Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 21

Table of Contents

I. Introduction............................................................................................................................3
II. Volcanic Environment of Epithermal gold deposit.................................................................3
1. Tectonic setting...................................................................................................................4
2. Mineralization.....................................................................................................................6
3. Alteration............................................................................................................................8
4. Geometrical controls.........................................................................................................12
5. Origin................................................................................................................................12
III. Type of epithermal sulfidation..............................................................................................12
1. High Sulfidation................................................................................................................12
2. Intermediate Sulfidation....................................................................................................14
3. Low Sulfidation................................................................................................................14
IV. Exploration guide..................................................................................................................16
I. Introduction
Epithermal Gold refers to a type of gold deposit formed from hydrothermal fluids
at shallow levels in the earth’s crust. “Epithermal” is an old term used to classify
hydrothermal deposits based on temperature and depth of deposition.
Lindgren (see Lindgren, 1922, 1933, p. 2110) created the term "epithermal" in an
effort to categorize hydrothermal mineral deposits using a similar rationale to that being
used to categorize metamorphic rocks. The regulating variables for metamorphic
assemblages were identified as pressure (or depth) and temperature, and both looked to
be able to be calculated for ore assemblages based on known information for mineral
stabilities and textures. Since depths are typically estimated from temperatures, it is
preferable to limit the term "epithermal" to a temperature range comparable to that of
explored geothermal fields (50 to 350°C). However, some authors (e.g. Schmitt, 1950)
have attempted to use the term to refer to or to infer a specified depth range.
This description of epithermal systems lacks the familiar (but typically unwieldy)
detail of many ore-deposit classifications, but it does have the advantage of focusing on
crustal environment and crustal process. Depending on the purpose of the categorization,
deposit types may be subdivided for convenience (e.g., when developing exploration
models) using additional criteria: gold deposits in sedimentary rock over thrust terranes
(Carlin-type deposits), for example, are examined in this volume by Berger and Bagby.
Likewise, volcanogenic huge sulphide deposits are a product of such a hydrothermal
environment, but they are distinguished by Para geneses and forms that represent seabed
depositional conditions. This chapter focuses on epithermal deposits that are temporally
and spatially linked to terrestrial volcanic rocks. The elements that contribute to the
formation of diverse epithermal deposit-settings are examined.
Epithermal gold deposits are one of the world's most important gold sources. These
deposits are usually found in volcanic rocks and are created by the deposition of gold and
other minerals from hot fluids that circulate through the rocks' fissures and faults. In
epithermal deposits, gold is frequently coupled with silver, and occasionally with other
metals such as copper, lead, and zinc.
Exploration and development of epithermal gold resources necessitate a thorough
understanding of the deposit's geological and mineralogical properties. This includes the
identification of the host rocks, the nature of the fluid that deposited the gold, and the
distribution of gold and other minerals within the deposit.

II. Volcanic Environment of Epithermal gold deposit


Epithermal gold deposits are often associated with recent or active volcanic activity.
They can be found in regions that have experienced volcanic activity in the past, or in
areas that are currently undergoing volcanic activity. The volcanic environment provides
the necessary heat and fluids for the formation of epithermal gold deposits.Volcanic
rocks such as rhyolite, andesite, and basalt are often associated with epithermal gold
deposits. These rocks can act as hosts for the mineralizing fluids that carry gold and other
minerals. The fractures and permeable zones in the surrounding rocks created by volcanic
activity allow the mineralizing fluids to flow through and deposit minerals such as gold.
In addition to volcanic rocks, epithermal gold deposits can also be associated with
sedimentary rocks or areas of structural deformation. These types of rocks can also
provide the necessary permeable zones for mineralizing fluids to flow through and
deposit minerals.

