Geothermal Systems Ancient and Modern: A Geochemical Review

Download as pdf or txt
Download as pdf or txt
You are on page 1of 50

Earth-Science Reviews, 19 (1983) 1--50 1

Elsevier Scientific Publishing Company, Amsterdam -- Printed in The Netherlands

Geothermal Systems Ancient and Modern:


A Geochemical Review
R.W. Henley and A.J. Ellis

ABSTRACT

Henley, R.W. and Ellis, A.J., 1983. Geothermal systems ancient and modern: a geochemi-
cal review. Earth-Sci. Rev., 19: 1--50.

Geothermal systems occur in a range of crustal settings. The emphasis of this review is
on those occurring in regions of active or recently active volcanism, where magmatic heat
at depths up to 8 km leads to convection of groundwater in the upper crust. Hot water
(and steam) flows are controlled by the permeability of the crust and recent data have
emphasised the dominance of secondary permeability, especially fractures. Drilling to
depths of up to 3 km in these systems encounters near-neutral pH alkali chloride waters
with temperatures up to about 350°C and chloride contents generally in the range 500 to
15,000 mg kg -1 although much higher salinities are encountered in some systems such as
in the Imperial Valley, California. Stable isotope studies indicate the predominance of a
meteoric source in the majority of geothermal systems although seawater predominates in
some regions, such as Reykjanes, Iceland. Mixing of waters from both sources also occurs
in some systems and some magmatic fluid may also be present.
The major element geochemistry of geothermal fluids is determined by a set of tempe-
rature-dependent mineral-fluid equilibria although chloride and rare gas contents appear
to be independent variables reflecting the sources of these components (sedimentary or
volcanic rocks, seawater, magmatic fluids, etc).
Boiling in the upper portion of geothermal systems is accompanied by the transfer of
acidic gases (CO2 and H2S) to the resultant steam which may penetrate the surface as
fumarolic activity or become condensed into shallow groundwaters giving rise, with oxi-
dation, to distinctive low pH sulphate bicarbonate water.
Fluid inclusion, stable isotope and mineral alteration studies have led to the recogni-
tion in many Tertiary hydrothermal ore deposits of physical and chemical environments
analogous to those encountered in the present-day systems. The vein-type gold-silver, Car-
lin-type gold and porphyry-type copper-molybdenum deposits of the western United
States are particularly well studied examples. Sub-ocean floor equivalents of the terres-
trial geothermal systems have been recognized in ocean floor spreading centres such as the
East Pacific Rise and deep-sea submersible vehicles have allowed visual observation of sea
floor hot springs actively depositing metal sulphides. These environments may parallel
those of the Cyprus-type massive sulphide depositing systems, while sub-sea floor systems
of the type responsible for Kuroko-type massive sulphide deposits may eventually be
encountered in island arc settings.

INTRODUCTION

Geo th er mal systems occur most f r equent l y in regions of active or recently


active volcanism such as t he Pacific margin. Many such systems are under-

0012-8252/83/0000--0000/$17.50 © 1983 Elsevier Scientific Publishing Company


2

going intensive exploration for energy utilization (e.g., Kenya, Philippines,


Japan, Iceland, Indonesia, Mexico, Central and South America and New Zea-
land) and a number of systems are already exploited for electricity genera-

Richard W. Henley received a BSc. in Geology in 1968


from the University of London and a PhD. in geochemistry
from the University of Manchester in 1971 following expe-
rimental studies of gold transport in hydrothermal solu-
tions and the genesis of some Precambrian gold deposits.
Dr. Henley was Lecturer in Economic Geology at the Uni-
versity of Otago, New Zealand from 1971 to 1975 and at
Memorial University, Newfoundland until 1977. Research
interests have focussed on the mode of origin of a number
of different types of ore deposit including post-metamor-
phic gold tungsten veins, porphyry copper, massive sul-
phide and placer gold deposits. He is currently Head of the
Geothermal Chemistry Section of the Department of
Scientific and Industrial Research at Wairakei, New Zea-
land, Chairman of the International Rock-Water Inter-
action Group's Interest Group on Active Geothermal Sys-
tems, and a visiting lecturer at the Auckland Geothermal
Institute. His present research includes a number of iso-
tope and chemical studies relating to the exploration and
development of geothermal systems and geothermal impli-
cations for the origin of ore deposits. Present address:
DSIR -- Geothermal Research Centre, Wairakei, Private
Bag, Taupo, New Zealand.

A. James Ellis is a graduate of Otago University, New Zea-


land, where he obtained a Msc. degree in chemistry in
1953 and a PhD. in geochemistry in 1957. He worked in
Chemistry Division, Department of Scientific and Indus-
trial Research, New Zealand from 1958 as a geochemist,
undertaking geothermal development work, high tempera-
ture and pressure water/mineral research, and geochemical
prospecting projects. He became Director of the Chemistry
Division in 1971, but continued with geothermal geochem-
istry projects in New Zealand and in many other countries.
In 1979 he was appointed Assistant Director-General of
the New Zealand Department of Scientific and Industrial
Research with responsibility for the chemistry, physics,
and engineering branches of the Department.
tion or directly for industrial energy. In these volcanic geothermal systems
temperatures may reach ~350°C at exploitable depths (<2.5 km). Recent
deep-ocean exploration has located superficially similar sub-sea floor systems
in sites of active ocean spreading such as the Mid-Atlantic Ridge, The East
Pacific Rise and Galapagos Spreading Centre (Green et al., 1981, and see p.
37). Hot springs are well known also in active tectonic zones associated with
plate convergence, as exemplified by the European Alps, Rocky Mountains,
Himalaya and the New Zealand Southern Alps (Barnes et al., 1978a). Here
meteoric waters penetrate into rocks, so rapidly uplifted and eroded that
their deep heat contents have not been dissipated by conduction to the sur-
face (Allis et al., 1979; Allis, 1981a). In more stable continental environ-
ments, in deep sedimentary basins such as in Hungary or Western Siberia, hot
water occurs at temperatures reflecting normal crustal geothermal gradients.
High pressure ('geopressured') reservoirs have been found associated with
thick sedimentary wedges at continental margins such as in the Gulf of
Mexico (Kharaka et al., 1978). Because of the wide range of environments,
rock types, temperatures and degree of exploration, we restrict this review
to the high temperature systems associated with volcanism. Evidence for the
existence of geothermal systems in the past and their characteristics as metal-
ore forming systems are briefly reviewed in relation to present-day systems.
The emphasis in this review is on the deep system rather than on the prop-
erties of natural surface discharges where complex mixing and steam losses
occur (e.g., Fournier, 1979a). No comprehensive attempt is made to review
the continuing development of stable isotope techniques in geothermal inves-
tigations as this was covered by Truesdell and Hulston (1979). Ellis and
Mahon (1977) gave an extensive general review of geothermal geochemistry
and Rybach and Muffler (1981) have compiled a useful set of source papers
covering a number of aspects of geothermal resource assessment and develop-
ment.

PHYSICAL CHARACTERISTICS OF GEOTHERMAL SYSTEMS

In regions of anomalously high heat flow, thermal convection dominates


the behaviour of groundwater in permeable crust (Elder, 1966, 1981) and
generates geothermal systems where hot upflow waters approach the surface.
Stable isotope studies and analyses of rare gases in geothermal discharges
(Mazor, 1975) have shown that the waters in these convective terrestrial geo-
thermal systems are largely derived from air-saturated meteoric water which
has penetrated the crust to the level occupied by cooling magmatic systems.
Electrical resistivity soundings have shown that these hot water systems
extend to depths in excess of 5 km. Fig. 1 shows the general features of a hy-
pothetical geothermal system. Analogous seawater-fed systems are now re-
cognised in submarine environments as well as hybrid seawater--meteoric-
water systems in oceanic islands and coastal regions, e.g., Matupi Harbour,
New Britain (Green et al., 1978).
Near
Meteoric Acid neutrot pH
water Sulphate& [htoride Ditute
I i Steaming bicarbonate boiling ChLoride

Approx
Scale

I km
HEATAND MASS (NoEl,tO2, 502, H20.)
TRANSFERFROMMAGMA SYSTEMS.
KEY
Pre-VoLcanic Basement Steam-hented Acid SOt _4- HEO~ waters

Intrusive Votconics F~TT~ SO~-[t" waters (fig 2)

Low Permeability Stratum Near neutral [htoride waters


eg. Mudstones (within 200° Isotherm opprox,)

Two Phase Region


Water Liquid + Steam (+Oos)

Fig. 1. Schema of the main features of a geothermal system typical of those in silicic vol-
canic terrane. Water recharge is provided by meteoric groundwaters and heat supplied, to-
gether with some gases, chloride, water and other solutes, by a deeply buried magmatic
system leading to a convective column of near-neutral pH chloride water. Two-phase
conditions occur in the upper level of the system. Steam separation processes give rise to
fumaroles and steam adsorption by groundwater, with oxidation of H2S at the water
table, gives rise to steam-heated acid sulphate -- bicarbonate waters. Mixing may occur
between deep near-neutral pH and steam-heated waters to give relatively oxidising
C1--S04 waters. Outflows from the deep system occur either directly as boiling alkaline
pH boiling springs generally associated with the formation of silica sinter or after dilution
as near-neutral pH chloride springs.
In these geothermal systems, continuous intrusive activity or cooling of
large plutons within the upper 10 km of the crust provides the thermal
energy. "1 Little direct evidence exists concerning the composition of these
rocks or their mechanism of heat loss, but the recent eruption of basalt sco-
ria from geothermal wells at Krafla, Iceland leaves little d o u b t about the vol-
canic heat source in this area (Larsen et al., 1979). Mineral alteration and
devitrification reactions (10--300 kJ kg-' ) and the decay of radioactive nu-
clides (6 • 10 -14 kJ kg -1 sec -1 , Fenn et al., 1978) may contribute a minor
fraction of the total heat flow of the systems. Stratigraphic evidence suggests
that the duration of geothermal activity is in the range l 0 s to 106 years
(Browne, 1979) while the heat flow, for example, of the undisturbed Waira-
kei system, 4.2 • l 0 s kJ sec-' (420 MW) is equivalent to the solidification of
about 0.1 m 3 sec-' of basalt, or for a 300,000 year life-span, about 700 km 3
basalt. The examination of deeply eroded fossil geothermal systems such as
p o r p h y r y copper or epithermal ore deposits can be instructive in considering
the deep thermal structure of terrestrial geothermal systems and this is dis-
cussed later.
Steady-state thermal convection dominates conductive heat flow in water-
saturated crust where the dimensionless Rayleigh Number (R a) for the sys-
tem exceeds a critical value (1 to 40) which depends on the criteria adopted
for the variation of fluid characteristics within the system (Horne and O'Sul-
livan, 1977a; Sorey, 1978; Straus and Schubert, 1977a):
K~gA 0 H
R a - - - -

K7
where K is the permeability of the system, ~, the thermal expansivity of the
fluid (about 103 cm 3 0 C-1 ), g acceleration due to gravity (103 cm sec -~ ), K,
thermal diffusivity of fluid saturated crust (about 3 × 10 -3 cm 2 sec -1 ), 7,
kinematic viscosity of fluid (about 12 10 -4 cm 2); H, depth of system
(about 8 • 10 s cm) and A0 is the vertical temperature difference across the
system (500°C). The Rayleigh Number is strongly dependent on the large-
scale permeability of the system so that, for the values inserted above, con-
ductive heat transfer dominates below permeabilities of about 10 -17 m 2
(0.01 millidarcy). Inserting typical values of K (10 -14 -- 10 -is m s ) found in
explored geothermal systems shows that the Rayleigh Number (500 to 5000)
greatly exceeds the critical value and vigorous convection is to be expected
(Elder, 1966).
As a guide to minimum crustal permeabilities, Norton and Knapp (1977)
examined the 'flow' or interconnected porosities of various rock types,
which were in the range 10 .3 to 10 -s compared to total porosities of 10 -1 to

"1 White and Guffanti (1979) and Simkin (1979) have reviewed recent geophysical
data, especially teleseismic P-wave delays, which are beginning to outline molten mag-
ma chambers beneath the active Yellowstone, Geysers-Clear Lake, Long Valley, Coso
and Krafla geothermal areas.
10 -2 . Their analysis suggested t h a t permeabilities of the order of 10 -1~ m 2
may characterise large regions of the earth's crust. The higher values of K in
geothermal areas reflect the dominance of fractures on fluid flow: e.g., K
values of 10 -14 -- 10 -is m 2 may be accounted for by a single crack about
0.04 m m wide, cutting each cubic metre of rock in the system.
Early fluorescein injection and more recent isotope tracer studies (Barry
et al., 1979) at Wairakei have confirmed the dominance of high permeability
paths such as fracture networks through the field. Grindley and Browne
(1975) noted the dominance of such secondary permeability as open fault
zones in the Broadlands field, and James (1975) examined fissure flow to
geothermal wells at Cerro Prieto and E1 Tatio. Batzle and Simmons (1976)
carefully d o c u m e n t e d the history of fracturing and sealing in the Dunes and
Raft River systems. McNabb and Henley (1979) examined the interaction of
fluids flowing in fractures with fluid contained in lower permeability wall
rocks and suggested that chemical and thermal exchanges occur by diffusion
between the two fluid regimes. They also pointed out that because heat dif-
fuses some 15--30 times faster than ions in solution, changes in the tempera-
ture and in the chemistry of solutions within a porous rock will occur at dif-
ferent rates when a cross-cutting fracture is invaded by a fluid which is out
of thermal and chemical equilibrium with the wall rock.
Donaldson (1962), Norton and Knight (1977), Cathles (1977), and H o m e
and O'Sullivan (1977b) used computational and laboratory methods to
model the patterns of fluid flow and temperature due to large-scale convec-
tion systems. These models each require certain boundary conditions and
assumptions concerning the homogeneity of the systems and nature of the
heat source, but they readily confirm the ability of the convective cooling of
plutons to maintain large crustal volumes (20--50 km 3 ) at temperatures up
to 500°C for around l 0 s years. The convective models of Donaldson (1962)
and Horne and O'Sullivan (1977b) adopt a constant-temperature heat source
such as might occur through repetitive intrusion, convective overturn of mag-
ma or, as advocated by Lister (1974, 1975), a process of steady state crack-
ing of the pluton during cooling which greatly enhances the surface area for
heat transfer from the pluton to the fluid. In contrast Norton and Knight
(1977) and Cathles (1977) introduced a rate-limiting step on the convective
system by considering that the pluton itself cooled conductively as is certain-
ly the case at the end of a magmatic episode.
Geothermal systems are often associated with caldera structures, e.g., sys-
tems in the R o t o r u a and Okataina Calderas, New Zealand, Yellowstone Val-
ley, Long Valley, Coso, U.S.A. and Krafla, Iceland. Many fossil ore-forming
hydrothermal systems are similarly associated, e.g., McDermitt Caldera, Ne-
vada, Creede and Silverton, Colorado and Vatukoula, Fiji. It may be that
deep penetration of meteoric fluids for the establishment of geothermal ac-
tivity is related to an extensive period of deep crustal fracturing related to
caldera formation.
Just as the nature of the heat source at depth in a geothermal system is
imprecisely known, so t o o is the character of fluids at temperatures > 350°C
below explored depths.
Wooding (1963) and Henley and McNabb (1978}, in discussing convection
in porous media, have emphasised the role of dispersive mixing of cool and
h o t fluids on the margin of upward moving convective plumes. The magni-
tude of the dispersion is related to the degree of fracturing in the medium so
that, in very permeable systems, the isotope and chemical compositions of
near surface geothermal waters may be dominated by the diluting ground-
waters rather than the deeper geothermal fluids.
Griffiths (1978) has discussed the possibility, raised by McNabb {1975)
that geothermal systems arise from diffusive heat and mass transfer above a
denser brine 'layer' at depth. Both these approaches suggest that brines of
salinity higher than near-surface geothermal waters may exist in the roots of
geothermal systems and be derived from a magmatic source or from deep
connate-metamorphic waters such as postulated in a number of ore-genesis
studies. In either case, dilution by lower-salinity waters during upflow pro-
gressively obliterates the chemical or isotope signature of the possible deep
fluid. Truesdell and Hulston (1979) stress the absence of isotope evidence
for contribution of magmatic water but note both evidence for magmatic gas
contributions and for non-magmatic origins of some other solutes.
Henley and McNabb (1978) in correlating the hydrothermal environments
of p o r p h y r y copper and geothermal systems, suggested that magmatic fluid * 1
may be injected from a cooling pluton in the depths of the system, although
the ultimate source of the water may be complex. A contribution of up to
10% magmatic fluid (originally at about 2 wt.% NaC1) could account for all
the chloride and up to 30% of the heat budget for a system like Wairakei.
Assuming that Wairakei discharged chloride water at the present natural rate
for the last 300,000 years requires the availability of 4 • 10 is g C1-. Ellis and
Mahon (1964, 1967) showed that 800 ppm C1- was available from rock by
leaching, so that some 2000 km 3 of rock would be required to be totally
leached to supply this c o m p o n e n t during the system lifetime. Even if a
steady state is not assumed (Ellis, 1970, Healy, 1975), about 70 km 3 would
be required during the 104 year heating period of the system during each
phase of its intermittent activity, a volume roughly equivalent to the present
extent of the system. As complete chloride extraction from the local system
rocks is unlikely, an additional deep or widespread source of components
seems necessary. Sigvaldason (1979) has also c o m m e n t e d on evidence for a
magmatic input of chloride. In high-salinity systems such as the Salton Sea,
seawater or evaporite sources of chloride dominate any contributory leach
or magmatic sources.

