Anda di halaman 1dari 19

6

Chemical Geothermometers and Their


Application to Formation Waters
from Sedimentary Basins
Yousif K. Kharaka and Robert H. Mariner

Abstract Introduction
Chemical geothermometers, based on the concentration Knowledge of subsurface temperatures is essential
of silica and proportions of sodium, potassium, lithium, to exploration for geothermal energy. Subsurface
calcium, and magnesium in water from hot springs and temperatures are also becoming an important tool
geothermal wells, have been used successfully to esti- in exploration for petroleum (see Hitchon, 1985;
mate the subsurface temperatures of the reservoir rocks. Meyer and McGee, 1985, for recent discussions).
Modified versions of these geothermometers and a new
This use of temperatures in the petroleum industry
chemical geothermometer, based on the concentrations
of magnesium and lithium, are developed to estimate the
stems from the fact, known for more than 100
subsurface temperatures (30C to 200C) in sedimen- years, that many oil and gas fields are associated
tary basins where water salinities and hydraulic pressures with mappable, positive geothermal anomalies at
are generally much higher than those in geothermal sys- producing levels (Ovnatanov and Tamrazyan,
tems. The new Mg-Li geothermometer, which can be 1970; Meyer and McGee, 1985). Also, all investi-
used to estimate subsurface temperatures as high as gations involving water-rock interactions require
350C for waters from sedimentary basins and geother- accurate subsurface temperatures.
mal systems, is given by: Oil and gas wells are the primary sources of tem-
2200 perature data in sedimentary basins. Unfortu-
t = log (~) + 5.47 -273,
nately, the accuracy of temperatures obtained from
oil wells varies widely (Meyer and McGee, 1985).
Many operators do not recognize the importance of
where t is temperature (0C) and Mg and Li concentra- temperatures in exploration and production of
tions are in mg/L. petroleum and make no attempt to gather accurate
Quartz, Mg-Li, Mg-corrected Na-K-Ca, and Na-Li data. Also, most temperatures are obtained from
geothermometers give concordant subsurface tempera- wells during drilling when the temperature distri-
tures that are within lOoC of the measured values for bution in the hole is under maximum thermal dis-
reservoir temperatures higher than about 70C. Mg-Li,
turbance. The most reliable temperatures are those
Na-Li, and chalcedony geothermometers give the best
results for reservoir temperatures from 30C to 70C.
obtained from static bottom-hole pressure and
Subsurface temperatures calculated by chemical geo- temperature surveys generally conducted in
thermometers are at least as reliable as those obtained by production wells. Temperatures obtained from
conventional methods. Chemical and conventional drill-stem tests are of intermediate accuracy.
methods should be used together where reliable temper- However, the majority of temperatures are
ature data are required. obtained from electric log headings; these give the

N. D. Naeser et al. (eds.), Thermal History of Sedimentary Basins


Springer-Verlag New York Inc. 1989
100 YK. Kharaka and R.H. Mariner

least reliable data and are generally lower than thebeen reviewed recently by Truesdell (1976), Ellis
true subsurface values. Meyer and McGee (1985) and Mahon (1977), Fournier (1981), and Henley et
showed that the discrepancy between log-header ai. (1984). The commonly used chemical geother-
and static bottom-hole test temperatures from the mometers (Table 6.1) include the silica geother-
Wattenberg field, Colorado, is large, ranging from mometers (Fournier and Rowe, 1966; Fournier,
11C to 61 0C. 1973), the Na-K-Ca geothermometer (Fournier
In this study we investigate the applicability ofand Truesdell, 1973), the Na-K geothermometers
chemical geothermometers, developed for geo- (White, 1965; Ellis, 1970; Truesdell, 1976; Four-
thermal systems, to formation waters from sedi- nier, 1979), and the Mg-corrected Na-K-Ca
mentary basins. The chemical geothermometers geothermometer (Fournier and Potter, 1979).
were applied to 88 water analyses; 54 ofthese were Silica geothermometry is based on solubility of
from sedimentary basins where subsurface tem- quartz, chalcedony, cristobalite, or amorphous sil-
peratures are known, and 34 were from geother- ica (Fournier and Rowe, 1966; Fournier, 1973).
mal systems. This data base is used to develop a The equations for quartz geothermometry closely
new chemical geothermometer based on the con- approximate the solubility of quartz at the vapor
centrations of magnesium and lithium that gives pressure of the solution (2C) in the temperature
the most reliable estimates of temperatures in range from OC to 250C; above 250C, the
sedimentary basins (30C to 200C). Results fur- equations depart from the experimentally deter-
ther show that chemical geothermometers com- mined solubility of quartz (Fournier, 1981). In
monly used for geothermal systems must be modi- applying the silica geothermometers, ambiguity
fied prior to their application to oil-field waters.may arise as to which silica mineral is controlling
The equations and parameters needed for these the dissolved silica concentration. Arn6rsson
modifications are derived and reported in this (1975) noted that in Iceland quartz controlled dis-
chapter. solved silica concentrations at temperatures of
more than about 180C, and chalcedony controlled
dissolved silica below about 110C. In the 180C
Chemical Geothermometers to 110C range, it was not possible to determine, a
priori, which silica mineral was controlling the dis-
for Geothermal Systems solved silica concentration. In granitic terrains,
quartz may control dissolved silica concentrations
Concentrations of dissolved constituents in geo- down to about 80C (Brook et aI., 1979). Also,
thermal fluids are a function of the temperature of mixing may sharply reduce the dissolved silica
the aquifer and the alteration mineral assemblage concentration, resulting in low calculated tempera-
(White, 1965; Ellis, 1970; Truesdell, 1976). Any tures from silica geothermometers in some
constituent whose concentration is controlled by a geothermal springs.
temperature-dependent reaction could, theoreti- It should further be noted that an assumption of
cally, be used as a geothermometer. However, to be silica geothermometry is that the dissolved silica is
useful as a geothermometer, additional conditions present in solution as silicic acid (H 4 SiOS). A nota-
must be met. These conditions include an adequate ble exception to this general assumption may occur
supply of the reactants, establishment of equili- in low-temperature geothermal systems in granitic
brium between water and minerals in the aquifer, rocks where the pH of water may be high (Mariner
and absence of additional interactions as the water et aI., 1983). In waters with pH values higher than
(gas) flows to the sampling point (Fournier et aI., about 8.5, a significant part of dissolved silica is
1974). Mixing of waters from aquifers with differ- present as H3 Si04 -. The actual concentrations of
ent temperatures could alter the concentrations of H 4 SiOS in these waters must be calculated for use
constituents used in geothermometers and require in silica geothermometers.
the application of specialized mixing models All geothermometers based on cation ratios are.
(Fournier and Truesdell, 1974; Fournier, 1981). empirical; that is, they are based on temperature-
Chemical geothermometers and the conditions dependent changes in the cation ratios of a large
for their application to geothermal systems have number of samples where aquifer temperatures are
6. Chemical Geothermometers and Formation Waters 101

TABLE 6.1. Equations for the most reliable chemical and isotope geothermometers
applied to geothermal waters.
Geothermometer Equation
1309
Quartz-no steam loss tOC = - 273
5. 19-1og SiO z
1522
Quartz-maximum steam loss tOC = - 273
5.75-log SiOz
1032
Chalcedony tOC = - 273
4.69-log SiO z
1217
Na-K (Fournier) tOC = - 273
log (NalK) + 1.483
885.6
Na-K (Truesdell) tOC = - 273
log (NalK) + 0.8573
1647
Na-K-Ca ~= -lli
log (Na/K) + 13[log(-./Ca/Na) + 2.06] + 2.47
13 = 4/3 for t < 100C; = 113 for t > 100C

Mg-Corrected Na-K-Ca (See text)


1700
A 180(SO. - HzO) tOC = - 273
(looolna + 4.1)'
1000 + /l 180(HSO -)
a = 4 and /1 180 in per mil
(1000 + /l 180(H zO)
Note: Concentrations are in mg/kg. Modified from Fournier (1981).