1. Tectonic setting

Epithermal gold deposits are typically associated with areas of recent or active
volcanic activity, which are often found near tectonic plate boundaries. For example,
many epithermal gold deposits are found along the Pacific Ring of Fire, which is a region
of intense volcanic and seismic activity that surrounds the Pacific Ocean. In general,
areas of active tectonic activity can provide the heat and fluids necessary for the
formation of epithermal gold deposits. Tectonic activity can create fractures and
permeable zones in the surrounding rocks, which can allow hot, mineral-rich fluids to
flow through the rocks and deposit minerals such as gold. The type of tectonic activity in
an area can also be important for the formation of epithermal gold deposits. For example,
areas of extensional tectonics, where the Earth's crust is being pulled apart, can create
pathways for mineralizing fluids to flow and deposit minerals such as gold. Areas of
compressional tectonics, where the Earth's crust is being squeezed together, can also be
important for the formation of epithermal gold deposits, as they can create zones of
deformation and hydrothermal alteration that can host mineralization.
The epithermal deposits occur in the same tectonic settings relative to porphyry
deposits, which are volcano-plutonic arcs, island arcs and Cordillera arcs associated with
subduction zones. They are mainly present along the pacific ring and
the alpine chain in Europe. They are very superficial deposits since they are deposited
between 2 km depth and surface [24]. They are therefore very sensitive to erosion, which
explains why most of the deposits still existing today are post Jurassic (Figure 1).
There are two types of epithermal deposits, which are distinguished primarily by the
difference in oxidation state of Sulphur in the related ore-forming fluids
[25]: we have “high sulfidation” deposits (also called acid epithermal) and “low
sulfidation” deposits (neutral epithermal). The “high sulfidation” epithermal
deposits form at temperatures between 150˚C and about 300˚C [24] from a very
acidic mineralizing fluid. Copper minerals, in particular the chalcopyrite, enargite and
luzonite, characterize the mineralization. This type of deposit is also very rich in pyrite; it
is the most common sulphide. These deposits are
hosted in volcanic andesitic to dacitic rocks belonging to the calc-alkaline series.
Like porphyry deposits, the “high sulfidation” epithermal deposits are marked
by zonation [24].Unlike the acidic epithermal deposits, “low sulfidation” deposits have
the particularity of being associated with hydrothermal fluids of neutral pH with
temperature not exceeding 300˚C. These fluids are dominated by meteoric waters,
but there is also evidence of the contribution of magmatic fluids and gases ([23]
[26]). These deposits have varied textures: different generations of breccias,
banded quartz, and chalcedony or calcite veins. The host rocks are magmatic or
volcano-sedimentary, generally of calc-alkaline composition similar to high sulphidation
epithermal deposits. Mineralization is essentially in the form of veins
and stockwerks, even though it can be disseminated ([24] [25]). The mineralogy
D. F. W. Nguimatsia et al.DOI: 10.4236/ojg.2017.711113 1696 Open Journal of Geology
of such deposits is dominated by pyrite but sphalerite, galena, arsenopyrite,
chalcopyrite and tellurides are equally found. The circulation of hydrothermal
fluids is the origin of alteration zones, zoning is less marked compared to acidic
epithermal deposits.
There is evidence showing the existence of a genetic and spatial link between
porphyry and high sulfidation epithermal deposits, these latter constitutes the
apex of porphyry systems [25]. The link between porphyry and low sulfidation
epithermal deposits is less obvious. Indeed, unlike acid epithermal, low sulphide
deposits do not occur in the immediate vicinity of a volcanic system but rather
within geothermal systems. The formation acid epithermal deposits results fromthe
release of vapours and magmatic fluids carrying metals during hydraulic
fracturing around the porphyry, which will migrate to the surface [27]. The deposition of
the mineralization will be caused either by a phenomenon of ebullition at the surface, a
mixture and a dilution of the magmatic fluid by meteoric
fluids, or a combination of these two phenomena ([27] [28]).
Hydrothermal fluids related to low sulphide epithermal deposits are a mixture
between magmatic fluids and meteoric waters, with the latter being dominant.
These fluids are less salty, thus poor in chlorine but particularly rich in gases
(H2S and CO2). Under surface conditions, the fluids will enter into ebullition,
phenomenon that will be accompanied by the loss of steam containing CO2 and
H2S [24]. This loss of sulphur will decrease the solubility of gold in the fluid
provoking its precipitation along the fluid’s circulation path.