* 1 We use t h e t e r m ' m a g m a t i c f l u i d ' to i n d i c a t e a fluid w h i c h has evolved t h r o u g h c h e m -


ical a n d i s o t o p e e q u i l i b r i u m w i t h a silicate melt, regardless of t h e p r e v i o u s h i s t o r y of
its c o m p o n e n t water.
Early exploration well data from Wairakei and Broadlands showed that
temperatures in the unexploited system lay close to the boiling point with
depth curve for water (Haas, 1971), although the physical behaviour of wa-
ter in such a two-phase convective upflow zone is not well understood. Well
discharges from the Broadlands system are characterised by carbon dioxide
contents an order of magnitude larger than at Wairakei. In this and similar
cases such as Ngawha (Giggenbach and Lyon, 1978) the boiling-point depth
function is unique to the initial CO2 content and enthalpy of the single-
phase deep fluid (Sutton and McNabb, 1977; Mahon et al., 1980a).
In some fields low porosity combined with high heat flow may cause a
steam zone to develop near the surface of a system either naturally (White
et al., 1971; Ellis and Mahon, 1977; Truesdell and White, 1973) or follow-
ing extensive well production (James, 1968; Weres et al., 1977). Principal
examples are the Geysers, California and Lardarello, Italy. D'Amore and
Truesdell (1979) discussed the complex interplay between 'reservoir'
steam, steam evaporating from the deep system and steam regenerated from
condensate during the exploitation of such systems and Allis (1981b) has
related changes in seismicity to the cumulative effects of steam withdrawal
from the Geysers field. Allis (1979a, b) and Henley (1979) described the
generation of such artificial steam-dominant zones in the Wairakei and
adjacent Tauhara geothermal systems consequent on the large-scale exploita-
tion of Wairakei.
Steam or steam and gas escaping from deep hot-water systems may appear
at the surface as fumarolic activity or may be absorbed by superficial
groundwaters, perhaps perched above a low permeability cap-rock, to form
distinctive sulphate--bicarbonate thermal waters with low chloride contents
or low pH sulphate waters (Fig. 1). At Broadlands and Kawah Kamojang,
Mahon et al., (1980b) have recognized the occurrence of sodium bicarbonate
waters where steam and CO2 derived from deep boiling is adsorbed into un-
mineralised groundwater. Each of the types of water mentioned may interact
with each other to produce hybrid waters such as the sulphate--chloride
water of the Frying Pan Lake, Waimangu and the chloride--bicarbonate wa-
ters encountered in some shallow Broadlands wells. Where steam flows are
focussed into low flow groundwater zones or pools of condensate, steady
state evaporation may lead to concentration of components and stable
isotope enrichment (Giggenbach, 1978; Henley and Stewart, 1982).
In volcanic regions of high topographic relief, extensive lateral flows
of h o t water may occur for up to 20 km from the centre of the geothermal
field, as at E1 Tatio, Chile and Ahuachapan, E1 Salvador (Healy and Hoch-
stein, 1973; Healy, 1975) (Fig. 2). Chloride springs may consequently be a
considerable distance from the exploitable geothermal reservoir, the locality
of which may be identified by fumarole activity or the presence of steam-
heated sulphate--bicarbonate waters higher on the volcanic slopes. Geo-
thermal activity associated with the Hakone Volcano, Japan (Oki and
Hirano, 1970), and the Kamojang system, Indonesia (Kartokusumo et al.,
1975) is expressed in this way.
o ~

N,-~ o
0 ~,.Q

7.~e
N2 ~

oO

_9
o \ -~ 0 ~

~u
z

0 0 0 0

r~

o " o o o o o o~ °

Ooo~o o ~ 0 m
o Oo° o ° o
o
o
o
o
o o~o~
L'C~o~O o°o o

._ ~.~

km
10

ROCK--WATER INTERACTIONS IN GEOTHERMAL SYSTEMS

Hydrothermal alteration

Hydrothermal alteration is a general term embracing the mineralogical,


textural and chemical response of rocks to a changing thermal and chemical
environment in the presence of h o t water, steam or gas. The limitations of
the term are poorly defined so that a complete transition exists through to
burial metamorphism (Zen, 1974). The criteria of distinction may well be
those discussed above, i.e., the extent of heat and mass transfer within the
system, so that low-grade burial metamorphic terrain would be characterised
by isochemical metamorphism and lower water/rock ratios than in geother-
mal systems.

Water~rock ratios

Sheppard et al. (1969, 1971), Taylor (1974), Magaritz and Taylor (1975)
and Criss (1981) showed that extensive oxygen isotope exchanges have oc-
curred on a large scale, implying the activity of extensive hydrothermal sys-
tems in and adjacent to cooling batholiths (e.g., Idaho Batholith, Coast
Range Batholith, British Columbia), as well as during the formation of por-
p h y r y ~ o p p e r deposits and a range of other ore-depositing systems.
In British Columbia, for example, meteoric waters appear to have affected
a huge volume of rock extending 300 km eastward from the quartz diorite
batholith. From an oxygen isotope balance calculated water/rock (w/r)
volume ratios ranged from about 0.001--0.003 for the batholithic rocks to
values of 0.1--3.0 for rocks to the east of the batholith. Reviewing data for
this and similar studies of ore deposits, Taylor (1974) showed that these
minimum ratios normally range up to 3 or 4. For the Wairakei and the Broad-
lands geothermal fields in New Zealand, from oxygen isotope measurements
Stewart (1978) found overall water/rock ratios of 2, and 0.7, respectively.
The difference between these t w o fields would agree with the higher perme-
ability observed in the Wairakei system. The Wairakei system may, in fact,
approach the maximum water/rock ratio occurring in convective geothermal
systems. The ratio is similar in value to the highest figure found in the Coast
Range (Magaritz and Taylor, 1975). Information of this t y p e from drilled
geothermal areas is biased toward the conditions in the most permeable part
of systems and should not necessarily be taken as representative of the peri-
pheral and deeper zones. Kendal (1976), Olson (1978), and Olson and Elders
(1978) showed that the degree of exchange of rocks from the Salton Sea sys-
tem reflected the primary permeability of the original lithology. Similarly,
Sheppard et al. (1969) showed that there was more extensive isotope ex-
change in the highly fractured ore depositing system at Butte, Montana than
in the Santa Rita p o r p h y r y copper deposit.
The lower limit of w/r is reached when all the free water is absorbed as hy-
11

drous minerals; this limiting value is a b o u t 0.04 for average conditions (Hat-
tori and Sakai, 1979). Hydration effects are not very significant at Wairakei
but may have affected the deuterium contents in the Broadlands system
which are slightly enriched relative to local cold surface waters.
A parameter related to rock/water ratio is the residence time of fluids in
the system which reflects the gross permeability and the relative contribu-
tions of long and short recharge paths to the system. The absence of tritium
(Wilson, 1963) suggested residence times for the dominant meteoric water of
> 1 0 0 years in Wairakei discharge. The absence of deuterium-enriched water
at Wairakei suggests an absence of interglacial waters perhaps indicative of a
residence time of < 1 2 , 0 0 0 years (Stewart, 1978) so that the slight enrich-
ment of Broadlands waters may reflect a contribution of older waters.
Hydrothermal alteration occurs through phase transformation, growth of
new minerals, mineral dissolution and precipitation and ion exchange reac-
tions. Although the original rock type has an influence on secondary min-
eralogy, its effect is less than those due to permeability, temperature and
fluid composition. Browne (1978) noted that at temperatures of 250--280 °
the c o m m o n assemblage quartz, albite, K-feldspar, chlorite, Fe-epidote, illite,
calcite, pyrite was found in basalts in Iceland, sandstones in the Imperial
Valley, rhyolites in New Zealand and andesites in Indonesia.
In fields dominated by fracture permeability there may be a zoning of
alteration ranging from high to low rank about a major flow zone (Steiner,
1977). Oxygen isotope ratios ( ' 8 0 / ' 6 0 ) in minerals within drill cores taken at
various depths in a field enable estimates to be made of the temperature pro-
file that existed within the country before drilling occurred. ~ 18 O measure-
ment in vein calcite has been the main technique used. Isotopic equilibrium
has been demonstrated between calcite and the water from which it is pre-
cipitated (Clayton et al., 1968; Elders et al., 1978), within fissure zones of
high permeability. Particularly effective applications of this m e t h o d were
made at Cerro Prieto by Elders et al. (1978) and at Kawerau by Blattner
(1979).
Providing that there is adequate permeability, at temperatures over about
100--150°C an exchange of oxygen isotopes occurs during the interaction
between water and rock minerals, particularly the fine-grained and layered
alumino-silicates, and calcite (Truesdell and Hulston, 1979). However, pri-
mary quartz is slow to exchange isotopes even at temperatures over 300 °
(Clayton et al., 1968). For many fine-grained minerals in permeable media,
isotopic equilibrium is approached at temperatures over 230--250°C, while
for secondary quartz and calcite deposited in fissures isotopic equilibrium
with solution is achieved (Blattner, 1975).

Alteration assemblages

A generalised summary of the temperature ranges over which various


alumino-silicate minerals have been observed in geothermal areas is given
12

I I I
Amorphous S!O,-
Quartz
K-Feldspar
Alblte
CalcIte w__
Montmorlllonlte
Mont - llllte
llllte -_--
Chlorite -__ -__-
D& ---

ElIotIts
Actlnolltf
Tremollte
Dlopslde
Garnet
Epldote
Prehnlte --_ - --
Heulandite -
Stllblte -

Ptllollte
Laumontlte -w-w
Wairoklte

I I I
1000 2000 300’ c
Fig. 3. Generalised summary of the temperature ranges over which alumino-silicate alter-
ation minerals have been observed. (Solid lines give an indication of the most commonly
observed temperature ranges). Temperatures of first appearance of marker minerals in
the Salton Sea, Cerro Prieto and Hawaiian areas are somewhat higher than in New Zea-
land. Chemical controls on mineral stabilities are discussed in the text. The three chlorite
stability ranges indicate the transition from swelling through mixed layer to non-swelling
chlorite with increasing temperature.

in Fig. 3. The diagram gives an indication of the minerals which may be


produced at different temperatures during hydrothermal alteration. The
actual presence or absence of a particular mineral depends on a number
of kinetic parameters as well as carbon dioxide and hydrogen sulphide
activities and solution pH.
Primary minerals also vary considerably in their susceptibility to alter-
ation. Volcanic glass is most reactive, frequently altering first to opal,
smectite, calcite or zeolite and then to mixed-layer clays. The presence
of such intermediate phases greatly influences local hydrogen ion and
silica activities and the succeeding reaction step (e.g., Keith et al., 1978). A
decreasing order of mineral susceptibility is approximately olivine, magne-
tite, hypersthene, horneblende, biotite = plagioclase (Steiner, 1968). Detrital
quartz is frequently inert to geothermal fluids at temperatures up to 300°C
but there is evidence from Cerro Prieto that recryslallisation occurs at higher
temperatures (Hoagland and Elders, 1978).
Steiner (1968) showed how successively higher temperatures affected the
formation of clay-mica group minerals following the alteration of silica-rich
13

volcanic glass by dilute geothermal waters. Alteration to calcium zeolites and


montmorillonite occurred at temperatures of 100--120°C. With increasing
temperature interstratified montmorillonite--illite occurred with the basal
spacings of the mineral becoming smaller until illite appeared at 230 °C. He
showed a trend in calcium zeolites from mordenite (ptilolite) at tempera-
tures of about 100°C, through laumonite from 150 ° to 200°C, to wairakite
at higher temperatures.
At temperatures above about 180°C, mineral assemblages in geothermal
systems are generally formed from solutions saturated with respect to
quartz, but at lower temperatures, particularly in glassy rocks, silica activi-
ties may be higher and approximate to the solubilities of chalcedony or
amorphous silica (Arnorsson, 1975). Consequently, low-temperature min-
eral assemblages, particularly zeolitic assemblages, are more variable (Browne,
1978), and favour the silica-rich zeolites such as mordenite.
The Reykjanes and Svartsengi, S.W. Iceland, geothermal areas provide
good examples of major readjustment of inflow water compositions by high-
temperature reaction with rock. These systems result from complex mixing
and evaporation sequences due to the penetration of meteoric water into hot
brine of seawater origin modified by fluid--basalt interactions. Sakai et al.
(1980) have discussed the disequilibrium distribution of 34S between sul-
phate and sulphide in these system waters, as well as in other Icelandic sys-
tems. With respect to seawater the Reykjanes fluid is depleted in magnesium
and sulphate ions and enriched in calcium, potassium, silica and minor ele-
ments including rubidium and arsenic. A montmorillonite--zeolite--calcite
alteration zone occurs to 230°C, followed by a mixed layer chlorite-mont-
morillonite and prehnite intermediate zone and a chlorite--epidote zone up
to about 300°C. An irregular distribution of anhydrite is probably due to
sulphate loss during successive incursions of seawater (Tomasson and Krist-
mannsdottir, 1972; Olafsson and Riley, 1978; Arnorsson, 1978a). Krist-
mannsdottir (1976) has carefully documented the clay mineral progression
from montmorillonites to non-swelling chlorites in this system.
Hydrothermal alteration of basaltic rocks by heated seawater also occurs
in oceanic spreading areas and largely has been described by reference to
dredged or deep drill samples or cross-reference to ophiolite sequences
(Miyashiro et al., 1971; Spooner and Fyfe, 1973; Liou and Ernst, 1979).
Jehl et al. (1977) confirmed that modified seawater was the active hydro-
thermal phase with up to five times seawater salinity and less than 1 mole %
CO2.
The brines of the Salton Sea geothermal system and other systems in the
Imperial Valley are derived from Colorado River water which has undergone
partial evaporation and exchange with sandstones and shales (Truesdell et al.,
1978). At Cerro Prieto, Mexico, diagenetic kaolinite--montmorillonite--do-
lomite--Fe-hydroxides--pyrite give way at 140 ° to 160°C to illite--chlorite
with overgrowths of quartz and feldspar or detrital clays, together with cal-
cite, pyrite and pyrrhotite. Above 250°C epidote appears with wairakite;
14

calcite being partially dissolved. Calcite is replaced entirely above 300 ° by


prehnite--actinolite and some diopside--hedenbergite, and illite--chlorite
and wairakite are almost completely lost with the incoming of biotite above
325°C (Hoagland and Elders, 1978). Kendall (1976) described a similar al-
teration sequence in the Magmamax wells in the Salton Sea and recognised
the occurrence of garnet, talc and tremolite at temperatures higher than the
biotite 'isograd'.
The general sequence of hydrothermal alteration with increasing depth in
a geothermal system relates through temperature to the boiling point with
depth relationship for the gas content and enthalpy of the deep single-phase
inflow to the system. In the two-phase region, loss of steam with CO2, H2 S
and other dissolved gases leads to increasing pH and the development of a
characteristic assemblage of calcite and alkali feldspar (Browne and Ellis,
1970), particularly in upflow zones. Elders et al. (1981) have shown how
integration of isotope and alteration studies may be used to develop a useful
flow model for the Cerro Prieto field.
Because of the appreciable concentrations of CO2 in geothermal fluids
and the low solubility of calcite at high temperatures, the m a x i m u m activity
of calcium in solution varies from one system to another. This control, to-
gether with the variation of silica activity in some lower-temperature waters,
accounts for the wide variety of calcium minerals being found in geothermal
systems. For example, wairakite and epidote are c o m m o n in the low carbon
dioxide Wairakei system but are rare in the high CO2 Broadlands system
where calcite is abundant. Similar difference exist in the Imperial Valley
with wairakite and prehnite present at Cerro Prieto but not in the Salton Sea
system (Elders et al., 1978).
Distinctive alteration assemblages occur in the near-surface steam-heated
zone of a geothermal system. Underground boiling, partition of dissolved
gases to the steam phase and possible oxidation leads to the formation
above the boiling zone of relatively acid condensate. Steiner (1977) des-
cribed the acid alteration (advanced argillic) assemblages at Wairakei, charac-
terised by kaolinite, alunite, gypsum, opal and hydrated iron oxides. In a
few areas, such as Matsukawa, Japan (Nakamura et al., 1970) and Matsao,
Taiwan (Chen, 1970), acid waters have been located at greater depths and
at temperatures of 250--280°C. These occurrences may be due to high-tem-
perature alteration of sulphur deposits or perhaps to an influx of acid gases
(SO2) at depth. Kaolinite, alunite, anhydrite, pyrophyllite, quartz and pyrite
occur in the central zone of these fields trending outward to less acid condi-
tions with the assemblages including montmorillonite, mica, quartz and an-
hydrite.
Temperature gradients at the margins and steam loss near surface suggest
that silica deposition should occur on the boundaries of geothermal systems.
Silicification is a c o m m o n feature of hydrothermal rock alteration and incor-
porates both the addition of silica (largely as a vein mineral by fluid flow)
and redistribution of the silica originally present as glass or cristobalite. Keith
15

et al., (1978) and White et al. (1975) described porosity reduction due to
silica precipitation in rocks penetrated by drilling in Yellowstone Park and
showed from pressure data between wells that there was effective self-sealing
of horizontal permeability. In contrast little physical evidence for self-sealing
has been obtained at Wairakei. Calcite and feldspar deposition in boiling
zones may also contribute to this process of self-sealing and Oki and Hirano
(1979) suggest that geothermal brines of the Hakone system are sealed
against seawater intrusion by precipitated sulphate and carbonate minerals.
However, sealing of vertical permeability is not well documented. Even
where rocks of relatively low permeability (e.g., mudstones) cap a system
chloride water and steam continue to penetrate to the surface through frac-
ture systems. The abundance of hydrothermal explosion features containing
high-chloride springs in the surface of geothermal systems (Lloyd, 1959,
1972; Muffler et al., 1971; Nairn and Wiradiradja, 1980) and hydrothermal
breccias (Grindley and Browne, 1975) suggests that systems actively main-
tain flow paths to the surface. Build-up of gas pressure as well as seismic dis-
turbance can initiate hydrothermal eruptions (Henley and Thomley, 1979).

Kinetics o f mineral alteration reactions


Few experimental studies have been concerned specifically with the kine-
tics and mechanism of mineral alteration reactions. Early theoretical studies
of the incongruent dissolution of feldspar (Helgeson, 1971) derived the con-
cept that in a closed system diffusion through a growing layer of new pre-
cipitate was the rate-determining step. Subsequent studies by Lagache (1976)
and Petrovic et al. (1976b) did not confirm this concept; dissolution rates
appeared to be controlled by the nucleation and growth at discrete sites of
intermediate phases in the sequence proposed by Helgeson (1971). Experi-
mental studies in the system CaO--MgO--SiO2--H20--HC1 (Sheppard, 1977)
support Petrovic's model. In a steady state open system where growth sites
for secondary minerals are developed, solution concentration and exposed
surface area are probably the major controls on mineral alteration rates.
Henley and Harper (1979) observed relative reaction rates during the
shallow reinjection of waste 95°C geothermal water at Broadlands. Among
the transient chemical effects observed during dispersive mixing of the inject-
ed and dilute formation water the adsorptions of caesium, rubidium and
potassium into clay minerals were most rapid while only partial re-equilibra-
tion occurred with respect to calcite and silica solubilities.

