Sei sulla pagina 1di 25

Miner Deposita (2014) 49:235–259

DOI 10.1007/s00126-013-0485-0

ARTICLE

Alteration patterns and structural controls of the El Espino


IOCG mining district, Chile
G. P. Lopez & M. W. Hitzman & E. P. Nelson

Received: 8 February 2012 / Accepted: 19 August 2013 / Published online: 14 September 2013
# Springer-Verlag Berlin Heidelberg 2013

Abstract The El Espino IOCG mining district is character- and 34 wt% NaCleq and temperature of approximately 425 °C
ized by several mineralized bodies the largest of which is the at an estimated depth of 3–4 km. Geochronological U–Pb and
40
El Espino deposit, which has an estimated geologic resource Ar/39Ar data indicate that hydrothermal alteration was coe-
of 123 Mt at 0.66 % Cu and 0.24 g/t Au. Mineralized bodies val with magmatic intrusive activity. One particular dioritic
are distributed in a 7×10 km2 area throughout a 1,000-m intrusion (88.5 Ma) preceded the calcic stage (88.4 Ma),
vertical section. They range from single veins to stockworks which was accompanied by iron oxide copper and gold min-
and breccias to manto-type deposits. The ore bodies are hosted eralization. Hydrolytic alteration, related to economic iron
primarily by volcanic, volcaniclastic, and sedimentary rocks oxide copper and gold mineralization, came immediately after
of the Early Cretaceous Arqueros and Quebrada Marquesa at 87.9 Ma.
formations, with a few mineralized zones within Late Creta-
ceous dioritic intrusions. The fault and vein architecture shows
that El Espino IOCG system was localized within a dilatation- Introduction
al jog along a major transtensional dextral fault system. Sodic
alteration (albite) is the most extensive style of alteration in the Iron oxide copper gold (IOCG) deposits are magnetite- and/or
district, and it is bounded by major NS–NNE trending faults. hematite-rich hydrothermal systems with economic copper
Sodic–calcic (epidote–albite) alteration occurs at deep to me- and gold. Their classification and genesis are still contentious
dium elevations (1,000–500 m) and grades inward into calcic (Haynes 2000; Hitzman 2000; Barton and Johnson 2000;
alteration. Calcic alteration surrounds dioritic intrusions of the Sillitoe 2003; Williams et al. 2005; Pollard 2006; Chiaradia
Llahuin plutonic suite. Significant iron oxides are associated et al. 2006; Groves et al. 2010). Andean-type IOCG deposits,
with later calcic alteration associations (actinolite–epidote– which are the youngest major province of the IOCG class,
hematite). The upper portions of the alteration system formed in an extensional to transtensional tectonic regime
(0–500 m) display hydrolytic alteration associations with within the Jurassic–Cretaceous magmatic arc along the con-
abundant hematite. Hydrolytic veins are feeders to zones of vergent margin of the Central Andes—from southern Peru to
manto-type alteration and mineralization within favorable north central Chile (Sillitoe 2003). Andean IOCG deposits
volcano-sedimentary lithologies that formed El Espino display variable morphologies and alteration styles and have a
deposit. Sulfides are largely confined to calcic and hydrolytic variety of relationships to possible causative intrusive rocks.
alteration associations. Hydrothermal fluids responsible for The El Espino deposit located at 31°23′ S in the Coastal Range
hematite and sulfide mineralization had salinities between 32 of Northern Chile is one of the southernmost known IOCG
systems in Chile (Fig. 1). Currently, the El Espino deposit has
a measured and indicated total resource of 145 Mt at 0.55 %
Cu and 0.22 g/t Au (Explorator Resources 2011). The El
Editorial handling: R. P. Xavier Espino deposit is located at the west margin of a much larger
mining district (El Espino mining district) that contains a
G. P. Lopez (*) : M. W. Hitzman : E. P. Nelson
number of small vein, stockwork, manto, and irregularly
Department of Geology and Geological Engineering,
Colorado School of Mines, Golden, CO 80401, USA shaped iron oxide-bearing copper–gold deposits within a
e-mail: gplopezo@gmail.com mixed sequence of Mesozoic volcanic and sedimentary rocks
236 Miner Deposita (2014) 49:235–259

