Sei sulla pagina 1di 13

Journal of Volcanology and Geothermal Research 173 (2008) 217–229

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research


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

Chemical transport in geothermal systems in Iceland


Evidence from hydrothermal alteration
Hjalti Franzson a,⁎, Robert Zierenberg b, Peter Schiffman b
a
Iceland GeoSurvey, 9 Grensásvegur, 108 Reykjavík, Iceland
b
Department of Geology, University of California, Davis, One Shields Avenue, 95616, CA, USA

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

Article history: This study focuses on the chemical changes in basaltic rocks in fossil low- and high-temperature
Received 1 July 2007 hydrothermal systems in Iceland. The method used takes into account the amount of dilution caused by
Accepted 28 January 2008 vesicle and vein fillings in the rocks. The amount of dilution allows a calculation of the primary concentration
Available online 4 March 2008
of the immobile element Zr, and by multiplying the composition of the altered rock by the ratio of Zr
(protolith)/Zr (altered rock) one can compute the mass addition caused by the dilution of the void fillings, and
Keywords:
also make a direct comparison with the likely protoliths from the same areas. The samples were divided into
Iceland
three groups; two from Tertiary fossil high-temperature systems (Hafnarfjall, Geitafell), and the third group
basalt
geothermal systems from a low temperature, zeolite-altered plateau basalt succession. The results show that hydrothermally
hydrothermal alteration altered rocks are enriched in Si, Al, Fe, Mg and Mn, and that Na, K and Ca are mobile but show either depletion
chemical transport or enrichment. The elements that are immobile include Zr, Y, Nb and probably Ti. The two high-temperature
isocon method systems show quite similar chemical alteration trends, an observation which may apply to Icelandic fresh
water high-temperature systems in general. The geochemical data show that the major changes in the altered
rocks from Icelandic geothermal systems may be attributed to addition of elements during deposition of pore-
filling alteration minerals. A comparison with seawater-dominated basalt-hosted hydrothermal systems
shows much greater mass flux within the seawater systems, even though both systems have similar alteration
assemblages. The secondary mineral assemblages seem to be controlled predominantly by the thermal
stability of the alteration phases and secondarily by the composition of the hydrothermal fluids.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction systems in which the high thermal gradient is due to shallow crustal
magmatic activity. The latter type is mostly confined to active volcanic
The location of Iceland as a subaerial part of the Mid-Atlantic Ridge, centres. However, fossil high temperature geothermal systems are
with a wealth of geothermal activity, makes it a unique area for the exposed by erosion allowing three dimensional access to the subsur-
study of water–rock interaction. Temperatures of fluid–rock interac- face portions of the hydrothermal systems (e.g. Fridleifsson, 1983,
tion range from about 2 to 3 °C in the groundwater systems to probable 1984). The majority of the hydrothermal systems in Iceland have
supercritical values in the high-temperature geothermal areas (Elders waters derived from local meteoric water, although seawater-domi-
and Fridleifsson, 2005). The rocks that host the geothermal reservoirs nated hydrothermal fluids occur in some near coastal geothermal
are of igneous origin, with basaltic compositions constituting about systems (Sveinbjornsdottir et al., 1986). The general similarity of both
90% of the volume with the remainder having more evolved com- source fluids and host rocks throughout much of Iceland allows
positions. The abundant geothermal resources are economically im- comparison among hydrothermal systems.
portant and widely utilized, and have consequently been extensively Rock properties play an important role in the petrophysical para-
studied from geochemical, structural, geophysical and geological meters of the individual systems and an increasing emphasis has been
points of view (e.g., Arnorsson et al., 1983; Bodvarsson, 1983; Marty placed on the study of reservoir characteristics. During the last
et al., 1991; Schiffman and Fridleifsson, 1991; Lonker et al., 1993; Riedel 13 years, Orkustofnun (the National Energy Authority of Iceland), with
et al., 2001). The geothermal resources range from low-temperature the financial support of Orkuveita Reykjavíkur (Reykjavik Energy), has
systems, in which shallow groundwater has gained heat in response to worked on a project with the main aim of defining the reservoir
the prevalent regional geothermal gradient, to high-temperature characteristics of the rocks in the geothermal systems in Iceland.
While active geothermal systems have the advantage of allowing
direct comparison of rock alteration to fluid composition, this com-
⁎ Corresponding author. Tel.: +354 528 1500; fax: +354 528 1699. parison only relates to a single stage, i.e., the present, in the evolution
E-mail addresses: hf@isor.is (H. Franzson). of the system. In contrast, sampling of fossil geothermal fields allows

0377-0273/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.jvolgeores.2008.01.027
218 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

examination of the time integrated history of rock alteration, in- has been mapped in considerable detail and the rocks are well
cluding determination of the temporal evolution of mineralogy and characterized chemically (Franzson, 1979). Twenty two samples of
porosity. basaltic lava flows were taken from this area.
The sampling strategy utilized in this study emphasized variably b) Hafnarfjall central volcano is a 5.5–4 myr old, deeply eroded cen-
eroded fossil geothermal systems, focussing on areas that had been tral volcano that has been mapped with respect to its structure,
previously mapped with respect to their geological structures and geochemical evolution, and locally, its hydrothermal alteration
hydrothermal alteration. The assumption is that rocks in these fossil (Franzson, 1979). The volcanic products range from basalts to
systems are close analogues of those in presently-active geothermal rhyolites which have been subjected to variable degrees of hydro-
reservoirs. Approximately 500 samples were collected for these stu- thermal alteration. Forty five samples are included in this study,
dies, a large proportion of which have been analyzed for total and and off these twenty seven are lava flows, five very scoraceous or
effective porosity, permeability, chemical composition, and petro- tuffaceous tops of lava flows and twenty intrusions (mostly fine
graphic characteristics. This extensive data set has been especially grained dykes or sills).
useful in determining the interrelationship between porosity and c) Geitafell central volcano was formed about 6 myr ago, is deeply
permeability of Icelandic rocks (e.g., Sigurdsson and Stefansson, 1994; eroded, and has been mapped in detail, both with respect to vol-
Sigurdsson et al., 2000; Stefansson et al., 1997), and the relationship of canic evolution and hydrothermal history (Fridleifsson, 1983). An
these physical properties to the degree of hydrothermal alteration investigation has also been completed on the geochemical
(e.g., Gudmundsson et al., 1995; Franzson et al., 1997, 2000; Franzson, evolution of the volcano (Thorlacius, 1991). Fifty five basalt sam-
1999). ples are included in the study, where thirty are lavas, five very
The main emphasis of this paper is to elucidate the nature of tuffaceous or scoraceous lava tops and twenty intrusions (mostly
chemical transport in Icelandic geothermal systems by correlating the fine grained dykes or sills).
bulk composition of rock samples with their petrographical char-
The three study areas are all Tertiary in age, and differ from younger
acteristics. Using the petrographically determined percentage infilling
volcanic successions that host active geothermal systems principally
in the rocks, we will show that it is possible to assess the amount of
by their lack of hyaloclastite formations, which occur almost exclu-
chemical dilution that has taken place. This allows the deduction of
sively in formations younger than the onset of glaciation at 3.3 myr (c.f.
the primary concentration of the immobile elements in the altered
Fig. 1). These older hydrothermal systems were subjected to an initial
rocks, and by correlating these with the relatively fresh volcanic rocks
phase of low-temperature alteration, followed by the main phase of
equivalents, we assess the chemical change that has taken place
high-temperature alteration, and subsequently overprinted by low-
during the hydrothermal alteration. Furthermore, using the “isocon”
temperature alteration formed during the demise of the central
method (Grant, 1986) for determining mass transport, we also eval-
volcanic complex and succeeding erosion. It has been demonstrated by
uate the chemical enrichment-immobility-depletion of various che-
extensive petrographic observations that mineral deposition during
mical components within individual samples.
the high temperature alteration regime in Icelandic hydrothermal
systems generally reduces rock permeability and the last low-tem-
2. Geological setting
perature alteration episode does not greatly affect the bulk rock com-
position. Petrography can provide important constraints on the
This study is based on data from approximately 130 basaltic rock
relative timing of alteration events. For example, minor late-stage
samples and focuses on three locations (Fig. 1) as described below:
zeolite deposition, present in some of our samples, has been tenta-
a) The Hvalfjordur basalt succession consists of ca. 4–3 myr old plateau tively interpreted as due to deposition from fluids trapped in relatively
basalts which have been subjected only to zeolitic alteration impermeable rock. In general calcite does not occur within the
(chabazite–thomsonite to mesolite–scolecite, e.g., as described epidote–amphibole zone in presently-active high-temperature sys-
from Eastern Iceland by Neuhoff et al., 1999). The Hvalfjordur area tems where temperatures exceed 290 °C. However, in the fossil