Composition of altered rocks


Changes in the major element composition of rocks during alteration re-
flect mineralogical changes, such as growth of K-feldspar and mica, or quartz
deposition. Minor elements may be taken into alteration mineral lattices, or
may form discrete accessory mineral phases such as oxides or sulphides dif-
ficult to detect in small amounts.
Comparing fresh and hydrothermally altered rhyolite rocks in the Waira-
1(3

kei system, Steiner (1977) showed a characteristic loss of sodium and cal-
cium and gain of potassium and silica through all alteration zones. In the
near-surface montmorillonite and intermediate micaceous clay-zeolite zones
the concentrations of Al, Fe, Mn, Mg, and P were little changed, but these
were depleted in the highest-rank alteration zones. The behaviour of Pb and
Ag was variable, probably due to their concentration into sulphides, which
occur in some of the alteration zones. Analysis of hydrothermal quartz
showed lead, arsenic and silver contents of 150, 300 and 50 ppm, respecti-
vely. Koga (1967) and Ellis and Sewell (1962) showed the depletion of zinc
and of boron in rocks during high-temperature alteration in the Wairakei
system.
Goguel (pers. comm., 1981) confirmed the leaching of calcium and so-
dium from hydrothermally altered rocks of the Wairakei and Broadlands
areas and showed the uptake of potassium, rubidium, caesium and lithium.
Analysis of individual minerals showed the transfer of rare alkalis from the
original ferromagnesian minerals and glass matrix of the volcanic rocks to
particular secondary minerals; for lithium mainly into chlorite (250 ppm Li)
and quartz (up to 400 ppm Li) (Barger et al. (1973) have located lepidolite
in shallow altered rocks in Yellowstone); for rubidium into illite and to a
lesser extent adularia; and for caesium into zeolites such as wairakite (250
ppm Cs) and to a lesser extent illite (45 ppm Cs). Lower temperatures corre-
lated with higher caesium concentrations in wairakite from a drill-core in the
E1 Tatio field, Chile (about 220°C). In non-zeolitic alteration zones caesium
concentrations were little different from those in the original rhyolitic rocks.
From the changes in water compositions in the fields from depth to the
surface and the integrated changes in rock composition during alteration,
Goguel deduced water/rock mass ratios of about 100, and for constant natu-
ral water outflow rates, field life-times of 6000000--106 years. This again
assumes that the rocks sampled by drilling were representative of overall
rock alteration in the field and that constant flow and chemical conditions
are maintained.
Naboko (1977) reviewed the changes in element compositions caused by
the hydrothermal alteration of dacite--andesite--basalt rocks with geother-
mal fields in Kamchatka. During hydrothermal alteration of a similar type
to the New Zealand fields, ranging from argillitic through zeolitic to quartz--
feldspar, the rocks were enriched in Rb, Sc, Ga, Ge, Sr, As, Sb, Ag, Au and
Sn. The elements Ti, Ba, Cu and B were depleted, while Li, Be, V, Cr, Co,
Ni, Zr, Mo and Pb showed variable behaviour but on average were little
changed. Zinc abundances showed a bimodal distribution suggesting leaching
at some levels and deposition at others. Lithium and rubidium concentra-
tions followed potassium, showing the general tendency to be deposited in
the rocks as the water temperatures decreased.
There are still insufficient data on the chemical composition of hydro-
thermally altered rocks to enable statistically significant trends to be shown
for m a n y elements. The statistical m e t h o d employed by Putnam (1975) for
17

mineralized granitic rocks may be appropriate.


Seawater composition may have been controlled through the earth's his-
tory by basalt--seawater interaction in ocean ridge geothermal systems (Wo-
lery and Sleep, 1976; Muehlenbachs and Clayton, 1976). A number of stu-
dies have attempted to quantify the extent and sequence of reactions oc-
curring in these systems and to examine alteration sequences preserved in
ophiolite terrain. Thompson and Humphris (1977) showed that altered ba-
salts from mid-ocean ridges had become relatively enriched in magnesium
and depleted in calcium, strontium and silica. B, Li, Cu, Zn, Mn and Fe
were also removed from the rock, together with some Co and Ni which were
subsequently fixed in sulphide segregations. Stakes (1978) examined the al-
teration chemistry and mineralogy of altered gabbros from the Mid-Atlantic
Ridge and East Pacific Rise as well as a metagabbro from Michigan. She
suggests that the absence of smectite from the Atlantic samples implied low
water--rock ratios in contrast to the occurrence of saponite in the East
Pacific Rise material which, by analogy with experimental studies, (Seyfried
and Mottl, 1977) was indicative of high wate--rock ratios. McGoldrick et al.
(1979) showed that the thallium content of deep submarine basalts was a
sensitive indicator of basalt--seawater interaction; as is the case for the
Broadlands system thallium behaves in a similar way to potassium but is
more highly enriched due to fixation in alteration silicates and iron-manga-
nese hydroxides. Albarede et al. (1981) have discussed the implications for
87 Sr/86 Sr of seawater--basalt interaction in spreading centre hydrothermal
systems.

Experimen tal studies


Several high-temperature experimental studies have demonstrated that
appreciable concentrations of solutes such as Li, Na, K, Rb, Mg, Ca, F, C1,
B, SO4 and NH3 can be brought readily into solution from fresh rocks (Ellis
and Mahon, 1964, 1967; Mahon, 1967; Kissen and Pakhomov, 1967; Ellis,
1968). Although the solute concentrations of dilute geothermal waters could
be explicable by a simple solution mechanism, the high salinities of some
waters cannot be accounted for by simple extraction from rocks in the likely
range of water--rock ratios operative. These experiments also confirmed that
the concentrations of silica, potassium, calcium, magnesium, iron, alumi-
nium, manganese, fluoride, carbonate, sulphate and many heavy metals were
controlled by pressure-temperature dependent equilibria between minerals
and solutions. Ewers (1977) observed the ready dissolution of As, Sb, S and
Se from greywackes between 100°C and 500°C. Relatively few elements (C1,
Br, B, Cs, Li, As and NH3 ) are conserved in solution in deep geothermal solu-
tions. Ratios between these components appear to reflect particular rock-
types in the geothermal system.
In experiments with salt solutions, heavy metals such as iron, manganese,
lead, zinc and copper are rapidly 'leached' from rocks (Ellis, 1968). With sea-
water reacting with basalt, Mg is first lost from solution into smectite-type
18

phyllosilicates, producing an acidic environment which leads to extraction


from the rock of K, Ca, Mn, Fe, Cu and Zn (Bischoff and Dickson, 1975;
Seyfried and Bischoff, 1977; Bischoff et al., 1981). Calcium sulphate is pre-
cipitated from solution. With low water--rock ratios {<10), after initial re-
moval of Mg from solution, the pH slowly rises in response to hydrogen ion
reaction with alumino-silicates. At high water--rock ratios (~50) there is less
Mg depletion of the fluid but the pH remains low and strong leaching of
the rock occurs with respect to Ca, K and Ba (Seyfried and Bischoff, 1979)
and heavy-metal concentrations in solution are higher. Seawater mixed with
'altered' seawater responds by the deposition of a complex magnesium sili-
cate. Mottl et al. (1979) showed that sulphide was readily leached from ba-
salt as well as being formed from reduction of seawater sulphate due to fer-
rous iron release from the basalt. They also showed that Fe and Mn concen-
trations increased with increasing reaction temperature. Menzies and Sey-
fried (1979) showed that rock depletion in K and Rb was a function of
temperature and water--rock ratio, although some altered ocean floor basalts
show no such depletion, perhaps due to retrograde reactions.
Dickson (1977) demonstrated that interaction of rhyolite with saline solu-
tions proceeded similarly but more rapidly. In these reactions anomalously
high silica concentrations occurred initially (Dickson et al., 1978a) which
subsequently decreased as zeolite nucleation and growth occurred (Rytuba
et al., 1978; Potter and Dickson, 1980). Similar behaviour was noted for
andesite as the solid reactant (Liou and Dickson, 1978).
GEOTHERMAL FLUID COMPOSITIONS
Mineral-fluid equilibria control the major element concentrations and pH
of deep geothermal fluids. The development of readily applicable thermody-
namic techniques for high temperature multi-component equilibria (Helge-
son, 1969, 1978a) has led to a greater realisation of the large-scale rock-
water equilibria controlling fluid compositions and alteration assemblages in
geothermal and fossil hydrothermal systems (Arnorsson et al., 1978; Bird
and Norton, 1981; Giggenbach, 1981).
C o m m o n solutes
The total salt concentrations of geothermal fluids range from about 0.05%
to 30%, but most commonly fall in the range 0.1--1%. Table I contains five
relatively dilute waters from systems containing volcanic rocks (Hveragerdi,
basalt; Otake, Tongonan and Ahuachapan, andesite; Wairakei and Broad-
lands, rhyolite). The Ngawha system contains siltstones and sandstones un-
derlain by argillites and greywackes, and the Cerro Prieto and Salton Sea sys-
tems comprise deltaic river sediments. The Reykjanes and Hakone systems
are recharged by seawater--meteoric water mixtures through basaltic rocks
and the Matsao system contains highly acidic chloride~sulphate waters in an
andesite--sandstone sequence. The alkali sulphate--chloride water at Cesano,
Italy was derived at greater than 250°C from brecciated limestones overlain
by potassic volcanics.
19

Features that are c o m m o n to most high-temperature waters include un-


usually high proportions of the rare alkalis, lithium, rubidium and caesium as
well as of silica, boron, arsenic, fluoride and ammonia. Except for acidic
fluids, iron, magnesium, heavy metals and sulphate--chloride ratios are
usually low.
Chloride ion concentration, along with temperature and local rock type,
is one of the principal independent variables defining the chemistry of geo-
thermal fluids (Ellis, 1970; Arnorsson et al., 1978). As discussed above, its
possible origins are diverse and most likely, in any system, it results from
mixing of two or more source fluids (White, 1974).
Since the ratios of the principal anions (K ÷ and Na ÷) to H ÷ are determined
by mineral alteration reactions such as:
K-feldspar + H ÷ = muscovite (illite) + K ÷ + SiO2
Na-feldspar + H ÷ = Na-montmorillonite + Na ÷ + SiO2
the pH of the geothermal system fluid is a function of the salinity of the sys-
tem. For example, Salton Sea waters are much more acidic than those from
Wairakei, and are noted for their corrosion potential and transport of heavy
metals. At lower temperatures the pH may be controlled by other alumino-
silicate mineral interactions. Arnorsson et al. (1978) suggested that up to
170°C the pH of some Icelandic waters was controlled by chlorite--smectite
--water reaction.
The pH of geothermal waters collected at atmospheric pressure is usually
higher than that of the original deep water due to the preferential loss of
acidic gases (CO2, H2S) into the steam phase. For waters of near neutral
pH, the change is frequently about 1--2 pH units. Arnorsson (1978b) gave
a detailed discussion of this process.
The Na/K ratio of geothermal waters has been used successfully as a geo-
thermometer in many geothermal systems, particularly for waters over about
180°C. Truesdell (1975) and Fourier (1979a) summarised empirical equa-
tions for the variation of Na/K with respect to temperature. The quotient
relates to the overall reaction and exchange of Na and K between albite and
K-feldspar, and decrease of Na/K with temperature reflects the net transfer
of potassium from deeper h o t t e r levels of a geothermal system to shallower
levels as may be recognized from whole rock analyses. Near surface extrac-
tion of potassium into clay minerals may affect the Na/K ratio of some
spring waters (Weissberg and Wilson, 1977).
Fournier and Truesdell (1973) have shown that Ca/Na and Na/K ratios
reflecting a wider range of mineral-solution equilibria, may be combined into
a useful Na K Ca geothermometer which they calibrated using waters from a
number of geothermal fields. For waters high in magnesium, Fournier and
Potter (1979) propose corrections which should be applied to the simple
NaCaK ratio temperature estimate. Paces (1975) showed that calculated tem-
peratures for some spring waters were higher than measured temperatures
and suggested a correction based on the carbon dioxide content of the water.
Nathenson (1981) has shown that for low flow springs from reservoirs
20

co t--
o

0 0

¢~ I o ~ o ~ I

¢,1 ¢~1

t- ¢~ ~ ~ ~. •

¢z

tO ,-~ t-

~.

CO

O9 ¢xl ~1 ¢~

. v l" ¢~
o
21

0~. O0

O'J ~J

O0

o0 ~

kO
• ~o ~
Cq

eo C~
v

~D

. ~ o~ ~o ~ ~ ~ ~ oo oo

,~ CD

~ ~ ~ ^ ~ ^ ~ ^ ~
o

~o o~
22

<I150°C apparent differences between the silica and alkali geothermometers


may be reconciled by allowing for conductive heat loss in the spring con-
duits.
The rare alkalis (lithium, rubidium and caesium) are present in unusually
high concentrations in high-temperature waters but this depends on local
rock types. The basaltic reservoir rocks of Iceland are associated with low
concentrations of rare alkalis in geothermal water, while the alkali--potassic
rocks of the Cesano area contain exceptional concentrations. Fouillac and
Michard (1979, 1981) suggested a general relationship between Na/Li ratio
and temperature, ranging from 100--200 at 50°C to 20--70 at 300°C. Dilute
waters and brines had a separate trend line.
The Na/Rb ratio in waters is of potential value as a geothermometer, par-
ticularly in revealing high temperatures at depth through the analysis of
spring waters, as low-temperature hydrothermal minerals incorporate potas-
sium in preference to rubidium (Goguel, pers. comm., 1981).
Calcium concentrations are related to the carbon dioxide content of the
system and the pH defined by the dominant alumino-silicate or solution buf-
fer system (Ellis, 1970). The controlling equilibrium is the solubility of cal-
cite, most waters being close to saturation, so that the maximum calcium
concentration is given by the following general equation:
aca2 + = K a~a÷ mco2-1
where K is a temperature-dependent constant. Calcite solubility increases
with falling temperature, so that highest calcium concentrations occur in
waters of low temperature, high salinity and low carbon dioxide concentra-
tions. Giggenbach (1981) showed that in the silicic volcanic terrane of the
Taupo Zone systems, the dominant reaction determining fluid chemistry
was the conversion of primary plagioclase by CO2 to calcite and clay min-
erals, while a set of subsequent secondary mineral equilibria control relative
CO2 -H2 S concentrations.
Ellis (1971) showed that in general aMg2+/a~+ ion ratios in geothermal
waters reflected the conditions for coexistence of chlorite, calcite and
calcium--magnesium montmorillonite.
Fluoride and sulphur are readily removed from most rock types by high
temperature water (Ellis and Mahon, 1964, 1967). The maximum concen-
trations of sulphate and fluoride are related to calcium concentrations
through the solubility equilibria for anhydrite and fluorite (Mahon, 1964;
Nordstrom and Jenne, 1977; Richardson and Holland, 1979). Although an-
hydrite is a c o m m o n alteration mineral, fluorite is not. Reviewing fluoride
contents of Icelandic waters, Arnorsson et al. (1978) suggested that fluoride
concentrations were limited by anion exchange with alumino-silicates con-
taining exchangeable hydroxyl groups.
Anhydrite has a low solubility in h o t water (24 ppm at 250°C, 8 ppm at
300°C) so that sulphate concentrations are low in high-temperature, near-
neutral pH waters. The Cesano brine, a sodium sulphate solution, occurs in a
unique chemical and redox situation. Acidic waters such as at Matsao and
23

near-surface steam-heated waters often contain much higher sulphate con-


contrations.
Boron and arsenic are also readily released from many rock types during
reaction with h o t water. The C1/B ratio in geothermal waters c o m m o n l y is
similar to that in the aquifer rocks (Ellis and Sewell, 1963; Mahon, 1967).
Waters with exceptional boron levels, such as Ngawha, are associated with
organic-rich sedimentary rocks.
Arsenic behaves as a soluble c o m p o n e n t in deep high-temperature fluids,
but is depleted in near-surface waters, probably due to destabilisation of its
sulphide solution complex, when H2S is removed by boiling (Ewers and
Keays, 1977; N a b o k o and Karpov, 1977).
The solubility of the various forms of silica increases rapidly with tem-
perature and it is observed that geothermal waters are saturated with silica at
the level for quartz solubility at temperatures above about 180°C but with
various amorphous or unstable crystalline forms of silica at lower tempera-
tures (Mahon, 1966; Fournier and Rowe, 1966; Arnorsson, 1975). Silica
concentrations in waters provide an excellent geothermometer, particularly
for geothermal wells where fluids reach the surface quickly and with mini-
m u m contact with rock surfaces at lower temperatures. Waters reaching the
surface through springs, even of high flow, may lose silica through partial re-
equilibration. Mixing of waters from deep levels with near-surface waters
occurs in most areas, and Fournier and Truesdell (1974), Truesdell and Four-
nier (1976) and Fournier (1979b) in particular have shown how 'mixing
models' for geothermal areas may be constructed to allow for this type of
effect.
R e d o x conditions
Seward (1974) reviewed the calculation of redox conditions in geothermal
fields, including estimation from pH and hydrogen concentrations (either
measured or calculated from ratios such as CH4/CO2, NH3/N2 ) and from
the coexistence of redox responsive minerals such as pyrite and pyrrhotite.
The hydrogen partial pressure in geothermal waters is usually in the range
0.02--0.1 bars (Ellis and Mahon, 1977), reflecting the conditions imposed by
the dominant mineral buffer and alteration controlled pH. In most systems
this is probably the equilibrium between pyrite and Fe-Mg/chlorite (Giggen-
bach, 1980). The redox potential (fo2) also controls the concentration of
sulphate (as free SO4e - or as ion pairs with sodium or potassium) and bisul-
phate with respect to hydrogen sulphide and bisulphide ion. At shallow
levels in geothermal systems there are marked changes in the redox environ-
ment due to mineral-fluid re-equilibration following steam loss or mixing
with near-surface air-saturated groundwaters or steam-heated waters. Naboko
and Karpov (1977) gave examples of changing Eh and pH and mineralogy
with depth in shallow wells and Pampura et al. (1975) discussed Eh-pH con-
ditions in Kamchatka well waters during their cooling by steam loss from
200°C.
24

Minor metals

Table II gives a selection of trace metal concentrations in a range of geo-


thermal fluids. In dilute, near-neutral pH geothermal waters typical of recent
volcanic areas, the concentrations of most heavy metals are low, even for the
c o m m o n rock-forming elements such as manganese and iron, and usually
less than the average for fresh surface water. Exceptions are those which
form complex species with sulphide in high temperature solutions, such as
gold, mercury, arsenic and antimony (Weissberg, 1969). Tungsten concentra-
tion may also be high locally. By contrast, in highly saline geothermal fluids
heavy metal concentrations may be high to the extent that economic recov-
ery is feasible. The Salton Sea brines discharged from exploration wells are
particularly notable b o t h for their high salinity and the abundance of
sulphide scales formed in discharge pipelines (Skinner et al., 1967).
McKibben (1979) has described the occurrence of hematite-dominated and
sulphide-dominated mineral assemblages, as well as synsedimentary pyrite/
marcasite mineralisation in the Salton Sea cores. Recrystallised early pyrite
was replaced by pyrrhotite, chalcopyrite, sphalerite assemblages at greater

TABLE II
Minor element concentrations in some natural and geothermal waters

Source and temperature pH C1 Mg Mn Fe


(o C) (cold) (ppm) (ppm) (ppm) (ppm)

Fresh water 7.8 4.1 0.3--300


Average seawater 19000 1350 0.002 0.01
Hot water, Reyholt,
Iceland (98 ° ) 9.1 90 0.03 0.006
Well 2, Broadlands, N.Z.
(260 ° ) 8.3 1180 0.01 0.009 0.23
Average Wairakei wells
(250 ° ) 8.2 1500 0.01 0.001 0.008
Aquifer, Cerro Prieto
wells, Mexicali (340 ° ) 7420--11750 6--33 0.64 0.2
Well H8, Reykjanes,
Iceland (275 ° ) 7.1 17912 7.0 1.9
Mean, 1 and 2 I.I.D. Wells,
Salton Sea (300--350 °) 4.7 155000 10--54 1400 2150
Well E-205, Matsao,
Taiwan (245 ° ) 2.9 9000 88 28 148
Tamagawa Springs,
Japan (98 ° ) 1.2 3240 83 4.2 105

(a) Hawkes and Webb (1962).