known. Historically, the relation between Na, K, Other chemical and isotope geothermometers
and aquifer temperatures was noted first by White have been suggested but have not been widely used.
(1965). Subsequently, slightly different curves Examples include the empirical Na-Li geother-
(equations) relating Na/K ratios to temperature mometer of Fouillac and Michard (1981); solubili-
were presented by Ellis (1970), Truesdell (1976), ties of minerals, such as anhydrite (Sakai and
and Fournier (1979). The Na-K geothermometer, Matsubaya, 1974); isotopic fractionation between
however, is useful only at temperatures of more dissolved constituents and water, such as sulfate
than about 150C. At lower temperatures, calcium and water (Mizutani, 1972; McKenzie and Trues-
usually makes up a significant fraction of the cat- dell, 1977); and various gas geothermometers
ions, and the Na-K geothermometer gives anomal- (Giggenbach, 1980; DJ\more and Panichi, 1980;
ously high temperature estimates in calcium-rich Nehring and DJ\more, 1984). The Na-Li geother-
waters. This led to the development of the Na-K- mometer is not widely utilized, perhaps because it
Ca geothermometer (Fournier and Truesdell, was developed later and has a chloride depen-
1973). In many low-temperature environments, dence. Isotopic geothermometers can give reliable
however, magnesium concentrations are high and temperatures, but their use is restricted because
the Na-K-Ca geothermometer gives excessively the mass spectrometer and isotope extraction lines
high temperature estimates. An empirical correc- required to prepare and determine the isotopic
tion for the magnesium concentration was deter- compositions are not widely available.
mined and applied to the Na-K-Ca geothermom- Finally, chemical geothermometers based on
eter for use in these waters (Fournier and Potter, solubilities of specific minerals other than quartz
1979). As a result, the Na-K-Ca geothermometer, or chalcedony have not generally been utilized,
with magnesium correction where appropriate, because chemical complexing and activity coeffi-
may be applied to waters with temperatures of 0 cients of the dissolved constituents must be calcu-
to 350C. lated at the aquifer-temperature. These calcula-
102 Y.K. Kharaka and R.H. Mariner

tions require complex chemical speciation pro- Chemical geothermometers based on the solu-
grams, such as SOLMNEQ (Kharaka and Barnes, bilities of silica minerals (Table 6.1) are derived
1973), which can calculate mineral saturation assuming a hydraulic pressure equal to the vapor
states at various temperatures. pressure of water at the specified temperatures.
This assumption introduces only small errors in the
case of high-enthalpy geothermal systems because
Application of Chemical hydraulic pressures in the reservoir rocks of these
Geothermometers to systems approximate the boiling-point curve (Ellis
and Mahon, 1977), resulting in pressures that are
Oil-Field Waters generally less than 100 bars (1,500 psi). However,
petroleum wells are generally deeper and can have
Oil-field waters differ from geothermal waters in much higher hydraulic pressures, especially in
several significant ways that affect the application geopressured systems where values greater than
of the chemical geothermometers discussed above. 1,000 bars (15,000 psi) may be attained (Kharaka
These differences are a function of the generally et aI., 1985).
higher pressures, lower temperatures, and higher Silica geothermometers (Table 6.1) must be
salinities of oil-field waters. Application of chemi- modified to account for the increased solubility
cal geothermometers to natural gas wells may be resulting from higher pressures encountered in oil-
complicated because chemical analyses of waters field waters. Solubility values, as a function of
from these wells may not represent the true chemi- pressure (0 to 1,000 bars) and temperatures (50C
cal composition of water from the production to 350C) were calculated using equations given by
zone. This complication arises because of dilution Fournier and Potter (1982). These calculations
by condensed water vapor produced with natural show (Fig. 6.1), that at any given temperature, the
gas, especially in the case of wells from geopres- increase in solubility is approximately linear with
sured geothermal systems. On the other hand, increasing pressure. The slope of this relationship
problems related to mixing of waters from differ- steepens at higher temperatures, showing that the
ent zones or to water-rock interactions prior to increase in solubility for a given increase in pres-
sampling are generally less severe than those sure is higher at higher temperatures.
encountered in geothermal systems. The increase in solubility of quartz per unit pres-
sure normalized to its solubility at the vapor pres-
Pressure Correction for Silica sure of water was fitted to the following equation:
Geothermometers y =a X e bt (3)
Silica geothermometers are based on the equilib-
where y is the relative increase in solubility per bar,
rium constant (K) for the dissolution of silica
t is temperature (50C to 350C), and a and bare
minerals given by the reaction:
regression coefficients. The values obtained are a
Si02 (s) + 2H20 -= H4 Si02 (I) = 7.862 X 10-5 and b = 3.61 X 10-3 An excellent
correlation coefficient (r) of 0.97 was obtained
The equilibrium constant is given by the following
when results calculated using Equation 3 were
equation:
compared with those given by Fournier and Potter
(1) (1982).
To correct for pressure effects, the concentra-
tions of silica in formation waters should be multi-
where a is the activity of the subscripted species. plied by a correction factor (Pi) given by:
Assuming unit activity for the silica mineral
reacted and replacing a H.SiO. with the product of pf = (1 - 7.862 X 10-5
its molality (m) and activity coefficient (y), Equa- X e(3.61 x 10-' x t) P) (4)
tion 1 becomes
where P is the hydraulic pressure in bars and t is
K = mH.SiO.o X YH.SiO.o (2)
the measured or calculated subsurface temperature
a H20 in 0c. Computed temperatures without pressure
2
6. Chemical Geothermometers and Formation Waters 103

corrections are always higher than measured sub- PRESSURE (BARS)( 10-2)
surface temperatures. The computed tempera- o 2 4 6 8 10
0
tures, for example, are higher than the true values c
>
by 5C at 100C and IrC at 200C in a reservoir 54:z1
u 200' C ;:j
at 1,000 bars. en
~ 0
~

52 ~
~ 290
Activity Coefficient of H4 Si0 4 ~
~

and Activity of Water ::::i ~


III 100 C 50
::> -t
..J
Chemical geothermometers based on solubilities 0270
en 8
N 480
of silica minerals (Table 6.1) are derived assuming ~

~
cr::
that the activity coefficient of H4 Si04 (YH.SiO. ) <{
::> 10
and the activity of water (aH,O) in Equation 2 are 250 ~____~____~~____~____~-J 46-

equal to unity. These assumptions introduce only o 4 8 12 16


PRESSURE (PSI )( 10-3)
minor errors in the computed temperatures for
FIGURE 6.1.Solubility of quartz as a function of pressure
geothermal systems where salinities are almost
at 100C and 200C.
always lower than that of seawater and tempera-
tures are higher than about 200C. The salinities
encountered in sedimentary basins, on the other We prefer the expression of Marshall (1980) and
hand, are generally much higher than seawater, Chen and Marshall (1982) for calculating the
reaching values greater than 350,000 mg/L of dis- activity coefficient ofH 4 Si04 o. The activity coeffi-
solved solids (White, 1965; Hitchon et aI., 1971; cient of H 4 Si04 was derived from data on the
Carpenter et aI., 1974; Kharaka et aI., 1985).
solubility of amorphous silica as a function of tem-
Values for the activity coefficient of H 4 Si04 and perature (OC to 350C) and salinity (0 to 6 molal).
activity of water (Equation 2) depart significantly
The expression is:
(> 5%) from unity at salinities greater than sea-
water (35,000 mg/L); silica geothermometers log (YH.SiOf) = (0.00978 X 1O(280IT) X m (6)
(Table 6.1) should be corrected for these depar- Sodium and chloride are by far the dominant
tures. No correction is necessary for the assump-
species in most formation waters from sedimen-
tion that H4 Si04 is the only silica species in tary basins (Kharaka et aI., 1985), and Equation 6
oil-field waters because the pH values of these
can be applied directly to these waters. However,
waters are almost always lower than about 8.0 the concentrations of other species, especially cal-
(Kharaka et aI., 1985).
cium, can be high, requiring modification of Equa-
There are several methods for calculating the tion 6. Data in Marshall (1980) and Chen and
activity coefficients of neutral species like H4 Si02 Marshall (1982) show that Equation 6 can be
that give different results. The activity coefficients
generalized to:
of all neutral species are generally assumed equal
to that of dissolved CO 2 in ~aCI solutions (Helge- log (YH.SiO.O) = (0.00489 X 1O("OIT)
son, 1969). The activity coefficients for CO 2 (7)
(Yeo,) are calculated using:
where Zi and mi are the charge and analytical
km molality of species (i). The summation covers all
Yeo, (1) = k (5)
the species in the formation water.
An estimate of the errors involved in calculating
where k and km are the Henry's law coefficients in subsurface temperatures assuming that YH.SiO. is
pure water and sodium chloride solutions of molal- equal to unity can be made from Equation 7. This
ity (m) at temperature Tin OK. Values for km and k equation shows that YH.SiO.o increases with in-
are available as a function of temperature (OC to creasing salinity of the water, but decreases with
350C) and molality of NaCI (0 to 6.0 molal) from increasing temperature. The activity coefficients
Ellis and Golding (1963) and Drummond (1982). of H4 Si04 at 100C are equal to 1.00, 1.21, and
104 Y.K. Kharaka and R.H. Mariner

