Sie sind auf Seite 1von 15

Clay Minerals (1988) 23, 471-485

CATION SITE OCCUPANCY IN CHLORITES AND


I L L I T E S AS A F U N C T I O N OF T E M P E R A T U R E

M. CATHELINEAU

CREGU and G. S. CNRS-CREGU, B.P. 23, 54501 Vandoeuvre-les-Nancy Cedex, France

(Received October 1987; revised 9 May 1988)

A B ST RA C T : The relationships between the composition and the crystallization temperature


of chlorites and illites have been investigated in different geothermal fields and in particular the
Los Azufres system in Mexico, considered to be a natural analogue to experimental laboratories,
as the main changes in physical and chemical conditions and mineralogy are related to
progressively increasing temperature with depth. Temperature was estimated from combined
geothermometric approaches, and especially from fluid inclusion studies on quartz coexisting
with clays. The Al(lv)content in the tetrahedral site of chlorites, and the K content and total
interlayer occupancy of illites increase with temperature. These chemical changes are mainly
related to the marked decrease in the molar fraction of the Si(lv)-rich end-members (kaolinite
for chlorites, and pyrophyllite for illites) which become negligible at ~ 300~ Other chemical
changes, such as the variation in Fe and Mg contents, are partly influenced by temperature, but
are strongly dependent on the geological environment, and consequently on the solution
composition. The empirical relationships between chemical variables and temperature were
calibrated from 150-300~ but extrapolations at lower and higher temperatures seem possible
for chlorites. Such geothermometers provide tools for estimating the crystallizationtemperature
of the clays, and are important for the study of diagenetic, hydrothermal and low-T meta-
morphic processes.

Clay minerals are the most a b u n d a n t minerals in most of the geological environments
submitted to temperatures from 50 ~ to 350~ where estimation of the crystallization
temperature of the clay minerals may be difficult. Most of the classical geothermometers
cannot be applied, fluid inclusions m a y be absent, and experimental data are scarce.
Nevertheless, such temperature estimations are of critical importance for geological studies
related to oil field exploration, as well as for diagenetic and low-T metamorphic processes.
1"here are three types of chemical or crystallographic change occurring in the clay fraction of
rocks which m a y indicate temperature of crystallization.
(1) The changes in the clay mineral assemblages described in studies of the burial
m e t a m o r p h i s m of sediments (Weaver, 1959; Dunoyer de Segonzac, 1970; Perry & Hower,
1970; Velde, 1977) and the thermal m e t a m o r p h i s m of rocks in geothermal fields (Schoen &
White, 1966; Steiner, 1968; Tomason & Kristmannsdottir, 1972; Cathelineau & Izquierdo,
1988). However, only the boundaries between two mineral assemblages m a y yield a
temperature estimate, largely restricting their use.
(2) The changes in the crystallographic features of a particular clay mineral, such as those
observed in the X-ray diffraction (XRD) traces of progressively ordered mixed-layered
minerals (Reynolds & Hower, 1970; Velde, 1977, 1985), or the crystallinity index of illite
(Kubler, 1967, 1969; Frey et al., 1980).

9 1988 The Mineralogical Society


472 M. Cathelineau
(3) The change in the chemical composition of a clay mineral considered as a solid solution
between different end-members. In this instance, the clay mineral may not result from any
mechanical or structural mixing of distinct clay minerals. The search for a geothermometric
relationship then consists of the identification of changes occurring in the chemical
composition of the clay as a function of temperature. Such variations in composition have to
be independent of fluctuations in the fluid composition.
The aim of this paper is to summarize and update the different data describing the
variations of the solid solution (illite and chlorite) composition with temperature. The
published data on chlorites and illites from the Los Azufres geothermal system (Cathelineau
& Nieva, 1985; Cathelineau & Izquierdo, 1988) have been completed by new analyses of
chlorites, whilst a critical re-examination of the temperature estimation has been carried out
in the light of new fluid inclusion data on the analysed samples. These results have been
compared with data from the literature, especially those from McDowell & Elders (1980) on
the Salton sea geothermal system, and Bishop & Bird (1987) on the Coso Host Springs.

MATERIALS AND METHODS

Geological setting of the geothermal fields


The Los Azufres system. Basic data were obtained in the Los Azufres geothermal field
(Gutierrez & Aumento, 1982; Cathelineau et al., 1985, 1987) in the Quaternary volcanic belt
in Mexico. This field has been considered as a natural experimental system, and has been
chosen for the thermometric calibration of clay geothermometers for many reasons. Firstly,
the temperature ranges from 100~ to 300~ and increases regularly with depth. Secondly,
rocks are either monotonous andesites or rhyolites, this giving two extreme lithogeochemical
environments but no strong variations in the rock composition or mineralogy within a given
unit. Furthermore, the study of a young hydrothermal system in a volcanic environment
obviates the difficulty of identification of authigenic minerals. Thus, as no inherited
phyllosilicate is observed, there is no ambiguity about the recent and hydrothermal origin of
the clay minerals studied.
Cuttings and cores were studied from more than 15 deep wells, where the following clay
assemblages were observed (Cathelineau & Izquierdo, 1988) with increasing temperature:
(i) in andesites--smectite, iUite + chlorite, and chlorites at greater depth;
(ii) in rhyolites--illites are dominant and are replaced by muscovite at depth.
XRD studies have shown that the minerals studied are pure illites or chlorites without any
detectable mixed-layered minerals.

