Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

PROCEEDINGS, Thirty-Eighth Workshop on Geothermal Reservoir Engineering

Stanford University, Stanford, California, February 11-13, 2013


SGP-TR-198

FLUID-MINERAL EQUILIBRIA IN GREAT BASIN GEOTHERMAL SYSTEMS:


IMPLICATIONS FOR CHEMICAL GEOTHERMOMETRY

Stuart F. Simmons

Geology and Geological Engineering, Colorado School of Mines


1500 Illinois St
Golden, CO, 80401, USA
e-mail: stuart@hotsolutions.co.nz

ABSTRACT quartz or one of its polymorphs can be applied to


geothermal reservoirs with quartz-bearing rock types
Partial to nearly full fluid-mineral equilibrium exists
(Fournier, 1981; 1985). The still widely used Na-K-
between common aqueous species in reservoir fluids
Ca geothermometer (Fournier and Truesdell, 1973)
and the mineral assemblage quartz (chalcedony), Na-
is an empirical geothermometer, which lacks
feldspar, K-feldspar, K-mica, Mg-chlorite
thermodynamic underpinning and functions as a
(clinochlore), calcite, and heulandite in Great Basin
complicated version of the Na-K geothermometer
geothermal systems. For chemical geothermometry,
(Giggenbach, 1988).
the silica geothermometer based on
quartz/chalcedony solubility appears to be the most
The results reported below represent preliminary
reliable indicator of reservoir temperature. Although
outcomes of a DoE funded project to identify new
Na/H vs K/H ratios plot on the Na-feldspar-K-
hydrogeochemical indicators of geothermal resources
feldspar equilibrium line (~250°C), application of the
in the Great Basin. One of the major goals is to
Na-K geothermometer overestimates reservoir
improve confidence in the application and
temperature, possibly because equilibration occurs at
interpretation of chemical geothermometers.
conditions that are hotter and deeper than existing
Accordingly, the states of fluid-mineral equilibria in
feed points in geothermal wells. A linear trend in
reservoirs were assessed, using public domain data
Na/K versus temperature permits formulation of a
sets of fluid compositions and reservoir temperatures
new empirical geothermometer
for Beowawe, Dixie Valley, Mammoth-Long Valley,
Raft River, and Roosevelt. The results build on
earlier work (Capuano and Cole, 1982; Tempel et al.,
[ ] 2011; White and Petersen, 1991) and were used to
determine the applicability of aqueous
that has yet to be tested and proven. Modest to low geothermometers. They also provide insights
confidence in the application of the K-Mg regarding the minerals that are likely to be
geothermometer is expected because the reservoir influencing the chemical compositions of deep
K/Mg ratios are scattered with respect to Mg- geothermal waters in the Great Basin.
chlorite-K-mica-K-feldspar-quartz equilibria.
DATA SETS
INTRODUCTION Geothermal reservoir water compositions (Table 1)
Chemical geothermometers have been available since that represent a spectrum of geothermal reservoirs of
the 1960s, providing a cost effective means of low to high temperature across the Great Basin were
estimating the temperature of a geothermal reservoir evaluated. All the waters are near neutral pH.
during exploration, development, and on going Chloride (350 to >3000 mg/kg Cl) is the dominant
management of a resource. The most widely used anion at Dixie Valley, Roosevelt, and Raft River,
cation geothermometers (i.e., Na-K, K-Mg) are based whereas bicarbonate (145->400 mg/kg HCO3) is the
on fluid-mineral equilibria, relating to hydrothermal dominant anion at Beowawe and Long Valley-
minerals that are commonly found in altered volcanic Mammoth. Sulfate is a minor anion, with
rocks (e.g., Fournier, 1981, 1991; Giggenbach, 1988, concentrations (30 to >200 mg/kg SO4) that are
1991). By contrast the geothermometers based on subequal with bicarbonate (30 to <500 mg/kg HCO3).
aqueous silica concentrations and the solubilities of
Temperature data (Table 1) represent maximum data from SUPCRT (Johnson et al., 1992). For these
measurements from well surveys or interpretations phase diagrams, certain zeolites, clays, feldspars and
from well enthalpy data. All reservoirs are assumed calcium alumino-silicates (i.e., beidellite, saponite,
to comprise a single phase liquid, with no free vapor clinoptilolite, margarite, grossular, high albite,
or gas. maximum microcline) were suppressed in order to
most closely match the minerals occurring in the
Mineralogical analyses (XRD, thin section reservoir rocks.
petrography) of drill cuttings from Beowawe, Dixie
Valley, Mammoth, and Roosevelt confirm that RESULTS
quartz, feldspars (Na-feldspar, K-feldspar,
plagioclase), illite (K-mica), chlorite, and calcite are A series of phase diagrams (Figs. 1-7), which show
common minerals in the reservoir rocks, despite the positions of mineral stability fields with respect to
variability in rock types and stratigraphic sequences reservoir water compositions, are used to assess the
that characterize each geothermal system. state of fluid-mineral equilibria. Silica phases and
calcite are evaluated first, followed by assessment of
COMPUTATIONS sodium, potassium, calcium, and magnesium
alumino-silicates. Na/K and K/Mg ratios are plotted
Speciation calculations were undertaken using the
as a function of well temperature to determine
computer code WATCH 2.0 (Iceland Geosurvey;
applicability of conventional geothermometers based
http://geothermal.is/software/software); the most
on equilibrium with hydrothermal feldspars, clays,
recent version is modified and updated from the one
and quartz/chalcedony.
that was initially written by Stefán Arnórsson, Sven
Sigurđsson and and Hörđur Svararson (e.g.,
Silica
Arnórsson et al., 1982, 1983a, 1983b). For Dixie
Since silicic acid (H4SiO4) is a neutral species in
Valley, gas and water data (Goff et al., 2002) were
solution (acid to near neutral pH), the computed
combined to formulate representative reservoir water
value of activities for SiO2(aq) are within a few
compositions using heat and mass balance constraints percent of the analyzed concentrations of SiO2(aq), and
(e.g., Henley et al., 1984). analytical data, corrected for steam loss, can be
plotted directly on phase diagrams to evaluate
WATCH input requires fluid analyses, pH,
equilibrium with respect to quartz, chalcedony, and
temperature/pressure of sampling, and well enthalpy
cristobalite (Fig. 1). Most of the waters plot on or
or reservoir temperature. The concentrations, activity
between chalcedony and quartz solubility curves,
coefficients, and activities of aqueous species at suggesting that these two phases are the main control
reservoir conditions are determined by an iterative on silica concentrations in geothermal reservoirs.
process that involves simultaneous solution of mass
Beowawe data are scattered with two analyses
balance and mass action equations, input from a
plotting near or above the cristobalite solubility
standard set of thermodynamic data, and a model of
curve; the only analysis from Roosevelt also plots
ion-ion interactions based on the extended Debye-
near the cristobalite solubility curve.
Hückel equation (e.g., Garrels and Christ, 1965). The
output gives concentrations of analytes at reservoir
Calcite
conditions, activity coefficients of aqueous species
As shown by Ellis (1959), calcite solubility at
(69 total), partial pressures of dissolved gas species,
hydrothermal conditions is mainly a function of
and reservoir pH, which are used to compute
temperature and aqueous CO2 concentration (Fig. 2).
activities and activity ratios of aqueous species. The activity products ([Ca++][HCO3-]2) for waters
These computed values are plotted on phase diagrams from all the geothermal reservoirs plot close to
showing the compositional and thermal stability
calcite saturation curves based on the concentrations
fields of silica polymorphs (quartz, chalcedony,
of reservoir aqueous CO2. All of the
cristobalite), calcite, and alumino-silicate minerals.
WATCH/SOLVEQ calculations indicate the waters
SOLVEQ (Reed and Spycher, 1984) is a computer
are supersaturated in calcite. There is nil likelihood of
code that is similar to WATCH, containing more forming a geothermometer based on calcite
aqueous and gaseous species in the database, but the solubility.
input files require more time to set up. A comparative
analysis of outputs from SOLVEQ and WATCH
Na-, K-, Ca-, & Mg-Alumino Silicates
using data from Dixie Valley showed that computed
Four similar phase diagrams (Fig. 3) portray the
activities of major species are within a few percent.
stability fields of alumino-silicate minerals as a
function of silica activity/concentration (log [SiO2])
Phase diagrams were computed using Geochemists and cation-hydrogen ion activities. The positions of
Workbench (Bethke, 1996) based on thermodynamic
Figure 1: Aqueous silica concentrations versus temperature with reference to the experimentally determined
quartz, chalcedony, and cristobalite solubility curves (Fournier, 1985).