TABLE 6.2. Activity of water in NaCI and CaCl 2 solutions at 25C to 200C.
From osmotic coefficientsb
Source of aH,O
and temp. (0C) All temperaturesa 25 100 200

mNaCI
0 1.00 1.00 1.00 1.00
0.2 0.993 0.993 0.994 0.994
0.5 0.983 0.984 0.984 0.985
1.0 0.966 0.967 0.967 0.967
2.0 0.932 0.932 0.932 0.937
3.0 0.898 0.893 0.894 0.900
4.0 0.864 0.852 0.854 0.871
5.0 0.830 0.807 0.812 0.837
6.0 0.796 0.760 0.768 0.802

mCaCl,
0 1.00 1.00 1.00 1.00
0.2 0.993 0.991 0.991 0.992
0.5 0.983 0.976 0.977 0.980
1.0 0.966 0.945 0.949 0.959
aFrom Equation 8.
bFrom Equation 9.

1.47 in NaCl solutions with salinities of 0, 3, and 6 using Equation 8. Equation 9, however, should be
molal, respectively. Assuming an activity coeffi- used for waters where the concentrations of diva-
cient of unity for H4 Si04 in these solutions results lent cations comprise more than about 20% of the
in calculated subsurface temperatures that are 9C total cations. Values for the osmotic coefficients of
and 17C lower than the true values in the 3- and NaCl, CaCh, and other electrolytes as a function
6-molal salinity samples. Calculations further of temperature and molality are given in Staples
show that these errors do not change appreciably and Nuttall (1977), Holmes et al. (1978, 1981),
with changes in temperatures from 50C to 200C. and Pitzer (1981).
The activity of water (Equation 2) can be calcu- An estimate of the errors involved in calculating
lated from the expression given by Garrels and subsurface temperatures assuming that aH,O = 1
Christ (1965) as can be made from Equations 8, 9, or Table 6.2.
These data show that aH,O decreases with increas-
aH,O = 1 - 0.017 Li mi (8)
ing salinity of water, but is essentially independent
The summation covers the molalities (mi) of all the oftemperature over the temperature range of 25C
species in solution. A more accurate value for the to 200C. Table 6.2 shows that aH,O ranges from
activity of water can be obtained from the expres- 1.00 to 0.80 as salinities increase from 0 to 6 molal
sion given by Helgeson et al. (1970) as NaCl. The activity of water is particularly impor-
tant because the solubility of quartz, as indicated
log (aH,O) = 0.00782 EVe me <Pe (9)
by Equation 2, is inversely proportional to aH,O.
where Ve is the number of moles of ions in the for- Assuming aH,O = 1 results in calculated subsur-
mula for the electrolyte (e) (e.g., vNaCI = 2, and face temperatures that are about 9C and 19C
vCaCl, = 3) and me and <Pe are the molality and lower than the true values in the 3- and 6-molal
osmotic coefficient of this electrolyte. Equations 8 NaCI solutions.
and 9 give approximately the same value for the The errors in calculated temperatures resulting
activity of water in NaCI solutions (0 to 6 molal) from assuming aH,O = 1 are additive to those
and temperatures of 25C to 200C (Table 6.2). assuming YH4SiO.O = 1 as indicated by Equation 2.
The activity of water is somewhat lower in CaCl 2 Thus, assuming aH,O = 1 and YH4Si040 = 1 results
solutions (Table 6.2). The activity of water can be in calculated temperatures that are about 18C and
approximated for the majority of oil-field waters 35C lower than the correct values in the 3- and
6. Chemical Geothermometers and Formation Waters 105

6-molal NaCl solutions at 100C. It should be 140 .----.-----.----,-----~--_,

noted that corrections for pressure effects will


somewhat reduce the magnitude of errors resulting
from assuming that aH,O and YH4Si040 are equal to
unity.
The possible total errors in estimated subsurface

...'
temperatures neglecting the effects of pressure on
the solubility of quartz and assuming that YH4Si040
and aH,O are equal to unity are shown in Figure
6.2. Using these assumptions, the calculated sub-
surface temperatures of 26 samples from the cen-
80 100 120 140 160
tral Mississippi Salt Dome basin (Kharaka et aI., MEASURED TEMPERATURE ' "C)
1986; unpublished data) and coastal Louisiana and
Texas (Kharaka et aI., 1978, 1979) are always FIGURE 6.2. Temperature calculated using uncorrected
lower than the measured values (Fig. 6.2). The quartz geothermometer versus measured subsurface
temperatures for brines from sedimentary basins. Note
selected samples have salinities that range from
that the calculated temperatures plot below the ideal line
100,000 mg/L to 330,000 mg/L of dissolved solids;
showing that they are always lower than the subsurface
the hydraulic pressures range from 200 bars to temperatures.
900 bars; and the corrected temperatures should
plot on the lines shown in Figure 6.2. The lower
calculated temperatures show that corrections for
waters is detailed in Fournier and Truesdell (1973).
the effects of YH4Si040 and aH,O are much greater
In the case of Mg-Li, an exchange reaction is writ-
in these waters than the pressure correction, which
ten of the type:
is in the opposite direction. The errors in calcu-
lated temperatures in samples with lower salini- Li+ + (0.5 Mg) Solid ;::t 0.5 Mg2+
ties, of course, will be lower than those indicated in
Figure 6.2. + (Li) Solid (II)

The equilibrium constant for this reaction at tem-


perature T (KT) is given by
New Mg-Li Geothermometer
KT = (mMg")o.s [ YK1~" a(Li) SOlid] (10)
It is generally recognized that the concentrations mLi' YLi' a(O.5 Mg) Solid
and proportions of magnesium in subsurface
waters are much lower than those in seawater and An assumption of chemical geothermometers is
generally decrease with increasing temperatures that the terms within the brackets is equal to unity,
(White, 1965; Fournier and Potter, 1979; Kharaka resulting in:
et aI., 1985). The concentrations and proportions
of lithium, on the other hand, increase with K _ (mMg)o.s
T- (11)
increasing temperatures (Fouillac and Michard, mLi
1981; Kharaka et aI., 1985), suggesting that the
magnesium-to-lithium ratio may be a sensitive The chemical geothermometers are based on the
indicator of temperature. The geochemical reasons van't Hoff equation given by
for this behavior are not fully understood (Ellis and
oInK = DJlo r
Mahon, 1977), but it is known that Mg2+ and Li+ (12)
commonly substitute for each other in amphiboles,
oT RT
pyroxenes, micas, and clay minerals. This substitu- which can be integrated (assuming !:J.Hor is con-
tion takes place mainly because the two cations stant) and simplified to
have almost identical crystalline ionic radii.
The methodology for developing chemical geo-
(13)
thermometers from chemical composition of
106 Y.K. Kharaka and R.H. Mariner

TEMPERATURE (OC) (mM )o.S) 1


350 200 100 log ( -,--=g- versus -T
3 mLi

Seawater that yields a line with a slope equal to B and an


Oil Field
2 Geothermal intercept equal to A. This geothermometer is
obtained by rearranging Equation 16 to yield:

:.J B
= - - - - - - - - - 273 (17)
~
tMg-Li
mMg)o.S)
log ( -A
CI mLi
0 0
...J

where t is temperature in 0c.