(b) Krauskopf (1967).
(d) Arnorsson (1970).
(d) D.S.I.R. Chemistry Division, New Zealand.
25

depth. Mineralisation is present b o t h as disseminated sulphides and as veins


with a mineral abundance pa t t e r n (in decreasing order); pyrite, hematite,
sphalerite, chalcopyrite, p y r r h o t i t e , marcasite, galena.
At least in some fields the heavy metals are derived by leaching of c o u n t r y
rock (Doe et al., 1966). Leeman et al. (1979) showed from Pb isotope com-
positions th at lead in h o t spring deposits in Yellowstone Park arose in part
from a sedimentary source. Weissberg et al. (1979} reviewed occurrences of
ore metals in h o t spring precipitates and well discharges from a n u m b e r of
geothermal systems. Vidal et al. (1978) recently described the submarine
occurrence o f f Baja California, or pyrite- and gypsum-bearing precipitates
enriched with As, Hg, Sb and T1.

Gas concentrations

The total gas c o n t e n t in the steam phase of geothermal well discharges


ranges from ab o ut 0.01% to several tens of percent. The gases are princi-
pally CO2 and H~S, but include also CH4, N:, H2, NH3, and traces of the
rare gases He, Ne, Ar, Kr and Xe. In liquid-dominated systems at deep

Ni Cu Pb Zn Sb W Ag Ref.
(ppb) (ppb) (ppb) (ppb) (ppb) (ppb) (ppb)

0.02--10 0.2--30 0.3--3 1--200 0.07--7 (a)


2 3 0.03 10 0.5 0.04 (b)

<1 <2 <2 (c)


0.1 0.6 0.8 0.6 130 87 0.5 (d)

0.7 1.3 3 1.5 70 (d)

2.2 5 4.6 6 400 4 (e)

8.6 10.4 66 0.4 0.7 (f)

5500 91000 520000 400 1400 (g)

35 5OO 88OO (d)


10 1000 2800 (h)

(e) Mercado (1967).


(f) Olafsson and Riley (1978).
(g) White (1968).
(h) Ozawa et al. (1973).
26

levels the gases are dissolved in the h o t water, b u t when steam is formed
by boiling the gases pass into the steam phase.
Giggenbach (1980, 1981) showed that gases in the system C--O--H--N
occur in equilibrium proportions in geothermal systems, while H2S con-
centrations are related to H2 and are controlled by the c o m m o n buffer
system pyrite and iron-bearing chlorite. The proportions of gases in geo-
thermal discharges provide an excellent means of deep temperature esti-
mation (D'Amore and Panichi, 1980; Giggenbach, 1980, 1981).
Carbon dioxide in geothermal systems may originate from decarbona-
tion reactions from carbonate rocks or from dispersed organic matter,
from solution in meteoric water or from magmatic sources. Panichi and
Tongiorgi (1975) and Truesdell and Hulston (1979) summarised carbon
isotope data for various geothermal areas and showed that the 13Cco 2
range overlapped with, b u t was slightly lighter than that for carbon in
marine limestones. This feature could be explained by mixing with lighter
organic carbon from local sediments. The carbon isotope range from --2
to --8°/°° 13C is also close to that assumed for juvenile carbon, derived
from magmatic systems originating in the mantle. Barnes et al. (1978b)
showed that the global distribution of carbon dioxide discharges, includ-
ing geothermal systems and spring waters, coincided with the major
belts of crustal seismicity and concluded that CO2 production was lar-
gely from regional or localised contact metamorphism of marine carbo-
nate sediments.
The methane associated with high temperature geothermal areas has a 13 C
compositional range much lighter than that of CO2 and reflects a partial ori-
gin from organic carbon. The slow equilibration of isotopes between CO2,
CH4, H2 and H20 (Sackett and Chung, 1979) allows CO2 and CH4 to con-
tinue to reflect their separate origins (Panichi and Gonfiantini, 1979). This
makes CH4--CO2 carbon isotope exchange geothermometry of dubious va-
lidity.
Giggenbach (1977) showed that sulphur in geothermal gases, and fixed as
sulphides within the systems, probably had a dual origin involving magmatic
assimilation of crustal material containing sulphides and the subsequent
evolution of a magmatic gas. By correlating the C/S ratio of volcanic and
geothermal gases, he suggests that only about 5% of the sulphur entering the
system with carbon dioxide may reach the exploitable zone of the system,
the remainder being fixed as sulphides in the deeper hotter zones. Arnorsson
(1974) correlated the sulphide concentrations and temperatures of Icelandic
geothermal systems and concluded that the sulphur was of magmatic origin.
Nitrogen may originate from the atmosphere dissolved in the meteoric-
water recharging geothermal systems as well as from magmatic sources, while
ammonia can be liberated by rock/water reactions at high temperatures. Loss
of ammonia into separated steam occurs as geothermal fluid moves toward
the surface, and it may, like rubidium, be absorbed into zeolites and clays,
occasionally forming distinct phases such as buddingtonite (Erd et al., 1964).
27

Rare gases
The ratios of the inert rare gases in m a n y geothermal waters below the
depth of boiling, suggest that they have an atmospheric origin and accom-
pany the meteoric-water supply (Mazor and Wasserburg, 1965; Mazor, 1975).
However, an excess of radiogenic argon and helium occurs in many geo-
thermal fluids, arising from rock leaching or from magmatic-fluid contribu-
tions.
Helium concentrations in geothermal gases can exceed those in air (5.2
ppm) by up to two orders of magnitude. The isotope ratio 3He/4He is about 3 •
10 -S for helium derived from the mantle, 3 , 10 -s for the earth's crust, and
1.4 • 10 -6 for the atmosphere. The light isotope 3 He is found in excess of
atmospheric proportions in gases in seawater above spreading ocean ridge
centres (Jenkins et al., 1978, 1980; L u p t o n et al., 1977, 1979) and from oce-
anic basalt glasses, while helium in volcanic and geothermal gases of Kam-
chatka, Kurile Islands, Japan, Iceland, Hawaii, California and New Zealand
also have a high 3 He/4 He ratio of about {0.5--2) • 10 -s . This may arise from
the leaching of recent basaltic rocks or by diffusion of helium from the
mantle. There appear to be only minor contributions from crustal or atmos-
pheric gases in these systems. The highest 3 He/4 He ratio of about 2.1 • 10 -s
was found in Kilauea volcanic gases (Gutsalo, 1975, 1976; Craig et al., 1978;
Thomas and Kroopnick, 1978; Naughton and Thomas, 1978; Nagao et al.,
1979; W.F. Giggenbach, pets. comm., 1981).
The helium isotopic ratio has the possibility of distinguishing the heat ori-
gins in geothermal systems, with the lowest ratio corresponding to crustal
conductive heating situations where radiogenic 4 He has accumulated, and
the highest ratios to areas heated by basaltic volcanism.
The ratio 40Ar/36 Ar in the atmosphere is 295.6, but higher ratios can
be expected in gases from deep levels in the earth's crust because of the
production of 4°Ar from the radioactive decay of 4°K. Several geothermal
areas, (e.g., Wairakei, Waiotapu and Kawerau in New Zealand, Monte Amiata
(Italy), and Yellowstone Park) have an 4°Ar/36Ar ratio in gases which differs
little from that in the atmosphere, indicating a meteoric origin (Ellis and
Mahon, 1977). On the other hand, in some areas of Kamchatka and of the
Kurile Islands, the gases have excess 4°Ar present in a proportion which in-
creases with the ratio 3He/4He, suggesting that radiogenic argon may accom-
pany 3He from a mantle origin (Gutsalo, 1976).
In crustal rock situations where little excess 3He is present, the ratio of
radiogenic to atmospheric argon in geothermal fluids may give information
on the fluid residence time in the system and the permeability to local
ground water inflow. For example, in Larderello steam, excess 4°Ar occur-
red, but the percentage excess decreased with time as the steam field was
exploited, suggesting a shallow recharge of meteoric waters (e.g., Wairakei
10--100, Broadlands 1--19, The Geysers about 20, Larderello 40--300 nCi/
kg). Radon originates from the radioactive decay of uranium-238 through
the intermediates thorium-230 and radium-226, and its release into geother-
28

mal fluids depends on the content of parent radioactive elements in the rock,
the rock mineralogy and the system-permeability, and whether liquid or gas
fills the rock pores. Radon may therefore be useful as a pathfinder in geo-
chemical exploration for geothermal energy (Wollenberg, 1975; Nielson,
1978; Cox, 1980). The short half-life of radon (3.83 days) means that mea-
surements of its concentration can give information on the immediate his-
tory of fluid before it is discharged from wells or natural vents, and reasons
for changes with time in well discharges. (D'Amore, 1975; Kruger et al.,
1976; D'Amore et al., 1978; Kruger, 1978; Whitehead, 1979).
The combination of radon measurements with determinations of other
rare gas isotopes, deuterium, 180, and tritium has a potential for investi-
gating the details of geothermal reservoir fluid dynamics.

FOSSIL HYDROTHERMALSYSTEMS

It has become clear from a host of recent geochemical studies that the ma-
jority of hydrothermal ore deposits hosted by volcanic rocks or their subja-
cent plutonic suites were formed within geothermal systems of similar size,
chemistry and behaviour to those we see active today. Indeed no other con-
clusion might be expected, should uniformitarianism be accepted as the
guide to geological interpretation. By the same token the study of such
fossil hydrothermal systems, which have been deeply dissected by erosion,
offers information on the deep characteristics of active geothermal systems
which is otherwise quite unattainable.
Here we attempt to highlight aspects of those types of deposit and their
environments of formation which correlate most strongly with the active
geothermal systems, rather than to provide a comprehensive review of the
full range of characteristics demonstrated by any particular family of hydro-
thermal ore deposits. Porphyry-type Cu--Mo, 'epithermal' type Au--Ag--Hg
and Cyprus-type massive sulphide deposits show the strongest correlations
and White (1967, 1974, 1981) has discussed the origin of these types of de-
posit in the context of active geothermal systems emphasizing active mer-
cury depositing systems such as Sulfur Bank, California and Ngawha, New
Zealand. The reader is referred to White's papers for full discussion of the
geothermal environment of mercury ore formation. The Hg deposits and
lithium-enriched tuffs of the McDermitt caldera are a particularly inter-
esting example of a fossil geothermal system (Rytuba, 1976). Other genetic
groups of hydrothermal deposits (e.g., Mississippi Valley type) may simi-
larly find their modern analogues but these must be sought in the geother-
mal environments of tectonic settings other than those of the volcanic
hosted systems (White, 1968).
Epithermal type gold--silver deposits occur in many Tertiary volcanic
regions, in association with base metal sulphides (e.g. Tayoltita and
Guanajuato, Mexico), with antimony, arsenic and mercury sulphides
(Carlin, Nevada) and as teUuride assemblages (Acupan, Philippines,
29

Vatukoula, Fiji, Thames, N.Z.). Many of these deposits are associated with
active warm or h o t springs, e.g., Acupan, Philippines (Sawkins et al., 1979).
It was early recognised (e.g., Lindgren, 1933; White, 1955) that clear paral-
lels exist between the near surface ( ~ 5 0 0 m) depositional environment of
these deposits and that of modern hot spring systems and these have been
emphasized by the results of recent exploration activity through the Western
U.S.A. and Canada (e.g., McLaughlin deposit, California).
Weissberg (1969) and Weissberg et al. (1979) reviewed the occurrence of
ore metals in hot spring environments such as the Champagne Pool, Waiota-
pu, New Zealand and the Ohaki Pool, Broadlands, New Zealand where gold
and silver reach 85 and 500 ppm, respectively, in the amorphous Au--Ag--Sb
--As--S precipitates. In both these geothermal systems galena, sphalerite and
minor chalcopyrite occur with pyrite and pyrrhotite at depths of 200--2000
m (Browne, 1969; J.W. Hedenquist, pers. comm., 1982) in the temperature
range 120 ° to 298°C. (Browne, 1969, also recorded minor telluride in these
assemblages.) These disseminated base metal sulphides occur in veins and
vugs often associated with adularia and calcite, both minerals indicative of
local boiling. This base metal sulphide deposition appears to be interpretable
as due to the combined effects of pH increase and H2 S loss due to boiling.
Ewers and Keays (1977) studied the zoning of metals within the Broadlands
field using whole rock samples, sulphide concentrates and silica sinters.
These data showed a crude metalliferous zoning with Sb, Au and T1 enriched
near the surface and Pb, Zn, Ag, Cu, Bi, Co, Te and Se at depth. Gold in the
dilute geothermal fluids of this field is transported as thiocomplexes and
the deep system fluids are greatly undersaturated with respect to the metal
(Seward, 1973). Its deposition in hot springs and sub-hot spring environ-
ments reflects loss of H2 S due to boiling or mixing with relatively oxidising
near surface waters (see below) and T.M. Seward (pers. comm., 1982) has
shown that colloidal As--Sb sulphides act as effective scavengers of gold and
other metals and may be kinetically instrumental in producing the ore grade
precipitates.
Large low-grade disseminated gold deposits of the Carlin-type occur with-
in the altered and silicified argillaceous, carbonaceous limestone of the
Roberts Mountain Formation, Nevada. Pyrite is abundant and the gold
occurs in stratiform bodies as part of an arsenic, antimony, thallium, mer-
cury assemblage, consisting of sulphates and sulfosalts and abundant organic
matter. The gold occurs as submicroscopic particles, mostly in pyrite and
arsenopyrite, and ranges in grade from 15 g tonne -1 in the Cortez deposit
to 29 g tonne -1 in the Carlin deposit (Wells and Mullens, 1973). Carlin
itself has produced over 108 g of gold. Dickson et al. (1978b, 1979) showed
that the ore fluids were dominantly meteoric in origin and that temperatures
for the gold--pyrite phase of deposition were in the range 160--180°C, while
later vein temperatures ranged from 235 ° to 300°C. Dickson et al. (1979)
and Radtke et al. (1980) consider that these deposits were formed as part of
the evolution of a geothermal system in which heat was derived from Ter-
30

tiary sub-volcanic intrusives, and ore and gangue components from deep
rock--water interactions at temperatures in excess of 350°C. They suggested
that ore formation was a response to a rapid change of temperature, perhaps
due to a sudden release of pressure or increase in vertical permeability.
Casadevall and O h m o t o (1977) discussed ore-forming conditions in the
Sunnyside Mine, Eureka District, Colorado where gold--silver--copper--lead
- z i n c - - c a d m i u m veins occur to depths of about 600 m. These formed
13--16.6 million years ago during resurgent volcanism within the 28 m.y.
old San Juan--Uncompahgue caldera complex. Stable isotope data showed
that the ore-forming fluid was dominantly meteoric in origin and they con-
cluded that the main stages of base metal ore deposition occurred at 250--
320°C, at fluid pressures of 110--220 bars. Carbon dioxide, ranged from
0.01 to 0.2 molal (0.04--0.9 wt.%), the higher values being dominant. R e d o x
(Po2 = 10-3° -- 10-32 bars) and pH (4.3--5.9) conditions were estimated
from mineral assemblages and thermochemical data.
In the majority of geothermal systems deep system waters are buffered
with respect to fo2 by pyrite-reduced iron assemblages (such as pyrite--
pyrrhotite, pyrite--chlorite; see p. 23) at a value far below that inferred for
ore deposition at Sunnyside. Here, as well as for a number of other epither-
mal deposits (e.g., Tonopah, Nevada, Pueblo Viejo, Dominican Republic)
(Kesler et al., 1981) the relatively high oxygen fugacity of the ore deposition
environment (where HSO4 > H2 S) and low pH with respect to deep system
waters is most likely the result of mixing with steam-heated acid sulphate
waters generated above the boiling zone of the system. Such interactions are
c o m m o n in volcanic geothermal systems; e.g., at Matsao, Taiwan acid sul-
phate waters occur with alunite in sandstone at depths of over 500 m, above
a normal chloride water zone at temperatures of around 250°C (Chen, 1975;
Lan et al., 1980). At Matsukawa, Japan, acid sulphate waters with low chlo-
ride content are associated with kaolinite--alunite--pyrophyllite to depths
of 700 m (Nakamura et al., 1970).
In Fig. 4 the solubility of gold as thiocomplexes has been recalculated as a
function of pH, fo2 and ES from the data of Seward (1973). These data
show that the gold content of deep system waters at Sunnyside could n o t
have been more than 50 ppm and by analogy with Broadlands may have
been considerably undersaturated. Up to 300°C the solubility of gold as
chloride complexes is imprecisely known but based on the data of Henley
(1973) is unlikely to exceed 5 ppb under the depositional conditions at
Sunnyside and thiocomplexing dominates the transport of gold in the deep
system (extrapolation of lower temperature data suggests far lower solubil-
ities in the region 5 • 10 -6 ppm). Mixing of relatively oxidised acid sulphate
waters with deep system H2S dominant waters would result in an inter-
mediate redox environment of the type shown b y the Sunnyside data. Under
these conditions the solubility of gold is a factor of at least 10 ~ less than in
the deep system waters. The precipitous drop in solubility close to the
HSO4 : H2S species boundary may lead to rapid bonanza-style deposition of
31

-26- uC[

o -32-

KAOLINITE MUSCOVITE K- FELDSPAR


I I
3 L. 5 6 7
pH
Fig. 4. Gold solubility contours for AuCI~ and Au(HS)~ in ppm as a function of oxygen
fugacity and pH at 300°C in relation to the stability fields of pyrite, pyrrhotite, magne-
tite, hematite, kaolinite, muscovite,
+
K-feldspar. The additional contribution of Au(HS) °
species has been omitted (mk = 0.02; m~l = 0.08; mH2 s = 0.05; data sources: Helgeson,
1969; Seward, 1973; Henley, 1973). The stippled region is the depositional environment
deduced for the Sunnyside mine by Casadevall and O h m o t o ( 1 9 7 7 ) a n d cross-hatched
area the chemical environment of a typical geothermal system in which fo2 is buffered
close to the pyrite--pyrrhotite phase boundar:~: evidence from Broadlands suggests that
such fluids are generally well undersaturated with respect to gold. The low pH and rela-
tively high fo2 for Sunnyside suggest that mixing of acid sulphate type steam-heated wa-
ters and deep near-neutral chloride waters may have led to ore deposition; gas loss due
to boiling results in temperature decrease and relative increase of pH and fo2, and may
contribute to ore deposition. (Use of Helgeson's data for the AuCl~ reaction in preference
to the experimental data gives solubilities 100 X less than those shown in the HSO~ field,
inferring an even greater solubility hiatus in the vicinity of the H2S--HSO~ boundary.)

some 90% of the gold in the mixing zone. As well as increasing pH, loss of
H~S by boiling of deep system fluid also affects a reduction in gold solubility
but for waters initially well below saturation the removal of 50% of the H2S
by boiling from 325 ° to 300°C may only render the solution just supersatur-
ated with effective deposition of no more than a few percent of the trans-
ported gold.
Barton et al. (1977), Bethke and Rye (1979) and Wetlaufer et al. (1979)
have described the ore forming environment of the silver--PbS--ZnS de-
:posits at Creede, Colorado in relation to present day geothermal systems.
Redox conditions in the vein assemblage are similar to Sunnyside, again sug-
L edvanced -
ergiU,c silica .~
o(ferofion sinter

, / F ~ ~ - ~ 7 - - ' ~,Au-Ag "-


Pb-Z0

KM 1

2 J

EPITHERMAL

J J;" . ~
~
--
-Ag
Pb -Zn

/ o o/ o
o o -" ~ .

