.54 452 6.44

Svartsengi
04–3003 SV-08 w 193 243 0.11 0.00 13,493 16.9 1.2 998,067 1689 3.43 0.69 237 2.89
04–3004 SV-18 w 193 240 0.10 0.00 14,255 12.3 1.3 998,597 1336 34.3 1.62 0.20 30.7 0.63

A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94


04–3006 SV-19 w 195 239 0.09 0.00 14,235 8.6 0.8 998,973 963 21.6 2.16 0.22 39.1 0.65
04–3029 SV-11 w 208 241 0.07 0.00 12,916 46.6 3.8 993,581 5720 61.8 66.1 1.60 562 7.45
04–3032 SV-09 w 196 244 0.10 0.00 13,230 6.9 1.2 998,892 1018 25.3 3.97 0.18 60.1 0.96
79–3008 SV-04 w 141 242 0.20 0.00 17,010 32.6 999,504 484 3.96 0.50 0.50 6.44
95–0394 SV-12 w 191 239 0.10 0.00 13,379 38.7 0.3 998,731 1237 22.4 1.39 0.13 8.29
95–0396 SV-11 w 200 246 0.10 0.00 13,583 47.9 1.1 997,311 2600 53.0 7.40 0.21 28.7

Theistareykir
1950–0013 f 269 995,769 3440 482 301 7.53
1982–0099 f 293 987,205 10,277 1210 703 169 423 11.9
1985–0017 f 255 996,643 2783 456 31.5 5.32 79.8 1.59
1997–0022 f 237 993,774 6070 23.4 3.63 0.42 126 1.51
2012–0002 f 289 994,553 3076 995 1292 17.1 66.6 0.96
2012–0004 f 266 998,269 1023 428 265 1.15 13.2 0.22

Torfajökull
1950–0016 f 254 997,588 2194 90.2 128
1982–0078 f 267 996,972 2493 206 299 6.31 20.7 1.74
1982–0092 f 326 972,070 22,995 1022 3759 17.7 131 6.65
1994–0001 f 268 994,943 4515 232 306 2.45 0.98
1997–0001 f 324 806,391 191,621 885 1101 0.83 1.31
1997–0002 f 247 999,052 788 81.7 77.2 1.12 0.072 0.090
2014–0062 f 271 994,548 4788 141 474 7.29 41.9 0.51

Vonarskard and Köldukvíslarbotnar


1984–0018 f 251 994,509 4468 96.0 96.5 49.5 767 14.5
1985–0013 f 275 987,158 7313 595 575 29.7 4183 146
2012–0016 f 280 993,585 5006 552 706 4.77 144 2.46
2012–0017 f 270 996,888 2253 398 331 3.10 126 2.13
a
Based on: Líndal and Hermannsson (1951), Ármannsson and Hauksson (1980), Gíslason et al. (1984), Ármannsson et al. (1985), Arnórsson and Gunnlaugsson (1985), Benjamínsson (1985), Benjamínsson et al. (1985), Arnórsson (1987)
Arnórsson et al. (1987, 1996), Ólafsson (1991), Poreda et al. (1992), Bjarnason (1996, 2002), Bjarnason and Ólafsson (2000), Gudmundsson and Arnórsson (2002, 2005), Hjartarson and Ólafsson (2005a, 2005b), Giroud (2008), Benjamínsson and
Hauksson (2010, 2011), Stefánsson et al. (2011, 2015), Kristinsson et al. (2013a, 2013b), Scott et al. (2014), Kaasalainen et al. (2015), Stefánsson and Barnes (2016).
b
For full ThermoFluid dataset (Stefánsson et al., 2016c), contact Ríkey Kjartansdóttir (rikey@hi.is).
c
w: two-phase well fluids; f: fumarole vapor.
d
Calculated with the aid of the WATCH program (Bjarnason, 2010); model#1 referred to liquid only reservoir and calculated enthalpy, model#2 refers to two-phase reservoir and measured discharge enthalpy.
e
The reservoir temperature for two-phase well discharges referes to quartz geothermometry (Gunnarsson and Arnórsson, 2000) and for fumaroles gas geothermometry (Arnórsson and Gunnlaugsson, 1985).
A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94 89

Fig. 3. The relationship between H2O, CO2 and S mole ratios in thermal fluids in Iceland. Also shown are the trends with adiabatic boiling of typical reservoir fluids of meteoric origin (typ-
ical Krafla fluid: KRA) and seawater origin (typical Reykjanes fluid: REY) from the reservoir temperature to surface at 100 °C. Saline and dilute reservoir thermal fluids have similar CO2
concentration whereas the latter are enriched in H2S, resulting in higher CO2/H2S ratio for saline thermal fluids relative to dilute thermal fluids. Upon boiling the gas phase becomes in-
creasingly enriched in H2O as observed by the modelled boiling trend and the fumarole vapor samples.