- 1
The data base for the development of the Mg-Li
and other geothermometers is shown in Table 6.3.
Reliable subsurface temperatures are generally not
-2 reported with chemical analyses of oil-field waters.
1 2 3 4 However, we have temperature and chemical data
l000fT IT In OK)
from more than 250 formation waters from about
FIGURE 6.3. Magnesium/lithium ratios as a function of 30 oil and gas fields located in Texas, Louisiana,
subsurface temperatures for oil field, geothermal, and
Mississippi, California, and Alaska (Kharaka et
seawater.
al., 1985). Fifty-four samples were selected mainly
from our own files based on: 1) reliability of
reported subsurface temperatures, 2) absence of
where T is temperature in oK, I'l.Ho r is the stan- dilution by condensed water in the case of samples
dard enthalpy of Reaction II in callmole. Equation from gas wells, and 3) relatively uniform tempera-
13 can be rearranged to give ture distribution over the range of temperatures of
oil-field waters (33C to 170C). For data from
1'l.H0r /'t;.HO r ( 1)
log KT = log K29B + 1,364 - 4.576 T (14) geothermal systems, only samples from wells were
selected. In the case of samples from wells where
Substituting for the value of log KT from Equation boiling had occurred, the reported in situ chemical
11 yields: compositions of waters were used.
The reliability of the Mg-Li geothermometer is
m M )o.S)
log ( g = log K29B + /'t;.HO r indicated by examination of Figure 6.3. An excel-
mLi 1,364 lent correlation coefficient (r) of 0.96 is obtained
for all the data in Figure 6.3. A very good correla-
(15) tion coefficient of 0.90 is obtained when the
regression is carried out using only samples from
Equation 15 can be simplified into a linear equa- oil-field waters. The correlation coefficients
tion by assuming that I'l.Hor is a constant. The obtained (see "Discussion and Recommendations")
equation with constants A and B is using other cation geothermometers are lower
than those for the Mg-Li geothermometer.
log Cm~~o.S) = A + B (~) (16) The least-squares line drawn through Figure 6.3
gives a slope of 2.20 and an intercept of -5.47.
I'l.Ho The Mg-Li geothermometer for all the data in this
where A is equal to log K298 + 1,36~ and B is equal study is given by:
I'l.Wr
to - 4.576.
2,200
tMg-Li = --(-;-(-C-M~g)-0.-C5)--- - 273 (18)
The Mg-Li geothermometer is developed from a log + 5.47
plot of CLi
?'
(')
::r-
et>
3
;:;.
e:..
TABLE 6.3. Cation concentrations (mglL) from selected oil-field and geothermal waters with known subsurface temperatures. 0
Sample no. Location Temperature (0C) Mg Ca Li Na K Reference S
::r-
et>
Seawater Average composition 8 1,350 400 0.17 10,500 380 Goldberg (1963)
.
30
83-TX-1 High Island, TX 55 869 1,370 2.0 28,400 172 Kharaka et at. (1985) 3et>
83-TX-6 do 53 499 679 .6 15,400 110 Do
G
83-TX-7 do 52 349 850 .6 13,300 103 Do
I
83-TX-9
.'"
do 49 1,010 2,250 2.2 31,800 167 Do ::s
p..
78-AK-54 Prudhoe Bay, AK 94 20 182 4.0 7,600 86 Do
78-AK-55 do 90 16 151 3.3 5,720 68 Do 61
79-00-201 Pleasant Bayou #2, TX 170 210 6,500 34 32,100 1,900 Kharaka et al. (1979) 3
79-00-204 do 154 660 9,100 39 38,000 840 Do !:?
81-00-51 Crown Zellerback #2, LA
o
::s
147 39 460 5.5 11,000 97 Kraemer and Kharaka (1986)
79-00-50 F.E Sutter #1, LA 132 670 7,670 19.0 48,300 990 Do ~
80-00-1 W. Oirouard #1, LA 134 17 115 3.5 8,700 43 Do G
79-00-251 B. Simon #2, LA 141 270 3,090 19.0 32,200 510 Do
.'"
77-00-58 E. Delcambre #1, LA 112 335 2,150 8.2 47,600 305 Do
77-00-55 do 114 270 1,850 7.2 40,800 260 Do
80-00-301 L. Koelemay #1, LA 127 6.2 27 1.8 7,280 32 Do
77-00-117 La Blanca, TX 148 3.3 150 1.2 2,680 46 Kharaka et at. (1980)
77-00-1 Erath, LA 77 800 2,570 2.1 47,000 200 Kharaka et at. (1978)
77-00-15 Bayou Sale, LA 159 920 14,800 35 51,700 467 Do
77-00-21 Weeks Island, LA 101 750 3,560 4.0 57,400 390 Do
WR-2YK74 Wheeler Ridge, CA 56 158 375 1.0 8,550 160 Kharaka (unpublished data)
WR-6YK74 do 50 106 370 1.6 4,200 81 Do
WR-8YK74 do 80 158 2,370 2.0 12,050 112 Do
WR-IOYK74 do 89 102 6,300 1.7 11,550 170 Do
81-NSV-2 Black Butte, CA 33 143 392 .19 2,220 12.1 Do
81-NSV-3 do 44 141 319 .30 7,380 28.8 Do
81-NSV-19 Orimes, CA 58 80 274 .36 8,100 40.3 Do
82-SSV-16 Suisun Bay, CA 65 76 130 1.30 5,800 36.5 Do
82-SSV-18 Rio Vista, CA 70 26 121 .44 5,250 43.5 Do
82-SSV-22 River Island, CA 49 50 215 .52 2,450 35 Do

.....
0
-..I
~

TABLE 6.3. Continued.


Sample no. Location Temperature (0C) Mg Ca Li Na K Reference
82-SSV-26 do 83 80 163 .37 2,650 65 Do
82-SSV-23 Lindsey Slough, CA 92 35 322 .94 7,400 83 Do
82-SSV-28 Union Island, CA 99 40 165 1.69 6,000 14.8 Do
82-SSV-3 do 66 49 100 .42 4,000 29 Do
82-SSV-4 do 71 17.8 76 .59 4,300 37.3 Do
82-SSV-5 Winters, CA 48 48.5 120 .32 4,550 25.5 Do
82-SSV-6 Lindsay Slough, CA 76 60 279 .62 4,600 76.5 Do
912-72 Kettleman North Dome, CA 81 206 797 1.43 13,200 88.4 Kharaka and Berry (1974)
912-97 do 98 109 1,650 2.41 12,200 174 Do
32-321 do 101 169 2,100 1.19 9,300 183 Do
912-45 do 122 2.43 88 3.26 3,800 66.5 Kharaka and Berry (1976)
912-2 do 136 3.42 30.7 3.05 3,760 94.2 Do
76-GG-2 Chocolate Bayou, TX 94 60 280 3.3 15,250 120 Do
76-GG-3 do 98 40 180 3.5 15,750 110 Do
76-GG-5 do 103 30 130 3.7 16,500 120 Do
76-GG-l do 102 35 160 3.2 14,000 90 Do
76-GG-24 Halls Bayou, TX 150 170 1,800 15 20,500 180 Do
76-GG-25 do 138 185 1,600 13 17,750 240 Do
76-GG-29 Alta Lorna, TX 119 90 700 6.0 19,500 190 Kharaka et al. (1977)
76-GG-51 White Point East, TX 61 445 2,500 5.1 29,500 215 Do
76-GG-52 do 46 475 3,040 2.6 22,750 77 Do
76-GG-54 do 68 340 2,880 5.4 25,250 125 Do
PM2a Paris Basin, France 80 141 581 1.39 4,090 74.3 Bouleque (1978) ~
CH 101 do 72 119 373 1.11 3,100 54.7 Do ~
CGEH (9) Coso, CA 205 .57 51 14.0 1,480 132 Fournier et al. (1980)
Well #1 EI Tatio, Chile 211 1.1 270 30.2 4,480 420 Cusicanqui et al. (1975)
~
~