Other geothermal fields


The Salton Sea geothermal system is located in the Salton Trough, an extension of the Gulf
of California rift valley filled with fluvial-deltaic sediments from the Colorado river
(Helgeson, 1968; Elders, 1979). A high thermal gradient and high salinity geothermal fluids
characterize the Salton Sea system. McDowell & Elders (1980) studied the mineralogy and
crystal-chemistry of authigenic chlorites and illites in samples from seven depths of the
borehole Elmore 1. Temperature data are mean temperatures of seven surveys over an eight
month period in the undisturbed well (Helgeson, 1968). These temperatures are considered
to represent a good approximation of the actual geothermal gradient, as well as the
Chlorite and illite geothermometers 473
crystallization temperature of the layer-silicates, and are in good agreement with fluid-
inclusion derived temperatures (Freckman, 1978). Bishop & Bird (1987) have published
results on illites from the Coso hot springs system, a geothermal system in the Sierra Nevada
mountains in a complex series of rhyolite domes affecting a crystalline basement of intrusive
and metamorphic rocks. Analyses have been made on four samples from the well 16-8, for
which a temperature estimation has been made from actual direct measurements.

Electron microprobe analysis


Systematic analyses of the minerals were carried out on an automatic electron microprobe
(Camebax) at the Nancy I University (France), using an acceleration voltage of 15 kV,
counting time of 6 s, and an excitation current of 6 nA. A ZAF Cor 2 correction program was
used, and the maximum analytical error was 3% of the total using the standard minerals
corundum (AI), albite (Na, Si), hematite (Fe), apatite (Ca), orthoclase (K), forsterite (Mg),
rutile (Ti), and rhodonite (Mn). The structural formulae were calculated on the basis of 11
and 14 oxygens, for illites and chlorites, respectively, (half formula). All K, Na, and Ca was
assigned to the interlayer.

Temperature estimation
The temperature estimation at Los Azufres has been obtained by combining the data from
different geothermometers, particularly those from the direct measurements from a thermal
probe (Kuster equipment), the data from fluid inclusions (Cathelineau et al., 1988) studied in
the minerals coexisting with the clays, and the data obtained by applying chemical
geothermometers (Fournier & Truesdell, 1973; Fournier & Potter, 1982) to the composition
of the geothermal fluid produced (Nieva et al., 1983). As the field is relatively young, the
microthermetric results obtained on fluid inclusions from quartz coexisting with illite or
chlorite are in a good agreement with the different estimates (the direct measurements
excepted) and give a rather good estimation of the temperature with an error < 10~
It is worth noting that there are marked difficulties in obtaining precise temperature
estimates from temperature records obtained only by thermal probe analysis when the
temperature reaches more than 150~ due to lack of thermal equilibrium in the borehole.
Repeated measurements made in boreholes under different analytical conditions (borehole
filled by the drilling mud or the geothermal fluid, borehole in/or not in production) and after
different times ranging from one day to several months, show that the temperature
distribution is highly modified in the geologic formations after drilling. Thus, hydrologic and
thermic features of the reservoir may be irreversibly affected, and are in any case disturbed
for a rather long time. As direct temperature measurements are time consuming, and are not
always correct, reliable data are often scarce in the literature. Temperature is frequently
underestimated above the drill-hole bottom, whilst errors in the temperature estimation may
range from 10-30~ and may reach higher values in the most unfavourable cases
(Cathelineau, 1987; Cathelineau et al., 1988). Thus, the data from the fluid inclusion studies
were considered more reliable in the estimation of the clay crystallization temperature. These
difficulties have frequently been underestimated, and may make uncertain a direct
comparison between data from the literature. In this paper, data from McDowell & Elders
(1980), and Bishop & Bird (1987), were compared with those from Los Azufres although these
authors did not provide any indication of the error in the temperature estimates, and very
little discussion on the reliability of the geothermometric approaches used.
474 M. Cathelineau

RESULTS

Table 1 shows revised data on chlorites and illites from Los Azufres (Cathelineau & Nieva
(1985) and Cathelineau & Izquierdo (1988)), along with data from the literature on other
geothermal systems (McDowell & Elders, 1980; and Bishop & Bird, 1987)9

Chlorites
Correlation lines obtained by a least-square fit show that some of the chemical variables
are strongly temperature dependent (McDoweU & Elders, 1980; Cathelineau & Nieva, 1985).
Correlation diagrams between some chemical variables and temperature are shown in Fig. 1.
At Los Azufres, the best positive correlation is given by A1 in the tetrahedral site of chlorites
which increases very regularly with temperature (Fig. la). Fe in the octahedral site of
chlorites show a positive correlation with temperature whilst the octahedral vacancy is
negatively correlated (Fig. lb and c). The first regression line AI(IV)/T(~ given in
Cathelineau & Nieva (1985) has been corrected by considering new analyses (Table 1) which
complete the high and low temperature ranges. The good regression coefficient (0.97) for the
whole data set, and a temperature range of 130-310~ argue in favour of a strictly linear
dependence of the AI tetrahedral occupancy with temperature.
Most of the AI values for the Salton Sea chlorites plot in good agreement with the Los
Azufres regression line. Only two analyses corresponding to the lowest temperature (190~
exhibit higher Si(IV) content than at Los Azufres. These analyses correspond to detrital