Methods and materials

Project goals were achieved through a combination of detailed


field mapping and laboratory studies. Geological, structural,
and alteration mapping were done throughout the 70-km2
study area at 1:20,000 scale during 2004 and 2005. Additional
information was gathered from mapping pits and underground
adits and from logging of drill core from the El Espino
deposit (Fig. 2). Field mapping and logging concentrated
on establishing the time–space relationships and structural
and stratigraphic controls of alteration and mineralization.
Samples collected during the fieldwork were examined
petrographically at Colorado School of Mines using trans-
mitted and reflected light microscopy and with scanning
electron microscopy (SEM) and a PGT Spirit analyzing
system attached to the SEM for X-ray energy dispersive
spectrometry. Detailed petrography was conducted on
samples for fluid inclusion analyses using the methodol-
ogy of Goldstein and Reynolds (1994) at Colorado School
of Mines.
Major and trace element geochemical analysis were
performed on relatively unaltered samples of volcanic rocks
(7) and intrusive (13) rocks, together with hydrothermally
altered and mineralized samples from throughout the district
to investigate chemical changes associated with hydrothermal
alteration. The samples were crushed and pulverized at ALS
Chemex, La Serena, Chile, and then were sent to ALS
Chemex, Vancouver, Canada, for geochemical analyses.
Whole rock analysis involved lithium metaborate/lithium
tetraborate (LiBO2/Li2B4O7) fusion and analysis by inductive-
ly coupled plasma (ICP) atomic emission spectroscopy. Trace
element analysis was conducted by a combination of ICP
mass spectrometry and ICP atomic absorption spectroscopy
after four acid near total digestions.
Fluid inclusions were analyzed on three double-sided
polished thick sections (50 μm) from the Espino and Romero
areas at Colorado School of Mines and Fluidinc, Denver. The
analyses were performed using a US Geological Survey-style
Fig. 1 Location of El Espino deposit in relation of other Chilean IOCG gas-flow heating/freezing stage mounted on an Olympus mi-
deposits. Two NS trending belts of ore deposits overlap along the Chilean
croscope equipped with a 40× objective (N.A.=0.55) and 10×
Iron Belt: magnetite–apatite deposits and IOCG deposits
oculars. The heating and freezing stages were calibrated at
374.1 and 0.0 °C using synthetic pure H2O fluid inclusions
and at −56.6 °C using synthetic CO2 fluid inclusions. Thermal
cycling was used to bracket the liquid–vapor homogenization
cut by intermediate composition intrusions. The El Espino temperatures (T h) and melting temperatures (T m) of the fluid
mining district has a vertical relief of approximately 1,000 m inclusion assemblages.
allowing investigation of alteration and mineralization styles Sulfur isotope analysis was performed by G. Lopez at
through a significant vertical section. This paper presents Colorado School of Mines on pyrite (13), chalcopyrite (22),
information on lateral and vertical zonation of alteration and and gypsum (1) separates derived from micro drilling of core
mineralization in the district, the structural setting of the and rock samples across the district. Approximately 25–
mineralization, the relative ages of igneous and hydrothermal 100 μg (weight dependent on the mineral analyzed) of sample
events, and the chemistry of the hydrothermal fluids inferred were combusted in a Eurovector 3000 elemental analyzer at
from petrographic and isotopic investigations. Colorado School of Mines, yielding sulfur dioxide that was
Miner Deposita (2014) 49:235–259 237