other cases a large range is observed, for example for CO2 concentra- equilibrium were attained or not. Such an approach is simply insensi-
tions at Torfajökull and Krafla. Given often a relatively narrow range in tive. Large uncertainties are related to calculation of reservoir fluid com-
temperature (~ 50 °C) for a given geothermal system and a generally position from two-phase well discharges. Moreover, data on fumarole
moderate slope with temperature of equilibrium gas concentrations vapor composition do not allow for reconstruction of total reservoir
predicted by the mineral-gas reactions, the former case may indicate a fluid properties, and distinguishing between reservoir characteristics
mineral-gas equilibrium. The latter case, however, is more characteristic from later stage reactions may be difficult. Finally, uncertainties are re-
of a source-controlled system, i.e. the gas concentrations reflect the var- lated in calculating equilibrium constants at elevated temperatures.
iable input of gas to the system and/or are affected by progressive reac-
tion that may influence the gas concentrations rather than the mineral- 6. Sources and reactions of gases in geothermal systems
gas equilibria.
The sources of the volatile gases can be many, including air, rock dis-
5.2. Gas-gas equilibria solution, organic matter decay and/or magma degassing. Nitrogen and
Ar are among the major gases in thermal fluids. Their source(s) have
Instead of looking at individual gas concentrations it is common to been obtained through the respective isotope ratios. The reported Ar
look at gas ratios reflecting gas-gas interactions (e.g., Giggenbach, isotope ratios (40Ar/36Ar) in thermal fluids in Iceland range from 295
1980, 1987; Chiodini and Marini, 1998). Gas-gas reactions involving to 313, i.e. close to the air ratio of 295 (Sano et al., 1985; Óskarsson
CO2, H2S, H2 and CH4 are shown as a function of reservoir temperature et al., 2015) suggesting air as the major source of Ar in thermal fluids
in Fig. 5, together with the calculated equilibrium values (Table 1). A in Iceland. For N2, reported isotope ratios (δ15N) range from −10.5 to
large scatter of data was observed in all cases. The agreement between + 3.0‰ (Sano et al., 1985; Marty et al., 1991; Óskarsson et al., 2015),
the measured and calculated equilibrium ratios were close, particularly considerably lower than air (0‰) but more similar to depleted MORB
for the reactions involving H2S and H2 (Fig. 5B and C), whereas the over- mantle (DMM) of –5 ± 2‰ and the range found in Icelandic basalts
all equilibrium values predicted between CO2, CH4 and H2 (Fig. 5A) (− 2.3 to + 5.7‰) (Halldórsson et al., 2016b). This suggest multiple
were ~ 5 log units offset relative to most data on vapor composition. sources of N2 in some thermal fluids, ranging from air to mantle N2,
Moreover, boiling may have considerable effects on gas concentrations, with the range of values reflecting heterogeneity of the Iceland plume
making it difficult to distinguish between reservoir conditions and sec- source (Halldórsson et al., 2016a).
ondary processes like boiling with fluid ascent. Ternary diagrams of N2-Ar-CO2-H2S relationships are shown in
Previously, Stefánsson and Arnórsson (2002) concluded that such Fig. 6. Many of the thermal gas samples indicate close to air saturated
gas-gas equilibria may not prevail in Icelandic geothermal systems N2/Ar ratio with a trend towards enrichment of CO2 and H2S. This is
based on comparing equilibrium and measured redox values (Eh) for taken to indicate air as the major source of N2 and Ar in these gases, to-
different half-reactions involving thermal gases. Such a comparison gether with the addition of CO2 and H2S. Inspection of the data in Fig. 6
does not imply that local equilibrium between some gases and/or gas- suggests systematic enrichment of N2 relative to Ar and air-saturated
reactions is not attained, simply that overall gas equilibria were not water. Either or both may indicate additional source(s) for N2, consis-
observed. tent with the δ15N values in some of these fluids and/or air contamina-
Based on the observations of individual gas concentrations and gas- tion during sampling. The latter is, however, considered highly unlikely
gas ratios, it cannot be distinguished what processes control the gas in the case of two-phase well discharges due to the high-discharge flow
concentrations in the vapor in Icelandic geothermal systems and if gas rates and general lack of measurable O2 in such samples.
90 A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94

0
A czo+cc+qtz+pre
czo+cc+qtz+gro

-1 boiling (dil. meas. h)


2
logx vCO

-2

-3

boiling (saline and


dil., calc. h)
-4

-2 B
boiling (dil. meas. h)
boiling (dil. calc. h)

-3
H2S
v
logx

-4

boiling (saline. calc. h)


-5 gro+pyr+pyrr+qtz+epi+wo
pyr+pre+mt+qtz+czo+anh

pyr+pyrr+mt
-6
C gro+pyr+pyrr+qtz+epi+wo
py+pyrr+pre+epi pyr+pyrr+mt
-1

-2
H2
logx v

boiling (dil. calc. h)


-3 boiling (dil. meas. h)

-4

-5
pyr+pre+mt+qtz+czo+anh
boiling (saline. calc. h)
-6
100 150 200 250 300 350
Temperature (°C)