Well #3 do 253 2.08 268 31.1 3,512 168 Do


Well #4 do 229 3.7 228 28.0 4,537 193 Do
~
~

Well #6 do 180 1.3 99 17.1 1,900 III Do Co


=
Well #5 (666) Cerro Prieto, Mexico 295 .50 505 22.9 8,000 1,900 Manon et al. (1977) :;g
fI:
::::
~
...
S
0>
...
?'
Ci
::r
~
8
o
e?.
0~
TABLE 6.3. Continued. 0
...::r
Sample no. Location Temperature ("C) Mg Ca Li Na K Reference ~
...,
80
M-25 do 256 .21 260 14.2 4,750 1,090 Fausto et al. (1979); Truesdell et al. (1979)
M-29 do 260 .85 314 15.5 4,830 850 Do 8~
~
M-19a do 304 .24 301 15.7 5,240 1,170 Do
......,
Well #2 Broadlands, New Zealand 260 .1 2.2 11.7 1,050 210 Mahon and Finlayson (1972) '"I>:>
::s
p..
Well #3 do 281 .8 3.0 12.2 1,045 213 Do
Well #9 do 294 .35 2.0 12.7 930 203 Do 61
Well #9 Otake, Japan 248 .19 12.3 5.2 936 131 Koga (1970) ae;.
Well #7 do 230 .025 9.9 4.5 846 105 Do
Well #1 Hatchobaru, Japan 300 .158 9.9 ILl 1,396 289 Do
o
::s
Well #1 Salton Sea, CA 340 54 28,000 215 50,400 17,500 Muffler and White (1969)
Well #2 do 300 10 28,800 210 53,000 16,500 White (1968)
~
G
...,
Utah State 14-2 Roosevelt, UT 268 .28 II 25 2,070 384 Capuano and Cole (1982)
Nathenson et al. (1982)
'"
Schmitt Well Raft River, ID 147 .24 50 1.4 535 22
7-0-009-81 Calistoga, CA 135 .5 2.0 1.95 206 9 Murray et al. (1985)
272 Klamath Falls, OR 148 .05 35 .38 230 7.4 Janik et al. (1985)
OT03RM79 Brady #7, NV 154 .1 41 2.0 900 67 Mariner (unpublished data)
OT06RM79 Brady #8, NV 154 .32 45 1.5 850 36 Do
Y-8 Yellowstone, WY 169 .04 1.2 2.6 360 15 Fournier (1981)
East Mesa 6-1 Imperial Valley, CA 190 15.2 1,020 45 9,130 1,180 Howard et al. (1978)
East Mesa 6-2 do 182 .3 13 4.0 1,400 125 Do
East Mesa 8-11 do 170 1.6 41 2.0 723 42 Do
Reynolds #1 Stillwater, NV 136 1.7 110 1.9 1,500 42 Mariner et al. (1974, 1975)
Chevron 1-29 Carson Desert, NV 199 .36 58 2.3 1,200 130 Olmsted et al. (1984)
Well #8 Kawerau, New Zealand 262 .39 Ll 5.5 740 130 Ellis and Mahon (1977)
Well #137 Rotorua, New Zealand 160 .22 1.0 1.4 565 31 Do
Well #1 Ngauha, New Zealand 225 1.4 29 10.7 900 78 Do
Well #24 Wairakei, New Zealand 250 .04 12 13.2 1,250 210 Do
Well #IA Kizildere, Turkey 200 .2 2.5 4.5 1,280 135 Do


110 Y.K. Kharaka and R.H. Mariner

TEMPERATURE (OC) TEMPERATURE (OC)


350 200 100 0
3 r-__~3T~~~2r OO~______0~~
OO~~lr 3 r---~---'~--~-------r~

Seawater Seawater
Oil Field Oil Field
Geothermal Geothermal

z'"
2 ... ~2
<.::l
o
..J
>C
III
+
Q
~1
<.::l
o
..J

O ~---- __ ~ ______ ~ ______ ~ O ~---- __~______~______~


1 2 3 4 1 2 3 4
l000/T IT in OK) l000/T (T in OK)

FIGURE 6.4. Sodium/potassium ratios as a function of FIGURE 6.5. Sodium/potassium/calcium ratios as a func-
subsurface temperatures for oil field, geothermal, and tion of subsurface temperatures for oil field, geother-
seawater. Note the scatter in data for waters from low- mal, and seawater. Note the large scatter in the data for
temperature environments. low-temperature environments.

where C is the concentration in mg/L of the sub- when this geothermometer is applied to geother-
scripted cation. A least-squares line was also mal waters alone (Fig. 6.5 and Table 6.4). The
drawn through samples from oil-field waters alone; Na-K-Ca geothermometer with a Mg correction
the equation is: (Fournier and Potter, 1979) gives excellent results
for all waters (r = 0.95) (Table 6.4). The calcu-
1,910
tMg-Li = ----(C-M-g'-)O-.s---- - 273 (19) lations of the corrections are somewhat cumber-
log + 4.63 some, but are critical in the case oflow-temperature
CLi waters with high concentrations of Mg.
The Mg-corrected Na-K-Ca geothermometer is
based on the temperature estimated from the Na-
Other Cation Geothermometers K-Ca geothermometer and a variable R given by:

The data in Table 6.3 were used further to test the R = CMg X 100 (20)
applicability of other chemical geothermometers
CMg + 0.61Cca + 0.31CK
to oil-field waters. Results are shown in Figures 6.4 where concentrations are expressed in mg/L. For R
to 6.8 and in Table 6.4. These results show that the values from 5 to 50 the correction is given by
Na-K geothermometer (Fig. 6.4) cannot be used to
calculate reliable subsurface temperatures for oil- ~tMg = 10.66 - 4.7415 R + 325.87 (log R)2
field waters. This geothermometer is much more - 1.032 X 105 (log R)2/T - 1.968
reliable when applied to geothermal systems, espe-
cially at temperatures higher than about 150C. X 107 (log R)2/P + 1.605
The Na-K-Ca geothermometer gives totally ran- X 107 (log R)3/P (21)
dom results (r = 0.00) when applied to oil-field
waters. Much better results are obtained (r = 0.88) and for R from 0.5 to 5
0-

n
::r
o
S.
o
~

~
o
g-
o
TABLE 6.4. Chemical geothermometers recommended for use in formation waters from sedimentary basins. a
o
Correlation coefficients (r) ao
(D
...,
Geothermometer Equation (concentrations in mg/L) Oil-field waters All waters Recommendations on
po
1309 5-
I. Quartz t = - 273
0.41 - log (K' pfJ 61
K and pi from Equations 2 and 4 (see text) 70C to 250C
a
a
o
::s
1032
2. Chalcedony t = - 273 30C to 70C
-0.09 - log (K' pfJ ~
2200 &
on
3. Mg-Li t = - 273 0.90 0.96 OC to 350C
10g(YMg/Li) + 5.47
1180
4. Na-K t = - 273 0.40 0.87 Do not use in oil-field waters
10g(Na/K) + 1.31
699
5. Na-K-Ca t = - 273 0.00 0.63 Do not use in oil-field waters
10g(Na/K) + 13 [log(v'CalNa) + 2.06] + 0.489

6. Mg-corrected Na-K-Ca Same as Na-K-Ca (Table 6.1) with Mg-corrections (see text) 0.83 0.95 OC to 350C
1120
7. (13 = 113) t= . - 273 0.58 0.86 Use only where no Mg data
10g(Na/K) + 113 [log(v'CalNa) + 2.06] + 1.32
1590
8. Na-Li t = - 273 0.80 0.91 OC to 350C
10g(Na/Li) + 0.779

.....
.....
112 Y.K. Kharaka and R.H. Mariner

when the regression is carried out using oil-field


waters alone.

p
300
..... An improved correlation (Fig. 6.7) is obtained
with the Na-K-Ca geothermometer when all the
w . samples are plotted using 13 = 0.333 instead of 13 =

.. ..
a:
::J 1.333 for the samples with calculated temperatures
l-
e{ less than 100C. A good correlation coefficient of

..
a: 200
w 0.86 (Table 6.4) is obtained for all the waters, but
Go
::E the correlation coefficient obtained for oil-field
w
I-
U
0> waters alone is still a poor 0.58.
~ t ..'&
The Na-Li geothermometers developed inde-
... 'I' _. . . . . .
~ 100 pendently by Fouillac and Michard (1981) and
g,
::E
\ t
. <
.6 . . . . ..