Figure 2: Product of calcium and bicarbonate activities versus temperature with reference to calcite saturation
curves as a function of aqueous CO2 concentration (after Ellis, 1959).
Figure 3: Log activities of Na/H, K/H, Ca/H, and Mg/H versus log activity of silica, showing compositions of
reservoir waters with respect to the stability fields of alumino-silicate minerals, quartz, and chalcedony
at 150°, 200°, and 250°C.

Figure 4: Log activities of Na/H versus K/H showing the compositions of reservoir waters with respect to the
stability fields of Na- and K-bearing feldspars and micas (quartz saturation) at 150°, 200°, and 250°C.
Figure 5: Log activity of K/H versus temperature showing the compositions of reservoir waters with respect to the
stability fields of K-bearing feldspars and micas.

Figure 6: Log activity of Na/K versus temperature showing the compositions of reservoir waters with respect to the
stability fields of Na- and K-bearing feldspars (solid line). The dashed line that is reported as equation 1
in the text represents a simple linear regression of the aqueous Na/K ratios. Modified from Giggenbach
(1988).
Figure 7: Log activity of Mg/K versus temperature showing the compositions of reservoir waters with respect to
the stability fields of K-feldspar and K-mica plus Mg-chlorite at quartz saturation. Modified from
Giggenbach (1988).