Fig. 4. The mole fractions (xi) of (A) CO2, (B) H2S and (C) H2 as a function of temperature.
Shown are the mole fractions of the various samples in reservoir fluids calculated from the
two-phase well discharges and for the two approaches applied, i.e. measured discharge Fig. 5. The relationship between gas ratios of (A) CH4-H2O-CO2-H2, (B) H2S-H2, (C) H2S-
enthalpy (meas. h) and calculated reservoir enthalpy assuming liquid only reservoir H2-H2O and temperature. Shown are the ratios of the various samples in reservoir fluids
(calc. h) as well as the mole fractions in fumarole vapor. The various mineral-gas calculated from the two-phase well discharges and for the two approaches applied, i.e.
equilibria are further shown (Table 1) and the effect of adiabatic boiling from the measured discharge enthalpy (meas. h) and calculated reservoir enthalpy assuming
reservoir conditions to surface at 100 °C for saline (N500 ppm Cl) and dilute (b500 ppm liquid only reservoir (calc. h) as well as the mole fractions in fumarole vapor. The
Cl) fluids. For explanation of symbols see Fig. 3. A) The concentrations of CO2 are various mineral-gas ratios equilibria are further shown (Table 1) and the effect of
observed to be scatter within ±1 log units for dilute and saline thermal fluids, the exact adiabatic boiling from the reservoir conditions to surface at 100 °C for saline (N500 ppm
composition mostly depending on the model applied to calculate the reservoir fluids. Cl) and dilute (b500 ppm Cl) fluids. For explanation of symbols see Fig. 3. A) The
Gas-mineral reactions show greater variability with temperature and between different reservoir fluids and fumarole vapors are not observed to correspond to the predicted
reactions suggesting that equilibrium of a given gas-mineral reaction may not control equilibrium conditions for the reaction CO2(g) + 4H2(g) = CH4(g) + 2H2O(g),
xCO2 in the thermal fluids. Boiling of reservoir fluids results in decreased xCO2 but these suggesting that an overall equilibrium between CH4, H2O, CO2 and H2 is not attained for
changes are insignificant relative to the scatter of the data. B) With respect to H2S, dilute Icelandic thermal fluids. In contrast, close to equilibrium conditions are observed between
thermal fluids have generally higher concentrations than saline fluids. Moreover, xH2S is H2(g), H2S(g) and H2O corresponding to the reactions B) pyrite + H2(g) =
observed to decrease with decreasing temperature, similar to the calculated equilibrium pyrrhotite + H2S(g) and C) 3pyrite + 2H2(g) + 4H2O(g) = magnetite + 6H2S(g). How-
values with respect to various gas-mineral reactions suggesting H2S to by fluid-mineral ever, it should be pointed out that in all cases large scatter of the data is observed and that
equilibrium. Boiling results in decreased xH2S, these changes being less than the scatter the calculated reservoir fluid composition from data on well discharges is in all cases very
of the data. C) For H2, dilute fluids generally display higher concentrations than saline dependent on the model applied making any conclusion if gas-gas and/or gas-mineral
fluids, similar to H2S. The predicted xH2 values by the gas-mineral buffers show great equilibrium controls gas composition of thermal fluids in Iceland.
variability for different reactions making it difficult to conclude if a fluid-mineral
equilibrium controls H2 concentration in the thermal fluids.
A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94 91

Fig. 6. The mole distribution of (A) CO2, N2 and Ar and (B) H2S, N2 and Ar in two-phase
well discharges and fumarole vapor. The source of Ar and N2 correspond to the ratios
between air and air saturated water, whereas the thermal fluids show progressive
enrichment with respect to Co2 and H2S suggesting addition of these gases to the fluids
to various degree.

The sources of CO2 and H2S may be the basalt and progressive fluid-
rock interaction and/or degassing of basaltic melts, either at great depth
upon partial melting within the upper mantle and lower crust or at
shallower levels within the crust. Both types of source(s) have been sug-
gested, particularly evidenced in the case of CO2 (e.g., Stefánsson et al., Fig. 7. The result of the gas-fluid-rock interaction modeling as a function of progressive
2016a), whereas H2S is considered to originate predominantly from ba- fluid rock interaction (ξ) for initial 0.1%, 1% and 10% volcanic gas mass input into
meteoric water. Shown are (A) pH, (B) CO2 concentrations, (C) H2S concentrations and
salt upon rock leaching (Stefánsson et al., 2015; Gunnarsson-Robin
(D) H2 concentrations. Also shown are the range of pH, CO2, H2S and H2 values in dilute
et al., 2017). reservoir fluids (b500 ppm Cl) hosted by basaltic rocks and assuming liquid only
In order to study the effects of progressive fluid-rock interaction and reservoir and calculated enthalpy. With progressive fluid-rock interaction the pH of the
magma gas addition, geochemical model calculations were conducted. solution increases due to consumption of H+ with basalt leaching with the exact pH
value determined by the input of volcanic gas (acid) and extent of reaction (ξ). H2S and
The calculations involve mixing non-thermal water with volcanic gas
H2 are observed to become controlled by metastable fluid-mineral equilibria at
of different ratios (0.1–10% volcanic gas) and reacting the liquid solution insignificant to moderate reaction progress (ξ = 0.01 to 0.5), the changes in the
with basaltic rock in steps, allowing secondary minerals commonly ob- metastable equilibria and mass of minerals resulting in insignificant variations for H2S
served in geothermal systems in Iceland to precipitate when saturated. concentrations but being important for H2 concentrations with increasing
The results of the geochemical calculations with respect to pH, CO2, H2S concentrations with increasing pH. In contrast, CO2 is not controlled by metastable fluid-
mineral equilibria except at a very high reaction progress, low gas input and alkaline pH,
and H2 are shown in Fig. 7. Additionally, the range of concentration of
thus suggesting that the CO2 concentrations are controlled by the ratio of volcanic gas to
gases of the reservoir fluids are shown (as box plot). With progressive source water in the geothermal reservoir in most cases.
fluid-rock interaction, the pH of the thermal solutions increases due to
consumption of H+ associated with basalt dissolution (e.g., Stefánsson,
2010; Gysi and Stefánsson, 2011). For fluids with a low-acid supply, is observed. At low volcanic gas input (b 0.1%), the H2S is rapidly miner-
i.e. low volcanic gas input, the pH increases rapidly to N 8 at low reaction alized into sulfides, resulting in a rapid decrease of H2S whereas with
progress (ξ b 0.l), whereas at high volcanic gas input, more basaltic rock considerable gas input, the H2S concentrations become buffered by
is needed to raise the pH. With respect to the gases, a very specific trend mineral-gas reactions at a low to moderate reaction progress (ξ 0.01
92 A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94