Figure 1: Distribution of different types of known gold deposit in geological time [3].
2. Mineralization

Isotopic studies have shown that the hydrothermal fluid in the low-sulfidation
environment is dominated by meteoric water, but some systems contain water and
reactive gases of magmatic origin (Hedenquist and Lowenstern, 1994). The fluids that
rise from great depth have equilibrated with their host rocks, so are reduced and have a
near-neutral pH (Giggenbach, 1992); this reaction results in NaCl, CO2 and H2S being
the principal species in the fluid (Fig. 1). Boiling at shallow depth generates a CO2 - and
H2S-rich vapor which may condense near the surface in the vadose zone, forming steam-
heated acid-sulfate waters from oxidation of H2S (pH 2-3 waters with a temperature
close to 100oC).
By contrast, in the high-sulfidation environment reactive components derived from an
oxidized magmatic source ascend to the near surface with little water-rock interaction at
depth. The HCl- and SO2 -rich vapor may become absorbed by groundwater (Rye, 1993),
resulting in a hot (200-300o C), highly acidic (pH 0-2) and oxidized fluid that reacts
extensively with and leaches the host rock at shallow depth. Thus one major difference
between these two styles of fluids is the degree to which they have equilibrated with their
host rocks below the level of ore deposition.

To simplify our brief discussion, we examine epithermal mineralization in which gold


is the principal economic metal, mainly hosted by or related to calc-alkalic or alkalic
volcanic rocks. Both styles are widespread (Table 1), principally in convergent tectonic
settings, and both have examples of major economic significance. A similar discussion
could be made for epithermal systems rich in silver, base-metals (Pb, Zn) or even tin.
The most basic characteristics of any ore deposit are its form, mineralogy, textures
and alteration zoning. Based on a variety of observations (Sillitoe, 1977; Buchanan,
1981; Heald et al., 1987; White et al., 1995), when we compare the low- and high-
sulfidation deposits we see that there is considerable overlap in characteristics, but there
are also many distinctive features. For example, the characteristic form of these two
styles of epithermal deposit is different (Table 2). Most low-sulfidation deposits consist
of cavity-filling veins with sharp boundaries, or stockworks of small veins. Veins may be
important in high-sulfidation deposits, but the majority consist of disseminated ores that
replace or impregnate leached country rock. Both styles are typically structurally
controlled, though in high-sulfidation deposits the disseminated form may conceal this.
Ore and Gangue The mineralogy of ores shows considerable overlap, but there are
several pronounced differences, based on a compilation of mineral data for more than 130
epithermal deposits (White et al., 1995; Table 3); these differences are mainly in the
sulfide mineralogy, which reflects the different redox conditions of the hydrothermal
fluid. One distinction is the common occurrence of sphalerite and arsenopyrite in low-
sulfidation deposits, whereas sphalerite is scarce and arsenopyrite rare in high-sulfidation
deposits (White et al., 1995). Unlike low-sulfidation examples, high-sulfidation deposits
commonly contain copper minerals, especially the high-sulfidation state sulfosalts
enargite-luzonite. Such sulfides, including the relatively high-sulfidation state minerals
tennantite-tetrahedrite (Barton and Skinner, 1979), are typically rare or absent in low-
sulfidation deposits. The total abundance of sulfide minerals (dominantly pyrite) is not
significant, as it can be high or low in either style. The gangue minerals associated with
the two styles of epithermal mineralization also show considerable overlap, but there are
clear differences as well (Table 4), differences that reflect the reactivity (pH) of the
altering fluid.
Quartz is common in both styles. Adularia and calcite, both indicating near-neutral
pH conditions, are common minerals in low-sulfidation deposits (the most common after
quartz; Buchanan, 1981), but are absent from high-sulfidation deposits. Minerals formed
under relatively acidic conditions, such as kaolinite and alunite (plus pyrophyllite,
diaspore and P-, Sr-, Pb- and REE-bearing phosphate-sulfate minerals), are common but
minor in high-sulfidation deposits. In low-sulfidation deposits kaolinite and alunite do not
occur as gangue, except as an overprint (Vikre, 1985).
Textures
The textures that characterise the two types of deposits are very different. Low-
sulfidation deposits show a wide variety of textures, including banded, crustiform quartz
and chalcedony veins, druse-lined cavities, and spectacular, multiple-episode vein
breccias. The latter reflect multiple episodes of mineral deposition and hydraulic
fracturing, followed by explosive pressure release which may be associated with
hydrothermal eruptions at the surface. Lattice-textured bladed calcite is common, formed
as a result of boiling (Simmons and Christenson, 1994), though it may be replaced by
quartz as the system cools. In areas that have experienced little erosion, distinctive silica
sinters deposited at the paleosurface by neutral-pH hot-spring waters may still be present
(Vikre, 1985; White et al., 1989). The sinters are rhythmically banded, with vertical
growth structures, and may contain plant fragments; they are easily distinguished from
siliceous replacement of bedded sediments.
By contrast, the typical textures of high-sulfidation deposits show relatively little
variation, with the most characteristic texture being massive bodies of vuggy quartz
typical of Nansatsu-type deposits, though locally veins and breccias may be
importanthosts to ore. The vuggy quartz is caused by acid leaching at pH < 2 (Stoffregen,
1987), which leaves open spaces and mainly silica behind; this residue then recrystallizes
to quartz, with additional quartz and pyrite deposited from solution. Massive to banded
sulfide veins consisting of pyrite and enargite may also cut the vuggy quartz bodies.
Silica sinters are never formed at the surface in this acidic environment because of kinetic
factors inhibiting the polymerization and precipitation of silica from acidic solutions. On
textural grounds alone, there is seldom much difficulty to distinguish the two styles of
deposit.