Kharaka et al. (1982) give better results (Fig. 6.8;


Table 6.4) with oil-field waters than with waters
from geothermal systems. A good correlation
coefficient of 0.80 is obtained with oil-field waters
100 200 300 alone using Kharaka et al.'s equation (Table 6.4).
MEASURED TEMPERATURE (OC)

FIGURE 6.6. Temperature calculated using magnesium-


corrected Na-K-Ca geothermometer versus measured Discussion and Recommendations
temperatures for all samples in the study.
Several chemical geothermometers can be used to
estimate the subsurface temperatures of formation
~tMg = 1.03 + 59.971 log R waters from sedimentary basins. The silica geo-
+ 145.05 (log R)2 thermometers are based on rigorous thermody-
namic principles and should give the best estimates
- 36711 (log R)2IT of subsurface temperatures. Silica geothermome-
- 1.67 X 107 log RIP (22) ters, however, may be ambiguous because the
water may be in equilibrium with quartz or one of
where ~tMg is the temperature correction in C its polymorphs. The quartz geothermometer is
that should be subtracted from the Na-K-Ca tem- generally applicable to waters from sedimentary
perature and T is the temperature in OK calculated basins with temperatures higher than about 70C
from Na-K-Ca geothermometer. The Mg-corrected (Kharaka et aI., 1977). At temperatures lower than
Na-K-Ca temperature (tMg) is given by 70C the chalcedony geothermometer gives the
best results for oil-field waters. A pressure correc-
tMg = tNa-K-Ca - ~tMg (23)
tion, given by Equation 4 and corrections for the
The Mg correction is negligible where R values activity coefficient of H 4 Si04 (Equation 7) and
are less than 0.5. No correction is available for activity of water (Equation 8 or 9), should be
waters with R values greater than 50, and the Na- applied to silica geothermometers where hydraulic
K-Ca geothermometer should not be applied to pressures are higher than 200 bars and salinities
these waters. Rarely, ~tMg may be negative; in this are higher than 10,000 mg/L of dissolved solids.
case, use the temperature calculated from the Na- Chemical geothermometers based on the con-
K-Ca geothermometer, ignoring the magnesium centrations of cations are all empirical and not
correction. related to the solubility of any single mineral. The
Temperatures calculated using Na-K-Ca geother- Mg-Li geothermometer can be applied success-
mometer with Mg correction are plotted in Figure fully to low-temperature systems, including the
6.6 against the subsurface temperatures (Table ocean, indicating that relatively fast exchange
6.3). The least-squares line through all the samples reactions involving clay minerals may control
gives an excellent correlation coefficient of 0.95. these ratios. For this chapter, however, we have
A greatly improved (compared to no Mg correc- made no attempt to identify the specific clay min-
tion) correlation coefficient of 0.83 is obtained eral(s) responsible for the Mg-Li geothermometer.
6. Chemical Geothermometers and Formation Waters 113

TEMPERATURE (OCI TEMPERATURE (OCI


350 200 100 o 350 200 100 0
3r----.---,----,--------.7I 5r----,----,-----r-------~__,

Seawater Seawater
Oil Field 011 Field
Geothermal Geothermal
z'" 4

~_2
Cl
g
)(

t!

o~~ ____ _________ ______


~ ~ ~

1 2 3 4 2 3 4
l000/T (T in KI l000/T (T In KI

FIGURE 6.7. Sodium/potassium/calcium ratios as a func- FIGURE 6.8. Sodiumllithium ratios as a function of sub-
tion of subsurface temperature for oil field, geothermal, surface temperatures for oil field, geothermal, and sea-
and seawater. Note the improved correlation obtained water.
for oil-field waters where 13 = 0.333 is used as compared
to that shown in Figure 6.5.
study (Table 6.4). A regression on temperatures
calculated by this procedure and those measured
The Mg-Li geothermometer gives improved re- gives a correlation coefficient of only 0.58.
sults over the Mg-corrected Na-K-Ca geothermom- However, the use of this geothermometer may be
eter for both geothermal and oil-field waters; the necessary in oil-field waters where Mg and Li con-
improvement is dramatic for oil-field waters. The centrations are not available.
improved results obtained with the Mg-Li geother- The Na-K geothermometer gives poor results for
mometer are for samples obtained from wells. oil-field waters. The Na-Li geothermometer (Table
Additional testing is needed to evaluate the appli- 6.4) , on the other hand, gives good results that are
cation of this geothermometer to samples from hot comparable to the Mg-corrected Na-K-Ca geother-
springs (Table 6.4). Temperatures calculated for mometer. The Na-Li geothermometer is recom-
oil-field samples (Table 6.3) using Equations 18 mended for use in both oil-field and geothermal
and 19 are slightly different. However, exactly the waters, but gives slightly better results for oil-field
same correlation coefficient (r = 0.90) was waters (Table 6.4).
obtained when these temperatures were regressed The ratios of cations (Table 6.3) were combined
against the subsurface temperatures. We suggest in equations other than those shown in Table 6.4 to
that Equation 18 be used for all subsurface waters investigate other possible geothermometers. The
because Equation 19 cannot be used reliably for best results were obtained by combining the mag-
geothermal waters with temperatures higher than nesiumllithium ratios with the magnesium/sodium
about 200C. ratios in the relationship
The uncorrected Na-K-Ca geothermometer can-
not be used to calculate subsurface temperatures (C )0.5)
lower than about 200C in sedimentary basins. log ( ~~i + 0.25 log
Improved, but still poor, results are obtained with
the Na-K-Ca geothermometer if 13 = 0.333 is The correlation coefficient obtained for the oil-
always used with the coefficients derived in this field samples using the above relationship, how-
114 Y.K. Kharaka and R.H. Mariner