to 0.2 at 1 to 10% volcanic gas input). In contrast, CO2 does not seem to values observed for CO2, and H2S and the variable extent of the reactions
be controlled by mineral-gas reactions but rather by the relative amount (ξ b0.01 to N1). The corresponding values for H2 indicate either some-
of gas input, except at low CO2 gas input, whereas the pH increases due what lower volcanic gas input and/or greater extent of reaction ξ N 1.
to fluid-rock interaction resulting in calcite formation and decreased The results of the simulations demonstrated that overall mineral-gas
CO2 concentration. In the case of H2, the thermal fluid concentration is equilibria of the systems do not control the concentrations of reactive
controlled by mineral-gas equilibrium, the minerals involved as well volatiles like CO2, H2S and H2 in geothermal systems. For multicompo-
as their relative masses depending on the pH of the solution. At alkaline nent and multiphase (solid, liquid and gas) systems where the compo-
pH, considerable H2 may form under hydrothermal conditions, such nents react within and between phases in an incongruent manner, the
conditions either produced by low volcanic gas input and insignificant concentrations of components may change as a function of the extent
to extensive fluid-rock interaction or high-volcanic gas input and exten- of reaction even though under given conditions the system is at meta-
sive fluid-rock interaction. stable equilibrium. An overall equilibrium will not be attained until
The results of the gas-fluid-rock interaction simulations clearly indi- the system is closed off to inputs from all sources and has reacted to
cate the key variables in controlling the concentrations of reactive vola- achieve its final thermodynamic stability. It follows that the concentra-
tiles in geothermal systems are degree of volcanic gas input and tions of components of such open systems reflect metastable equilibria
progressive fluid-rock interaction or fluid to rock ratio. An overall gas- for some components like H2 and H2S and non-equilibrium source con-
gas and mineral-gas equilibrium is not attained; instead the system is trolled concentrations for other components like CO2, the metastability
in metastable equilibrium along the reaction progress (e.g., Anderson, of the system being controlled by the ratio of the entry of elements into
2002). The concentrations of H2S and H2 seem to reflect close to the system and the extent of reactions occurring within the system.
mineral-gas metastable equilibria at low to high-volcanic gas input,
the absolute concentration of the gases, however, depending on the 7. Summary and conclusions
mass movement in the system controlled by the gas input and extent
of fluid-rock interaction. In contrast, the concentrations of CO2 do not The available data on gas concentrations in thermal fluids in Iceland
seem to be controlled by fluid-rock equilibria but rather the ratio of vol- have been summarized. The samples included two-phase well dis-
canic gas to source water. For the Icelandic systems, a close to ~1% vol- charges and single phase vapor fumaroles, for a total of 862 samples
canic gas and 99% source water seems to correspond closely to the from 18 geothermal areas. The vapor was dominated by H2O, generally

Fig. 8. A summary of the sources and processes affecting volatile concentrations in thermal fluids in Iceland.
A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94 93