1,5

R = 0,97 . J
[] Los A z u f r e s
AI(IV) []~ 9 9 [] a
Salton Sea
1.11

~ - 61,9229 + 321,9772 (AIIV)

o,s i
100 200 300 400
T~
3 ~ b 0,6" W
c

I 0,4
[]

I [ ] m []
[] m
m "~
O 0,2 "

[] []
[]

[]
[]
[]
m
[] 9
I
1 . . ~ 0,0 .
100 200 300 400 100 200 300 400
Toc T~
FIG. 1. Plot ofAl(IV), Fe andoctahedral vacancyvs, temperature for chlorites from Los Azufres
and Salton Sea. The regression line in the AI(IV)/T diagram has been calculated with the Los
Azufres data and the mean values of the Salton Sea data (mean values of the analyses given by
McDowell & Elders, 1980, for a given temperature, with rejection of the 190~ data).
Chlorite and illite geothermometers 475

o o 6 ~ o ~ 6 ~ 6 6 6 6 6 6

6 o ~ 6 ~ 6 ~ o 6 6 ~ 6 6 6

. . . ~ . . ~ o o o . .

II

II
. . . ~ ~ @ ~

~m II

14 8
~ 6 6 8 6 6 ~ o 6 6 6 6 6 6 6 6
II

I II

~5
?
476 M . Cathelineau

chlorites replaced by authigenic chlorites within dolomite-bearing sandstones, whilst other


analyses correspond mostly to authigenic chlorites crystallized in a calcite-bearing sandstone.
The high A1 content of the 190~ chlorites may result from differences in the rock lithology
and from a possible inheritance of the chemical features of the early detrital chlorites. The
regression line obtained for all temperatures above 200~ has a relatively good regression
coefficient (0.88). The simultaneous plot of the Salton Sea and Los Azufres data shows the
perfect agreement of the data obtained on the two geothermal fields. The coefficient of the
regression line is 0.97, and the relation between T and AI(IV) is the following:
T (~ = -61-92 + 321.98 (Alw)

lllites

The plots of chemical variables vs. temperature have been made for three sets of analyses:
Los Azufres (Cathelineau & Izquierdo, 1988), Coso hot springs (Bishop & Bird, 1987), and
Salton sea (McDowell & Elders, 1980). Some variables such as Fe or Mg are correlated with
temperature, but trends are different from one geothermal field to the other. The Fe content,
for instance, increases with temperature in most fields, but correlation coefficients are not
good, and the slopes of the correlation lines are different in the three cases considered. The
Mg is negatively correlated with temperature at Salton Sea, whilst it is positively correlated
with temperature at Los Azufres in illites crystallized in altered andesites. Other chemical
variables such as the AI(IV), or the AI(VI) show less clear trends with temperature. The
interlayer occupancy, and especially the K content, exhibit strong temperature dependence
(Fig. 2). However, there is a significant break between the data obtained on illites crystallized

1,0 1,0 " o~

0,9 o,9-
/ u 9 .. n , h ..'

K 0,8' a ,/ 0,8" e p . . u .~i . .


I.C
n " ."o~
0,7 ' 0,7 " /"

0,6 /. / 0,6 "

0,5 , , 0,5 , , i
100 200 300 400 100 200 300 4O0
ToC T~

0,4

0,3
[] L o s A z u f r e s
Fe
0,2 n oim " ~ Salton sea
.-.......
.~o~, 9 [] Coso
0,1 /
u...D.~...-u"
t
0,0
100 200 300 400
T~

FIG. 2. Plot of K , F e and the interlayer charge ( I . C . ) vs. temperature for illites from the Los
Azufres, Salton Sea, and Coso geothermal systems (analytical data from Cathelineau &
Izquierdo, 1988; McDowell & Elders, 1980, Bishop & Bird, 1987).
Chlorite and illite geothermometers 477
at temperatures below 200~ and those above. Slopes of the regression lines in the diagram K
vs. T are very similar whatever the field, and correlation coefficients calculated for each line
are good, > 0.96, in the range 200-320~ (Fig. 2). However, the K content at a given
temperature is higher at Los Azufres than at Coso, or at Salton Sea. Thus, the three regression
lines are separated by significant shifts. Such differences in the K content at a given
temperature are due, at least partly, to a variation in the Na or Ca content, at a constant
interlayer charge. However, they may also result partly from analytical problems in the
determination of alkalis. Thus, K or Na losses may occur during illite analysis by electron
microprobe, and may lead to difficulties in comparing results from different laboratories.