/' i o
o
o o'-q o
o o°
o °
~"-
"%
2000

\" ......... j~r"I'1 'lii


f"
,/
, i i
PORPHYRY
KM %%g
omo Cu-Mo
0

//
B
33

gesting t h a t m i x i n g as well as boiling m a y h a v e p l a y e d a significant role in


ore d e p o s i t i o n in this n e a r - s u r f a c e e n v i r o n m e n t . Base m e t a l sulphides (gale-
na, s p h a l e r i t e ...) o c c u r in a n d b e l o w A u - - A g ore z o n e s at T a y o l t i t a a n d
G u a n a j u a t o in M e x i c o . A t T a y o l t i t a o v e r 283 t o n n e s o f gold a n d 1 4 0 0 0
t o n n e s o f silver h a v e b e e n p r o d u c e d . D e p o s i t i o n o c c u r r e d o v e r a 300 m
vertical interval at an i n f e r r e d d e p t h o f 5 0 0 - - 6 0 0 m , t e m p e r a t u r e s w e r e
in t h e range 2 6 0 - - 2 9 0 ° C a n d salinities w e r e up to 8.4 wt.% NaC1 equiv.
(Clarke a n d A l b i n s o n , pers. c o m m . , 1 9 8 0 ) . A t G u a n a j u a t o d e p o s i t i o n t e m -
p e r a t u r e s w e r e a r o u n d 2 3 0 ° C , fluid salinities were o f the o r d e r 3 wt.%
NaC1 equiv, a n d t h e f o r m a t i o n d e p t h was a b o u t 3 4 0 m ( B u c h a n a n , 1 9 8 0 ) . In
b o t h d e p o s i t s d e p o s i t i o n o c c u r r e d f r o m boiling fluids, t h a t at G u a n a j u a t o
p e r h a p s b e i n g higher in dissolved CO2 leading t o calcite d o m i n a n c e over epi-
d o t e in the ore a s s e m b l a g e ( B r o w n e a n d Ellis, 1 9 7 0 ) . T h e d e p o s i t i o n a l envi-
r o n m e n t s o f t h e s e d e p o s i t s are clearly a n a l o g o u s t o t h o s e at B r o a d l a n d s a n d
W a i o t a p u w i t h lead a n d zinc and p e r h a p s s o m e gold d e p o s i t e d d i r e c t l y d u e
t o boiling. Apparent salinities of t h e B r o a d l a n d s fluids d e t e r m i n e d b y fluid
inclusion m e a s u r e m e n t s are higher t h a n t h o s e m e a s u r e d d i r e c t l y f o r reser-
voir fluids t h r o u g h t h e p r e s e n c e o f u p t o 3 wt.% CO2 in s o l u t i o n ( B r o w n e et
al., 1974). T h e p r e s e n c e o f dissolved gas m a y at least in p a r t a c c o u n t s f o r the
apparent salinity d i f f e r e n c e b e t w e e n o r e - f o r m i n g fluids and t h o s e a c t u a l l y
p r e s e n t in t h e m a j o r i t y o f p r e s e n t - d a y g e o t h e r m a l systems.
G e n e t i c studies o n these e p i t h e r m a l t y p e d e p o s i t s f r e q u e n t l y a d h e r e to
t h e gross c o n v e c t i v e c h a r a c t e r i s t i c o f t h e Wairakei g e o t h e r m a l s y s t e m w h e n
closer a t t e n t i o n t o t h e n e a r - s u r f a c e t o p o g r a p h i c a l l y c o n t r o l l e d h o t w a t e r
h y d r o l o g y a n d its i n t e r a c t i o n w i t h t h e d e e p e r s y s t e m fluids m a y be m o r e
a p p r o p r i a t e as an e x p l o r a t i o n guide. As in the p r e s e n t - d a y s y s t e m s , frac-
t u r e s or h y d r o t h e r m a l e x p l o s i o n vents m a y f o c u s d e e p s y s t e m fluids i n t o
t h e surface d e p o s i t i o n a l e n v i r o n m e n t s a n d act as e f f e c t i v e fluid m i x i n g
e n v i r o n m e n t s as is t h e case f o r t h e C h a m p a g n e Pool, W a i o t a p u . Fig. 5A
m i m i c s t h e s c h e m a t i c f l o w m o d e l o f a t y p i c a l g e o t h e r m a l s y s t e m (Fig. 1) and
a t t e m p t s t o e m p h a s i z e t h e d e p o s i t i o n a l e n v i r o n m e n t s o f e p i t h e r m a l ores in
r e l a t i o n t o t h e n e a r - s u r f a c e h y d r o l o g y o f t h e s y s t e m . E v i d e n c e f o r direct
i n v o l v e m e n t o f m a g n e t i c fluids in e p i t h e r m a l ore genesis t h r o u g h i n t r o d u c -
t i o n o f m e t a l s i n t o t h e u n d e r l y i n g g e o t h e r m a l s y s t e m s is scant; stable i s o t o p e

Fig. 5. Environments of hydrothermal ore deposition in typical geothermal systems asso-


ciated with: A. a silicic volcanic terrane; and B. a calc--alkaline stratovolcano. The sys-
tems' structure and symbols for fluid compositions are as for Fig. 1. Advanced argillic
alteration corresponds to the zone of acid sulphate (steam-heated) waters and argillic
through propylitic alteration to the chloride water zone. Higher alteration grades would
occur in the deeper zones, closer to any 'magmatic' fluid input and may be associated
with skarn type mineralisation. (The limits of the porphyry ore zone in B are schematic
only; the existence of equivalent mineralisation in the subsilicic terrane is a matter of spe-
culation.) Epithermal ores may be associated with typical near surface features such as
silicification and silica sinters, and telescoping of metal zonation may occur. Ore deposits
occur in close association with major flow controls such as deep fractures ('veins') or
hydrothermal explosion vents.
34

data from Creede and from Comstock (O'Neil and Silberman, 1974) merely
hint at this possibility and it is most likely that any deep system input is
highly dispersed by the time the near surface environment is r e a c h e d - as is
the case for active geothermal system fluids. Leaching of part or all the ore
metals from country rock in the system is an equally unproven alternative.
'Porphyry copper' deposits, ranging from 50 to 500 million tonnes of
ore at over 0.4% copper, are abundant in the Late Cretaceous--Tertiary vol-
canic belts of the Circum-Pacific and Alpine zones. In the southwest United
States, typified by the Bingham deposit, Utah, these large disseminated sul-
phide deposits are associated with felsic rocks (quartz monzonites and por-
phyries), b u t in British Columbia and the Southwest Pacific with diorites
and andesite. Sillitoe (1973) recognized that these deposits were formed at
depth in subvolcanic hydrothermal systems, which are perhaps most simi-
lar to those currently active in the Central and South American volcanic
regions and the Philippines. Wood (1980) has described the c o m m o n occur-
rence of chalcopyrite, with galena and sphalerite at temperatures of 300--
320°C in the Mahiao--Tongonan geothermal field, Philippines. The South-
ern Negros geothermal system is close to an actively explored porphyry
copper deposit, while porphyry prospects and deposits are found for ex-
ample in Chile, Panama, Mexico and Fiji near to active geothermal systems.
Particularly detailed studies of the E1 Salvador, Chile and Butte, Montana
porphyry copper deposits are by Gustafson and H u n t (1975) and Brimhall
(1979), respectively.
Lowell and Guilbert (1970) and Guilbert and Lowell (1974) have stressed
the c o m m o n large-scale pattern of alteration developed in p o r p h y r y systems;
a potassic assemblage coincides with low-grade ore in the core of the deposit
which is mantled by a higher-grade ore shell associated with a pyrite--phyllic
alteration assemblage and this, in turn, by a less altered argillic--propylitic
zone containing veins of gold, silver, copper, zinc and lead. Roedder (1971)
and others have shown that fluid inclusions from the periphery of the de-
posits contain fluids of less than 6% salinity, while in the core high-salinity
fluids ( ~ 5 0 wt.%) coexist with a lower salinity vapour phase in a boiling rela-
tionship. Homogenisation temperatures range from about 200 ° to 350°C in
the periphery and from 500 ° to 700°C in the core. Sheppard et al. (1969,
1971) reported a zoning of stable isotopes in alteration minerals which
they interpret as due to the presence of meteoric groundwaters in the pe-
ripheral zone and mixed meteoric--magmatic waters in the core. Estimates
of depths of formation range down to about 5 km for Bingham. At E1
Salvador and elsewhere, late-stage advanced argillic alteration is recognised,
representing acid-type near-surface water interaction.
White et al. {1971) attempted to model the ore-forming environment
by analogy with that of vapour-dominated hydrothermal systems, but
Sillitoe (1973) could n o t reconcile their model with the occurrence of
highly saline solutions in porphyry copper inclusions or with the absence
of the typical p o r p h y r y alteration in drilled vapour-dominated systems.
35

Henley and McNabb (1978) suggested that p o r p h y r y ore deposits are


formed where a rising plume of high temperature magmatic fluid pene-
trates and mixes with low salinity groundwater or deep formation fluids.
Consequent composition, redox and temperature changes lead to the con-
centric alteration sequence and ore deposition. They suggested that, at the
pressures and temperatures deduced for the system, magmatic fluid in the
system NaC1--H20-+ CO2 (Sourirajan and Kennedy, 1962) must be evolved
from the magmatic system as a low-salinity vapour which would condense
to a high-salinity liquid coexisting with vapour as pressure decreased during
upward b u o y a n t flow. Dilution with low-salinity groundwaters would occur
during subsequent convective ascent through the crust.
Henley and McNabb's model may offer a glimpse of the deep roots of
active geothermal systems, as shown in Fig. 1. Geothermal systems in the
Taupo Volcanic Zone occur in deep pyroclastic sequences overlying a
greywacke basement within which active magmatic systems may be oper-
ating and evolving heat both by conductive and convective heat loss. The
hidden presence of such a deep magmatic system could account for an
additional supply of chloride and sulphur. An interesting facet of these
fossil geothermal systems is the preservation of a record of early magmatic
fluid injection into the system followed by waning stage redistribution of
ore components and superimposition of lower temperature mineral assem-
blages. This pattern may characterise present-day geothermal systems but
may not be demonstratable at the shallow levels reached by drilling. Giggen-
bach (1977) has suggested that more than 95% of the sulphur introduced at
deep levels in a geothermal system may be fixed at depth as metal sulphide.
Fig. 5B shows the geothermal environment inferred for p o r p h y r y ore for-
mation. It is of interest to consider whether epithermal and p o r p h y r y envi-
ronments are mutually exclusive or gradational. The Comstock Lode, Neva-
da occurs within the eroded remnant of a typical andesite stratovolcano and
is regionally associated with porphyry ore environments. The epithermal
ores of silicic volcanic terrane appear to show no such proximity to por-
p h y r y ores b u t this may relate to erosion level rather than to any funda-
mental difference in the nature of the geothermal system, its fluid phase
or host rocks. Uplift and ½-1 km of erosion from a stratovolcano may
readily intersect both the p o r p h y r y and epithermal environment, while the
equivalent erosion or dissection of silicic terrane remains essentially in the
epithermal zone.
Submarine massive sulphide deposits contain a metal suite similar to
that of p o r p h y r y deposits and occur in a similar volcanic environment
(Hutchinson and Hodder, 1972). The deposits are best known in the Kuroko
districts of Japan (Lambert and Sato, 1974; Ishihara, 1974) and the Ar-
chaean greenschist belts of Canada (Sangster, 1972). They form as massive
fine-grained sulphide bodies on the contemporary sea-floor atop or adjacent
to mineralised stockwork zones in altered felsic volcanics (Lambert and Sato,
1974). Solomon (1976), Andrews and Fyfe (1976) and MacGeehan (1978)
36

recognised that the environment of ore deposition was that of submarine


discharge from sub-sea floor geothermal systems. Recent isotope studies con-
firm that the ore-forming fluids originate in this way although Ohmoto and
Rye (1974) and Hattori and Sakai (1979) have shown that some meteoric
water may be incorporated in the ore fluids in some deposits suggesting
that close analogy may exist with the active Hakone geothermal system,
Japan. Henley and Thornley (1979, 1981) suggest that sub-sea floor geother-
mal systems in these environments were supplied by sulphur and ore metals
by both the leaching demonstrated by experimental studies and magmatic
vapour by analogy with the porphyry ore model. Ore deposition may have
been consequent on hydrothermal eruptions through the sea-floor and subse-
quent higher power discharge through the eruption vents until resealing oc-
curred by mineral deposition. Examples of this sequence are at Buchans,
Newfoundland and Kosaka, Japan as well as a number of deposits in the
Noranda district. The possibility of magmatic fluid involvement is also sug-
gested by recent isotope data from the Kosaka and Buchans deposits
(Hattori and Muehlenbachs, 1980; Kowalik et al., 1981). Correlatives of the
Kuroko environment have yet to be located; a task of some difficulty in view
of the active volcanism and rapid sedimentation of their island arc submarine
setting. The nearest equivalents are the metalliferous sediments off Santorini
and Vulcano and those in Matupi Harbour, New Britain (for review see
Cronan, 1980).
The sub-sea floor geothermal systems of the Red Sea (Degens and Ross,
1969) have attracted much attention as analogues of sea floor mineralisa-
tion, correlating perhaps with Precambrian stratiform base metal sulphide
deposits, such as Mt. Isa, Queensland. Bignell et al. (1976) have carefully
documented the chemistry and mineralogy of the metalliferous muds which
occur within sea floor depresssions in contact with hot brine. Although the
water itself may have a complex origin, it is generally accepted that the high
salinity is due to dissolution of Miocene evaporites. Mixing of brine and
Red Sea waters appears to account for the range of salinities occurring in
the Atlantis II and Discovery Deeps (Danielsson et al., 1980). Shanks and
Bischoff (1977) proposed that the metals (dominantly Cu, Zn, Pb) were
derived by brine--basalt interaction at 150--250°C and that metal pre-
cipitation occurs largely in response to dissociation of chloride complexes
during cooling; sulphur being provided by inorganic reduction of sulphate.
Cyprus-type massive sulphide deposits are characterised by a simpler
sulphide assemblage (chalcopyrite, pyrite and some sphalerite) and an
association with pillow basalts in the Tethyan Troodos ophiolite sequence
in Cyprus and Palaeozoic Appalachian ophiolites of Newfoundland (Upad-
hayay and Strong, 1973). The sulphide ores are overlain by an amorphous
deposit (umber) consisting of Fe, Mn oxides with silica and alumina, while
the deposits themselves undergo submarine weathering to a Mn-poor, Fe-
rich sediment, or 'ochre' (Constantinou, 1976). Spooner and Fyfe (1973),
Spooner et al. (1974, 1977) and Chapman and Spooner (1977) showed that
37

the deposits formed above large-scale hydrothermal systems which operated


in the oceanic lithosphere. At the same time deep drilling and dredging in
the areas of ocean floor spreading of the East Pacific and North Atlantic
identified the presence of such systems operating in present-day crust. These
systems are characterised by metalliferous sediments over a wide area, thick-
ened manganese encrustations (Scott et al., 1976; Natland and Rosendahl,
1979; Corliss et al., 1978; Grill et al., 1981) and abundant organic growth
close to geothermal discharge points. Natland et al. (1979} suggest that the
Galapagos sub-sea floor system is at least 30,000 years old and that water--
rock ratios were initially low, leading to manganese poor sedimentation but
subsequently increased to the stage of localised formation of the hydrother-
mal mounds of manganese-rich material.
The close similarity between the tectonic settings of exposed fossil ore
depositing systems of Cyprus and Newfoundland and of the East Pacific, and
North Atlantic spreading centres may have led to the anticipation of the dis-
covery of actively forming Cyprus-type ore bodies and one of the most ex-
citing recent developments in ore deposition geochemistry has been the direct
observation of active sulphide-depositing vents (black 'smokers') by deep
diving on the East Pacific Rise (lat. 21°N) (Franchetau et al., 1979; Hekinian
et al., 1980). Water temperatures in the vents exceed 350°C. In the Galapa-
gos Rift, vents of similar chemistry but lower discharge temperature (~ 10°C)
occur, perhaps indicative of greater dilution and sub-sea floor ore deposition
(Corliss, 1979). Studies to date have confirmed the strong analogies between
these active ore-depositing systems and the Cyprus-type ore deposits while
highlighting certain important distinctions, such as the size and rate of de-
struction by oxidation, which may infer that the Cyprus ores owe their pre-
servation to their scale and this in turn to relatively faster spreading rates
during their formation (East Pacific Rise Study Group, 1981). Discussion
also focusses on whether the ore metals are leached by convecting seawater
or whether this medium actually intersects sulphide segregations in shallow
magma chambers.

ACKNOWLEDGEMENTS

In this review we have a t t e m p t e d to compile useful source references, as


well as to cover and integrate a wide range of chemical, geological and geo-
physical topics. In covering the vast a m o u n t of literature currently available
some personal bias is inevitable. Our thanks go, however, to all concerned
with the complex task of unravelling the geochemistry and behaviour of
geothermal systems, ancient and modern.