accounting for N 99 mol%. Carbon dioxide, H2S and H2 were the other Benjamínsson, J., Hauksson, T., 2011. Kröflusvædi and Bjarnarflag: environmental moni-
toring 2010. Report LV-2011/119 (in Icelandic).
dominant gases, followed by N2, CH4, and Ar. An overall gas-gas and Benjamínsson, J., Sigurdsson, K.H., Hjaltadóttir, V., Tómasson, S., 1985. Krafla: composition
mineral-gas equilibrium is not attained within geothermal systems in of gas in fumaroles, June 1985. Report OS-85-060/JHD-25 (in Icelandic).
Iceland; instead the thermodynamic system occurs at various metasta- Bjarnason, J.Ö., 2010. The chemical speciation program WATCH, version 2.4. ÍSOR –
Iceland Geosurvey, Reykjavík, Iceland (Accessible at: http://www.geothermal.is/
ble equilibria along a reaction progress. The concentrations of H2S and software).
H2 seem to reflect close to mineral-gas metastable equilibria, whereas Bjarnason, J.Ö., Ólafsson, M., 2000. Torfajökll: chemistry of steam and thermal water. Re-
the concentration of CO2 does not seem to be controlled by either port OS-2000/030 (in Icelandic).
Bjarnason, J.Ö., 1996. Svartsengi, chemical monitoring, 1988–1995. Report OS-96082/
mineral-gas or gas-gas reactions but rather by the ratio of volcanic gas JHD-10 (126 pp).
to source water with the input of magma gas corresponding to N 0.1 to Bjarnason, J.Ö., 2002. Reykjanes, chemistry of geothermal seawater and steam, 1971–
b5 mass%. An overall equilibrium is not reached in the geothermal sys- 2001. Report OS-2002/038 (69 pp).
Bödvarsson, G., 1961. Physical characteristics of natural heat resources in Iceland. Jökull
tems until they are closed off to inputs of all sources and have reacted to
11, 29–38.
its final thermodynamic stability. Moreover, the concentrations of gases Chiodini, G., Marini, L., 1998. Hydrothermal gas equilibria: the H2O-H2-CO2-CO-CH4 sys-
controlled by metastable equilibria may vary as a function of extent of tem. Geochim. Cosmochim. Acta 62, 2673–2687.
reaction, for example progressive fluid-rock interaction. With fluid as- Chiodini, G., Cioni, R., 1989. Gas geobarometry for hydrothermal systems and its applica-
tion to some Italian geothermal areas. Appl. Geochem. 4, 465–472.
cent to surface, boiling and condensation may occur, further changing D'Amore, F., Nuti, S., 1977. Notes on the chemistry of geothermal gases. Geothermics 6,
the gas concentrations and hence surface fumaroles may not reflect 39–45.
the reservoir fluid characteristics but rather these secondary processes D'Amore, F., Panichi, C., 1980. Evaluation of deep temperature of hydrothermal systems
by a new gas-geothermometer. Geochim. Cosmochim. Acta 44, 549–556.
(Fig. 8). Fernandez-Prini, R., Alvarez, J.L., Harvey, A.H., 2003. Henry's constants and vapor–liquid
distribution constants for gaseous solutes in H2O and D2O at high temperatures.
J. Phys. Chem. Ref. Data 32, 903–916.
Fridleifsson, I.B., 1979. Geothermal activity in Iceland. Jökull 29, 47–56.
Acknowledgments Füri, E., Hilton, D.R., Halldórsson, S.A., Barry, P.H., Hahm, D., Fischer, T.P., Grönvold, K.,
2010. Apparent decoupling of the he and ne isotope systematics of the Icelandic man-
I would like to thank Jóhann Gunnarsson-Robin for his help in com- tle: the role of he depletion, melt mixing, degassing fractionation and air interaction.
Geochim. Cosmochim. Acta 74, 3307–3332.
piling the data and assisting in sample collection and analysis. Gauthier, P.J., Sigmarsson, O., Gouhier, M., Haddadi, B., Moune, S., 2016. Elevated gas flux
Gudmundur Sverrisson and Hjörtur Andrason are also thanked for their and trace metal degassing from the 2014–2015 fissure eruption at the Bárdarbunga
help in the field. The ThermoFluid dataset can be obtained from Ríkey volcanic system, Iceland. J. Geophys. Res. Solid Earth http://dx.doi.org/10.1002/
2015JB012111.
Kjartansdóttir (rikey@hi.is). I would also like to thank the guest editor,
Giggenbach, W.F., 1975. A simple method for the collection and analysis of volcanic gas
Yuri Taran and two reviewers, Jacob Lowenstern and an anonymous re- samples. Bull. Volcanol. 39, 132–145.
viewer for their constructive comments which have improved this Giggenbach, W.F., 1980. Geothermal gas equilibria. Geochim. Cosmochim. Acta 44,
contribution. 2021–2032.
Giggenbach, W.F., 1987. Redox processes governing the chemistry of fumarolic gas dis-
charges from White Island, New Zealand. Appl. Geochem. 2, 143–161.
References Giggenbach, W.F., 1991. Chemical techniques in geothermal exploration. In: D'Amore, F.
(Ed.), Application of Geochemistry in Geothermal Reservoir Development. UNITAR,
Anderson, G.M., 2002. Stable and metastable equilibrium: The third constraint. In: Water- pp. 119–144.
rock interactions, ore deposits and Env. Geochem A tribute to D.A. Crerar, Hellmann, Giggenbach, W.F., 1992. Isotopic shift in waters from geothermal and volcanic systems
R., Wood, S.A. (eds.). Geochem. Soc. Spec. Publ. 7, 181–189. along convergent plate boundaries and their origin. Earth Planet. Sci. Lett. 113,
Ármannsson, H., 2015. The fluid geochemistry of Icelandic high temperature geothermal 495–510.
areas. Appl. Geochem. 66, 14–64. Giggenbach, W.F., 1993. Redox control of gas compositions in Philippine volcanic-
Ármannsson, H., Benjamínsson, J., Sigurdsson, K.H., 1985. Krafla: chemical composition of hydrothermal systems. Geothermics 22, 575–587.