Fig. 1. A simplified geological map of Iceland showing the Tertiary, Plio-Pleistocene and the volcanic zones. The areas sampled for the petrophysical study are dark shaded, and the
locations focussed on in this study (Hafnarfjall, Hvalfjordur and Geitafell) are also marked.
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 219

Fig. 2. The alteration zones used in Iceland, their dependence on temperature and the main alteration features of the primary basaltic rock components.

systems, calcite is occasionally superimposed on epidote–actinolite contrasts in the degree of alteration within intrusive rocks and their
zone mineral assemblages, implying carbonate deposition either surrounding extrusive host rocks should be expected and may depend
during cooling of the high-temperature system by inflow of colder on factors other than differences in permeability and porosity.
fluids or during the succeeding low-temperature episode. Hence,
changes in the chemical composition of rocks relative to their fresh 3. Petrographic data
counterparts are believed to have occurred up to the highest tempe-
rature stage of alteration that the rock has undergone, with possible The rock samples are all of basaltic composition, ranging from
effects due to precipitation of late-stage calcite. olivine tholeiite to quartz normative tholeiite. They would dominantly
The samples are all either volcanic rocks which have been gra- be holocrystalline, but with some glass fraction in the most scoracious
dually buried by accumulation of younger volcanics, or intrusive rocks. part of the lavas. The main primary components in basalts are relatively
The age of the accumulated sequence is progressively older and more Ca-rich plagioclase, clino-pyroxene (augite) and opaques (magnetite–
altered as it is more deeply buried. However, the same cannot be ilmenite), and with subordinate amount of olivine, especially in the
inferred about the intrusive rocks, as their minimum age is generally more primitive basalts. Hydrothermal alteration is essentially made up
unconstrained. This may be of importance when considering the of two components: (a) replacement of primary components in the
timing of alteration, especially within the most intense alteration rocks by alteration minerals, and (b) precipitation of alteration mi-
zones, as intrusions may occur during any stage of the alteration. nerals into voids in the rock. Hydrothermal alteration in active
Rocks intruded at a later stage of alteration may only show a minor Icelandic geothermal systems is systematically zoned with respect to
alteration effect compared to those intruded earlier, even though they temperature (Fig. 2). The top zone, which contains relatively fresh
were subjected to similar temperatures and pressures. Therefore, rocks, includes a sequence that does not show any indication of

Fig. 3. Range in primary porosity in Icelandic rocks deduced from petrographic study of the rock samples (n = 127).
220 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

Fig. 4. The relation between degree of infilling and the total percentage of alteration minerals (i.e., % alteration minerals divided by the total mode) of a sample. Samples are grouped
according to their alteration zone (from Franzson et al., 2000).

geothermal interaction and may at the most show minor oxidation due vesicles and minor fractures), and to assess how much of that porosity
to groundwater circulation. The temperature dependency, as seen in had been filled by deposition of alteration minerals. Two hundred
Fig. 2, affects both the sequence of alteration minerals that fill voids, as points were counted on each rock thin section. Fig. 3 shows the
well as the ones that replace primary minerals. Thus olivine and glass distribution of primary porosity in Icelandic rocks, where porosity
are completely altered near the upper boundary of the mixed layer clay ranges from zero up to about 70%. As alteration proceeds void space in
zone. Plagioclase and opaques are more resistant to alteration and may the rock is progressively filled with alteration minerals. Fig. 4 shows
be only partially altered to the minerals shown on Fig. 2, especially the relationship between the degree of pore filling and the extent of
when taking into account that some of the rock samples are low alteration, defined as the percentage of alteration minerals relative to
porosity intrusions. Pyroxene shows in general similar resistance to primary minerals. Rocks with a primary porosity, as petrographically
alteration as plagioclase. It is mainly seen altering into clays and then determined, below approximately 15% have variable degrees of
actinolite at deeper levels. infilling (Fig. 5). In contrast, rocks that initially exceeded the 15%
Petrographic examination of the rock samples collected for this primary porosity threshold tend to show near complete filling of the
study used point counting (200 points) to quantify primary porosity, primary pore-space. This implies a non-linear relationship between
i.e. the original open space in rock prior to alteration (dominantly permeability and porosity in the Icelandic basaltic samples.

Fig. 5. The “rate” of deposition of alteration minerals into vesicle filling in relation to alteration zones and primary rock porosity. Note that there is apparently more rapid (i.e.,
efficient) deposition in rocks with N 15% primary porosity, as determined petrographically (from Franzson et al., 2000).
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 221