Table 1: Water analyses for Beowawe, Dixie Valley, Mammoth-Long Valley, Roosevelt, and Raft River.
the water compositions plot within or near the K-mica, K-feldspar, Mg-chlorite, and quartz, which
stability fields of K-feldspar, K-mica, albite, forms the basis for the K-Mg aqueous
heulandite, and Mg-chlorite, and within the range of geothermometer (Giggenbach, 1988). The array of K-
quartz-chalcedony solubility for respective reservoir Mg water data are scattered. Dixie Valley and
temperatures. Roosevelt Log [Mg++]/[K+]2 values, however, plot
close or on the equilibrium line. Overall the results
Na/H & K/H Activity Ratios are mixed, and they offer modest to poor confidence
Figure 4 shows the positions of geothermal waters in the application of a K-Mg chemical
with respect to the stability fields of Na-feldspar geothermometer.
(albite), K-feldspar, K-mica, Na-mica (paragonite),
and kaolinite. The waters form a linear trend that DISCUSSION & CONCLUSIONS
coincides with the Na-feldspar-K-feldspar boundary,
suggesting that Na/K ratios are controlled by fluid- These preliminary results show that despite
mineral equilibria involving these feldspars at variability in the rock types and stratigraphic columns
~250°C. This is true even for Mammoth where of geothermal reservoirs across the Great Basin,
because the [H+] ion activity is elevated, the waters quartz (chalcedony), Na-feldspar, K-feldspar, and K-
plot in the K-mica field; a similar trend has been mica influence the compositions of thermal waters
noted in high temperature geothermal reservoirs from via fluid-mineral equilibria. Accordingly, the quartz-
New Zealand (e.g., Simmons and Browne, 2000). silica and Na-K geothermometers should yield
reliable temperatures, even if they reflect different
Figure 5 shows the positions of geothermal waters temperatures and depths of equilibration.
with respect to the stability fields of K-feldspar, K-
mica, and kaolinite. The data plot about the K- The preceding analysis assumes the water data and
feldspar-K-mica boundary, suggesting these two temperature profiles contain minimum analytical
minerals moderate reservoir fluid pH. errors. Some of the data are several decades old, and
the details of measurements are missing or poorly
Na/K & Mg/K Activity Ratios documented. For example, some of the aqueous silica
Figure 6 shows the positions of geothermal waters data from Roosevelt and Beowawe are too high with
with respect to the stability fields of Na-feldspar and respect to quartz solubility. Other sources of potential
K-feldspar as a function of temperature, which is the error might lie in pH measurements, and/or analyses
basis for the Na-K geothermometer (e.g. Giggenbach, of Mg and CO2-HCO3 concentrations. New samples
1988). The data form a low angle linear trend that and analyses are required to confirm the trends
plot well within the K-feldspar stability field, identified in this report.
intersecting the Na-feldspar-K-feldspar boundary
around 250°C close to where the Dixie Valley data Some of the future avenues for research include
plot. This Na/K trend seems to contradict the results investigation of: 1) the origin(s) of the linear Na/K vs
of Figure 4, and it is possible that Great Basin feed temperature relationship; 2) the control of depth and
temperatures and depths may be cooler and shallower temperature on equilibration points for the various
than the environment in which equilibrium is mineral-aqueous ion reactions; 3) the role that
established. Alternatively, geothermal waters are out shallow saline ground waters play in modifying
of equilibrium with Na-feldspar and K-feldspar. geothermal water compositions during fluid
Regardless of the correct explanation, a linear movement to the surface.
regression of the Na/K vs temperature trend yield an
empirical relationship that can be tentatively used as ACKNOWLEDGMENTS
a substitute geothermometer:
Bridget Ayling and Joe Moore are thanked for
supplying the geochemical data on Raft River in
(1) advance of publishing their results. This work was
[ ]
sponsored by Geothermal Program DOE contract