steam in fumaroles. July 1984. Report OS-85058/JHD-23 (in Icelandic). Giroud, N., 2008. A Chemical Study of Arsenic, Boron and Gases in High-Temperature
Ármannsson, H., Gíslason, G., Hauksson, T., 1982. Magmatic gases aid the mapping of the Geothermal Fluids in Iceland. (Ph.D. thesis). University of Iceland.
flow pattern in a geothermal system. Geochim. Cosmochim. Acta 46, 167–177. Gíslason, G., Johnsen, G.V., Ármannson, H., Torfason, H., Árnason, K., 1984. Theistareykir.
Ármannsson, H., Hauksson, T., 1980. Krafla: chemical composition of steam in fumaroles. Surface study of a high-temperature geothermal area. Report OS-84089/JHD-16 (in
Report OS80027/JHD16 (in Icelandic). Icelandic).
Árnason, B., 1977. Hydrothermal systems in Iceland traced by deuterium. Geothermics 5, Gíslason, S.R., Stefánsdóttir, G., Pfeffer, M.A., Barsotti, S., Jóhannsson, Th., Galeczka, I., Bali,
125–151. E., Sigmarsson, O., Stefánsson, A., Keller, N.S., Sigurdsson, Á., Bergsson, B., Galle, B.,
Arnórsson, S., 1995. Geothermal systems in Iceland: structure and conceptual models - I. Jacobo, V.C., Arellano, S., Aiuppa, A., Jónasdóttir, E.B., Eiríksdóttir, E.S., Jakobsson, S.,
High-temperature areas. Geothermics 24, 561–602. Gudfinsson, G.H., Halldórsson, S.A., Gunnarsson, H., Haddadi, B., Jónsdóttir, I.,
Arnórsson, S., Björnsson, S., Muna, Z.W., Bwire-Ojiambo, S., 1990. The use of gas chemistry Thordarson, Th., Riishuus, M., Högnadóttir, Th., Dürig, T., Pedersen, G.B.M.,
to evaluate boiling processes and initial steam fractions in geothermal reservoirs with Höskuldsson, Á., Gudmundsson, M.T., 2015. Environmental pressure from the 2014-
an example from the Olkaria field, Kenya. Geothermics 19, 497–514. 15 eruption of Bárdarbunga volcano, Iceland. Geochem. Persp. Lett. 1, 84–93.
Arnórsson, S., 1987. Gas chemistry of the Krísuvík geothermal field, Iceland, with special Gudmundsson, B.Th., Arnórsson, S., 2002. Geochemical monitoring of the Krafla and
reference to evaluation of steam condensation in upflow zones. Jökull 37, 31–48. Námafjall geothermal areas, N-Iceland. Geothermics 31, 195–253.
Arnórsson, S., Barnes, I., 1983. The nature of carbon dioxide waters in Snæfellsnes, west- Gudmundsson, B.Th., Arnórsson, S., 2005. Secondary mineral-fluid equilibria in the Krafla
ern Iceland. Geothermics 12, 171–176. and Námafjall geothermal systems, Iceland. Appl. Geochem. 20, 1607–1625.
Arnórsson, S., Bjarnason, J.Ö., Giroud, N., Gunnarsson, I., Stefánsson, A., 2006. Sampling Gunnarsson, I., Arnórsson, S., 2000. Amorphous silica solubility and the thermodynamic
and analysis of hydrothermal fluids. Geofluids 6, 203–216. properties of H4SiO4 in the range of 0 to 350°C at Psat. Geochim. Cosmochim. Acta
Arnórsson, S., Fridriksson, Th., Gunnarsson, I., 1996. Krafla and Námafjall: gases in wells 64, 2295–2307.
and fumaroles in 1995–1996. Report RH 12–96 (in Icelandic). Gunnarsson-Robin, J., Stefánsson, A., Ono, S., Torssander, P., 2017. Sulfur isotopes in
Arnórsson, S., Gunnlaugsson, E., 1985. New gas geothermometers for geothermal Icelandic thermal fluids. J. Volcanol. Geotherm. Res. 346, 161–179.
exploration–calibration and application. Geochim. Cosmochim. Acta 49, 1307–1325. Gysi, A.P., Stefánsson, A., 2011. CO2-water–basalt interaction. Low temperature experi-
Arnórsson, S., Ívarsson, G., Cuff, K.E., Sæmundsson, K., 1987. Geothermal activity in the ments and implications for CO 2 sequestration into basalts. Geochim. Cosmochim.
Torfajökull field, South Iceland: summary of geochemical studies. Jökull 37, 1–10. Acta 81, 129–152.
Arnórsson, S., Sigurdsson, S., Svavarsson, H., 1982. The chemistry of geothermal waters in Halldórsson, S.A., Barnes, J.D., Stefánsson, A., Hilton, D.R., Hauri, E.H., Marshall, E.W.,
Iceland. I. Calculation of aqueous speciation from 0° to 370°C. Geochim. Cosmochim. 2016a. Subducted lithosphere controls halogen enrichments in the Iceland mantle
Acta 46, 1513–1532. plume source. Geology 44, 679–682.
Arnórsson, S., Stefánsson, A., Bjarnason, J.Ö., 2007. Fluid-fluid interaction in geothermal Halldórsson, S.A., Hilton, D.R., Barry, P.H., Füri, E., Grönvold, K., 2016b. Recycling of crustal
systems. Rev. Min. Geochem. 65, 259–312. material by the Iceland mantle plume: new evidence from nitrogen elemental and
Barry, P.H., Hilton, D.R., Füri, E., Halldórsson, S.A., Grönvold, K., 2014. Carbon isotope abun- isotope systematics of subglacial basalts. Geochim. Cosmochim. Acta 176, 206–226.
dance systematics of Icelandic geothermal gases, fluids and subglacial basalts with Hilton, D.R., Fischer, T.P., Marty, M., 2002. Noble gases and volatile recycling at subduction
implications for mantle plume- related CO2 fluxes. Geochim. Cosmochim. Acta 134, zones. Rev. Min. Geochem. 47, 319–370.
74–99. Hilton, D.R., Grönvold, K., Sveinbjörnsdóttir, A.E., Hammerschmidt, K., 1998. Helium iso-
Benjamínsson, J., 1985. Krafla: comparison of gases in fumaroles between 1979 and 1984/ tope evidence for off-axis degassing of the Icelandic hotspot. Chem. Geol. 149,
85. Report OS-85061/JHD-26 (in Icelandic). 173–187.
Benjamínsson, J., Hauksson, T., 2010. Kröflusvædi and Bjarnarflag: environmental moni- Hjartarson, Á., Ólafsson, M., 2005a. Kerlingarfjöll: Study and mapping og a high-
toring 2009. Report LV-2010/110 (in Icelandic). temperature geothermal area. Report ISOR-2005/012 (in Icelandic).
94 A. Stefánsson / Journal of Volcanology and Geothermal Research 346 (2017) 81–94