3. Alteration

In addition to gangue mineralogy, the mineralogy and zonation of hydrothermal


alteration assemblages is another distinguishing characteristic. Many alteration minerals
are stable over limited temperature and/or pH ranges, and thus provide important
information to reconstruct the thermal and geochemical structure of the hydrothermal
system. Because of the near-surface origin of these deposits, and the dynamic setting in
which they form (including the possibility of significant erosion during hydrothermal
activity; Reyes, 1990; Sillitoe, 1994), alteration that overprints the system must be
distinguished from that associated with ore.

The ore-associated alteration in low-sulfidation deposits is produced by near-neutral


pH thermal waters, with temperature decreasing both with decreasing depth and with
increasing distance from the conduit of fluid flow. In active systems alteration
mineralogy and temperature are directly measured, thus indicating the range of thermal
stability of temperature-dependant minerals (Henley and Ellis, 1983; Reyes, 1990; Fig.
2). During exploration of epithermal prospects this information allows paleoisotherms to
be deduced from the distribution of alteration minerals, which in turn helps to locate
conduits of paleoflow, and to determine the level of erosion. The former is significant
because major ore accumulations must occur in conduit zones. The latter is significant as
most epithermal ore is deposited over the range of 180 to 280oC, equivalent to a depth
below the paleowater table from about 100 m to 800-1500 m (Hedenquist and Henley,
1985). Prospects with paleotemperatures low in that range are encouraging, whereas
indications of temperatures >280o C suggest the prime epithermal potential has been
eroded.
Variations in the basal spacing of clay minerals, which dominate alteration in the
low-sulfidation 5 environment (Table 5) are among the best indicators of
paleotemperature. With increasing temperature smectite (stable at 220o C (Reyes, 1990).
This progression in thermal stability commonly results in a clear upward and outward
zonation of minerals from low-sulfidation ore bodies (Table 6). The ore zone contains
minerals indicating the highest pH (adularia and calcite), as boiling in the conduits causes
CO2 loss and consequent increase in pH,though these minerals are relatively temperature
insensitive (Fig. 2). Other temperature-sensitive minerals include zeolites (most stable at
<220oC, except for wairakite) and Ca-silicates such as epidote (stable above 200-240oC);
hydrothermal biotite andamphiboles form at temperatures above about 280oC, near the
base of the epithermal environment.
Figure 2: Temperature stability of hydrothermal minerals common in the epithermal
environment (from Henley and Ellis, 1983; Reyes, 1990; E. Izawa and M. Aoki, personal
communication, 1994). Some workers disagree on the absolute temperature of first
appearance of some minerals, and the temperature of transition from one clay to another,
but the relative stability is similar in geothermal systems throughout the world. It is
useful to identify zones of mineral assemblages, as these may be more meaningful in
indicating paleotemperatures than single minerals (Reyes, 1990). In low-sulfidation
systems the principal gangue minerals are quartz, calcite, and adularia; in high-sulfidation
systems the principal gangue mineral is quartz.
Izawa et al. (1990) deduced the paleotemperature-sensitive suite of hydrothermal
minerals that are characteristic of the Hishikari district, Japan, and based on these
alteration minerals the paleoisotherms were mapped using surface and drillhole samples.
The presence of primary cristobalite or tridymite (from high-temperature devitrification
of volcanic glass) indicates the least altered rocks affected by the lowest temperature
hydrothermal fluid. Progressively higher temperature zones were mapped based on the
presence of kaolinite, smectite, interstratified clays, and finally (beneath the present
surface) chlorite abundance exceeding that of interstratified clays. These zones form an
elongate halo centered on the vein 6 system, and in cross section drape over the vein
system, the pattern expected for hydrothermal fluids ascending along fractures. In
contrast to neutral-pH alteration, minerals such as kaolinite, dickite, pyrophyllite,
diaspore and alunite are stable under acidic conditions (Hemley et al., 1969, 1980; Reyes,
1990), and several of these minerals are also temperature sensitive (Fig. 2). Pyrophyllite
may form at a temperature 200oC. Zunyite, topaz and andalusite also indicate acidic
conditions and high temperatures of alteration, >260oC. These minerals comprise the
advanced argillic alteration assemblage formed by high-temperature acidic fluids in high-
sulfidation deposits. The host rock most commonly consists of leached silicic alteration
(Meyer and Hemley, 1967), with advanced argillic alteration giving way outward to
argillic alteration halos (Steven and Ratte, 1960); the illite or smectite becomes stable as
the acidic water is progressively neutralized by reaction with the host rock away from the
conduit (Figure 6).