4. Chemical data

The chemical analyses were made by two commercial chemical


laboratories, The Caleb Brett Laboratory in England and McGill
University in Canada. Both used standardized XRF techniques. Values
for samples analyzed by both laboratories are generally within
analytical error. The samples were analyzed for major, minor, and
several trace elements. Loss on ignition (LOI) was measured in all the
samples, and in some samples, CO2 and Stotal were specifically ana-
lyzed. The trace elements analyzed included Zr, Y, Zn, Cu, Rb, Sr, Nb,
Ga, Ce, V, Pb, U, Th and As. All the analyses presented here have been
recalculated to 100%, without LOI, to allow direct comparison with
unaltered protoliths. The compositional range of relatively unaltered
rocks has been acquired from other sources for comparison with the
data on altered rocks presented here. Analyses describing the evolu-
tion trend within the Hafnarfjall central volcano and the Hvalfjordur
plateau basalt succession are from Franzson (1979), and analyses from
Fig. 7. A sketch showing a simplified model for the dilution of a chemical rock com-
the Geitafell central volcano are from Thorlacius (1991). Chemical
ponent with the introduction of vesicle filling. The Y-component in the rock is
analysis from the Krafla central volcano (Karl Gronvold, unpublished symbolised in black and becomes relatively reduced when vesicle fillings are added to
data) and the Reykjanes–Langjokull volcanic zone (Sveinn Jakobsson, the sample.
unpublished data, 1999) have been used as reference samples for
comparison with the altered samples where relatively unaltered sam-
ples are unavailable from the respective areas. Specifically, the methodology arises from the key observation that
Loss on ignition values range from zero up to about 13%. Fig. 6 much of the alteration process entails the filling of primary pores by
shows the relationship between the % of remaining primary rock secondary minerals, effectively diluting the chemical composition of
component (i.e. = 100-(rock alteration + void filling)) in basaltic lava the protolith by mass addition (Fig. 7). This mass addition will dilute
flows and LOI, where the samples have been grouped into alteration most of the components in the original rock and thus whole rock
zones. The figure clearly shows a strong correlation between LOI and analyses will show lower concentrations of immobile elements even
the extent of alteration in these samples. It also shows an inverse though there has been no mass flux of these elements. The extent of
correlation between the percentage of primary component and the dilution is known through point-counting as discussed above. By
minimum LOI for samples as a function of alteration grade. Smectite– identifying components in the rock that have remained immobile
zeolite grade samples with LOI b 1% have a much higher percentage during alteration and taking into account this dilution effect, the
(N80%) of primary components than epidote–amphibolite samples primary composition of the immobile element in the sample may be
with similar LOI values. This is probably due to more abundant of calculated.
alteration minerals with low-water contents in the higher grade One way of assessing the effect of chemical dilution on the rock
samples. LOI consists mainly of H2O+ and CO2. The latter was measured composition is to plot the elemental concentration against the rock
separately in only part of the sample group. The results show that CO2 dilution (Fig. 8). In this diagram the concentration from z to x
is typically b1% in samples with less than 60% alteration intensity with represents the concentration range measured for element A in
higher CO2 values (b6.5%) only present where alteration is more unaltered basalt. The tie lines from 100% dilution (point R) to points
intense. A good correlation is between CO2 and calcite in the samples. z and x represent the amount of dilution of z and x, respectively,
assuming that the dilutant mineral (or minerals) contains no chemical
5. Methodology for calculating mass transfer component A. If element A is immobile and has y concentration in the
sample, the dilution will cause the apparent decrease of that com-
The methodology used in this study is based upon the petrographic ponent along the line towards R. One may therefore expect that all
characteristics of Icelandic basalts altered in geothermal systems. immobile elements within the basalt range will, with increasing
dilution, be contained within the field labelled “a”. A mobile element

Fig. 8. Diagram showing effect of dilution on the concentration of hypothetical chemical


Fig. 6. Loss on ignition (LOI) plotted against the % primary rock component (i.e., 100-rock component A during alteration of basaltic rocks. Fields labelled “a”, “b”, and “c”:
alteration and void filling). Also shown are the best fit lines belonging to each of the represent, respectively, the range of diluted compositions following no mass transfer,
alteration groups. addition of component A, and loss of component A. See text for more details.
222 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

Fig. 9. A plot showing the relation between SiO2% and % mineral deposition (dilution)
deduced from petrography for samples grouped into alteration zones. The lines
demarcate the basalt compositional field (c.f. lines z–R–x in Fig. 8). Deviation of the
Fig. 11. A plot showing the relation between Zr and % mineral deposition. Most of the
samples outside the basalt field indicates apparent SiO2 enrichment.
samples fall within the basalt field, as defined by the lines, which indicates that the
element has remained relatively immobile.
may not follow the same trend. If vesicle fillings contain the mobile
element, which has been derived from outside the sample, the diluted
compositions will shift to the right of the line y–R (e.g., field b). If, on the samples are normal basalts and that more evolved basaltic com-
the other hand, a chemical component is being removed from the positions with higher Zr contents are less abundant. These deductions
rock, the dilution will move compositions to the left of line y–R, and further imply that dilution must play the largest role in modifying the
these samples will group within the field indicated by c. Samples for chemical composition of basaltic rocks within the alteration zones
which the extent of addition or depletion of an element is small studied here. It is interesting that there are no apparent progressive
relative to the variation of that element in the unaltered protolith may gradients in chemical transport discernable between samples in the
not be distinguishable by this method. However, rocks that have smectite–zeolite to the epidote–amphibole zones, which may indicate
undergone significant mass flux should be readily identified. that such changes, if present, are small compared to the chemical
In Fig. 9, the SiO2 concentrations of many of the altered samples dilution effect.
plot to the right of the primary basaltic compositional field, indicating An important indication of the immobile elements is that their
a strong SiO2 enrichment. This enrichment is seen in samples from all ratios will remain constant by definition during dilution and depletion
of the alteration zones. Fig. 10 on the other hand, generally shows of other chemical components (e.g. Barrett and MacLean, 1994; Leitch
indications of K2O depletion during hydrothermal alteration and dilu- and Lentz, 1994). Fig. 12 shows the Zr/Y ratios of the basalt samples
tion. Zircon is an example of an immobile element (Fig. 11), where a from successive alteration zones compared with values from the least
majority of the samples are confined within the field of primary altered rocks analyzed from the Hvalfjordur and Reykjanes–Langjokull
basaltic composition. Some samples plotting to the right of the pri- volcanic zones (Sveinn Jakobsson unpublished data, and Karl Gronvold
mary basalt compositional field have basaltic andesite compositions unpublished data) and the Geitafell (Thorlacius, 1991) and Hafnarfjall
and are retained for our purposes as they help define the fractionation (Franzson, 1979) central volcanoes. The figure shows that the Zr/Y
trend of the sample group. Fig. 11 shows (a) the strict confinement of ratios in altered rocks have a similar range to those in their fresh
the samples within the basalt compositional range, as previously counterparts and further confirms Y as also an immobile element in
mentioned, and (b) that most samples plot in the left side of the these systems.
basaltic field, in accordance with the observation that the majority of If an element is immobile, its original concentration prior to dilu-
tion can be found by extrapolating along the dilution line to the
original concentration of the element at 0% dilution, as shown along