INTER-SITE CATION SUBSTITUTIONS AND THE END-MEMBER


MOLAR FRACTION CALCULATION

Algebraic model for clay solid-solutions


From the thermodynamical point of view, the difference between a clay solid-solution
composed of different minerals, and a simple mixing of the same minerals, is the mixing
energy. The Gibbs free energy of formation of the solid-solution containing different molar
fractions of each mineral is represented by the sum of the different energies plus a mixing
energy. If negative, the solid-solution is more stable than the separate, pure end-members.
The equilibrium between the solid-solution and a fluid is characterized by the n equilibrium
relationships between the n end-members and the fluid. Each end-member participates in the
solid solution as a function of its saturation state in the fluid as demonstrated by Fritz (1981).
There are many models of solid-solutions in the literature, in particular those of Stoessel for
illites and chlorites (1979, 1984), and those of Aagard & Helgeson (1983), Fritz (1981) and
Tardy & Fritz (1981) for the smectite-illite series. The last gives probably the most powerful
solution to the modelling of a natural solid solution, because the end-members are not chosen
arbitrarily, but the maximum number of possible end-members are considered.
However, calculation of the equilibrium conditions for the stability of a given clay solid-
solution requires at least the composition of the solution in equilibrium with the clay, and the
temperature. For minerals crystallized under unknown conditions, processes are nearly the
reverse. It is first necessary to describe the chemical features of the natural clay, and then to
compare these data with temperature calibrated variations of reference clay compositions.
For this purpose, an algebraic model was used, which calculates the contribution of each of
the chosen theoretical end-members. Substitutions are given in a form similar to that
proposed by Bragg (Bragg, 1937; Bragg & Claringbull, 1965) and Thompson (Thompson &
Thompson, 1976; Thompson et al., 1982). Following these authors, a solid-solution may be
described by a reference composition and substitution vectors. The substitution vector
expresses the algebraic relations which describe the addition or the subtraction of one end-
member to a reference mineral. These algebraic relations are dependent on the mass and
charge balance in the mineral. As an example, the reaction:
K(A1Si3)A12010(OH)2 + 4H § + H4SiO4 = Si4AlzOlo(OH)2 + K + + A13+ + 4HzO (1)
muscovite pyrophyllite
may also be expressed by:
K(AISi3)AlzOlo(OH) 2 + Si 4+ = Si4AI201o(OH)2 + K + + A13+ (2)
muscovite (Mu) pyrophyllite (Pyr)
478 M . Cathelineau

In vectorial space, muscovite and pyrophyllite are represented by two points with the
following coordinates:
Mu Pyr
K 1 0
Ca 0 0
Na 0 0
Si(IV) 3 4
AI(IV) 1 0
AI(VI) 2 2
R2+ 0 0
The reaction (2) may be represented by a vector L the coordinates of which are given in Fig.
3. The vector I may be written following Thompson's formalism: K_I(Si+IAI_I)w. Any
intermediate composition between muscovite and pyrophyllite may be expressed by the
equation:
V=x/
where V is the vector describing the compositional changes between the reference muscovite
and the clay. The reference composition may be chosen arbitrarily but not the substitution
vectors. Such a procedure simplifies the equations and reduces the number of variables e.g., a
given illite composition may be roughly described by a reference mineral such as muscovite,
and two substitution vectors as shown in Fig. 3. For a complete description, more vectors are
needed, but they have to be linearily independent. The matrical representation of the system
is given in Fig. 4 and the representative substitution vectors are listed in Table 2. Other
vectors may be used, but they are linear combinations of those listed in Table 2. The mass and
electric balance expressed by these vectors give the general correlations between chemical
variables. Correlations coefficients, for instance, are necessarily + 1 between K and AI(IV),
and - 1 between Si and K, and Si and AI(IV), when substitutions are fixed by vector L The
coefficients (xi) of the substitution vectors are the molar fraction of theoretical end-members.
Their values are computed using a program called N O R M I N (Cathelineau, 1987), which is
able to solve systems of n equations and p unknowns simultaneously.

TABLE2. Main correlations between chemical variables in chlorites


and illites expressed by the substitution vectors. I: interlayer charge,
I-]: octahedral vacancy

Illites Chlorites

Origin: reference mineral


Muscovite Talc-3 brucite

Substitution Vectors
I-I(Si+IAI_0w ($1+1AI-1)Iv(A1-1R+ 1)vi
9 2+

K-xNa+x (AI+1R25/2[] 1/2)VI


K-1Ca+I/2 (or (Si_IAI+I)Iv(AI_1R2+$[7 _ 1)v0
(Sx+IAI_I)Iv(AI_1R+I)Vl
9 2+
FeMg_t
FeMg_a
Chlorite and illite geothermometers 479

~t~,-"
' . ,~2+++X~
+ I k ',Ix t -,r P P I
l.+~.+~l~X.+Or
' + + ~ +.~ K o -I
Na 0 0
~ Ca 0 0
~ ~ ~.~-~ V Sl +1 +1
AI(IV) - 1 - 1
Mu ~ ~ ~ ~ ~ AI(VI) - 1 0
R2+ +1 0
x2

"1) I

V=XlP+X2I
FIG. 3. Vectorial representation of the two main substitutions in illites. The reference
composition is muscovite (Mu), and the two substitution vectors P ("phengitic") and I ("illitic")
describe the changes occurring from muscovite to celadonite, and muscovite to pyrophyllite,
respectively. Coefficients of the substitution vectors are given at the right. V represents a given
illite, and xl, x2 are the molar fraction of the end-members.

Temperature dependence of the intersite substitutions


Main correlation coefficients between the different chemical variables have been
calculated, and two examples of the correlation matrix for illites and chlorites from Los
Azufres are shown in Table 3. Matrices show that the variables are partly multicorrelated,
thus justifying the choice of vectors using different chemical variables linked by electric and
mass balance relationships. The differences in the values of the regression coefficients from
one set of analyses to the other are therefore explained by the relative importance of one
substitution over the others. Thus, a better expression of composition changes with
temperature is given by the values of the molar fraction of the theoretical end-members.