Figure 2. Temperature stability of hydrothermal minerals common in the epithermal


environment (from Henley and Ellis, 1983; Reyes, 1990; E. Izawa and M. Aoki, personal
communication, 1994). Some workers disagree on the absolute temperature of first appearance of
some minerals, and the temperature of transition from one clay to another, but the relative
stability is similar in geothermal systems throughout the world. It is useful to identify zones of
mineral assemblages, as these may be more meaningful in indicating paleotemperatures than
single minerals (Reyes, 1990). In low-sulfidation systems the principal gangue minerals are
quartz, calcite, and adularia; in high-sulfidation systems the principal gangue mineral is quartz.
Figure 3. Distribution of hydrothermal alteration associated with high and low
sulfidation deposits. The alteration mineralogy varies both vertically and laterally.
Quartz is stable in all areas. Propylitic alteration occurs in regions of low water:rock
ratios, i.e., outside conduit zones, and its mineralogy is controlled by rock composition.
Typical minerals include albite, calcite, chlorite, epidote and pyrite. Steam-heated
overprint can occur in either low or high sulfidation environments; though in the latter,
the hypo gene and steam-heated mineralogies are similar. The effects of the steam-heated
overprint are most apparent in the low sulfidation environment, as the alteration
mineralogy is markedly different from that produced by hypogene fluids. In the low-
sulfidation environment. Silica sinters may form where the neutral-pH thermal water
discharges at the surface; in areas of high relief, however, this hot water may flow
laterally a great distance before reaching the surface, becoming entrained in ground water
and thus not liable to precipitate silica sinter. In areas of little erosion and low relief,
veins may terminate at the paleo surface in sinter (e.g., McLaughlin, USA); by contrast,
in areas of high paleo relief, the veins may pinch out upwards, perhaps into argillic-
altered rock, deep below the surface (e.g., Hishikari, Japan). In the high-sulfidation
environment, ore mineralization (if present) occupies part of the intensely acid-altered
core of the alteration system. This core is typically very irregular in shape, but commonly
pinches out above and below the main ore body.