Hjartarson, Á., Ólafsson, M., 2005b. Hveravellir: study and mapping of a high-temperature Stefánsson, A., Arnórsson, S., 2002. Gas pressures and redox reactions in geothermal fluids
geothermal area. Report ISOR-2005/014 (in Icelandic). in Iceland. Chem. Geol. 190, 251–271.
Johnson, J.W., Oelkers, E.H., Helgeson, H.C., 1992. Supcrt92 – a software package for calcu- Stefánsson, A., Arnórsson, S., Gunnarsson, I., Kaasalainen, H., Gunnlaugsson, E., 2011. The
lating the standard molal thermodynamic properties of minerals, gases, aqueous spe- geochemistry and sequestration of H2S into the geothermal system at Hellisheidi.
cies, and reactions from 1 bar to 5000 bar and 0°C to 1000°C. Comput. Geosci. 18, J. Volcanol. Geotherm. Res. 202, 179–188.
899–947. Stefánsson, A., Barnes, J.D., 2016. Chlorine isotope geochemistry of Icelandic thermal
Kaasalainen, H., Stefánsson, A., 2012. The chemistry of trace elements in surface geother- fluids: implications for geothermal system behavior at divergent plate boundaries.
mal water and steam, Iceland. Chem. Geol. 330–331, 60–85. E. Planet. Sci. Lett. 449, 69–78.
Kaasalainen, H., Stefánsson, A., Giroud, N., Arnórsson, S., 2015. The geochemistry of trace Stefánsson, A., Gunnarsson, I., Giroud, N., 2007. New method for the direct determination
elements in geothermal fluids, Iceland. Appl. Geochem. 62, 207–223. of dissolved inorganic, organic and total carbon in natural water by reagent-free ion
Kristinsson, S.G., Fridriksson, Th., Ólafsson, M., Gunnarsdóttir, H.G., Níelsson, S., 2013a. chromatography and inductively coupled plasma atomic emission spectrometry.
High-temperature areas at Theistareykir, Krafla and Námafjall: monitoring of surface Anal. Chim. Acta 582, 69–74.
activity and groundwater. Report LV-2013-091 (in Icelandic). Stefánsson, A., Keller, N.S., Gunnarsson-Robin, J., Ono, S., 2015. Multiple sulfur isotope sys-
Kristinsson, S.G., Óskarsson, F., Ólafsson, M., Óladóttir, A.A., Tryggvason, H.H., Fridriksson, tematics of Icelandic geothermal fluids and the source and reactions of sulfur in vol-
Th., 2013b. High-temperature areas at Theistareykir, Krafla and Námafjall: monitor- canic geothermal systems at divergent plate boundaries. Geochim. Cosmochim. Acta
ing of surface activity and groundwater in 2013. Report LV-2013-132 (in Icelandic). 165, 307–323.
Kristmannsdóttir, H., 1978. Alteration of bedrock in the Krafla geothermal system. Nation- Stefánsson, A., Sveinbjörnsdóttir, Á.E., Heinemeier, J., Arnórsson, S., Kjartansdóttir, R.,
al Energy Authority Report (OS-JHD-7854). Kristmannsdóttir, H., 2016a. Mantle CO2 degassing through the Icelandic crust evi-
Líndal, B., Hermannsson, S., 1951. Chemical Analysis of hot Springs. Jarðboranir ríkisins denced from carbon isotopes in groundwater. Geochimica. Cosmochimica Acta 191,
(95 pp.). 300–319.
Lonker, S.W., Franzson, H., Kristmannsdóttir, H., 1993. Mineral–fluid interaction in the Stefánsson, A., Arnórsson, S., Kjartansdóttir, R., Gunnarsson-Robin, J., Sveinbjörnsdóttir,
Reykjanes and Svartsengi geothermal systems, Iceland. Am. J. Sci. 293, 605–670. Á.E., Kaasalainen, H., Keller, N.S., 2016c. GeoFluids database 2016. Chemical composi-
Lowenstern, J.B., Bergfeld, D., Evans, W.C., Hunt, A.G., 2015. Origins of geothermal gases at tion of Icelandic fluids and gases. Sci. Inst. Report RH-10-2016.
Yellowstone. J. Volcanol. Geotherm. Res. 302, 87–101. Stefánsson, A., Hilton, D.R., Sveinbjörnsdóttir, Á.E., Torssander, P., Heinemeier, J., Barnes,