ever, was lower (r = 0.85 versus r = 0.90) than Capuano, R.M., and Cole, D.R. 1982. Fluid-mineral
that obtained using the magnesium/lithium ratios equilibria in a hydrothermal system, Roosevelt Hot
alone. Springs, Utab. Geochimica et Cosmochimica Acta
Chemical geothermometers should not be ap- 46: 1353-1364.
plied to chemical analyses of waters from gas wells Carpenter, A.B., Trout, M.L., and Pickett, E.E. 1974.
Preliminary report on the origin and chemical evolu-
as these waters are often diluted by condensed
tion of lead- and zinc-rich brines in central Missis-
water vapor produced with natural gas. Kharaka sippi. Economic Geology 69:1191-1206.
and associates (1977, 1985) have shown that chem- Chen, c.A., and Marshall, W. L. 1982. Amorphous silica
ical analysis from many gas wells, especially those solubilities: IV. Behavior in pure water and aqueous
from geopressured geothermal systems, are diluted sodium chloride, sodium sulfate, magnesium chloride,
and do not represent the true chemical composi- and magnesium sulfate solutions up to 350C. Geo-
tion of formation water from that production zone. chimica et Cosmochimica Acta 46:279-287.
The criteria to distinguish such samples are given Cusicanqui, H., Mahon, W.A.1., and Ellis, A.1. 1975.
by these authors, but temperatures calculated from The geochemistry of the EI Tatio geothermal field,
these samples are erratic and generally lower than Northern Chile. Proceedings of the Second United
the measured temperatures. Nations Symposium on the Development and Use of
Geothermal Resources, San Francisco, vol. 1, pp.
All the chemical geothermometers recom-
703-711.
mended in this study are incorporated in a modi-
D'Amore, F., and Panichi, C. 1980. Evaluation of deep
fied version ofthe geochemical model SOLMNEQ temperature of hydrothermal systems by a new gas
(Kharaka and Barnes, 1973). The program com- geothermometer. Geochimica et Cosmochimica Acta
putes the activity of H 20 and H4 Si02 and makes 44:549-556.
the required pressure correction in the case of sil- Drummond, S.E., Jr. 1982. Boiling and mixing of hydro-
ica geothermometers. Models such as this can be thermal fluids: Chemical effects on mineral precipita-
used to accurately determine subsurface tempera- tion. Ph.D. thesis, Pennsylvania State University,
tures from the chemical composition of formation University Park, PA, 380 pp.
water and solubilities of individual minerals known Ellis, A.1. 1970. Quantitative interpretation of chemical
to be present in the aquifer. characteristics of geothermal systems. Geothermics
2:516-528.
Ellis, A.J., and Golding, R.M. 1963. The solubility of
carbon dioxide above 100C in water and in sodium
Acknowledgments. We would like to thank D.I chloride solutions. American Journal of Science 261:
Specht who worked with the senior author on an 47-60.
earlier version of the Mg-Li geothermometer. We Ellis, A.1., and Mahon, W.A.1. 1977. Chemistry and
would also like to thank our colleagues R.o. Four- Geothermal Systems. New York, Academic Press,
nier, ID. Hem, N.D. Naeser, and D.K. Nordstrom 392 pp.
for a thorough review of this manuscript. Fausto, 1.1., Sanchez, A.A., Jimenez, M.E.S., Esquer,
I.P., and Ulloa, F.H. 1979. Hydrothermal geochemis-
try ofthe Cerro Preito geothermal field. Second Sym-
References posium on the Cerro Prieto Geothermal Field, Baja
California, Mexico, pp. 199-223.
Arnorsson, S. 1975. Application of the silica geother- Fouillac, c., and Michard, G. 1981. Sodiumllithium
mometer in low temperature hydrothermal areas in ratio in water applied to geothermometry of geother-
Iceland. American Journal of Science 275:763-784. mal reservoirs. Geothermics 10:55-70.
BouIegue, 1. 1978. Metastable sulfur species and trace Fournier, R.D. 1973. Silica in thermal water: Laboratory
metals (Mn, Fe, Cu, Zn, Cd, Pb) in hot brines from and field investigations. In: Proceedings of the Inter-
the French Dogger. American Journal of Science national Symposium on Hydrogeochemistry and Bio-
278: 1394-1411. geochemistry, Japan, 1970: Vol. 1. Washington, DC,
Brook, C.A., Mariner, R.H., Mabey, D.R., Swanson, The Clark Company, pp. 122-139.
1.R., Guffanti, M., and Muffler, L.1.P. 1979. Hydro- Fournier, R.D. 1979. A revised equation for the Na/K
thermal convection systems with reservoir tempera- geothermometer. Geothermal Resources Council
tures ~90C. In: Muffler, L.1.P. (ed.): Assessment of Transactions 3:221-224.
Geothermal Resources of the United States- 1978. Fournier, R.O. 1981. Application of water chemistry to
U.S. Geological Survey Circular 790, pp. 18-85. geothermal exploration and reservoir engineering. In:
6. Chemical Geothermometers and Formation Waters 115

Rybach, L., and Muffler, L.lP. (eds.): Geothermal Hitchon, B., Billings, G.K., and Klovan, lE. 1971.
Systems: Principles and Case Histories. New York, Geochemistry and origin of formation waters in the
Wiley, pp. 109-143. western Canada sedimentary basin: III. Factors con-
Fournier, RD., and Potter, R.W., II. 1979. A magnesium trolling chemical composition. Geochimica et Cos-
correction for the Na-K-Ca geothermometer. Geo- mochimica Acta 35:567-598.
chimica et Cosmochimica Acta 43: 1543-1550. Holmes, H.F., Baes, e.F., Jr., and Mesmer, R.E. 1978.
Fournier, R.o., and Potter, R.W., II. 1982. An equation Isopiestic studies of aqueous solutions at elevated tem-
correlating the solubility of quartz in water from 25 peratures: I. KCI, CaCI 2 , and MgCI 2 Journal of Chem-
to 900C at pressures up to 10,000 bars. Geochemica ical Thermodynamics 10:983-996.
et Cosmochimica Acta 46:1969-1973. Holmes, H.F., Baes, e.F., Jr., and Mesmer, R.E., 1981.
Fournier, R.o., and Rowe, J.l 1966. Estimation of Isopiestic studies of aqueous solutions at elevated tem-
underground temperatures from the silica content of peratures: III. (1-y) NaCI + yCaCl 2 Journal of Chem-
water from hot springs and wet-steam wells. Ameri- ical Thermodynamics 13:101-113.
can Journal of Science 264:685-697. Howard, lH., Apps, lA., Benson, S.M., Goldsten,
Fournier, R.o., Thompson, J.M., and Austin, e.F. 1980. N.E., Graf, A.N., Haney, J.P., Jackson, D.D., Jupra-
Interpretation of chemical analyses of waters collected sert, S., Majer, E.L., McEdward, D.G., McEvilly,
from two geothermal wells at Coso, California. Jour- T.V., Narasimhan, T.N., Schechter, B., Schroeder,
nal of Geophysical Research 85:2405-2410. R.e., Taylor, P.e., van de Kamp, P.C., and Wolery, T.l
Fournier, R.o., and Truesdell, A.H. 1973. An empirical 1978. Geothermal resource and reservoir investiga-
Na-K-Ca chemical geothermometer for natural wa- tions of U.S. Bureau of Reclamation Leasehold at East
ters. Geochemica et Cosmochimica Acta 37:1255- Mesa, Imperial County, California. Berkeley, CA,
1275. Lawrence Berkeley Laboratory, University of Califor-
Fournier, R.o., and Truesdell, A.H. 1974. Geochemical nia, Report LBL-7094, 305 pp.
indicators of subsurface temperatures: Part 2. Estima- Janik, C.J., Truesdell, A.H., Samrnel, E.A., and White,
tion of temperature and fraction of hot water mixed A.F. 1985. Chemistry of low-temperature geothermal
with cold water. Journal of Research of the U.S. Geo- waters at Klamath Falls, Oregon. Geothermal Re-
logical Survey 2:263-270. sources Council Transactions 9(1):325-331.
Fournier, R.o., White, D.E., and Truesdell, A.H. 1974. Kharaka, YK., and Barnes, I. 1973. SOLMNEQ:
Geochemical indicators of subsurface temperatures: Solution-mineral equilibrium computations. Spring-
Part 1. Basic assumptions. Journal of Research of the field, VA, U.S. Department of Commerce, NTIS
U.S. Geological Survey 2:259-262. Report PB 215-899, 81 pp.
Garrels, R.M., and Christ, e.L., 1965. Solutions, Kharaka, YK., and Berry, F.A.F. 1974. The influence of
Minerals, and Equilibria. New York, Harper & Row, geological membranes on the geochemistry of subsur-
450 pp. face waters from Miocene sediments at Kettleman
Giggenbach, w.F. 1980. Geothermal gas equilibria. North Dome, California. Water Resources Research
Geochimica et Cosmochimica Acta 44:2021-2032. 10:313-327.
Goldberg, E. D. 1963. Chemistry - the oceans as a chem- Kharaka, Y.K., and Berry, F.A.F.1976. The influence of
ical system. In: Hill, M.N. (ed.): The Sea: Vol. 2. geological membranes on the geochemistry of subsur-
Composition of Seawater, Comparative and Descrip- face waters from Eocene sediments at Kettleman
tive Oceanography. New York, Interscience, pp. 3-25. North Dome, California: An example of effluent-type
Helgeson, H.e. 1969. Thermodynamics of hydrother- waters. In: Cadek, l, and Paces, T. (eds.): Proceed-
mal systems at elevated temperatures and pressures. ings of the International Symposium on Water-Rock
American Journal of Science 267:729-804. Interaction, Prague, 1974. Prague, Czechoslovakia,
Helgeson, H.e., Brown, T.H., Nigrini, A., and Jones, The Geological Survey, pp. 268-277.
T.A. 1970. Calculation of mass transfer in geochemi- Kharaka, YK., Brown, P.M., and Carothers, W.W.
cal processes involving aqueous solutions. Geochim- 1978. Chemistry of waters in the geopressured zone
ica et Cosmochimica Acta 34:569-592. from coastal Louisiana: Implications for the geother-
Henley, RW., Truesdell, A.H., and Barton, P.B., Jr. mal development. Geothermal Resources Council
1984. Fluid-mineral equilibria in hydrothermal sys- Transactions 2:371-374.
tems. Reviews in Economic Geology: Vol. 1. Chelsea, Kharaka, YK., Callender, E., and Carothers, W.W.
MI, Society of Economic Geologists, 267 pp. 1977. Geochemistry of geopressured geothermal
Hitchon, B. 1985. Geothermal gradients, hydrodynam- waters from the Texas Gulf Coast. Proceedings, Third
ics, and hydrocarbon occurrences, Alberta, Canada. Geopressured-Geothermal Energy Conference, Uni-
American Association of Petroleum Geologists 68: versity of Southwestern Louisiana, Lafayette, Lou-
713-743. isiana: Vol. I, pp. G1121-G1165.
116 Y.K. Kharaka and R.H. Mariner