III. Type of epithermal sulfidation


1. High Sulfidation
High sulfidation (HS) gold systems are widely distributed in volcanic arcs
worldwide. Deposits range from structurally controlled and deeper seated examples, such
as El Indio, to shallow host-rock or breccia controlled examples such as Yanacocha,
Pierina, and Pueblo Viejo (Figure 4; Sillitoe, 1999). At the regional scale, HS systems lie
within calc-alkaline volcanic arcs dominated by andesitic volcanism. They form in the
upper parts of Cu (Au, Mo) porphyry systems, which themselves do not necessarily
contain economic mineralization. The giant HS deposits of northern Peru and the central
Andes of Argentina and Chile are all Mid- to Upper-Miocene in age, and are inferred to
have formed above flat or flattening subduction zones, and temporally coincident with
compression and shortening in the upper crust. As with porphyry systems, the giant HS
systems appear to be localized at intersections of arcparallel with arc-transverse crustal-
scale structures. Locally, giant HS systems are associated with felsic subvolcanic or
volcanic rocks, often within igneous centers showing prolonged activity. They can be
hosted either in volcanic rocks, as at Yanacocha and Pierina, or in their basement, as at
Veladero, Pascua-Lama, and Alto Chicama, the latter case reflecting compression-driven
uplift. The HS deposits lie within large volumes of advanced argillic alteration formed
through the mixing of acidic magmatic vapors and groundwater above mineralized
porphyry intrusives (Hedenquist et al., 1998). Typically these advanced argillic alteration
zones show characteristic zoning from proximal vuggy silica through advanced argillic
assemblages containing alunite, pyrophyllite, dickite and kaolinite to distal argillic
alteration. The central siliceous alteration zones are the main hosts to ore. The nature of
the host rock can produce variations from these typical alteration assemblages and zoning
patterns.

Figure 4: Schematic model of a dome-related HS system above an underlying parent porphyry


system. Alteration and Cu sulfide mineral assemblages vary with depth below the paleosurface,
which is marked by acid-leached rock of steam-heated origin. Adapted from Sillitoe (1999).

Mineralization in HS deposits comprises pyrite-rich sulfide assemblages including


high sulfidation-state minerals like enargite, luzonite and covellite. Mineralization post-
dates the formation of the advanced argillic lithocap described above. The mineralizing
fluid is much less acid than the fluid responsible for forming the advanced argillic
alteration zones which host mineralization (Jannas et al., 1990; Arribas, 1995).
Fluctuations from enargite to tetrahedrite-tennantite are a common feature during the
evolution of HS deposits and indicate changes in sulfidation state and pH of the
mineralizing fluid during the life of the hydrothermal system (Sillitoe and Hedenquist,
2003, Einaudi et al., 2003). Minor gold can occur with early enargite mineralization but
most gold is introduced with paragenetically later tennantite-tetrahedrite-low Fe
sphalerite mineralizing events (Einaudi et al., 2003). The giant systems comprise
disseminated Au-Ag mineralization often in mushroom-shaped ore bodies with narrower
structural roots (Figure 4). Permeability contrasts between aquitards and permeable
lithologies can be important controls on the distribution of gold. Additionally, breccias
are usually abundant and host ore in some systems. Phreatomagmatic breccias are present
in all giant HS deposits underlining the genetic connection with an underlying intrusive.
Mineralization can occur over vertical intervals of 100s of meters below the paleosurface,
from disseminated Au -Ag immediately below surficial steam-heated alteration to deeper
structurally controlled Au-enargite at depth. Supergene oxidation, often to considerable
depths in permeable silicified rock, generates oxide gold mineralization amenable to
recovery by cyanide leaching.
2. Intermediate Sulfidation
Intermediate sulfidation (IS) gold systems also occur in mainly in volcanic
sequences of andesite to dacite composition within calc-alkaline volcanic arcs. Large IS
Au deposits are found in compressional as well as in extensional magmatic arcs. Some
Au-rich IS systems are spatially associated with porphyry systems (e.g. Rosia Montana,
Baguio) while others adjoin coeval HS systems (Victoria, Chiufen-Wutanshan).
Additionally, several larger Au-rich IS deposits are associated with diatremes which
further emphasizes a magmatic connection. At the deposit scale, mineralization occurs in
veins, stockworks and breccias. Veins with quartz, manganiferous carbonates and
adularia typically host the Au mineralization. Gold is present as native metal and as
tellurides together with a variety of base metal sulfides and sulfosalts. Low-Fe sphalerite,
tetrahedrite-tennantite and galena often dominate these assemblages. IS Au veins can
show classical banded crustiform,colloform textures in the veins. Permeable lithologies
within the host sequence may allow development of large tonnages of lowgrade
stockwork mineralization. Alteration minerals in IS Au deposits are zoned from quartz ±
carbonate ± adularia ± illite proximal to mineralization through illite-smectite to distal
propylitic alteration (Simmons et al., 2005). Breccias may be common and can show
evidence for repeated brecciation events.