REFERENCES

Albarede, F., Miehard, A., Minster, J.F. and Michard, G., 1981.87Sr/S6Sr ratios in hydro-
thermal waters and deposits from the East Pacific Rise at 21°N. Earth Planet. Sci.
Lett.. 55: 229--236.
38

Allis, R.G., 1979a. Changes in heat flow associated with exploitation of the Wairakei
field. Proc. N.Z. Geothermal Workshop, 1979. University of Auckland, pp. 11--13.
Allis, R.G., 1979b. Thermal history of the Karapiti Area, N.Z.D.S.I.R., Geophys. Div.
Rep., 137, 38 pp.
Allis, R.G., 1981a. Continental underthrusting beneath the Southern Alps of New Zea-
land. Geology, 9: 303--307.
Allis, R.G., 1981b. Comparison of mechanisms proposed for induced seismicity at the
Geysers geothermal field, California. Proc. N.Z. Geothermal Workshop, Univ. Auck-
land, pp. 57---62.
Allis, R.G., Henley, R.W. and Carman, A.F., 1979. The thermal regime beneath the
Southern Alps. In: R. Walcott and M. Cresswell, Origin of the Southern Alps. R. Soc.
N.Z., Bull., 18: 79--85.
Andrews, A.J. and Fyfe, W.S., 1976. Metamorphism and massive sulphide generation
in oceanic crust. Geosci. Can., 3: 84--94.
Arnorsson, S., 1970. Geochemical studies of thermal waters in the southern lowlands of
Iceland. Geothermics (Spec. Iss. 2), Pt. 2: 547--552.
Arnorsson, S., 1974. The composition of thermal fluids in Iceland and geological features
related to the thermal activity. In: Geodynamics of Iceland and the North Atlantic
Area. D. Reidel Publ. Co., Dordrecht, pp. 307--323.
Arnorsson, S., 1975. Application of the silica geothermometer in low temperature hydro-
thermal areas of Iceland. Am. J. Sci., 275: 763--784.
Arnorsson, S., 1978a. Major element chemistry of the geothermal sea-water at Reykjanes
and Svartsengi, Iceland. Miner. Mag., 42: 209--220.
Arnorsson, S., 1978b. Precipitation of calcite from flashed geothermal waters in Iceland.
Contrib. Miner. Petrol., 66: 21--28.
Arnorsson, S., Grhnvold, K. and Sigurdsson, S., 1978. Aquifer chemistry of four high-
temperature geothermal systems of Iceland. Geochim. Cosmochim. Acta, 42: 523--
536.
Bargar, K.E., Beeson, M.H., Fournier, R.O. and Muffler, L.J.P., 1973. Present-day deposi-
tion of lepidolite from thermal waters in Yellowstone National Park. Am. Miner., 58:
901--904.
Barnes, Ivan, Downes, C.J. and Hulston, J.R., 1978a. Warm springs, South Island, New
Zealand and their potential to yield laumontite. Am. J. Sci., 278: 1412--1427.
Barnes, Ivan, Irwin, W.P. and White, D.E., 1978b. Global distribution of CO2 discharges
and major zones of seismicity. U.S. Geol. Surv., Water-Resour. Invest. Open-File Rep.,
pp. 78--139.
Barnett, P.R., 1979. Geochemistry of the Mahiao sector of the Tongonan geothermal
field. N.Z. Geothermal Workshop, 1979. University of Auckland, Part 2.
Barry, B.J., McCabe, W.J. and Manning, M.R., 1979. Radioisotope tracing of water
movement at Walrakei. Proc. New Zealand Geothermal Workshop, Univ. of Auckland,
pp. 102--103.
Barton, P.B., Jr., Bethke, P.M. and Roedder, E., 1977. Environment of ore deposition in
the Creede Mining District, San Juan Mountains, Colorado, Part III. Progress towards
interpretation of the ore forming fluid for the OH vein. Econ. Geol., 72: 1--24.
Batzle, M.L. and Simmons, G., 1976. Microfractures in rocks from two geothermal areas.
Earth Planet. Sci. Lett., 30: 71--93.
Bethke, P.M. and Rye, R.O., 1979. Environment of ore deposition in the Creede Mining
District, San Juan Mountains, Colorado, Part IV. Source of fluids from oxygen, hydro-
gen and carbon isotope studies. Econ. Geol., 74: 1832--1851.
Bignell, R.D., Cronan, D.S. and Tooms, J.S., 1976. Red Sea metalliferous precipitates.
Geol. Assoc. Can. Spec. Pap., 14: 147--179.
Bird, D.K. and Norton, D.L., 1981. Theoretical prediction of phase relations among
aqueous solutions and minerals: Salton Sea geothermal system. Geochim. Cosmo-
chim. Acta, 45: 1479--1493.
39

Bischoff, J.L. and Dickson, F.W., 1975. Seawater--basalt interactions at 200°C and 500
bars. Earth Planet. Sci. Lett., 25: 385--397.
Bischoff, J.L., Radtke, A.S. and Rosenbauer, R.J., 1981. Hydrothermal alteration of
greywacke by brine and seawater: Rholes of alteration and chloride complexing on
metal solubilization at 200 ° and 350°C. Econ. Geol., 76: 659--676.
Blattner, P., 1975. Oxygen isotopic composition of fissure-grown quartz, adularia
and calcite from Broadlands geothermal field, New Zealand. Am. J. Sci., 275:
785--800.
Blattner, P., 1979. Oxygen isotope composition of downhole calcite crystals, Kawerau
geothermal field. Open-file Report PB-101, N.Z. Geological Survey, Lower Hutt,
9 pp.
Brimhall, G.H., 1979. Lithological determination of mass transfer mechanisms of multiple
stage porphyry copper mineralization at Butte, Montana - - Vein formation by hypo-
gene leaching and enrichment of potassium silicate protore. Econ. Geol., 74: 556--
589.
Browne, P.R.L., 1969. Sulfide mineralization in a Broadlands geothermal drillhole, Taupo
Volcanic Zone, New Zealand, Econ. Geol., 64: 156--159.
Browne, P.R.L., 1978. Hydrothermal alteration in active geothermal fields. Ann. Rev.
Earth Planet. Sci., 6: 229--250.
Browne, P.R.L., 1979. Minimum age of the Kawerau Geothermal Field, North Island,
New Zealand. J. Volcanol. Geothermal Res., 6: 213--215.
Browne, P.R.L. and Ellis, A.J., 1970. The Ohaki-Broadlands hydrothermal area, New
Zealand: Mineralogy and related geochemistry. Am. J. Sci., 269: 97--131.
Browne, P.R.L., Roedder, E. and Wodzicki, A., 1974. Comparison of past and present
geothermal waters from a study of fluid inclusions, Broadlands, New Zealand. Int.
Water-Rock Interaction Symp., Prague, pp. 140--149.
Buchanan, L.J., 1980. Ore controls of vertically stacked deposits, Guanajuato, Mexico.
Soc. Min. Eng., AIME. In press.
Calamai, A., Cataldi, R., Dall'Aglio, M. and Ferrara, G.C., 1975. Preliminary report on the
Cesano hot brine deposit (Northern Latium, Italy). 2nd U.N. Symp. Devel. Use Geo-
thermal Resour., San Francisco, Abstract, II-6.
Casadevall, T. and Ohmoto, H., 1977. Sunnyside Mine, Eureka Mining District, San Juan
Country, Colorado: Geochemistry of gold and base metal ore deposition in a volcanic
environment. Econ. Geol., 72: 1285--1320.
Cathles, L.M., 1977. An analysis of the cooling of intrusives by groundwater convection
which includes boiling. Econ. Geol., 72: 804--826.
Chapman, M.H. and Spooner, E.T.C., 1977. 878r enrichment of ophiolitic sulphide de-
posits in Cyprus confirms ore formation by circulating seawater. Earth Planet. Sci.
Lett., 35: 79.
Chen, C.H., 1970. Geology and geothermal power potential of the Tatun volcanic region.
Geothermics (Spec. Iss., 2), (Pt. 2), pp. 1134--1143.
Chen, C.H., 1975. Thermal waters in Taiwan, a preliminary study. Proc. Int. Assoc.
Hydrol. Sci., Grenoble, Publ., 199: 79--88.
Choussy, M.E. and Penate, T.S., 1976. Campo geotermico de Ahuachapan despues de un
ano de explotacion. Simposio Int. sobre la Energia Geotermica en America Latina,
E1 Salvador, Oct. 1976.
Clayton, R.N., Muffler, L.J.P. and White, D.E., 1968. Oxygen isotope study of calcite
and silicates of the River Ranch No. 1 Well, Salton Sea geothermal field, California.
Am. J. Sci., 266: 968--979.
Constantinou, G., 1976. Genesis of the conglomerate structure, porosity and collomorphic
textures of the massive sulphide ores of Cyprus. Metallogeny and plate tectonics. In:
D.F. Strong (Editor), Geol. Assoc. Can., Spec. Pap., 14: 187--210.
Cortiss, J.B., 1979. Metallogenesis at oceanic spreading centres. J. Geol. Soc. Lond., 136:
621--626.
40

Corliss, J.M., Lyle, M., Dymond, J. and Crane, K., 1978. The chemistry of hydrothermal
mounds near the Galapagos Rift. Earth Planet. Sci. Lett., 40: 12--24.
Cox, M.E., 1980. Ground Radon Survey in Hawaii. Geophys. Res. Lett., 7: 283--286.
Craig, H., Lupton, J.E. and Horibe, Y., 1978. A mantle helium component in circum-
Pacific volcanic gases. Adv. Earth Planet. Sci., 3: 3--16.
Criss, R.E., 1981. An 180/160, D[H and K/Ar study of the southern half of the Idaho
batholith. PhD thesis, California Inst. of Technology, 368 pp.
Cronan, D.S., 1980. Underwater Minerals. Academic Press, London, 362 pp.
D'Amore, F., 1975. Radon survey in Larderello geothermal field Italy. Geothermics, 4:
96--108.
D'Amore, F., Sabroux, J.C. and Zettwoog, P., 1978. Determination of characteristics of
steam reservoirs by radon-222 measurements in geothermal fluids. Pure Appl. Geo-
phys., 117: 253--261.
D'Amore, F. and Truesdell, A.H., 1979. A model for steam chemistry at Larderello and
The Geysers. H.J. Ramey and P. Kruger (Editors), Proc. Fifth Workshop Geothermal
Reservoir Eng. Stanford Univ. SGP-TR40, 283--298.
D'Amore, F. and Panichi, C., 1980. Evaluation of deep temperatures of hydrothermal
systems by a new gas geothermometer. Geochim. Cosmochim. Acta, 44: 549--556.
Danielsson, L., Dyrssen, D. and Gran~li, A., 1980. Chemical investigations of Atlantis
II and Discovery Brines in the Red Sea. Geochim. Cosmochim. Acta, 44: 2051--
2065. ~
Degens, E.T. and Ross, D.A. (Editors), 1969. Hot Brines and Recent Heavy Metal De-
posits in the Red Sea. Springer, New York, 600 pp.
Dickson, F.W., 1977. Role of rhyolite--seawater reaction in the genesis of Kuroko ore
deposits. Proc. 2nd Water--Rock Interaction Conf., Strasbourg, 1977.
Dickson, F.W., Potter, J., Rytuba, J. and Radtke, A.S., 1978a. Reaction of rhyolite with
H 2 O and NaC1-H 2 O at 300°C and 500 bars (abs). Am. Geophys. Union, San Francisco,
1978.
Dickson, F.W., Rye, R.O. and Radtke, A.S., 1978b. The Carlin gold deposit as a product
of rock-water interactions. Proc. Symp. Int. Assoc. Genesis Ore Deposits (IAGOD),
5th, Snowbird, Utah, pp. 101--108.
Dickson, F.W., Radtke, A.S. and Peterson, N., 1979. Ellisite, T13AsS3, a new mineral
from the Carlin gold deposit, Nevada and assemblages of associated sulfides and sul-
fosalts. Amer. Miner., 64: 701--707.
Doe, B.R., Hedge, C.E. and White, D.E., 1966. Preliminary investigation of the source
of lead and strontium in deep geothermal brines. Econ. Geol., 61: 462--483.
Donaldson, I.G., 1962. Temperature gradients in the upper levels of the earth's crust
due to convective water flows. J. Geophys. Res., 67 : 3499--3559.
East Pacific Rise Study Group, 1981. Crustal processes of the Mid-Ocean Ridge. Science,
213: 31--40.
Elder, J.W., 1966. Heat and mass transfer in the earth, hydrothermal systems. N.Z.D.S.
I.R. Bull., 1 6 9 , 1 1 5 pp.
Elder, J.W., 1981. Geothermal Systems. Acad. Press, New York, N.Y., 632 pp.
Elders, W.A., Hoagland, J.R. and Olson, E.R., 1978. Hydrothermal mineralogy and iso-
topic geochemistry in Cerro Prieto geothermal field, III. Practical applications. Geo-
therm. Resour. Counc. Trans., 2: 177--179.
Elders, W.A., Williams, A.E. and Hoagland, J.R., 1981. An integrated model for the natu-
ral flow regime in the Cerro Prieto hydrothermal system, B.C., Mexico based upon
petrological and isotope chemical criteria. Third Cerro Prieto Field Symposium, San
Fransico. Geothermics. In press.
Ellis, A.J., 1968. Natural hydrothermal systems and experimental hot-water/rock interac-
tions: Reactions with NaC1 solutions and trace metal extraction. Geochim. Cosmo-
chim. Acta, 32: 1356--1363.
Ellis, A.J., 1970. Quantitative interpretation of chemical characteristics of hydrothermal
systems. Geothermics (Spec. Iss. 2), 2 (Pt. 1), pp. 516--528.
41

Ellis, A.J., 1971. Magnesium ion concentrations in the presence of magnesium chlorite,
calcite, carbon dioxide, quartz. Am. J. Sci., 271: 481--489.
Ellis, A.J. and Mahon, W.A.J., 1964. Natural hydrothermal systems and experimental
h ot-water/roc k interactions. Geochim. Cosmochim. Acta, 28: 1323--1357.
Ellis, A.J. and Mahon, W.A.J., 1967. Natural hydrothermal systems and experimental hot-
water/rock interactions (Part II). Geochim. Cosmochim. Acta, 31: 519--539.
Ellis, A.J. and Mahon, W.A.J., 1977. Chemistry and Geothermal Systems. Academic
Press, New York, N.Y., 392 pp.
Ellis, A.J. and Sewell, J.R., 1963. Boron in rocks and waters of New Zealand hydrother-
mal areas. N.Z.J. Sci., 6: 589--606.
Erd, R.C., White, D.E., Fahey, J.J. and Lee, D.E., 1964. Buddingtonite, an ammonium
feldspar with zeolitic water. Am. Miner., 49: 831--850.
Ewers, G.R., 1977. Experimental hot-water/rock interactions and their significance to
natural hydrothermal systems of New Zealand. Geochim. Cosmochim. Acta, 41:
143--150.
Ewers, G.R. and Keays, R.R., 1977. Volatile and precious metal zoning in the Broad-
lands geothermal field, New Zealand. Econ. Geol., 72: 1337--1354.
Fan, P.F., 1978. Mineral assemblages of hydrothermal alteration of basalts from Hawaii.
Geotherm. Resour. Counc. Trans., 2 : 185--187.
Fehn, U., Cathles, L.M. and Holland, H.D., 1978. Hydrothermal convection and uranium
deposits in abnormally radioactive plutons. Econ. Geol., 73: 1556--1566.
Fouillac, C. and Michard, G., 1979. Un geothermometre empirique: le rapport Na/Li des
eaux. CR. Acad. Sci. Paris, 288: 123--126.
Fouillac, C. and Michard, G., 1981. Sodium/lithium ratio in water applied to geothermo-
metry of geothermal reservoirs. Geothermics, 10: 55--70.
Fournier, R.O., 1979a. A revised equation for the Na/K geothermometer. Geotherm.
Resour. Counc. Trans., 3: 221--224.
Fournier, R.O., 1979b. Geochemical and hydrologic considerations and the use of enthal-
py-chloride diagrams in the prediction of underground conditions in hot spring sys-
tems. J. Volcanol. Geotherm. Res., 5: 1--16.
Fournier, R.O. and Potter, R.W., 1979. A magnesium correction for the Na--K--Ca geo-
thermometer. Geochim. Cosmochim. Acta, 43: 1543--1550.
Fournier, R.O. and Rowe, J.J., 1966. Estimation of underground temperatures from
the silica content of water from hot springs and wet-steam wells. Am. J. Sci., 264:
685--697.
Fournier, R.O. and Truesclell, A.H., 1973. An empirical Na--K--Ca geothermometer for
natural waters. Geochim. Cosmochim. Acta, 3 7 : 1 2 5 5 - - 1 2 7 5 .
Fournier, R.O. and Truesdell, A.H., 1974. Geochemical indicators of subsurface tempera-
ture, 2. Estimation of temperature and fraction of hot water mixed with cold water.
U.S. Geol. Surv. J. Res., 2: 263--270.
Fournier, R.O., Sorey, M.L., Mariner, R.H. and Truesdell, A.H., 1979. Chemical and iso-
topic prediction of aquifer temperature in the geothermal system at Long Valley,
California. J. Volcanol. Geotherm. Res., 5: 17--34.
Franchetuau, J., Needham, H.D., Choukroune, P., Juteau, T., Sequet, M., Ballard, R.D.,
Fox, P.J., Normark, W., Carranza, A., Cordoba, D., Guerrero, J., Rangin, C., Bougault,
It., Cambon, P. and Hekinian, R., 1979. Massive deep-sea sulphide ore deposits dis-
covered on the East Pacific Rise. Nature, 277 : 523--528.
Giggenbach, W.F., 1977. The isotopic composition of sulphur in sedimentary rocks bor-
dering the Taupo Volcanic Zone. In: A.J. Ellis (Editor), Geochemistry 1977. N.Z.
D.S.I.R. Bull., 218: 57--64.
Giggenbach, W.F., 1978. The isotopic composition of waters from the E1 Tatio geother-
mal field, Northern Chile. Geochim. Cosmochim. Acta, 42: 979--988.
Giggenbach, W.F., 1980. Geothermal gas equilibria. Geochim. Cosmochim. Acta, 44:
2021--2032.
42