Fig. 12. The Zr/Y ratios in the rock sample and a comparison with fresh volcanic rocks
(vertical bars) from Reykjanes–Langjokull volcanic zone (Sveinn Jakobsson, unpub-
Fig. 10. A plot showing the relation between K2O % and % mineral deposition. Lines lished data, 1999) and the least altered samples from the Hvalfjordur basalt, Hafnarfjall
demarcate the basalt field (c.f. Fig. 8). Note that samples more commonly plot to the left and Geitafell central volcanoes (Franzson, 1979; Thorlacius, 1991). See text for further
of the basalt field, indicating apparent depletion. explanation.
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 223

line R to y in Fig. 8. Barrett and MacLean (1994) suggested that by change affecting the altered rocks. It is of course not always possible to
comparing the concentration of immobile elements prior to and after determine the composition of the unaltered protolith. However, since
hydrothermal alteration the mass changes that have taken place Zr can be shown to be essentially immobile in these rocks, we can
during the hydrothermal alteration event can be calculated. The bracket the potential compositional variations in primary composition
calculation of mass change “removes the misleading effects of closure if we can determine the fractionation relationship between Zr and the
(constant sum of 100%) on the relations between elements in un- element of interest in unaltered rocks. By plotting the corrected Zr
treated samples” (Barrett and MacLean, 1994). After determining the concentration against another component and comparing it with the
primary concentration of a given immobile element (in this case Zr) by fractionation trend of the respective volcano one can assess whether
the method described above, we can then calculate the reconstructed enrichment, depletion or no apparent change has occurred during
composition for each rock component by multiplying the original alteration. The extensive data available on the composition of fresh
value by a factor equal to the concentration of Zr in the precursor basaltic rocks from Iceland provides constraints on the fractionation
divided by the concentration of Zr in the altered sample. In this way, trends with respect to Zr. The limited range of fractionation, from the
we have now resurrected the “volume” of each chemical component spread of Zr concentrations in unaltered basalts, provides constraints
to what it was in the primary rock. If dilution represented the only on the possible original variations in other elements and allows us to
change in rock composition, then all elements other than those con- recognize significant mass flux.
tained in the pore-filling precipitates would reveal their original For this exercise, the samples have been subdivided into 3 groups as
values by the correction and would plot on the one-to-one line when described above, i.e. the Hafnarfjall and the Geitafell central volcanoes
compared to an unaltered protolith. This allows us to test the hypo- and the Hvalfjordur plateau basalt succession. Although Icelandic vol-
thesis that mass dilution during pore filling is the dominant chemical canoes follow, in general, a “Thingmuli” tholeiitic evolutionary trend

Fig. 13. Chemical components plotted against Zr. Xs are least altered rocks from Hafnarfjall and Geitafell central volcanoes (Franzson, 1979; Thorlacius, 1991). Diamonds are rocks of
0–33% intensity alteration, squares are rocks of 33–66% alteration and triangles are rocks of 66–100% alteration. The line delineates the primary compositional field of the respective
volcanoes.
224 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

Fig. 13 (continued ).

(Carmichael, 1964), individual centers may show small deviations from which may exceed 1 myr. The Hvalfjordur basalt succession, on the
that trend. Grouping the altered samples according to their petrogen- other hand, has only been subjected to low-temperature burial
esis allows more direct comparison to their primary counterparts and alteration with the regional temperature gradient resulting in
results in a better resolution of compositional changes that have taken temperatures ranging from about 30–100 °C, and is apparently devoid
place during the hydrothermal alteration. An additional advantage is of major geothermal anomalies.
that this allows a comparison of enrichment/depletion between differ- The results from the central volcanoes can be divided into four
ent geothermal environments, and gives a scope for speculation of main categories:
heterogeneity in chemical alteration trends in Icelandic geothermal
(a) TiO2, Ce, Nb, Y and Zr show clear indications of being immobile
systems.
as they fall very near the fractionation line of the least altered
For plotting purposes, the samples have been divided into three
samples.
groups based on their extent of alteration (Fig. 4) at 0–33%, 33–66%
(b) The elements that most consistently show signs of enrichment
and 66–100% alteration (vesicle filling + rock alteration), in order to
are SiO2, Al2O3, FeO, and to a lesser extent, MnO, MgO, Na2O,
show gradual chemical changes, if any, from less to more altered
CaO and possibly Cu.
samples. Figs. 13 and 14 are examples of some of the chemical plots
(c) K2O and Sr are enriched in some samples, but are depleted in
where the relation between individual components is shown against
others.
Zr concentration. Relatively unaltered rocks from the same area are
(d) Other elements do not show a clear trend. However, a lack of
also plotted on these diagrams to indicate the compositional range of
comparison with fresh rock equivalents prevents an assessment
the protoliths of the altered rocks. The overall results are summarized
of elements such as Th, U, Pb, As, S, and partly Rb.
in Table 1. It must be emphasized that the results depict the integrated
chemical changes that have taken place during the history of the The hydrothermal chemical changes between the two central
hydrothermal areas. In particular, the central volcanoes have probably volcanoes are similar. The chemical changes observed in the
undergone several geothermal episodes throughout their lifetime, Hvalfjordur plateau basalts are much more subtle, as expected, due
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 225

Fig. 14. Chemical components plotted against Zr. + are fresh rocks from the Reykjanes–Langjokull volcanic zone (Sveinn Jakobsson, unpublished data, 1999), x are least altered basalt
lavas from Hvalfjordur lava succession, squares are lavas from Krafla central volcano (Karl Gronvold, unpublished data), and triangles are the rock samples of this study.

to its lower geothermal gradient. These do, however, indicate SiO2, the disadvantage of requiring sample by sample comparison to specific
and K2O enrichment and subtle enrichments in Al2O3, FeO, MgO, Na2O precursor rock compositions. A distinct advantage of the isocon
and possibly Sr. Comparison with some of the trace elements is method is that it allows direct evaluation of the mass flux of elements
difficult due to lack of data from fresh rock equivalents. in a specific sample, allowing easier recognition of coupled geochem-
ical behaviour between elements (e.g., K and Na). Our approach in
6. Isocon plots using the isocon plots was to first correct the altered rock geochemical
data for the known effects of dilution and to plot these extrapolated
The data presented above allows us to evaluate the general beha- values against appropriate precursor compositions to identify mass
viour of individual chemical components within a group of samples for flux variations that are not due to simple rock dilution by alteration
an entire hydrothermal field. The method also allows easy comparison minerals.
between the magnitude of compositional variation related to igneous The isocon diagrams (Fig. 15) compare the relationship between the
fractionation trends relative to those that can be attributed to hydro- various components within four representative samples, two chosen
thermal alteration. The isocon method of determining mass flux has from each of the central volcanos. Two of the samples have a primary
226 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

Table 1
Chemical changes in samples subjected to hydrothermal alteration from the Hvalfjordur basalt succession and Hafnarfjall and Geitafell central volcanoes

Chemical Hvalfjordur Hafnarfjall Geitafell


component
Depletion Unchanged Enrichment Depletion Unchanged Enrichment Depletion Unchanged Enrichment
SiO2 X x x
TiO2 X (x) (x)
Al2O3 (x) (x) (x) x x
FeO(tot) X (x) x x
MnO X x (x) (x)
MgO X (x) x (x) x x
CaO X (x) (x) x x (x) x x
Na2O X X (x) x x x x
K2O X (x) x (x) (x) x (x)
P2O5 X (x) x x
Ce X x x
Cu (x) x x (x)
V X ? ?
Zn x x x
Ga X ? ?
Nb X x x
Pb ? ? ?
Rb X ? ? ? (x)? (x)? (x)?
Sr (x) (x) x (x)? ?
Th ? ? x? ? ?
U ? ? ?
Y X x x
Zr X x x
As ? ?
S ? ? ?

x-affirmative, (x)-probable, (x)?-possible, ?-uncertain.