3. Low Sulfidation

Low sulfidation (LS) epithermal gold deposits of the alkalic and subalkalic
subtypes share a number of characteristics (Table 1) and are described together. Differing
characteristics of the less common alkalic LS deposits are highlighted where appropriate.
Most LS gold deposits are found in intra-arc or back-arc rifts within continental or island
arcs with bimodal volcanism (Table 1). Rifts may form during or after subduction or in
post collisional settings. Additionally, some LS deposits are found in andesite-dacite-
rhyolite volcanic arcs, but only in clearly extensional settings (Sillitoe and Hedenquist,
2003). Deposits of the alkalic subset of low sulfidation epithermal deposits are
specifically associated with alkaline magmatic belts but share an extensional setting with
their calk-alkaline counterparts (Table 1; Jensen and Barton, 2000). At the deposit scale,
LS gold deposits are typically hosted in volcanic units, but can also be hosted by their
basement. Vein development in the basement does not reflect syn-mineral uplift, as is the
case in HS and IS systems, but rather the intersection of the hydrothermal system with
rheologically more favorable basement host rocks. Syn-mineral mafic dykes are common
in these deposits (Sillitoe and Hedenquist, 2003). Both low-grade disseminated and
structurally controlled high-grade deposits can form, such as Round Mountain and
Hishikari, respectively (Figure 5). Calc-alkalic LS deposits have restricted vertical
continuity, generally < 300whereas alkalic LS deposits such as Porgera and Cripple
Creek can extend in excess of 1 km vertically. Mineralization in subalkalic LS systems
generally has high silver (Au:Ag ratio<1), and low base metal content and gold is
associated with pyrite – high-Fe sphalerite ± pyrrhotite ± arsenopyrite. In contrast, LS
alkalic mineralization commonly contains abundant telluride minerals, has elevated
Au:Ag ratios, and less voluminous quartz gangue (Jensen and Barton, 2000). Alteration
mineralogy in LS systems shows lateral zoning from proximal quartz-chalcedony–
adularia in mineralized veins, which commonly display crustiform-colloform banding
and platy, lattice-textured quartz indicative of boiling; through illitepyrite to distal
propylitic alteration assemblages (Figure 5). Vertical zoning in clay minerals from
shallow, low temperature kaolinite-smectite assemblages to deeper, higher temperature
illite have also been described (Simmons et al., 2005). As with HS and IS systems, host
rock composition can also cause variations in the alteration mineral zoning pattern in LS
systems. Alteration assemblages in alkalic LS deposits commonly contain roscoelite, a V-
rich white mica, and abundant carbonate minerals (Jensen and Barton, 2000).

Figure 5: Schematic section showing typical alteration and mineralization patterns in a low
sulfidation system. Modified from Hedenquist et al. (2000).