Giggenbach, W.F., 1981. Geothermal mineral equilibria. Geochim. Cosmochim. Acta, 45:
393--410.
Giggenbach, W.F. and Lyon, G.L., 1978. The chemical and isotopic composition of wa-
ter and discharges of the Ngawha geothermal field, Northland. N.Z.D.S.I.R., Chemis-
try Div., Open-File Rep.
Green, D.C., Hulston, J.R. and Crick, I.H., 1978. Stable isotope and chemical studies of
volcanic exhalations and thermal waters, Rabaul Caldera, New Britain, Papua, New
Guinea. B.M.R.J. Aust. Geol. Geophys. 3: 233--239.
Green, K.E., Van Herzen, R.P. and Williams, D.L., 1981. The Galapagos Spreading Centre
at 86°W: a detailed geothermal field study. J. Geophys. Res., 86: 979--986.
Griffiths, R.W., 1978. Layered double diffusive convection in a porous medium. PhD
thesis, Australian Nat. Univ., 55 pp.
Grill, E.V., Chase, R.L., MacDonald, R.D. and Murray, J.W., 1981. A hydrothermal de-
posit from Explorer Ridge in the northeast Pacific Ocean. Earth Planet. Sci. Lett., 52:
141--150.
Grindley, G.W. and Browne, P.R.L., 1975. Structural and hydrological factors controlling
the permeabilities of some hot-water fields. Proc. 2nd U.N. Syrup. Devel. Use Geo-
therm. Resour., San Francisco, 1: 377--386.
Guilbert, J.M. and Lowell, J.D., 1974. Variations in zoning patterns in porphyry ore de-
posits. Can. Mining Metall. Bull., 67 (742): 99--109.
Gustafson, L.B. and Hunt, J.P., 1975. The porphyry copper deposit at E1 Salvador, Chile.
Econ. Geol., 70: 857--912.
Gutsalo, L.K., 1975. Helium isotopic geochemistry in thermal waters of the Kurile Islands
and Kamchatka. Proc. 2nd U.N. Symp. Develop. Use Geotherm. Resour., San Francis-
co, 1: 745--749.
Gutsalo, L.K., 1976. On the sources and distribution of helium and argon isotopes in the
thermal waters of the Kurile Islands and Kamchatka. Geochem. Int., pp. 167--175.
Haas, J.L., 1971. The effect of salinity on the maximum thermal gradient of a hydro-
thermal system at hydrostatic pressure. Econ. Geol., 66: 940--946.
Hattori, K. and Muehlenbachs, K. 1980. Marine hydrothermal alteration at a Kuroko ore
deposit, Kosaka, Japan. Contrib. Miner. Petrol., 74: 285--292.
Hattori, K. and Sakai, H., 1979. D/H ratios, origins and evolutions of the ore forming
fluids for the Neogene veins and Kuroko deposits of Japan. Econ. Geol., 74: 535--555.
Hawkes, H.E. and Webb, J.S., 1962. Geochemistry in Mineral Exploration. Harper and
Rowe, New York, N.Y., 415 pp.
Haymon, R.M. and Kastner, M., 1981. Hot spring deposits on the East Pacific Rise at
21°N: preliminary description of mineralogy and genesis. Earth Planet. Sci. Lett.,
53: 363--381.
Healy, J., 1975. Geothermal fields in zones of recent volcanism. 2nd U.N. Syrup. Devel.
Use Geotherm. Resour., San Francisco, pp. 415--422.
Healy, J. and Hochstein, M.P., 1973. Horizontal flows in geothermal systems. J. Hydrol.
(N.Z.), 12: 71--82.
Hekinian, R., Fevrier, M., Bischoff, J.L., Picot, P. and Shanks, W.C., 1980. Sulfide de-
posits from the East Pacific Rise near 21°N. Science, 207: 1433--1444.
Helgeson, H.C., 1969. Thermodynamics of hydrothermal systems at elevated tempera-
tures and pressures. Am. J. Sci., 267: 729--804.
Helgeson, H.C., 1971. Chemical interaction of feldspars and aqueous solutions. In: W.S.
Mackenzie and J. Zussman (Editors), The Feldspars. Manchester University Press, pp.
184--217.
Helgeson, H.C., 1978. Summary and critique of the thermodynamic properties of rock
forming minerals. Am. J. Sci., 278-A: 1--229.
Henley, R.W., 1973. Solubility of gold in hydrothermal chloride solutions. Chem. Geol.,
11: 73--87.
43

Henley, R.W., 1979. Devolution of the Tauhara geothermal system. Proc. N.Z. Geo-
thermal Workshop, Univ. Auckland, 1979, pp. 7--9.
Henley, R.W. and Harper, R.T., 1979. Water--rock interaction during reinjection of
flashed BR 2 water at BR 34. D.S.I.R. Chem. Div., Open-File Rep.
Henley, R.W. and McNabb, A., 1978. Magmatic vapor plumes and ground water interac-
tion in porphyry copper emplacement. Econ. Geol., 73 : 1--20.
Henley, R.W. and Stewart, M.K., 1982. Chemical and isotopic changes in the hydrology
of the Tauhara Geothermal Field due to exploitation at Wairakei. Jour. Volcanol.
Geotherm. Res. In press.
Henley, R.W. and Thornley, P., 1979. Some geothermal aspects of polymetallic massive
sulfide formation. Econ. Geol., 74:1600--1612.
Henley, R.W. and Thornley, P., 1981. Low grade metamorphism and the geothermal
environment of massive sulphide ore formation, Buchans, Newfoundland. In: E.A.
Swanson, D.F. Strong and J.G. Thurlow (Editors), The Buchans Orebodies. Geol.
Assoc. Can. Spec. Pap., 22: 205--228.
Hoagland, J.R. and Elders, W.A., 1978. Hydrothermal mineralogy and isotopic geo-
chemistry of Cerro Prieto geothermal field Mexico, I. Hydrothermal mineral zonation.
Geotherm. Resour. Counc. Trans., 2: 283--286.
Home, R.N. and O'Sullivan, M.J., 1974. Oscillatory convection in a porous medium
heated from below. J. Fluid. Mech., 66: 339--352.
Horne, R.N. and O'Sullivan, M.J. 1977a. Convection in porous medium heated from be-
low: The effect of temperature-dependent viscosity and thermal expansion coeffi-
cient. Am. Soc. Mech. Eng. A.I. ChE -- A.S.M.E. Heat Transfer Conference, Salt Lake
City, Utah.
Horne, R.N. and O'Sullivan, M.J., 1977b. Numerical modelling of the Wairakei geother-
mal reservoir. Am. Inst. Min. Metall. Petrol. Eng. Calif. Reg. Mtg., Bakersfield, Calif.,
Pap. No. SPE 6536.
Hutchinson, R.W. and Hodder, R.W., 1972. Possible tectonic and metallogenic relation-
ships between prophyry copper and massive sulphide deposits. Can. Inst. Mining
Metall. Trans., 65: 16--22.
Imai, H., 1978. Geological Studies of the Mineral Deposits in Japan and East Asia. Uni-
versity of Tokyo Press, 389 pp.
Ishihara, S. (Editor), 1974. Geology of Kuroko deposits. Mining Geol., Spec. Iss. 6, Soc.
Mining Geol. Jpn., 437 pp.
James, R., 1968. Wairakei and Lardarello: Geothermal power systems compared. N.Z.J.
Sci., 11: 706--719.
James, R., 1975. Drawdown tests differentiate between crack flow and porous bed per-
meability. 2nd U.N. Symp. Devel. Use Geotherm. Resour., pp. 1693--1696.
Jehl, V., Poty, B. and Weisbrod, A., 1977. Hydrothermal metamorphism of the oceanic
crust in North Atlantic Ocean. Bull. Soc. Geol. Fr., 19: 1213--1221.
Jenkins, W.J., Edmond, J.M. and Corliss, J.B., 1978. Excess 3He and 4He in Galapagos
submarine hydrothermal waters. Nature, 272 : 156--158.
Jenkins, W.J., Rona, P.A. and Edmond, J.M., 1980. Excess SHe in the deep water over
the Mid-Atlantic Ridge at 26°N: evidence of hydrothermal activity. Earth Planet. Sci.
Lett., 49: 39--44.
Kartokusumo, W., Mahon, W.A.J. and Seal, K.E., 1975. Geochemistry of the Kawah
Kamojang geothermal system, Indonesia. U.N. Syrup. Devel. Use Geotherm. Resour.,
1: 575--579.
Keith, T.E.C., White, D.E. and Beeson, M.H., 1978. Hydrothermal alteration and self-
sealing in Y-7 and Y-8 drillholes in northern part of Upper Geyser Basin, Yellowstone
National Park, Wyoming. U.S. Geol. Surv. Prof. Pap., 1054-A, 26 pp.
Kendall, C., 1976. Petrology and Stable Isotope Geochemistry of Three Wells in the But-
tes area of the Salton Sea Geothermal Field Imperial Valley, California, U.S.A. PhD
thesis, University of California, Riverside, Calif., 210 pp.
44

Kesler, S.E., Russell, N., Seaward, M., Rivera, J., McCurdy, K., Cumming, G.L. and Sutter,
J.F., 1981. Geology and geochemistry of sulfide mineralization underlying the Pueblo
Viejo gold--silver oxide deposit. Dominican Republic. Econ. Geol., 76 : 1096--1117.
Kharaka, Y.K., Brown, P.M. and Carothers, W.W., 1978. Chemistry of waters in the geo-
pressurized zone from coastal Louisiana -- implications for geothermal development.
Geotherm. Resour. Counc. Trans., 2: 371--374.
Kissen, I.G. and Pakhomov, S.I., 1967. Effect of high temperatures on the chemical com-
position of ground waters. Geochem. Int., 4: 295--308.
Koga, A., 1967. Germanium, molybdenum, copper and zinc in New Zealand thermal wa-
ters. N.Z.J. Sci., 10: 428--446.
Koga, A., 1970. Geochemistry of the waters discharged from drillholes in the Otake and
Hatchobaru areas. Geothermics, Spec. Iss., 2: 1422--1425.
Kowalik, J., Rye, R. and Sawkins, F.J., 1981. Stable isotope study of the Buchans Poly-
metallic Sulphide Deposits. In: E.A. Swanson, D.F. Strong and J.G. Thurlow (Editors),
The Buchans Orebodies: Fifty years of Geology and Mining. Geol. Assoc. Can. Spec.
Pap., 22: 229--268.
Krauskopf, K.B., 1967. The source of ore metals. Geochim. Cosmochim. Acta, 35: 643--
659.
Kristmansdottir, H., 1976. Types of clay minerals in hydrothermally altered basaltic
rocks, Reykjanes, Iceland. Jokull, 26: 30--38.
Kruger, P., 1978. Radon in geothermal reservoir engineering. Geothermal Resour. Counc.
Trans., 2: 383--385.
Kruger, P., Stoker, A. and Amana, A., 1976. Radon in geothermal reservoir engineering.
Geothermics, 5: 13--20.
Lagache, M., 1976. New data on the kinetics of the dissolution of alkali feldspars at
200°C in CO2 charged water. Geochim. Cosmochim. Acta, 4 0 : 1 5 7 - - 1 6 1 .
Lambert, I.G. and Sato, T., 1974. The Kuroko and associated ore deposits of Japan: a
review of their features and metallogenesis. Econ. Geol., 6 9 : 1 2 1 5 - - 1 2 3 6 .
Lan, C.Y., Liou, J.G. and Seki, Y., 1980. Investigations of drillhole core samples from
Tatun geothermal area, Taiwan. Proc. 3rd Int. Water--Rock Interaction Symp., Ed-
m o n t o n , pp. 183--185.
Larsen, G., Gronvold, K. and Thorarinsson, S., 1979. Volcanic eruption through a geo-
thermal borehole at Namafjall, Iceland. Nature, 278: 707--710.
Leeman, W.P., Doe, B.R. and Whelan, J., 1979. Radiogenic and stable isotope studies of
hot-spring deposits in Yellowstone National Park. Geochim. J. In press.
Lindgren, W., 1933. Mineral Deposits. 4th ed. McGraw Hill, New York, N.Y. 929 pp.
Liou, J.G. and Dickson, F.W., 1978. Interaction of NaC1 solution and seawater with ande-
site 200--400°C, 500--1000 bars. Abstract. Special Water--Rock Interaction Syrup.,
San Francisco, Calif.
Liou, J.G. and Ernst, W.G., 1979. Oceanic ridge metamorphism of the East Taiwan ophio-
lite. Contrib. Miner. Petrol., 62: 335--348.
Lister, C.R.B., 1974. On the penetration of water into hot rock. Geophys. J. R. Ast. Soc.;
39: 465--509.
Lister, C.R.B., 1975. Qualitative theory on the deep end of geothermal systems. 2nd U.N.
Syrup. Devel. Use Geotherm. Resour., 1: 459--463.
Lloyd, E.F., 1959. The hot springs and hydrothermal eruptions of Waiotapu. N . Z . J .
Geol. Geophys., 2: 141--176.
Lloyd, E.F., 1972. Geology and hot springs of Orakeikorako. N.Z. Geol. Surv. Bull., 85:
164 pp.
Lowell, J.D. and Guilbert, J.M., 1970. Lateral and vertical alteration mineralization
zoning in porphyry ore deposits. Econ. Geol., 65: 373--408.
Lupton, J.E., Weiss, R.F. and Craig, H., 1977. Mantle helium in the Red Sea brines. Na-
ture, 266: 244--246.
45

Lupton, J.E., Craig, H. and Klinkhammer, G., 1979. Helium-3 and manganese at the 21°N
EPR hydrothermal site. EOS, 60: 864.
MacGeehan, J.P., 1978. Geochemistry of altered volcanic rocks at Matagami, Quebec; a
geothermal model for massive sulphide genesis. Can. J. Earth Sci., 15: 551--570.
McGoldrick, P.J., Keays, R.R. and Scott, R.B., 1979. Thallium: a sensitive indicator of
rock--seawater interaction and of sulfur saturation of silicate melts. Geochim. Cosmo-
chim. Acta, 43: 1303--1311.
McKibben, M.A., 1979. Ore minerals in the Salton Sea geothermal system, Imperial
Valley, California, U.S.A. MSc. thesis, Univ. Calif., Riverside, Calif., 90 pp.
McNabb, A., 1975. Geothermal physics. N.Z.D.S.I.R., Appl. Math. Div. Tech. Rep., 32:
35 pp.
McNabb, A. and Henley, R.W., 1979. Water--rock interaction during reinjection of geo-
thermal waster water. Proc. N.Z. Geothermal Workshop, Univ. Auckland, N.Z., pp.
68--71.
Magaritz, M. and Taylor, H.P., 1975. lso/1~O and D/H studies of igneous and sedimen-
tary rocks along a 500 km traverse of the Coast Range batholith into British Columbia.
Geol. Soc. Am. Ann. Mtg., 7 : 1 1 8 5 (Abstract).
Mahon, W.A.J., 1964. Fluorine in the natural thermal waters of New Zealand. N.Z.J. Sci.,
7: 3--28.
Mahon, W.A.J., 1966. Silica in hot water discharged from drillholes at Wairakei, New Zea-
land, N.Z.J. Sci., 9: 135--144.
Mahon, W.A.J., 1967. Natural hydrothermal systems and the reaction of hot water with
sedimentary rocks. N.Z.J. Sci., 10: 206--221.
Mahon, W.A.J., McDowell, G.D. and Finlayson, J.B., 1980a. Carbon dioxide: its role in
geothermal systems, N.Z.J. Sci., 23: 133--148.
Mahon, W.A.J., Klyen, L.E. and Rhode, M., 1980b. Neutral sodium/bicarbonate/sulphate
hot waters in geothermal systems. Chinetsu (J. Jpn. Geotherm. Energy Assoc.), 17:
11--24.
Mazor, E., 1975. Atmospheric and radiogenic noble gases in thermal waters: their poten-
tial application to prospecting and steam production studies. Proc. 2nd U.N. Symp.
Devel. Use Geotherm. Resour., San Francisco, 1: 793--802.
Mazor, E. and Wasserburg, G.J., 1965. Helium, neon, argon, krypton and xenon in gas
emanations from Yellowstone and Lassen Volcanic National Parks. Geochim. Cosmo-
chim. Acta, 29: 443--454.
Menzies, M. and Seyfried, W.E., 1979. Basalt~sea water interaction: trace element
and strontium isotopic variations in experimentally altered glassy basalt. Earth. Planet.
Sci. Lett., 44: 463--472.
Mercado, S., 1967. Geoquimica hidrothermal en Cerro Prieto B.C. Mexico. Comision
Federal de Electricidid, Mexicali, 31 pp.
Mercado, S., 1976. Cerro Prieto geothermal field Mexico: wells and plant operation. Int.
Congress on Thermal Waters Geothermal Energy and Volcanism of the Mediterranean
area.
Miyashiro, A., Shido, F. and Ewing, M., 1971. Metamorphism on the Mid-Atlantic Ridge
near 24 ° and 30°N. Phil. Trans. R. Soc. Lond., Ser. A, 268: 589--603.
Monin, A.S., Plakhin, E.A., Podrazhansky, A.M., Sagalevich, A.M. and Sorokhtiu, O.G.,
1981. Visual observations of the Red Sea hot brines. Nature, 291 : 222--225.
Mottl, M.J., Holland, H.D. and Corr, R.F., 1979. Chemical exchange during hydrothermal
alteration of basalt by seawater, II. Experimental results for Fe, Mn and S species.
Geochim. Cosmochim. Acta, 43: 869--884.
Muehlenbachs, K. and Clayton, R.N., 1976. Oxygen isotope composition of the oceanic
crust and its bearing on seawater. J. Geophys. Res., 81: 4365--4369.
Muffler, L.J.P. and White, D.E., 1969. Active metamorphism of Upper Cenozoic sedi-
ments in the Salton Sea geothermal field and the Salton Trough, southeastern Califor-
nia. Geol. Soc. Am. Bull., 80: 157--182.
46