Zr content of about 150 ppm (H-99, G-62), while the others have about volcanic complexes. Most of the hydrothermal waters have low
300 ppm (H-54, G-28) representing relatively primitive and evolved salinity with chloride b200 ppm (Arnorsson and Andresdottir, 1995),
basalt compositions, respectively. The error bars attached to each of consistent with heating of local groundwaters. Exceptions are the
the components represent the range of primary compositional values high-temperature systems on the Reykjanes Peninsula which show
expected at each Zr concentration value within the individual volcano salinity approaching that of seawater (Sveinbjornsdottir et al., 1986).
(Fig. 13). Although error limits on the recalculated compositions are Fluid inclusion studies indicate, however, that much of the alteration
difficult to quantify, one may expect that the error would increase with in the Reykjanes systems may have occurred during an initial
increasing dilution. This is easily observed in Fig. 8 where a small error freshwater stage during the last glacial period (e.g., Franzson et al.,
in petrographic estimation at high dilution could divert the extrapola- 2002). Studies on the alteration zones in several fields show quite
tion line considerably down towards the base line. If the samples have similar distributions of alteration minerals (Fig. 2). There also seems, in
all been properly corrected with respect to mass addition, then the general, to be a close correlation between alteration zones and the
median line should represent the line of no chemical change. Indeed, progressive breakdown of the primary rock and the alteration
the immobile elements (Ce, Zr, Y, Nb) tend to lie on this line. An products formed. The alteration is, to a large extent dependent on
exception is Y in the Hafnarfjall samples, which may possibly be due to temperature, and the maximum temperature at any depth in the
a systematic analytical error in the primary trend in the volcano, as explored systems is constrained by the boiling curve. The common
deduced from the Zr/Y plot (Fig. 13). In contrast, Fig. 15 indicates a occurrence of hyaloclastites in the presently-active systems is not
tendency for slight Al2O3 and FeO enrichment in the altered samples, shared by the fossil hydrothermal systems developed in Tertiary rocks,
as is also clearly indicated in Fig. 13. MgO can be either enriched or which are dominantly in holocrystalline volcanic rocks erupted in the
depleted, as are Na2O and K2O. One might expect that Na2O and K2O absence of a glacial icecap. Volcanic glass is very readily altered during
would generally show similar geochemical behaviour and Fig. 15 geothermal activity, and this will increase the availability of chemical
confirms that the alkalis are either both depleted or both enriched in components (e.g., SiO2) into thermal fluids. Therefore the chemical
any given sample. Na2O and K2O are rather easily leached from rocks flux in the presently-active geothermal systems may be higher than it
during hydrothermal alteration and many rocks are in fact depleted in was for the Tertiary systems.
them. One reason for enrichment may be due to remobilization from It must be emphasized that this study describes the overall changes
nearby more felsic volcanic and intrusive rocks. This hypothesis could that have taken place during the whole lifetime of the hydrothermal
be tested by looking at field relations of the altered basalt samples that systems. Mineralogical studies of vesicle and vein fillings at Geitafell
show alkali enrichment. The isocon plots generally confirm the central volcano show distinct episodes of alteration (Fridleifsson, 1983,
conclusions that have been attained through plots shown in Figs. 9– 1984). The same pattern of chronologically distinct hydrothermal epi-
14 and support the proposition that changes in chemical composition sodes has been recognized in many of the explored active high-tem-
of the altered rocks is mainly due to addition of mass to the rock by perature systems (e.g., Lonker et al., 1993). The number of mineral
deposition of alteration minerals in primary pores. species in individual vesicles and veins in the high-temperature sys-
tems at Reykjanes, Nesjavellir, Olkelduhals and Svartsengi (Franzson,
7. Discussion 1983, 2000; Franzson et al., 2002), based on about 2000 observations,
is very similar. About 70% show two void-filling minerals, about 25%
Geothermal systems in Iceland have a number of common features. contain three and about 5% show four or more void-filling minerals.
The rocks are dominantly of basaltic composition, with minor The proportion of mono-mineralic veins and vesicles has not been
amounts of rocks having more evolved compositions within central checked specifically. It is interesting to note that many minerals seem
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 227

Fig. 15. Isocon plots for four samples from Hafnarfjall (H-54, H-94) and Geitafell (G-28, G-62) central volcanoes. The samples were selected with about 150 and 300 ppm Zr, and from
the 66–100% alteration group and moderate SiO2 enrichments. See text for further explanation.

to grow to a certain threshold grain size, after which they are suc- ridges and ophiolites, are different from the subareal eruptions in Ice-
ceeded by another mineral (e.g. clays, epidote, prehnite, wairakite, land (Batiza and White, 2000).
chalcedony and sometimes quartz). Other minerals like calcite, an- Seafloor eruptions produce more abundant glassy volcanics and
hydrite (the latter in saline systems), and sometimes quartz, are more hyaloclastites than subaerial eruptions in the Tertiary, which should
rarely constrained by size and tend to fill the available space. Whatever enhance the alteration potential of seafloor rocks. However, Pleisto-
the reason, the limit of mineral size does increase the probability that cene eruptions in Iceland often occurred beneath ice cover and resul-
one could find at least two minerals in any one vesicle. Calcite is the ted in extensive formation of hyaloclastite. Seafloor pillow lavas and
mineral that is most likely to fill available space. However, calcite sheet flows, which characteristically show drain-back features, have
generally precipitates at a late stage in the mineral sequence after the very high intrinsic, large scale porosity and permeability, but the
deposition of the other minerals. The above observations imply that permeability drops significantly in older, more deeply buried lavas
typical rocks in a given high temperature system retain open pore due to collapse and infilling by later eruptions (Perfit and Chadwick,
space for sufficient time to record the effects of more than one geo- 1998; Fornari et al., 2004). However, submarine lavas are rarely vesi-
thermal episode. Thus the enrichment/depletion trends tend to accu- culated. In contrast, flow top and bottom breccias and highly vesi-
mulate over the lifetime of the system. It has also been noted that culated flow tops in subaerial lavas provide laterally continuous zones
changes in mineral deposition sequences can be correlated regionally of high porosity that results in anisotropic permeability that favors
within individual geothermal fields (e.g. Franzson, 2000), which im- lateral flow and restricts vertical convection (Fornari and Embley,
plies that mineral deposition is controlled by large scale hydrological 1995; Fornari et al., 2004).
changes in the geothermal system. One major compositional difference between subarial and sub-
It is instructive to compare the low salinity Icelandic geothermal marine lavas is in the concentration of volatile elements, particularly
systems with basalt-hosted seawater-derived hydrothermal fluids (Alt sulfur. Sulfur concentrations of MORBs are typically about 1000 to
and Teagle (2000) and references therein and Butterfield (2000) and 1500 ppm, in contrast to values of about 100 ppm in subarial lavas that
references therein), as evidenced in ophiolitic rocks and “black smoker” have degassed during eruption (Wallace and Anderson, 2000). Because
hydrothermal systems. The overall similarity in the composition of ba- potentially acid-generating volatiles (e.g. SO2, HCl, CO2) are lost during
salts from the seafloor, ophiolites and subaerial Iceland facilitates com- eruption, subsequent hydrothermal leaching will not involve these
parison of these systems. Their physical volcanology does show some species and the resultant hydrothermal fluids may not attain low pHs
differences that could lead to differences in fluid–rock interaction. that facilitate rock alteration. This is particularly true for hydrothermal
Extensional tectonics and rift zone volcanism are common to each of systems developed in plateau basalts, such as at Hvalfjordur. Magmatic
these settings, but the deep water eruptions, common on mid-ocean degassing during intrusive/eruptive events can result in short lived
228 H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229