Chlorites
The statistical treatment of chemical variables of the Los Azufres chlorites has shown that
together with the increases in AI in site IV, an increase in divalent cations, mainly Fe, and a
decrease in the number of vacancies and in AI(VI) are observed in the oetahedral site. Such a
result means that the electrical deficit caused by AI replacing Si is balanced by a Fe-AI
substitution in the octahedral site of chlorites. These chemical changes m a y be easily
described by the substitution vector (Si_IAI+I)(AI_IFe~2[I]_1), and by the increasing
chamosite molar fraction and decreasing pyrophyllite-2 gibbsite molar fraction with

f(V2) , ..., f(Vn)

In
al I=i,i , =1,2' .... ' = l , n
(V) = M.(X) = a2 = =2.~i ' =2,2' .... " = 2 , n

p ~=p,l ' =p,2' .... ' =p,n

FIG. 4. Matrix of the relationships between the stoichiometric coefficients (a~)of a chemical
element in the structural formula, the molar fraction (x~)of the different end-members, and the
substitution vectors (fV~) represented by the matrix of their coefficients (%,,). The x~ are the
unknowns and are calculated by solving this system of n equations with p unknowns.
480 M. Cathelineau

TABLE 3. Correlation matrix of chemical variables for chlorites (data from Cathelineau & Nieva, 1987), and
illites (Well A23, rhyolitic environment) from Los Azufres. (I.C. : interlayer charge, O.C. : octahedral charge)

Siw Allv Alvi Fevi Mgvi R2+vl K Ca Na I.C.


Silv
AIIv - 1"00 Illites
Alvl -0.23 0.23
Fevl 0.33 -0.33 -0.85
Mgvl -0.12 0.12 0.01 -0'39
R2+Vl 0"34 --0"34 -0"91 0"90 -0"01
K -0"51 0"51 -0"49 0"31 -0"06 0"35
Ca -0"03 0"03 0"23 -0"28 -0'17 -0"42 --0"42
Na -0.60 0.60 0.19 -0"38 -0"02 -0'47 0'21 0"65
I.C. -0.63 0.63 -0.34 0"05 0"00 0"05 0"83 0"12 -0"47
O.C. 0.08 -0.08 -0.83 0"50 0"01 0"54 0"59 0"00 0"19 0"63

Siw Allv Alvl Fevl Mgvl R2+v1


Silv
Allv - 1.00 Chlorites
Alvl 0.31 -0.31
Fevl - 0.64 0.64 -0"30
Mgvi 0-01 - 0.01 - 0.68 - 0.46
R2+vI -0"56 0.56 -0.96 0.45 0"59
6-Z 0.81 -0.81 0'80 -0"58 -0.41 -0.94

increasing temperature (Cathelineau & Nieva, 1985). The pyrophyllite-2 gibbsite molar
fraction exhibits the same variations at Salton Sea as at Los Azufres, as the AI(IV) and
octahedral vacancy have the same behaviour in the two fields.

Illites

Results for illites are presented in Fig. 5. Figure 5a shows that the K_ISi_IAI_I vector
coefficient which represents the pyrophyllite molar fraction decreases with temperature from
0.35-0.4 at 190~ down to zero at ~ 300~ Other theoretical end-members exhibit less
intense changes in their molar fractions. Fe substitutes preferentially to Mg with increasing
temperature in the R 2§ site, yielding to the increase of the Fe-celadonite molar fraction (Fig.
5b). The Fe-celadonite molar fraction does not depend only on temperature, but also on the
rock lithology, as shown by the higher value observed in illites crystallized in andesites than
in rhyolites. Chemical changes in the interlayer lead to the decrease of the Ca-paragonite, and
the increase of the paragonite molar fractions (Fig. 5c and d).

Discussion

The main results are summarized in Table 4. The decrease of the molar fractions of the
Si0v)-rich end-members, the pyrophyllite-2 gibbsite in chlorites (Cathelineau & Nieva, 1985),
and the pyrophyllite in illites (Fig. 5), may be considered as the most common changes with
increasing temperature for the layer-silicates studied. Other changes are probably related
essentially to the composition of fluids in equilibrium with rocks. Such a result indicates that
the Si(iv)-rich end-members, such as kaolinite (expressed as pyrophyllite-2 gibbsite) and
Chlorite and illite geothermometers 481

~
0,6 ~ a 0,300

y = 0,7928 - 0,0025x II
0,4 O,2OO'
R 0,82
= ~ ,~ I
e~
II u 9 ,~ j~" ./

0,2 :~ ~ o,,oo. n
b n... ~ . . - .-'li

0,0 , , , 0,000
100 200 300 400 100 200 300 400
T~C ~C
0,08" y = - 0,0802 + 4,244e-4x 9149 0,03" d
II

R =0,77 /
0,06"
0,02"
9 .$
0,04"
9 V.
0,01
0,02" = "q
iii 9'.
[] a :nlnm 9 9
0,00 1 0,00 i