Muffler, L.J.P., White, D.E. and Truesdell, A.H., 1971. Hydrothermal explosion craters in
Yellowstone National Park. Geol. Soc. Am. Bull., 82: 723--740.
Naboko, S.I., 1977. The metal behaviour by interaction of thermal water with rock in the
bowels of Kamchatka hydrothermal systems. Proc. 2nd Int. Symp. Water--Rock Inter-
action, Strasbourg, III: 19--29.
Naboko, S.I. and Karpov, G.A., 1977. Pore solution metasomatism, sulphide formation.
Proc. 2nd Int. Symp. Water--Rock Interaction, Strasbourg, III: 46--53.
Nagao, K., Takaoka, N. and Matsubayashi, O., 1979. Isotopic anomalies of rare gases in
the Nigorikawa Geothermal area, Hokkaido, Japan. Earth Planet. Sci. Lett., 44:
82--90.
Nairn, I.A. and Wiradiradja, S., 1980. Late Quaternary hydrothermal explosion brec-
cias at Kawerau geothermal field, New Zealand. Bull. Volcanol., 43: 1--13.
Nakamura, H., Sumi, K., Katagiri, K. and Iwata, T., 1970. The geological environment of
Matsukawa geothermal area, Japan. Geothermics (Spec. Iss., 2) 2 (Pt. 1): 221--231.
Nathenson, M., 1981. Low temperature geothermal. Geotherm. Resour. Counc. Bull.,
(August 1981), pp. 3--7.
Natland, J.H., Rosendahl, B. et al., 1979. Galapagos hydrothermal mounds: stratigraphy
and chemistry revealed by deep sea drilling. Science, 204 : 613--616.
Naughton, J.J. and Thomas, D., 1978. Helium in fumarole and well gases as an index of
long-term geothermal potential. Geotherm. Resour. Counc. Trans., 2(2): 479--482.
Nielson, D.L., 1978. Radon emanometry as a geothermal exploration technique; theory
and an example from Roosevelt hot springs, Utah. Rep. Earth Sci. Lab., Univ. Utah
Res. Inst., ESL-14, 31 pp.
Nordstrom, D.K. and Jenne, E.A., 1977. Fluorite solubility equilibria in selected geother-
mal waters. Geochim. Cosmochim. Acta, 41: 175--188.
Norton, D. and Knapp, R., 1977. Transport phenomena in hydrothermal systems: the
nature of porosity. Am. J. Sci., 277: 913--936.
Norton, D. and Knight, J., 1977. Transport phenomena in hydrothermal systems: cooling
plutons. Am. J. Sci., 277: 937--981.
Ohmoto, H. and Rye, R.O., 1974. Hydrogen and oxygen isotopic composition of fluid
inclusions in the Kuroko deposits, Japan. Econ. Geol., 69: 947--953.
Oki, Y. and Hirano, T., 1970. The geothermal system of the Hakone volcano. Geo-
thermics (Spec. Iss.), 2: 1157--1166.
Oki, Y. and Hirano, T., 1979. Geochemistry of the Hakone, Yugawara and Atami hydro-
thermal systems (abstr.). Am. Geophys. Union, San Francisco.
Olafsson, J. and Riley, J.P., 1978. Geochemical studies on the thermal brine from Reyk-
janes, Iceland. Chem. Geol., 21: 219--237.
Olson, E.R., 1978. Oxygen isotope studies of the Salton Sea geothermal field: new
insights. N.Z.D.S.I.R. Bull., 220: 121--126.
Olson, E.R. and Elders, W.A., 1978. Hydrothermal mineralogy and isotopic geochemistry
in the Cerro Prieto geothermal field, Mexico, II. Isotopic geochemistry. Geotherm.
Resour. Counc. Trans., 2: 513--515.
O'Neill, J.R. and Silberman, M.L., 1974. Stable isotope relations in epithermal Au-Ag
deposits. Econ. Geol., 69: 902--909.
Ozawa, T.M., Kamada, M., Yoshida, M. and Sanemasa, I., 1973. Genesis of acid h o t
springs. Proc. Syrup. Hydrogeochemistry and Biogeochemistry, 1, Clark Co., Wash.,
pp. 105--121.
Paces, T., 1975. A systematic deviation from the Na-K-Ca Geothermometer below 75°C
and above 10 -a atm. Pco2. Geochim. Cosmochim. Acta, 39: 541--544.
Pampura, V.D., Karpov, I.K. and Kazmin, L.A., 1975. Modelling of the equilibrium com-
ponent compositions and properties of NaC1 hydrothermal solutions of the Pauzhetsk
type. Proc. 2nd U.N. Syrup. Devel. Use Geothermal Resources, San Francisco, 1:
809--814.
47

Panichi, C. and Gonfiantini, R., 1977. Environmental isotopes in geothermal studies. Geo-
thermics, 6: 143--162.
Panichi, C. and Tongiorgi, E., 1975. Carbon isotopic composition of CO2 from springs,
fumaroles, mofettes, and travertines of central and southern Italy. Proc. 2nd U.N.
Syrup. Devel. Use Geothermal Resources, San Francisco, 1 : 815--825.
Petrovic, R., Berner, R.A. and Goldhaver, M.B., 1976a. Rate control in dissolution of
alkali feldspars, I. Study of residual feldspar grains by X-ray photoelectron spec-
troscopy. Geochim. Cosmochim. Acta, 40: 537--548.
Petrovic, R., Berner, R.A. and Goldhaver, M.B., 1976b. Rate of control in feldspar disso-
lution, II. The protective effect of precipitates. Geochim. Cosmochim. Acta, 40:
1509--1521.
Potter, J.M. and Dickson, F.W., 1980. Role of solution composition in the alteration of
rhyolite glass at 300°C (abstr.). Proc. 3rd Int. Syrup. Water--Rock Interaction, Ed-
monton.
Putnam, G.W., 1975. Base metal distribution in granitic rocks, II. Three dimensional vari-
ation in the Lights Creek Stock, California. Econ. Geol., 70: 1225--1241.
Radtke, A.S., Rye, R.O. and Dickson, F.W., 1980. Geology and stable isotope studies of
the Carlin gold deposit, Nevada. Econ. Geol., 75: 641--672.
Richardson, C.K. and Holland, H.D., 1979. Fluorite deposition in hydrothermal systems.
Geochim. Cosmochim. Acta, 43: 1327--1335.
Roedder, E., 1971. Fluid inclusion studies of the porphyry-type ore deposits at Bingham,
Utah, Butte, Montana and Climax, Colorado. Econ. Geol., 66: 98--120.
Rona, P.A., 1980. TAG hydrothermal field; Mid-Atlantic Ridge crest at latitude 26°N. J.
Geol. Soc. Lond., 137: 385--402.
Rybach, L. and Muffler, L.J.P., 1981. Geothermal systems: Principles and Case Histories.
Wiley, New York, N.Y., 359 pp.
Rytuba, J.J., 1976. Geology and ore deposits of the McDermitt caldera, Nevada--Oregon.
U.S. Geol. Surv., Open-File Rep., 76--535, 9 pp.
Rytuba, J.J., Potter, J.M., Dickson, F.W. and Radtke, A.S., 1978. Experimental alteration
of rhyolite glass at 300°C; implication for silicate mineral zoning in the McDermitt
mercury deposit, Nevada. (abstr.). Am. Geophys. Union, San Francisco.
Sackett, W.M. and Chung, H.M., 1979. Experimental confirmation of the lack of carbon
dioxide exchange between methane and carbon oxides at high temperatures, Geochim.
Cosmochim. Acta, 43: 273--276.
Sakai, H., Gunnlaugsson, E., Tomasson, J. and Rouse, J.E., 1980. Sulfur isotope sys-
tematics in Icelandic geothermal systems and influence of seawater circulation at
Reykjanes. Geochim. Cosmochim. Acta, 44: 1223--1231.
Sangster, D.F., 1972. Precambrian volcanogenic massive sulphide deposits in Canada: a
review. Geol. Surv. Can. Pap., 7 2 - - 1 2 2 : 4 4 pp.
Sawkins, F.J. and Kowalik, J., 1981. The source of ore metals at Buchans: Magmatic
versus leaching models. In: E.A. Swanson, D.F. Strong and J.G. Thurlow (Editors).
The Buchans Orebodies. Geol. Assoc. Can. Spec. Pap., 22: 205--228.
Sawkins, F.J., O'Neil, J.R. and Thompson, J.M., 1979. Fluid inclusion and geochemical
studies of vein gold deposits, Baguio District, Philippines. Econ. Geol., 74: 1420--
1434.
Scott, R.B., Malpas, J., Rona, P.A. and Udintsev, G., 1976. Duration of hydrothermal
activity at an oceanic spreading centre, Mid-Atlantic Ridge (at 26°N). Geology, 4:
233--236.
Seward, T.M., 1973. Thio complexes of gold and the transport of gold in hydrothermal
ore solutions. Geochim. Cosmochim. Acta, 37: 379--399.
Seward, T.M., 1974. Equilibrium and oxidation potential in geothermal waters at Broad-
lands, New Zealand. Am. J. Sci., 274: 190--192.
Seyfried, W.E. and Bischoff, J.L., 1977. Hydrothermal transport of heavy metals by sea
water: The role of sea water/basalt ratio. Earth Planet. Sci. Lett., 34: 71--77.
48

Seyfried, W.E. and Bischoff, J.L., 1979. Low temperature basalt alteration by seawater:
an experimental study at 70°C and 150°C. Geochim. Cosmochim. Acta, 43: 1937--
1947.
Seyfried, W.E. and Mottl, M.J., 1977. Origin of submarine metal-rich hydrothermal solu-
tions: experimental basalt--seawater interaction in seawater dominated system. Proc.
2nd Int. Symp. Water--Rock Interaction, I.A.G.C., Strasbourg, pp. 173--179.
Shanks, W.C. and Bischoff, J.L., 1977. Ore transport and deposition in the Red Sea geo-
thermal system: a geochemical model. Geochim. Cosmochim. Acta, 41: 1507--1519.
Sheppard, D.S., 1977. Hydrothermal Chemistry of Some Silicates. PhD thesis, University
of Otago, Dunedin, 179 pp.
Sheppard, S.M.F., Nielsen, R.L. and Taylor, H.P., 1969. Oxygen and hydrogen isotope
ratios of clay minerals from porphyry copper deposits. Econ. Geol., 64: 755--777.
Sheppard, S.M.F., Nielsen, R.L. and Taylor, H.P., 1971. Hydrogen and oxygen isotope
ratios in minerals from porphyry copper deposits. Econ. Geol., 66: 515--542.
Sigvaldason, G.E., 1979. Fluids in volcanic and geothermal systems. Proc. Nobel Symp.
1979, Chemistry and Geochemistry of Solutions at High Temperature and Pressure,
Karlskoga, Sweden. Pergamon, London.
Sillitoe, R.H., 1973. The tops and bottoms of porphyry copper deposits. Econ. Geol., 68:
799--815.
Simkin, T., 1979. Volcanology. Rev. Geophys. Space Phys., 17: 872--886.
Skinner, B.J., White, D.E., Rose, H.J. and Mays, R.E., 1967. Sulfides associated with the
Salton Sea geothermal brine. Econ. Geol., 62: 316--330.
Solomon, M., 1976. Volcanic massive sulfide deposits and their host rocks -- a review and
an explanation. In: K.H. Wolf (Editor) Handbook of Strata-Bound and Stratiform Ore
Deposits, II. Regional Studies and Specific Deposits. Elsevier, Amsterdam.
Sondergeld, C.H. and Turcotte, D.L., 1977. An experimental study of two phase convec-
tion in a porous medium with application to geological problems. J. Geophys. Res.,
82: 2045--2053.
Sondergeld, C.H. and Turcotte, D.L., 1978. Flow and visualisation studies of two-phase
thermal convection in a porous layer. In: L. Rybach and L. Stegena. Geothermics and
Geothermal Energy. Contributions in Current Research in Geophysics, 7. Birkhauser
Verlag, Basel, pp. 321--330.
Sorey, M.L., 1978. Numerical modelling of liquid geothermal systems. U.S. Geol. Surv.,
Prof. Pap., 1044-D: 25 pp.
Sourirajan, S. and Kennedy, G.C., 1962. The system H20-NaC1 at elevated temperatures
and pressures. Am. J. Sci., 260: 115--141.
Spooner, E.T.C. and Fyfe, W.S., 1973. Sub-sea floor metamorphism, heat and mass trans-
fer. Contrib. Miner. Petrol., 42: 287--304.
Spooner, E.T.C., Beckinsale, R.D., Fyfe, W.S, and Smewing, J.D., 1974. 1SO enriched
ophiolitic metabasic rocks from E. Linguria (Italy), Andos (Greece) and Troodos
(Cyprus). Contrib. Miner. Petrol., 47: 41--62.
Spooner, E.T.C., Chapman, M.J. and Smewing, J.D., 1977. Strontium isotopic contamina-
tion and oxidation during ocean floor hydrothermal metamorphism of the ophiolitic
rocks of the Troodos Massif, Cyprus. Geochim. Cosmochim. Acta, 41 : 873.
Stakes, D.S., 1978. Mineralogy and geochemistry of the Michigan Deep Hole Metagabbro
compared to seawater hydrothermal alteration. J. Geophys. Res., 83: 5820--5824.
Steiner, A., 1968. Clay minerals in hydrothermally altered rock at Wairakei, New Zea-
land. Clays Clay Miner., 16: 193--213.
Steiner, A., 1977. The Wairakei geothermal area, North Island, New Zealand. N.Z. Geol.
Surv. Bull., 9 0 : 1 3 6 pp.
Stewart, M.K., 1978. Stable isotopes in waters from the Wairakei geothermal area, New
Zealand. Stable Isotopes in the Earth Sciences. N.Z.D.S.I.R. Bull., 220: 113--120.
Straus, J.M. and Schubert, G., 1977a. Thermal convection of water in a porous medium:
49

effects of temperature and pressure dependent thermodynamic and transport proper-


ties. J. Geophys. Res., 82: 325--333.
Straus, J.M. and Schubert, G., 1977b. Two-phase convection in a porous medium. J. Geo-
phys. Res., 82: 3411--3421.
Styrt, M.M., Brackman, A.J., Holland, H.D., Clark, B.C., Pisutha-Arnond, V., Eldridge,
C.S. and Ohmoto, H., 1981. The mineralogy and isotopic composition for sulfur in hy-
drothermal sulfide-sulfate deposits on the East Pacific Rise, 21°N latitude. Earth
Planet. Sci. Lett., 53: 382--390.
Sutton, F.M. and McNabb, A., 1977. Boiling curves at Broadlands geothermal field, New
Zealand. N.Z.J. Sci., 20: 333--337.
Taylor, H.P., 1974. Application of oxygen and hydrogen isotope studies to problems of
hydrothermal alteration and ore deposition. Econ. Geol., 69: 843--883.
Thomas, D. and Kroopnick, P.M., 1978. Isotopes and gases in a Hawaiian geothermal sys-
tem. Geotherm. Res. Counc. Trans., 2: 653--654.
Thompson, G. and Humphris, S.E., 1977. Seawater--rock interaction in the oceanic base-
ment. Proc. 2nd Int. Symp. Water--Rock Interaction, Strasbourg, III: 13--18.
Tomasson, J. and Kristmannsdottir, H., 1972. High-temperature alteration minerals and
thermal brine, Reykjanes, Iceland. Contrib. Miner. Petrol., 36: 123--137.
Truesdell, A.H., 1975. Summary of Section III -- geochemical techniques in exploration.
Proc. 2nd U.N. Symp. Devel. Use Geothermal Resources, 1: 53--63.
Truesdell, A.H. and Fournier, R.O., 1976. Calculation of deep temperatures in geother-
mal systems from the chemistry of boiling spring waters of mixed origin. Proc. 2nd
U.N. Symp. Devel. Use Geothermal Resources, San Francisco, 1: 837--844.
Truesdell, A.H. and Hulston, J.R., 1979. Isotopic evidence on environments of geother-
mal waters. In: P. Fritz and J.Ch. Fontes (Editors), Handbook of Environmental Iso-
tope Geochemistry, Vol. 1, Elsevier, Amsterdam, pp. 179--226.
Truesdell, A.H. and White, D.E., 1973. Production of superheated steam from vapor-
dominated geothermal reservoirs. Geothermics, 2: 154--173.
Truesdell, A.H., Rye, R.O., Pearson, F.J. Jr., Olson, E.R., Nehring, N.L., Whelan, J.F.,
Huebner, M.A. and Coplen, T.B., 1979. Preliminary isotopic studies of fluids from the
Cerro Prieto geothermal field. Geothermics, 8: 223--230.
Upadhayay, H.D. and Strong, D.F., 1973. Geological setting of the Betts Cove copper
deposits -- an example of ophiolitic sulphide mineralization. Econ. Geol., 68: 161--
167.
Vidal, V.M.V., Vidal, F.V. and Isaacs, J.D., 1978. Coastal submarine hydrothermal activ-
ity off Northern Baja California. J. Geophys. Res., 83: 1757--1774.
Weissberg, B.G., 1969. Gold--silver ore-grade precipitates from New Zealand thermal wa-
ters. Econ. Geol., 64: 95--108.
Weissberg, B.G. and Wilson, P.T., 1977. Montmorillonites and the Na/K Geothermometer,
N.Z.D.S.I.R. Bull., 218: 31--34.
Weissberg, B.G., Browne, P.R.L. and Seward, T.M., 1979. Ore metals in active geothermal
systems. In: H.L. Barnes (Editor). Geochemistry of Hydrothermal Ore Deposits. Wiley
Interscience, New York, N.Y., 2nd ed., pp. 738--780.
Wells, J.D. and Mullens, T.E., 1973. Gold-bearing arsenian pyrite determined by micro-
probe analysis, Cortez and Carlin Gold Mines, Nevada. Econ. Geol., 68: 187--201.
Weres, O., Tsao, K. and Wood, B., 1977. Resource, technology and environment at the
Geysers. Lawrence Berkeley Laboratory, Report LBL-5231.
Wetlaufer, P.H., Bethke, P.M. and Barton, P.B., 1979. The Creede Ag-Pb-Zn-Cu-Au dis-
trict, Central San Juan Mountains, Colorado -- a fossil geothermal system. Int. Assoc.
Genesis of Ore Deposits (IAGOD), 5th Symposium, Snowbird, Utah. Proc. Volume,
2: 159--164.
White, D.E., 1955. Thermal springs and epithermal ore deposits. In: A.M. Bateman
(Editor). Econ. Geol. 50th Anniv. Volume, pp. 99--154.
50

White, D.E., 1967. Mercury and base-metal deposits with associated thermal and mineral
waters. In: H.L, Barnes (Editor), Geochemistry of Hydrothermal Ore Deposits. Holt,
Rinehart, Winston, New York, N.Y., 1st ed., pp. 575--631.
White, D.E., 1968. Environments of generation of some base-metal ore deposits. Econ.
Geol., 63: 301--335.
White, D.E., 1974. Diverse origins of hydrothermal ore fluids. Econ. Geol., 69: 954--973.
White, D.E., 1981. Active geothermal systems and hydrothermal ore deposits. Econ.
Geol., 75th Anniv.
White, D.E. and Guffanti, M., 1979. Geothermal systems and their energy resources, Rev.
Geophys. Space Phys., 17: 877--902.
White, D.E., Muffler, L.J.P. and Truesdell, A.H., 1971. Vapor-dominated hydrothermal
systems compared with hot water systems. Econ. Geol., 66: 75--97.
White, D.E., Fournier, R.O., Muffler, L.J.P. and Truesdell, A.H., 1975. Physical results
for research drilling in thermal areas of Yellowstone Park, Wyoming. U.S. Geol. Surv.
Prof, Pap., 8 9 2 : 7 0 pp.
Whitehead, N.E., 1979. Radon measurements at Wairakei and Broadlands. Institute of
Nuclear Sciences, D.S.I.R., Open-File Rep. No. 1000.
Wilson, S.H., 1963. Tritium determinations on bore waters in the light of chloride en-
thalpy relations. In: E. Tongiorgi (Editor), Nuclear Geology on Geothermal Areas,
Spoleto, Italy. Consiglio Nazionale della Ricerohe Laboratorio di Geologia Nucleate,
Pisa, pp. 173--184.
Wolery, T.J. and Sleep, N.H., 1976. Hydrothermai circulation and geothermal flux at
mid-ocean ridges. J. Geol., 84: 249--275.
Wollenberg, H.A., 1975. Radioactivity of geothermal systems. 2nd U.N. Symp. Devel. Use
of Geothermal Resources, 2: 1283--1292.
Wood, C.P., 1980. Mineralogy in the Mahiao geothermal field, Philippines. N.Z. Geochem.
Group Newslett., 56: 91.
Wooding, R.A., 1963. Convection in a saturated porous medium at large Rayleigh number
or Peclet number. J. Fluid Mech., 15: 527--544.
Zen, E-an, 1974. Burial metamorphism. Can. Mineral., 12: 445--455.

(Received July 7, 1981; accepted February 3, 1982)

You might also like