compositional changes in geothermal systems due to acidification approaching chlorite with SiO2 as low as 35 wt.% (Humphris et al.,
reactions (Armannsson et al., 1982). In hydrothermal systems devel- 1998) and FeOT + MgO N 25 wt.% (Zierenberg et al., 1988). Thus the
oped in longer-lived central volcanoes (e.g. Hafnarfjall, Geitafell), late most significant difference between seawater and meteoric water
stage intrusive rocks can supply both heat and acid-generating volatile recharged basalt-hosted hydrothermal systems is the mass flux from
species that can change the pattern of alteration of the older, degassed the system, not the mineral assemblages which seem to be controlled
lava flows. Degassing of sulfur from basalt can affect the mobility of predominantly by the thermal stability of the alteration phases and
other elements as well, particularly the base metals. For example, secondarily by the composition of the hydrothermal fluids.
submarine basalts erupted at higher pressures due to the weight of the
overlying water column and therefore do not degass sulfur as readily. 8. Conclusions
Copper in submarine basalts is typically concentrated as chalcopyrite
and is generally only mobilised by high temperature (N330 °C) This paper reports a method to compare hydrothermally altered
hydrothermal fluids due to the low solubility of chalcopyrite in basaltic rocks and their protoliths that facilitates correction for mass
sulfide-bearing fluids. In contrast, copper in subareally erupted basalts additions to the rock, during hydrothermal alteration. This method
that have degassed sulfur, chalcopyrite may not be stable leaving takes into account the amount of mineral deposition into the rock, as
copper in a form that is easily leached and transported by low determined by petrography, and recalculates the mass addition using
temperature, oxidized fluids (Lincoln, 1981). the immobile element concentration. The derived recalculated com-
The upper temperature limits in shallow geothermal fields are position of the altered rocks can then be compared with the least
typically set by the boiling temperature curve of the fluid. In deeper, altered protoliths from the same areas. The comparison showed that
higher salinity systems rapid changes in the physical and chemical hydrothermal alteration causes pronounced Si-enrichment as well as
properties of the fluid at temperatures approaching the critical point a considerable Al, Fe, Mg and Mn enrichment. Elements such as Ca, Na,
provide similar temperature constraints. The higher pressures and K, are also mobile and can be either depleted or enriched. Immobile
salinities inherent in submarine geothermal systems therefore favour elements include Zr, Y, Nb, Ce, and probably Ti. An important con-
higher ultimate temperatures of water–rock reaction in the deeper clusion from this work is that the predominant process leading to
portions of these system. It is interesting that the general zonation of chemical flux in these rocks is deposition of alteration minerals in
alteration minerals in Icelandic geothermal systems (Fig. 2) is similar primary pore space, but mass flux due to recrystallization and replace-
to those observed in drill holes throughout the oceanic crust (Alt and ment reactions are of secondary importance. This contrasts with
Teagle, 2000 and references therein; Wilson et al., 2006) and seawater-dominated hydrothermal systems where metasomatic reac-
ophiolites (Schiffman et al., 1991; Gillis and Banerjee, 2000). tions result in extensive mass flux. Another important conclusion is
The major difference in subarial geothermal systems recharged by that similar chemical changes were experienced in the two Tertiary
meteoric water and submarine systems recharged by seawater is the central volcanoes studied, suggesting that these may represent a ge-
salinity of the fluid. Chloride complexing is significant for many cat- neral trend in Icelandic high-temperature systems. A study of a
ions and strongly enhances the solubility of most minerals, especially sample suite of flood basalts within a low-temperature zeolite alter-
at elevated temperatures. Higher concentrations of major element ation environment of Tertiary age also shows chemical changes akin to
cations in seawater-recharged hydrothermal fluids will also affect the the former, but the changes are more subtle. It is also significant that
stability of alteration minerals making it all the more surprising that the general zonation of alteration minerals in Icelandic alteration
the zonation of alteration minerals in seafloor hydrothermal systems zones is similar to those in seawater-dominated hydrothermal
is so similar to that in Iceland. Although the general pattern of systems in spite of the differences in the extent of metasomatism.
alteration mineral zonation is similar, the extent of hydrothermal This implies that temperature, not fluid composition, provides the first
alteration and mass flux from high temperature seafloor systems is order control on mineral assemblages in basalt-hosted hydrothermal
significantly different from that of the Icelandic geothermal systems. systems.
The geochemical data presented in this paper document that the
major changes in the compositions of altered rocks from Icelandic Acknowledgements
geothermal systems result from the addition of elements during
deposition of pore-filling alteration minerals. Isocon plots of altered We thank Howard Day and Shuwen Liu for very helpful discussion on
samples, corrected for dilution, show relatively small mass fluxes for various aspects of this work, and the late Valgardur Stefansson for
most elements, even at high degrees of alteration. This general critically reading the manuscript and his vigorous support. Patrick R. L.
observation is true across the range of alteration zones even in rocks Browne and an anonymous reviewer are thanked for critically reviewing
that show epidote–amphibole alteration. In contrast, basalt/seawater the paper and suggesting a number of improvements. This paper was
reaction results in significant metasomatism of the rocks (Seyfried and written during the sabbatical leave of the first author at the Geological
Ding, 1995). As seawater is heated by circulating in the oceanic crust Department, University of California, Davis in the USA. The financial
on the recharge limb of hydrothermal convection cells, metasomatic support of Orkustofnun to complete this work is acknowledged.
addition of seawater-derived Mg and Ca into basalts releases hydrogen
ions lowering the fluid pH (measured at 25 °C) into the 3–4.5 range References
that is typical of black smoker fluids. Calculated in situ pH at high
temperature are about one unit lower than neutral (Ding and Seyfried, Alt, J.C., Teagle, D.A.H., 2000. Hydrothermal alteration and fluid fluxes in ophiolites and
1992). This low pH fluid has significant capacity for hydrogen meta- oceanic crust. In: Dilek, Y., Moores, E., Elthon, D., Nicolas, A. (Eds.), Ophiolites and
somatism in the deep reaction zone and the rising limb of the con- Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program.
Geol. Soc. Am. Spec. Paper, vol. 349, pp. 273–282.
vective circulation. Acid alteration of the basalts releases a significant
Armannsson, H., Gislason, G., Hauksson, T., 1982. Magmatic gases in well fluids aid in the
flux of cations into the hydrothermal fluids resulting in highly cation- mapping of the flow pattern in a geothermal system. Geochim. Cosmochim. Acta
leached alteration zones, enriched in aluminum and silica, in focussed 46, 167–177.
Arnorsson, S., Andresdottir, A., 1995. Processes controlling the distribution of boron and
upflow zones where integrated water–rock ratios are high (Humphris
chlorine in natural waters in Iceland. Geochim. Cosmochim Acta 59, 4125–4146.
et al., 1998; Zierenberg et al., 1988, 1995, 1998; Teagle and Alt, 2004). Arnorsson, S., Gunnlaugsson, E., Svavarsson, H., 1983. The chemistry of geothermal
Silicified basalts in hydrothermal upflow zones can have silica con- waters in Iceland. II. Mineral equilibria and independent variables controlling water
tents in excess of 80 wt.% (e.g. Zierenberg et al., 1995). Alternatively, Fe compositions. Geochim. Cosmochim. Acta 47, 547–566.
Batiza, R., White, J.D.L., 2000. Submarine lavas and hyaloclastites. In: Sigurdsson, H.,
and Mg metasomatism is hydrothermal upflow zones can result in Houghton, B.F., McNutt, S.R., Rymer, H., Stix, J. (Eds.), Encyclopedia of Volcanoes.
pervasive chloritization of basalt resulting in bulk compositions Academic Press, San Diego, CA, pp. 361–381.
H. Franzson et al. / Journal of Volcanology and Geothermal Research 173 (2008) 217–229 229