By definition, epithermal systems form close to the paleosurface and therefore each
of the systems described above may lie beneath steam-heated alteration blankets formed
above the paleowater table (Figure 5). As the name indicates, this alteration is formed by
the acidification of cool meteoric waters by acidic vapors derived from boiling, ascending
hydrothermal fluids. Steam-heated alteration typically comprises fine-grained, powdery
cristobalite, alunite and kaolinite and has a morphology which mimics the
paleotopography. Massive opaline silica layers mark the water table. Siliceous sinters can
also form, marking outflow zones where the paleowater table intersected topography, but
sinters will only form above or lateral to LS systems where the upwelling fluid has near-
neutral pH (Simmons et al., 2005).
IV. Exploration guide
The last decade has seen a significant decline both in the number of major gold deposits
discovered (>2.5 Moz Au) and the amount of gold contained in these deposits, compared
to the early to mid ‘90s (Metals Economics Group, 2006). Of the 44 major gold
discoveries in the last decade, 32 were made during 1996-2000 and only 12 during 2001-
2006. Of these 44 major gold discoveries, 31 were attributable to greenfield exploration
and only 13 to brownfield exploration, but the near mine discoveries have not declined at
the same rate as greenfield discoveries. Such data attest to the continued value of regional
exploration and to the importance of near-mine exploration in the strategy of any mid- to
large-size gold producer. In addition to declining discovery rates, future success will have
to be achieved in a context of increasing costs, increasing pressure for yearly replacement
of reserves/ resources, and increasing minimum size of deposits that really impact the
bottom line in larger companies. A review of the main discovery methods of gold
deposits found in the last 10 years indicates that geological understanding was the key
element in the discovery process in both the greenfield and brownfield environments (e.g.
Sillitoe and Thompson, 2006). Geochemistry in support of geology plays an important
role in cases particularly where deposits are exposed, and geophysics aided discovery in
some cases where the discoveries were concealed (Sillitoe and Thompson, 2006). A clear
lesson from this analysis is that geology should remain an important underpinning of
future gold exploration programs. Accordingly, a key element of success for the
explorationist will be to understand and detect the different types of gold deposit and
their favorable geologic settings and controls at the regional to local scales, and
increasingly in covered terrains. So is an understanding of erosion levels relative to depth
of formation of the explored systems, of the environments where Robert, F., et al. Models
and Exploration methods for Major Gold Deposit Types 703 they can be best preserved.
Another element will be the wise application of proven and emerging detection
techniques, in close integration with geology. The successful strategy should emphasize
as much the detection of geological features typical of favorable settings, as that of
hydrothermal manifestations of deposits, such as alteration and mineralization, and their
dispersion products in the surficial environment. In addition, the exploration approach
also needs to take the unusual and unique into account so that deposits which do not
conform closely to current models or that occur in unusual settings, are not overlooked
(e.g. Sillitoe 2000b). Exploration is now supported by a variety of advanced data
integration and processing tools, from advanced 2D GIS platforms, which have the ability
to display drilling data, to full blown 3D data modeling, processing and visualization
packages. 3D packages are more suited to near mine or data rich environments, whereas
2D GIS platforms have become an essential tool in regional exploration. However, any
approach should focus on detecting a footprint, or elements of a footprint of a
mineralized system, at both regional and local scales. Finally, people are the backbone of
any good exploration approach. Not only do team members need to have ability and
experience, they must also understand the characteristics of the gold deposits they are
searching for and be given sufficient field time to adequately test their targets. Other
factors such as an excellent understanding of proven exploration methods, effective use
of technology, enthusiastic and responsible leadership, confidence in corporate direction,
and attracting and training young professionals, are also important.
REFERENCE
[23] Hedenquist, J.W. and Lowenstern, J.B. (1994) The Role of Magmas in the
Formation of Hydrothermal Ore Deposits. Nature, 370, 519-527.
https://doi.org/10.1038/370519a0
[24] White, N.C. and Hedenquist, J.W. (1995) Epithermal Gold Deposits: Styles,
Characteristics and Exploration. SEG Newsletter, 23, 9-13.
[25] Hedenquist, J.W., Izawa, E., Arribas, A. and White, N.C. (1996) Epithermal Gold
Deposits: Styles, Characteristics, and Exploration. Society of Resource Geology
Ressource Geology Special Publication No. 1, 17 p.
[26] Simmons, S.F. (1995) Magmatic Contribution to Low-Sulfidation Epithermal
Deposits. Mineralogical Association of Canada Short Course Series: Magmatic, Fluids,
and Ore Deposits, 23, 455-477.
[27] Arribas, A. (1995) Characteristics of High-Sulfidation Epithermal Deposits, and
Their Relation to Magmatic Fluids. Mineralogical Association of Canada Short
Course Series: Magmatic, Fluids, and Ore Deposits, 23, 419-454.
[28] Henley, R.W. and Berger, B.R. (2011) Magmatic-Vapor Expansion and the
Formation of High-Sulfidation Gold Deposits: Chemical Controls on Alteration and
Mineralization. Ore Geology Reviews, 39, 63-74.
https://doi.org/10.1016/j.oregeorev.2010.11.003
researchgate.net/figure/Schematic-model-of-a-dome-related-HS-system-above-an-
underlying-parent-porphyry-system_fig4_284574136
https://www.researchgate.net/publication/
268274871_Epithermal_gold_deposits_Styles_characteristics_and_exploration
https://link.springer.com/chapter/10.1007/3-540-27946-6_148

You might also like