Marty, B., Gunnlaugsson, E., Jambon, A., Oskarsson, N., Ozima, M., Pineau, F., Torssander, P., D.J., Ono, S., Halldórsson, S.A., Fiebig, J., Arnórsson, S., 2017a. Isotope systematics of
1991. Gas geochemistry of geothermal fluids, the Hengill area, southwest rift zone of Icelandic thermal fluids. J. Volcanol. Geotherm. Res. 337, 146–164.
Iceland. Chem. Geol. 91, 207–225. Stefánsson, A., Stefánsson, G., Keller, N.S., Barsotti, S., Sigurdsson, Á., Thorláksdóttir, S.B.,
Ólafsson, M., 1991. Geothermal fluids in Krýsuvík: Sampling of fumaroles, spring 1990 Re- Pfeffer, M.A., Eiríksdóttir, E.S., Jónasdóttir, E.B., Löwis, S., Gíslason, S.R., 2017b. Major
port MÓ-91-06. impact of volcanic gases on the chemical composition of precipitation in Iceland dur-
Óskarsson, F., Fridriksson, Th., Thorbjörnsson, D., 2015. Geochemical monitoring of the ing the 2014–2015 Holuhraun eruption. J. Geophys. Res. Atmos. 122, 1971–1982.
Reykjanes geothermal reservoir 2003 to 2013. Proc. World. Geotherm. Cong. 1-9. Stefánsson, A., Keller, N.S., Gunnarsson-Robin, J., Kaasalainen, H., Björnsdóttir, S.,
Parkhurst, D.L., Appelo, C.A.J., 1999. User's guide to PHREEQC (Version 2) - A computer Pétursdóttir, S., Jóhannesson, H., Hreggvidsson, G.Ó., 2016b. Quantifying mixing, boil-
program for speciation, batch-reaction, one-dimensional transport, and inverse geo- ing, degassing, oxidation and reactivity of thermal waters at Vonarskard, Iceland.
chemical calculations. Report 99-4259, U.S. Geological Survey, Water-Resources J. Volcanol. Geotherm. Res. 309, 53–62.
Investigations. Sveinbjörnsdóttir, Á.E., 1992. Composition of geothermal minerals from saline and dilute
Pope, E.C., Bird, D.K., Arnórsson, S., Giroud, N., 2015. Hydrogeology of the Krafla geother- fluids—Krafla and Reykjanes, Iceland. Lithos 27, 301–315.
mal system, northeast Iceland. Geofluids 16, 1–23. Sveinbjörnsdóttir, A.E., Coleman, M.L., Yardley, B.W.D., 1986. Origin and history of hydro-
Poreda, R.J., Craig, H., Arnórsson, S., Welhan, J.A., 1992. Helium isotopes in Icelandic geo- thermal fluids of the Reykjanes and Krafla geothermal fields, Iceland: a stable isotope
thermal systems: I. 3He, gas chemistry, and 13C relations. Geochim. Cosmochim. Acta study. Contrib. Mineral. Petrol. 94, 99–109.
56, 4221–4228. Taran, Y.A., 1986. Gas geothermometers for hydrothermal systems. Geochem. Int. 23,
Sano, Y., Urabe, A., Wakita, H., Chiba, H., Sakai, H., 1985. Chemical and isotopic composi- 111–126.
tions of gases in geothermal fluids in Iceland. Geochem. J. 19, 135–148. Taran, Y.A., Hedenquist, J.W., Korzhinsky, M.A., Tkachenko, S.I., Shmulovich, K.I., 1995.
Scott, S., Gunnarsson, I., Arnórsson, S., Stefánsson, A., 2014. Gas chemistry, boiling and Geochemistry of magma gases from Kudryavy volcano, Itrup, Kuril Island. Geochim.
phase segregation in a geothermal system, Hellisheidi, Iceland. Geochim. Cosmochim. Cosmochim. Acta 59, 1749–1761.
Acta. 124, 170–189. Taran, Y.A., 2011. N2, Ar and He as a tool for discriminating sources of volcanic fluids with
Sigvaldason, G.E., Elísson, G., 1968. Collection and analysis of volcanic gases at Surtsey. application to Vulcano, Italy. Bull. Volcanol. 73, 395–408.
Iceland. Geochim. Cosmochim. Acta 32 (797-80).
Stefánsson, A., 2010. Low-temperature alteration of basalts–the effects of temperature,
acids and extent of reaction on mineralization and water chemistry. Jökull 60,
165–184.

You might also like