Barrett, T.J., MacLean, W.H., 1994. Chemostratigraphy and hydrothermal alteration in Lincoln, T.N., 1981. The redistribution of copper during low-grade metamorphism of the
exploration for VHMS deposits in greenstones and younger volcanic rocks. In: Lenz, Karmutsen Volcanics, Vancouver Island, British Columbia. Econ. Geol. 76, 2147–2161.
D.R. (Ed.), Alteration and Alteration Processes Associated with Ore-forming Sys- Lonker, S.W., Franzson, H., Kristmannsdottir, H., 1993. Mineral–fluid interactions in the
tems. Geological Association of Canada, Short Course Notes, vol. 11, pp. 433–467. Reykjanes and Svartsengi geothermal systems, Iceland. Am. J. Sci. 293, 605–670.
Bodvarsson, G., 1983. Temperature/flow statistics and thermomechanics of low-tem- Marty, B., Gunnlaugsson, E., Jambon, A., Oskarsson, N., Ozima, M., Pineau, F., Torssander,
perature geothermal systems in Iceland. J. Volcanol. Geotherm. Res. 19, 255–280. P., 1991. Gas geochemistry of geothermal fluids, the Hengill area, southwest rift
Butterfield, D.A., 2000. Deep ocean hydrothermal vents. In: Sigurdsson, H., Houghton, B.F., zone of Iceland. Chem. Geol. 91, 207–225.
McNutt, S.R., Rymer, H., Stix, J. (Eds.), Encyclopedia of Volcanoes. Academic Press, San Neuhoff, P.S., Fridriksson, Th., Arnorsson, S., Bird, D.K., 1999. Porosity evolution and
Diego, CA, pp. 857–875. mineral paragenesis during low-grade metamorphism of basaltic lavas at Teigar-
Carmichael, I.S.E., 1964. The petrology of Þingmúli, a Tertiary volcano in eastern Iceland. horn, eastern Iceland. Am. J. Sci. 299, 467–501.
J. Petrol. 5, 435–460. Perfit, M.R., Chadwick Jr., W.W., 1998. Magmatism at mid-ocean ridges, constraints from
Ding, K., Seyfried Jr., W.E., 1992. Determination of Fe–Cl complexing in the low pressure volcanological and geochemical investigations. Am. Geophys. Union Monogr. 106,
supercritical region (NaCl fluid): iron solubility constraints on pH of subseafloor 59–115.
hydrothermal fluids. Geochim. Cosmochim. Acta 56, 3681–3692. Riedel, C., Schmidt, M., Botz, R., Theilen, F., 2001. The Grimsey hydrothermal field
Elders, W.A., Fridleifsson, G.O., 2005. The Iceland Deep Drilling Project—scientific offshore North Iceland, crustal structure, faulting and related gas venting. Earth
opportunities. Proc. World Geothermal Congress. Antalya, Turkey. 7 pp. Planet. Sci. Lett. 193, 409–421.
Fornari, D.J., Embley, R.W., 1995. Tectonic and volcanic controls on hydrothermal Schiffman, P., Fridleifsson, G.O., 1991. The smectite–chlorite transition in drillhole NJ-15,
processes at the mid-ocean ridge; an overview based on near-bottom and submer- Nesjavellir geothermal field, Iceland: XRD, BSE and electron microprobe investiga-
sible studies. In: Humphris, S.E., Zierenberg, R.A., Mullineaux, L.S., Thomson, R.E. tions. J. Metamorph. Geol. 9, 679–696.
(Eds.), Seafloor Hydrothermal Systems: Physical, Chemical, Biological and Geolo- Schiffman, P., Evarts, R.C.E., Williams, A.E., Pickthorn, W.J., 1991. Hydrothermal
gical Interactions, Am. Geophys. Union Monograph, vol. 91, pp. 1–46. metamorphism in oceanic crust from the Coast Ranges ophiolite of California:
Fornari, D., Tivey, M., Schouten, H., Perfit, M., Yoerger, D., Bradley, A., Edwards, M., fluid–rock interaction in a rifted island arc. In: Peters, T.J., Nicolas, A., Coleman, R.G.
Haymon, R., Scheirer, D., Von Damm, K., Shank, T., Soule, A., 2004. Submarine lava (Eds.), Ophiolite Genesis and Evolution of the Oceanic Lithosphere. Kluwer
flow emplacement at the East Pacific Rise 9° 50′ N: implications for uppermost Academic Publishers, Dordrecht, pp. 399–425.
oceanic crust stratigraphy and hydrothermal fluid circulation. In: German, C.R., Lin, Seyfried Jr., W.E., Ding, K.G., 1995. Phase equilibria in subseafloor hydrothermal
J., Parson, L.M. (Eds.), Mid-Ocean Ridges: Hydrothermal Interactions Between the systems; a review of the role of redox, temperature, pH and dissolved Cl on the
Lithosphere and Oceans. Amer. Geophys. Union Monograph, vol. 148, pp. 187–217. chemistry of hot spring fluids at mid-ocean ridges. Am. Geophys. Union Monogr. 91,
Franzson, H. 1979. Structure and petrochemistry of the Hafnarfjall-Skarðsheiði central 248–272.
volcano and the surrounding basalt succession, W-Iceland. PhD thesis, University of Sigurdsson, O., Stefansson, V., 1994. Forðafræðistuðlar. Mælingar á sýnum. (Reservoir
Edinburgh, United Kingdom. parameters. Measurements of rock samples). Report Orkustofnun OS-94049/JHD-
Franzson, H., 1983. The Svartsengi high temperature field, Iceland. Subsurface geology 28 B (in Icelandic), 35 pp.
and alteration. Geothermal Resources Council, Transactions, vol. 7, pp. 141–145. Sigurdsson, O., Gudmundsson, A., Fridleifsson, G.O., Franzson, H., Gudlaugsson, S.Th.,
Franzson, H. 1999. Forðafræðistuðlar. Samband ummyndunar og efnaflutninga í basalti á Stefansson, V., 2000. Database on igneous rock properties in Icelandic geothermal