100 200 300 400 100 200 300 400


Toc T~
FIG. 5. Plot of the molar fractions of pyrophyllite (coefficient of the substitution vector
(K_ISi+IAL1), Fe-celadonite, paragonite and Ca-paragonite vs. temperature, for the Los
Azufres, Salton Sea, and Cos 9 geothermal systems (symbols as for Fig. 2). In diagram b, the lines
a and b correspond to two different sets of analyses made on illites crystallized in rhyolites and
andesites, respectively, from the Los Azufres geothermal system.

pyrophyllite, become less and less stable in the solid-solution with increasing temperature.
The decrease in the activity of the kaolinite as an end-member in the solid-solution indicates
a decreasing Qj !onic product of kaolinite in solution, and consequently a lower saturation
rate at higher temperatures. It should be noted that the properties of the kaolinite as a mineral
in equilibrium with a solution or as an end-member in a solid-solution are essentially the
same. Thus, various authors have demonstrated from the analysis of natural waters and from
theoretical predictions that the saturation state of kaolinite decreases with increasing
temperature (e.g. Fritz, 1981).

TABLE4. Summary of the main changes in chlorites and illites with increasing
temperature. (1) Application domain, (2) variation of the cation content, (3)
variations of the end-member molar fractions.

Illites Chlorites

(1) 170-300 T~ 100-350


(2) K : Na,~ Fe,~ (Allv)/ Allv7 Fe.: F e / M g :
A l v ~ Ca"., Oct. Vacancy N, Alv1*
(3) I_l(Si+l Al-x)iv N, (Si_tAI+I)w(AI_IR2+ [] -1) /
X pyrophyllite'~ X pyrophyllite - 2 gibbsite',,
X paragonite.~ X chamosite d
X Ca-paragonite
482 M. Cathelineau
USE AND M I S U S E OF T H E G E O T H E R M O M E T R I C
RELATIONSHIPS

The numerical correlations between chlorite composition and temperature are the same in
the two geothermal fields for which detailed data are available--the Salton Sea and the Los
Azufres systems. It is clear that the revised linear expression (1) could be used as a
geothermometer in other fields. Numerous data from the literature support this assumption.
In Iceland, for instance, (Sveinbjornsdottir, 1986), the AI(IV) content exhibits a rather
regular increase with temperature. At Larderello, the chlorites from the deeper parts of the
field (Cavaretta et al., 1982) have an AI(IV) content ~ i.25, and crystallized at a temperature
~ 350~ data are in agreement with the extrapolation of the regression line(l), and
could indicate a possible use of the geothermometer at temperatures higher than 350~ Such
an hypothesis is supported by the data from Aud~oud (1982) who described, in the
metamorphic series of Shaba (Za'ire), Al(IV)-rich chlorites, characterized by an AI(IV)
content ranging from 1.2-1-4, and a probable crystallization temperature ~ 350~
Different authors described chlorites with a very low A1 content in the range 0.4-0.7 in
sedimentary basins (Meunier, 1984, in the Salt Wash basin, Colorado, and Aagard et al.,
1988, in the Norwegian North Sea sandstone reservoirs). These chlorites are considered to
have crystallized in the range 100-140~ on the basis of fluid inclusion studies and thermal
reconstruction of the oil fields. Such data are in good agreement with the extrapolated
regression line (1). In addition, Aagaard et al. (1988) described an increase in AI(!V) from
125-150~ in chlorites, thereby supporting the use of the chlorite geothermometer at such a
low temperature.
Thus, the application domains of the linear correlation are not restricted to chlorites
crystallized in geothermal fields characterized strictly by the same features as those of Los
Azufres or Salton Sea. The chlorite geothermometer could probably be applied successfully to
chlorites from diagenetic, hydrothermal or metamorphic origin, as the AI content in the
tetrahedral site seems to be independent of the rock lithology or the fluid composition.
The composition of illite, especially the Fe and Mg content, is mainly dependent on the
rock lithology, i.e. the composition of fluids in equilibrium with the rocks. The layer
occupancy, which depends mostly on temperature, exhibits differences which are probably
related to differences either in the rock lithology or in the fluid characteristics, particularly
salinity and the K, Na, Ca contents. However, relatively good correlations have been
obtained between K, Na, Ca, and interlayer charge and temperature above 200~ At lower
temperatures, breaks in the composition trends may indicate the beginning of a possible
mixing, or undetected mixed-layering, with other clay particles.
From a more general point of view, care must be exercised in the use of clay
geothermometers, as well as for any other geothermometer. It is important to note that the
chlorites and illites studied are authigenic minerals crystallized through water-rock
interaction processes characterized by high water/rock ratios in environments where quartz
is present. Thus, as the behaviour of the minerals in closed systems is not known precisely, the
correlation lines may nofbe used in any geological example without caution. As the minerals
studied have common compositions, it may be difficult to use the correlations for unusual
minerals with Fe, Zn, and Cr end-members. It is also necessary to work with pure minerals
without any mechanical or structural mixing with other sheet silicates at a small scale ( < 1
#m). Smectites have Si/AI(IV) ratios rather different from chlorites, and may affect the
chlorite chemistry if intimately associated, even in small amounts.
Chlorite and illite geothermometers 483

It is highly recommended to check whether the clay has formed from the alteration of a
previous layer-silicate or not. Authigenic chlorites formed from biotite alteration and illites
formed from muscovites may inherit chemical features of the parent mineral if alteration has
not been complete. It is also necessary to determine the composition ranges of a given mineral
in order to obtain reasonable error ranges for the temperature estimates.
Clay geothermometers provide an estimate of temperature which is considered to
represent the temperature of crystallization of the clay. The good agreement between the
fluid inclusion data and the clay data obtained on fossil hydrothermal systems seem to
confirm this. However, late thermal events may play a role in the chemical adjustment of the
clay particles to new physical and chemical conditions, and so the question is whether the
estimated temperature corresponds to the early or to the late thermal events which could have
affected the clay. Kinetics of clay transformation with temperature are not sufficiently well
known to answer this question.