However, for a small range of Na/K values (e.g., due DE-EE0005522.
to analytical error or natural variation) there is a large
of range of corresponding temperatures that could REFERENCES
easily exceed several tens of degrees (°C), hence the Ayling, B. and Moore, J. (in press) “Fluid
usefulness of this empirical relationship needs to be geochemistry at the Raft River geothermal field,
proven. Idaho, USA: New data and hydrogeological
implications,” Geothermics.
Figure 7 shows the positions of geothermal waters
with respect to an equilibrium condition controlled by
Arnórsson, S., Gunnlaugsson, E. and Svavarsson, H. Giggenbach, W. F. (1988) “Geothermal solute
(1983a) “The chemistry of waters in Iceland.II. equilibria. Derivation of Na-K-Mg-Ca
Mineral equilibria and independent variables geoinidicators,” Geochimica Cosmochimica
controlling water compostions,” Geochimica Acta, 52, 2749-2765.
Cosmochimica Acta, 47, 547-566.
Giggenbach, W. F. (1991) “Chemical techniques in
Arnórsson, S., Gunnlaugsson, E. and Svavarsson, H. geothermal exploration,”: in Applications of
(1983b) “The chemistry of waters in Iceland.III. Geochemistry in Geothermal Reservoir
Chemical geothermometry in geothermal Development, UNITAR-UNDP (ed. F.
investigations,” Geochimica Cosmochimica D'Amore), p. 119-144.
Acta, 47, 567-577.
Goff, F., Bergfeld, D., Janik, C.J., Counce, D., and
Arnórsson, S., Sigurŏsson, S. and Svavarsson, H. Murell, M. (2002) “Geochemical data on waters,
(1982) “The chemistry of waters in Iceland.I. gases, scales, and rocks from Dixie Valley
Calculation of aqueous speciation from 0 to 370°,” region, Nevada,” Los Alamos National
Geochimica Cosmochimica Acta, 46, 1513-1532. Laboratory Report LA-13972-MS, Los Alamos,
NM.
Bethke, C. M. (1996) Geochemical Reaction
Modeling Concepts and Applications, Oxford Henley, R.W., Truesdell, A.H., and Barton, P.B. Jr.
University Press, 416 pp. (1984) Fluid-mineral equilibria in hydrothermal
systems: Reviews in Economic Geology, 1, 267
Capuano, R. M., Cole, D. R. (1982) “Fluid-mineral
p.
equilibria in a hydrothermal system,”
Geochimica Cosmochimica Acta, 46, p1353- Johnson, J.W., Oelkers, E.H., and Helgeson, H.C.
1364. (1992) “SUPCRT92: a software package for
calculating standard molal thermodynamic
Cole, D.R., and Ravinsky, L.I. (1984)
properties of minerals, gases, aqueous species,
“Hydrothermal Alteration Zoning in the
and reactions from 1 to 5000 bar and 0 to
Beowawe Geothermal System, Eureka and
1000°C,” Computers in Geosciences, 18, 899-
Lander Counties, Nevada,” Economic Geology,
947.
79, 759-767.
Reed, M. (1989) “Thermodynamic calculations of
Ellis, A.J. (1959) “The solubility of calcite in carbon
calcium carbonate scaling in geothermal wells,
dioxide solutions,” American Journal of Science,
Dixie Valley geothermal field, USA,”
257, 354-365.
Geothermics, 18, 269-277.
Fournier, R. O. (1981) “Application of water
Reed, M. H., and Spycher, N. (1984 “Calculation of
geochemistry to geothermal exploration and
pH and mineral equilibria in hydrothermal
reservoir engineering,” chapter 4 in Geothermal
waters with application to geothermometry and
Systems: Principles and Case Histories (L.
studies of boiling and dilution,” Geochimica
Rybach and L. P. J. Muffler, eds.), Wiley New
Cosmochimica Acta, 48, 1479-1492.
York, 109-143.
Simmons, S. F. and Browne, P. R. L. (2000)
Fournier, R. O. (1985) “The behavior of silica in
“Hydrothermal minerals and precious metals in
hydrothermal solutions,” Reviews in Economic
the Broadlands-Ohaaki geothermal system:
Geology, 2, 45-62.
Implications for understanding low-sulfidation
Fournier, R. O. (1991) “Water geothermometers epithermal environments,” Economic Geology,
applied to geothermal energy,” in Applications of 95, 971-999.
Geochemistry in Geothermal Reservoir
Tempel, R. N., Sturmer, D. M., ad Schilling, J.
Development, UNITAR-UNDP (ed. F.
(2011) “Geochemical modeling of the near-
D'Amore), 37-69.
surface hydrothermal system beneath the
Garrels, R. M., and Christ, C. L. (1965) Solutions, southern moat of Long Valley caldera,
Minerals, and Equilibria, Harper & Row (New California,” Geothermics, 40, 91-101.
York), 450 pp.

You might also like