Íslandi. (Petrophysics. The relationship between hydrothermal alteration and systems, status and unexpected results. Proc. World Geothermal Congress, Kyushu,
chemical transport in basaltic rocks in Iceland). Report Orkustofnun OS-99108 Tohuku, Japan, pp. 881–2886.
(in Icelandic), 31 pp. Stefansson, V., Sigurdsson, O., Gudmundsson, A., Franzson, H., Fridleifsson, G.O.,
Franzson, H., 2000. Hydrothermal evolution of the Nesjavellir high-temperature Tulinius, H., 1997. Core measurements and geothermal modelling. Second Nordic
system, Iceland. World Geothermal Congress, Japan, May 28–June 10, 2000, p. 6. Symposium on Petrophysics. Fractured reservoir. Nordic Petroleum Series, vol. 1, pp.
Franzson, H., Fridleifsson, G.O., Gudmundsson, A.,Vilmundardottir, E.G. 1997. Forðafræðist- 198–220.
uðlar. Staða rannsóknar í lok 1997 (Reservoir parameters. Status of petrological studies Sveinbjornsdottir, A.E., Coleman, M.L., Yardley, B.W.D., 1986. Origin and history of
by the end of 1997). Report Orkustofnun OS-97077 (in Icelandic), 57 pp. hydrothermal fluids of the Reykjanes and Krafla geothermal fields, Iceland. Contrib.
Franzson, H., Gudlaugsson, S.Th., Fridleifsson, G.O., 2000. Petrophysical properties of Mineral. Petrol. 94, 99–109.
Icelandic rocks. Sixth Nordic Symposium on Petrophysics. Nordic Petroleum Teagle, D.A.H., Alt, J.C., 2004. Hydrothermal alteration of basalts during formation of the
Technology Series VI. ISBN: 82-994330-6-1. 14 pp. Bent Hill massive sulfide deposits, Middle Valley, Juan de Fuca Ridge. Econ. Geol. 99,
Franzson, H., Thordason, S., Bjornsson, G., Guðlaugsson, S.Th., Richter, B., Fridleifsson, G.O., 561–584.
Thorhallsson, S., 2002. Reykjanes high-temperature field, SW-Iceland. Geology and Thorlacius, J.M., 1991. Petrological Studies of the Geitafell Central Volcano, S.E. Iceland.
hydrothermal alteration of well RN-10. Proceedings, 27th Workshop on Geothermal MSc thesis, University of Edinburgh, United Kingdom.
Reservoir Engineering. Stanford University, pp. 233–240. Wallace, P., Anderson Jr., A.T., 2000. Volatiles in magmas. In: Sigurdsson, H., Houghton,
Fridleifsson, G.O., 1983. The geology and alteration history of the Geitafell central B.F., McNutt, S.R., Rymer, H., Stix, J. (Eds.), Encyclopedia of Volcanoes. Academic
volcano, SE Iceland. PhD thesis, Edinburgh University, United Kingdom. Press, San Diego, CA, pp. 149–170.
Fridleifsson, G.O., 1984. Mineralogical evolution of a hydrothermal system II: heat Wilson, D.S., Teagle, D.A.H., Alt, J.C., et al., 2006. Drilling to gabbro in intact ocean crust.
sources, fluid interactions. Trans. Geothermal Resources Council, vol. 8, pp. 119–123. Science 312, 1016–1020.
Gillis, K.M., Banerjee, N.R., 2000. Hydrothermal alteration patterns in supra-subduction Zierenberg, R.A., Shanks, W.C., Seyfried, W.E., Koski, R.A., Strickler, M.D., 1988.
zone ophiolites. In: Dilek, Y., Moores, E., Elthon, D., Nicolas, A. (Eds.), Ophiolites and Mineralization, alteration, and hydrothermal metamorphism of the ophiolite-
Oceanic Crust: New Insights from Field Studies and the Ocean Drilling Program. hosted Turner–Albright sulfide deposit, southwestern Oregon. Journal of Geophy-
Geol. Soc. Am. Spec. Paper, vol. 349, pp. 283–297. sical Research 93 (B5), 4657–4674.
Grant, J.A., 1986. The isocon diagram—a simple solution to the Gresens equation for Zierenberg, R.A., Schiffman, P., Jonasson, I.R., Tosdal, R.M., Pickthorn, W.J., McClain, J.S.,
metasomatic alteration. Econ. Geol. 81, 1976–1982. 1995. Alteration of basalt hyaloclastite at the off-axis Sea Cliff hydrothermal field,
Gudmundsson, A., Franzson, H., Fridleifsson, G.O., 1995. Forðafræðistuðlar. Söfnun sýna Gorda Ridge. Chem. Geol. 126, 77–99.
(Reservoir parameters. Sample collection). Report Orkustofnun OS-95017/JHD-11 B Zierenberg, R.A., Fouquet, Y., Miller, D.J., Bahr, J.M., Baker, P.A., Bjerkgarden, T., Brunner, C.A.,
(in Icelandic), 72 pp. Duckworth, R.C., Gable, R., Gieskes, J.M., Goodfellow, W.D., Groeschel-Becker, H.M.,
Humphris, S.E., Alt, J.C., Teagle, D.A.H., Honnorez, J.J., 1998. Geochemical changes during Henrike, M., Guerin, G., Ishibashi, J., Iturrino, G.J., James, R.H., Lackschewitz, K.S.,
hydrothermal alteration of basement in the stockwork beneath the active TAG Marquez, L.L., Nehlig, P., Peter, J.M., Rigsby, C.A., Schultheiss, P.J., Shanks III, W.C.,
hydrothermal mound. Proc. ODP Sci. Results 158, 255–276. Simoneit, B.R.T., Summit, M., Teagle, D.A.H., Urbat, M., Zuffa, G.G., 1998. The deep
Leitch, C.H.B., Lentz, D.R., 1994. The Gresens approach to mass balance constraints of structure of a sea-floor hydrothermal deposit. Nature 392, 485–488.
alteration systems: methods, pitfalls, examples. In: Lenz, D.R. (Ed.), Alteration and
Alteration Processes Associated with Ore-forming Systems. Geological Association
of Canada, Short Course Notes, vol. 11, pp. 161–192.

Potrebbero piacerti anche