CONCLUSIONS

This study of the correlation between the composition of clay solid-solution and temperature
demonstrates that:
(i) The cation occupancy of a given site is not independent of the other ones, as shown by
the multicorrelations among chemical variables, and the good description of the chemical
changes by substitution vectors describing simultaneous exchanges of 3 or 4 variables. This
indicates that substitutions are not independent from an energy point of view. Consequently,
the chemical changes of the clays are described better by substitution vectors and molar
fractions of the end-members than by single chemical variables.
(ii) The agreement between the data obtained on natural clays considered as end-members
in a solid-solution or as minerals, is important for the solid-solution model. On the other hand,
the data from natural examples may help greatly in refining the calculation of Gibbs free
energy of the clay minerals used in these models, as thermodynamic data are frequently
lacking, even for the end-members.
(iii) The empirical relationships between clay composition and temperature provide tools
for the estimation of the temperature of crystallization of the clays in environments for which
the use of other geothermometric approaches is difficult. This has a practical importance for
the study of diagenetic, hydrothermal and low-T metamorphic processes.

ACKNOWLEDGMENTS
The author wishes to express his gratitude to two Clay Minerals' reviewers for their critical comments, and
especially to DE Bain for help in improving the presentation.

REFERENCES
AAGAARDP. & HELGESONH. C. (1983)Activity/compositionrelations amongsilicates and aqueous solutions.
II: Chemical and thermodynamic consequences of ideal mixing of atoms on homological sites in
montmorillonites, illites and mixed-layer clays. Clays Clay Miner. 31, 207-217.
AAGAARDP., ROALDSETE. & WELHAVENJ. E. (1988) Diagenetic observations from North Sea sandstone
reservoirs: a comparisonwith formationwater chemistry. "Clay diagenesis in hydrocarbonreservoirs and
shales" Meeting, Cambridge, March 1988.
484 M . Cathelineau