Kharaka, Y.K., Hull, R.w., and Carothers, w.w. 1985. Nevada and Oregon. U.S. Geological Survey Open-
Water-rock interactions in sedimentary basins. In: File Report, 27 pp.
Relationship of Organic Matter and Mineral Diagene- Mariner, R.H., Rapp, lB., Willey, L.M., and Presser,
sis. Society of Economic Paleontologists and Mineral- T.S. 1974. Chemical composition and estimated mini-
ogists Short Course 17. Center for Energy Studies, mum thermal reservoir temperatures of the principal
University of Southwestern Louisiana, pp. 79-176. hot springs of northern and central Nevada. U.S. Geo-
Kharaka, Y.K., Lico, M.S., and Carothers, W.w. 1980. logical Survey Open-File Report, 32 pp.
Predicted corrosion and scale-formation properties of Marshall, w.L. 1980. Amorphous silica solubilities: III.
geopressured-geothermal waters from the northern Activity coefficient relations and predictions of solu-
Gulf of Mexico Basin. Journal of Petroleum Technol- bility behavior in salt solutions, 0-350C. Geochem-
ogy 32:319-324. ica et Cosmochimica Acta 44:925-931.
Kharaka, Y.K., Lico, M.S., and Law, L.M. 1982. McKenzie, W.E, and Truesdell, A.H. 1977. Geother-
Chemical geothermometers applied to formation mal reservoir temperatures estimated from the oxygen
waters, Gulf of Mexico and California basins (abst.). isotope compositions of dissolved sulfate and water
American Association of Petroleum Geologists Bulle- from hot springs and shallow drill holes. Geothermics
tin 66:588. 5:51-62.
Kharaka, Y.K., Lico, M.S., Wright, V.A., and Caro- Meyer, H.J., and McGee, H.W. 1985. Oil and gas fields
thers, W.W. 1979. Geochemistry of formation waters accompanied by geothermal anomalies in Rocky
from Pleasant Bayou No.2 well and adjacent areas in Mountain region. American Association of Petroleum
coastal Texas. In: Dorfman, N.H., and Fisher, W.L. Geologists Bulletin 69:933-945.
(eds.): Proceedings, Fourth United States Gulf Coast Mizutani, Y. 1972. Isotope composition and under-
Geopressured-Geothermal Energy Conference: Re- ground temperature of the Otake geothermal water,
search and Development. Austin, TX, University of Kyushu, Japan. Geochemical Journal 6:67-73.
Texas at Austin, pp. 178-193. Muffler, L.lP., and White, D.E. 1969. Active metamor-
Kharaka, Y.K., Maest, A.S., Fries, T.L., Law, L.M., phism of upper Cenozoic sediments in the Salton Sea
and Carothers, w.w. 1986. Geochemistry of lead and geothermal field and the Salton Trough, southeastern
zinc in oil field brines: Central Mississippi Salt Dome California. Geological Society of America Bulletin
basin revisited. Proceedings of Conference on the 80:157-182.
Genesis of Stratiform Sediment-Hosted Pb-Zn Depos- Murray, K.S., Jonas, M.L., and Lopez, C.A. 1985.
its, Stanford University, California, pp. 50-54. Geochemical exploration of the Calistoga geothermal
Koga, A. 1970. Geochemistry of the waters discharged resource area, Napa Valley, California. Geothermal
from drillholes in the Otake and Hatchobaru Areas. Resources Council Transactions 9(1):339-344.
Geothermics (Special Issue 2),2(2):1422-1425. Nathenson, M., Nehring, N.L., Crosthwaitie, E.G.,
Kraemer, T.E, and Kharaka, Y.K. 1986. Uranium Harmon, R.S., Janik, C.l, and Borthwick, J. 1982.
geochemistry in U.S. Gulf Coast geopressured- Chemical and light-stable isotope characteristics of
geothermal systems. Geochimica et Cosmochimica waters from the Raft River Geothermal Area and envi-
Acta 50: 1440-1455. rons, Cassia County, Idaho; Box Elder County, Utah.
Mahon, W.A.l, and Finlayson, lB. 1972. The chemis- Geothermics 11(4):215-237.
try ofthe Broadlands geothermal area, New Zealand. Nehring, N.L., and DJ\more, E 1984. Gas chemistry
American Journal of Science 232:48-68. and thermometry of the Cerro Prieto, Mexico, geo-
Manon, A.M., Mazor, E., Jimenez, M.E.S., Sanchez, thermal field. Geothermics 13:75-89.
A.A., Faustu, ll, and Zenizo, C. 1977. Extensive Olmstead, EH., Welch, A.H., Van Denburgh, A.S., and
geochemical studies in the geothermal field of Cerro Ingebritsen, S.E. 1984. Geohydrology, aqueous
Prieto, Mexico. Berkeley, CA, Lawrence Berkeley geochemistry, and thermal regime of the Soda Lakes
Laboratory, University of California, Report LBL- and Upsal Highback geothermal systems, Churchill
7019, 113 pp. County, Nevada. U.S. Geological Survey Water
Mariner, R.H., Brook, C.A., Reed, M.l, Bliss, lD., Resources Investigations Report 84-4054, 166 pp.
Rapport, A.L., and Lieb, R.l 1983. Low-temperature Ovnatanov, S.T., and Tamrazyan, G.P. 1970. Thermal
geothermal resources in the western United States. In: studies in subsurface structural investigations, Apse-
Reed, M.l (ed.): Assessment of Low-Temperature ron Peninsula, Azerbaijan, USSR. American Associa-
Geothermal Resources of the United States-1982. tion of Petroleum Geologists Bulletin 54: 1677-1685.
U.S. Geological Survey Circular 892, pp. 31-50. Pitzer, K.S. 1981. Characteristics of very concentrated
Mariner, R.H., Presser, T.S., Rapp, lB., and Willey, aqueous solutions. In: Rickard, D.T., and Wickman,
L.M. 1975. The minor and trace elements, gas and EE. (eds.): Chemistry and Geochemistry of Solutions
isotope compositions of the principal hot springs of at High Temperatures and Pressures. Physics and
6. Chemical Geothermometers and Formation Waters 117

Chemistry of the Earth: Vol. 13-14. New York, Perga- cisco: Vol. 1. Berkeley, CA, University of California,
mon Press, pp. 249-272. pp.53-78.
Sakai, H., and Matsubaya, O. 1974. Isotope geochemis- Truesdell, A.H., Thompson, J.M., Coplen, T.B., Nehr-
try of the thermal waters of Japan and its bearing on ing, N.L., and Janik, c.J. 1979. The origin of Cerro
the Kuroko ore solutions. Economic Geology 69: Prieto geothermal brine. Second Symposium on the
974-991. Cerro Prieto Geothermal Field, Baja California, Mex-
Staples, B.R., and Nuttall, R.L. 1977. The activity and ico, pp. 224-240.
osmotic coefficients of aqueous calcium chloride at White, D.E. 1965. Saline waters of sedimentary rocks.
298.l5K. Journal of Physical Chemistry Reference In: Young, A., and Galley, G.E. (eds.): Fluids in Sub-
Data 6:385-407. surface Environments. American Association of Pe-
Truesdell, A.H. 1976. Geochemical techniques in troleum Geologists Memoir 4, pp. 342-366.
exploration, summary of section III. Proceedings of White, D.E. 1968. Environments of generation of some
the Second United Nations Symposium on the Devel- base-metal ore deposits. Economic Geology 63:
opment and Use of Geothermal Resources, San Fran- 301-335.

Anda mungkin juga menyukai