AUDI~OUDD. (1982) Les min~ralisations urani#res et leur environnement ~ Kamoto, Kambove et Shinkolobwe
(Shaba, ZaYre) petrographie, g~ochimie et inclusions fluides. Thesis, Univ. Lyon, France.
BISHOPB. P. & BIRD D. K. (1987) Variation in sericite compositions from fracture zones within the Coso hot
springs geothermal system. Geochim. Cosmochim. Acta 51, 1245-1256.
BXAGG W. L. (1937) Atomic Structure of Minerals. Ithaca: Cornwell University Press.
BRAGG L. & CLARINGBULLG. F. (1965) Crystal Structures of Minerals. G. Bell & Sons, Ltd., London.
CATHELINEAUM. & NIEVAD~ (1985) A chlorite solid solution geothermometer. The Los Azufres geothermal
system (Mexico). Contrib. Mineral. Pet. 91, 235-244.
CATHELINEAUM., OLIVERR., NmvA D. & GAmrAS A. (1985) Mineralogy and distribution of hydrothermal
mineral in Los Azufres (Mexico) geothermal field. Geothermics 14, 49-57.
CATHELINEAUM. (1987) Les interactions entre fluides et min~raux : thermometric et mod~lisation. L'exemple d'un
systbme g~othermique actif (Los Azufres, Mexique) et d' alt~rations fossiles dans la Chaine Varisque. Doct.
Thesis, I.NP.L. Nancy, France.
CATHELINEAUM., OLIVERR. & NIEVAD. (1987) Quaternary volcanic series of the Los Azufres geothermal
field (Mexico). Special vol. on Mexican Volcanic Belt. Part. 3 (S. P. Verma, editor), Geol. Int. 26, 273-290.
CXTHI~LI~;EAtrM. & IZQUmRDOG. (1988) Temperature-composition relationships for authigenic clay minerals
in the Los Azufres geothermal system. Contrib. Mineral. Pet. (in press).
CATHELINEAUM., IZQUIERDOG. & NI~VAD. (1988) Thermobarometry of hydrothermal alteration in the Los
Azufres geothermal system: significance of fluid inclusion data. Chem. Geol. (in press).
CAV~ETTA G., GIANELLI G. & PUXEDDUM. (1982) Formation of authigenic minerals and their use as
indicators of the chemico-physical parameters of the fluid in the LardereUo-Travele geothermal field.
Econ. Geol. 77, 1071-1084.
DUNOYER DE SEGONZXr G. (1970) The transformation of clay minerals during diagenesis and low-grade
metamorphism. A review. Sedimentology 15, 281-346.
ELDERSW. A. (1979) The geological background of the geothermal field of the Salton Trough. Pp. 1-19 in:
Geology and Geothermics of the Salton Trough. (W. A. Elders, editor) Geol. Soc. Am. Guidebook No 7, San
Diego.
FOURNIER R. O. & TRUESDELLA. H. (1973) An empirical Na-Ca-Ca geothermometer for natural waters.
Geochim. Cosmochim. Acta 47, 579-586.
FOURNmR R. O. & POTTERR. W. (1982) A revised and expanded silica (quartz) geothermometer. Geochim.
Cosmochim. Acta 43, 1543-1550.
FREer,MAN J. T. (1978) Fluid inclusion and oxygen isotope geothermometry of rock samples from Sinclair 4
and Elmore 1 boreholes, Sahon Sea Geothermal Field, Imperial Valley, California, USA. Inst. Geophys.
Planetary Phys., Univ. Calif., Riverside, Rep. 78/5, 66pp.
FREY M., TEICHMULLERM., TEICHMULLERR., MULLIS J., KUNZI B., BREITSCHMIDA., GRUNER V. &
SCHWIZERB. (1980) Very low grade metamorphism in external part of the Central part of the Central
Alps: Illite cristallinity, coal rank, and fluid inclusion data. Eclogae geol. Heir. 73, 173-203.
FRITZ B. (1981) Etude thermodynamique et mod61isation des r~actions hydrothermales et diag6n&iques. Sci.
Geol. Mere. 65, 197 pp.
GUTIERREZA. N. & AUMENTOF. (1982) The Los Azufres, Michoacan, Mexico, geothermal field. J. Hydrol. 56,
137-162.
HELGESONH. C. (1968) Geologic and thermodynamic characteristics of the Salton sea Geothermal system.
Am. d. Sci. 266, 129-166.
KUBLER B. (1967) La cristallinit~ de l'illite et les zones tout ~ fait sup6rieures du m&amorphisme. Etages
tectoniques. Coll. Neuchatel 105-122.
KUBLERB. (1969)]~valuation quantitative du metamorphisme par la cristallinit6 de l'illite. Bull. Centre Rech.
Pau, SNPA 212, 385-397.
McDOWELL S. D. & ELDERSW. A. (1980) Authigenic layer silicate minerals in borehole Elmore 1, Salton Sea
geothermal field, California, USA. Contrib. Mineral. Pet. 74, 293-310.
MEUNIER J. D. (1984) Les ph6nom6nes d'oxydo-r6duction dans un gisement urano-vanadif6re de type
tabulaire: les gr6s du Salt-Wash (Jurassique Sup6rieur), district minier de Cottonwood-Wash (Utah, Etats
Unis). Geol. Geoch. Uranium, Mere. Nancy 4, 200 pp.
NIEVAD., QUIJANOL., GARFIASA., BARRAGANR. M. & LAREDOF. (1983) Heterogeneity of the liquid phase,
and vapor separation in the Los Azufres (Mexico), geothermal reservoir. Proc. 9th Workshop Geothermal
Res. Eng. Stanford Univ. SGP-TR-74.
PERRY E. A. & HOWERJ. (1970) Burial diagenesis in Gulf Coast pelitic sediments. Clays Clay Miner. 18, 165-
178.
Chlorite and illite geothermometers 485

REYNOLDSR. C. & HOWER J. (1970) The nature of interlayering in mixed-layered illite-montmorillonites.


Clays Clay Miner. 18, 25-36.
SCHOENR. & WHITla.D. E. (1966) Hydrothermal clay minerals in granodiorite in the Main Terrace Steambot
Springs, Nevada. Clays Clay Miner. 13, 121-122.
STEINERA. (1968) Clay minerals in hydrothermally altered rocks at Wairakei, New Zealand. Clays Clay Miner.
16, 193-213.
STOESSELR. K. (1979) A regular site-mixing model for illites. Geochim. Cosmochim. Acta 43, 1151-1159.
STOESSELR. K. (1984) Regular solution site-mixing model for chlorites. Clays Clay Miner. 32, 205-212.
SVEIBJOI~NSDOT'rlRA. (1986) The chemical characteristics of the hydrothermal fluids at the Krafla and
Reykjanes systems, as inferred from the coexisting mineralogy. Proc. 5th Int. Sym. WaterRock lnteraction,
Reykjavik, Iceland 546-549.
TARDYY. & FRITZ B. (1981) An ideal solid solution model for calculating solubility of clay minerals. Clay
Miner. 16, 361-373.
THOMPSONJ. B. JR. & Tnoral, sor~ A. B. (1976) A model system for mineral facies in pelitic schists. Contrib.
Mineral. Pet. 58, 3-55.
THOMPSONJ. B., LAIRDJ. & THOMPSONA. B. (1982) Reactions in amphibolite, greenschist and blueschist. J.
Pet. 1, 1-27.
TOMASON J. & KRISTMAr,q~SOOT'rIRH. (1972) High temperature alteration minerals and thermal brines,
Reykjanes, Iceland. Contrib. Mineral. Pet. 36, 123-134.
VELDEB. (1977) Clays and clay minerals in natural and synthetic systems. Developmentsin Sedimentology, 21.
Elsevier, Amsterdam.
VELDE B. (1985) Clay minerals. A physico-chemical explanation of their occurrence. Developments in
Sedimentology, 40. Elsevier, Amsterdam
WEAVERC. E. (1959) The clay petrology of sediments. Clays Clay Miner. 6, 154-187.

Das könnte Ihnen auch gefallen