Fig. 2 Geological framework and location of Espino district. Regional map is simplified from the 1:1,000,000 scale Chilean Geological Survey map
(Servicio Nacional de Geología y Minería, 2002). Rectangle shows location of Fig. 3

delivered to a Isoprime mass spectrometer using continuous- outer arcs developed in the present Coastal Cordillera and in
flow techniques, with helium as the carrier gas. Repeat anal- the Chile–Argentina border, respectively (Mpodozis and
yses of a lab working standard (Colorado School of Mines Ramos 1990). The El Espino deposit is located at the eastern
Barium Sulfate) yield a precision of 0.2 ‰. The isotopic data margin of this inner magmatic arc.
are reported using the δ notation in units of per mile, relative The Early Cretaceous times were marked by an extensional
to the Cañón Diablo Troilite standard. regime, marine transgression, decreased plutonism in the Coast-
al Cordillera, and extensive subaerial volcanism—with marine
intercalations—associated with intra-arc rifting (Åberg et al.
Tectonic and geological setting 1984; Vergara et al. 1995). The volcanic rocks are high-K,
calc-alkaline to shoshonitic basaltic andesite to andesite. Stron-
The segment between 27° and 33° S in Chile is characterized tium and neodymium isotopic studies suggest that the volcanic
by a Mesozoic–Cenozoic magmatic arc built mostly over a rocks and coeval intrusive rocks were derived from depleted
Late Paleozoic to Triassic accrecionary prism and arc along a mafic magmas and metasomatized subducted sediments with
subducting margin. The Mesozoic Andean evolution was inclusion of partially melted Jurassic plutonic rocks (Morata
characterized by a Jurassic–Early Cretaceous magmatic arc and Aguirre 2003). These rocks underwent high subsidence and
accompanied by a sedimentary back-arc basin and an aborted low-grade burial metamorphism concurrent with extension-
marginal basin followed by a Late Cretaceous magmatic arc related plutonism (Levi and Aguirre 1981; Parada et al. 2005).
with a fold and thrust belt to the east (Fig. 2). The Jurassic– At approximately 100 Ma, the tectonic regime shifted from
Early Cretaceous magmatic arc comprised coeval inner and extensional to compressional between 32 and 33° lat. S,
238 Miner Deposita (2014) 49:235–259

resulting in crustal shortening, basin closures, eastward mi- Cretaceous intrusions are exposed over an area of approx-
gration of magmatism, and uplift (Arancibia 2004; Parada imately 3,000 km2 along the Coastal Range between 31° and
et al. 2005). The change from extensional to compressional 32° S (Fig. 2). This area contains three large plutonic bodies
tectonics has been attributed by Arancibia (2004) to higher (Fig. 3) and a number of smaller bodies of similar composition
spreading rates in both the southeast Pacific and the south (Rivano and Sepulveda 1991). The plutons include diorite,
Atlantic (Wilson 1992). Mid-Cretaceous uplift of the Coastal tonalite, amphibole and pyroxene granodiorite, and amphibole
Range in this portion of Chile has been documented by apatite monzodiorite. Potassium–argon geochronology available in-
fission track data from Jurassic and Early Cretaceous plutonic dicates that the Quilatapia–El Durazno pluton (Fig. 3) ranges
rocks that cooled to approximately 80–125 °C by 110–90 Ma in age from 134 to 108 Ma, whereas the Illapel–Caimanes
(Gana and Zentilli 2000; Parada et al. 1999). pluton (Fig. 3) ages range from 130 to 86 Ma and young to the
The Early Cretaceous Arqueros and Quebrada Marquesa east (Rivano and Sepulveda 1991).
formations were deposited in the Coastal Range region between
30° and 32° S. The Neocomian Arqueros Formation consists of
lava, volcanic breccia, tuff, and agglomerate with lenticular El Espino district geology
intercalations of conglomerate, sandstone, and locally thin fos-
siliferous limestone. It has an estimated thickness of 3,500– The El Espino mining district contains the Cretaceous volcano-
4,000 m and is interpreted to have formed near a continental sedimentary Arqueros, Quebrada Marquesa, and Salamanca
margin with associated volcanic activity and recurrent marine formations. Only the upper portion of the Arqueros Formation
transgressions (Aguirre and Egert 1965). The Barremian to and the basal portion of the Salamanca Formation are exposed
Albian Quebrada Marquesa Formation consists of mixed in the area (Fig. 3).
volcanic and sedimentary rocks (Aguirre and Egert 1965). The These formations are intruded by a series of Late Cretaceous
formation was subdivided in two members: the lower Espino intermediate stocks of granodioritic to dioritic composition
Member and the upper Quelen Member (Rivano and Sepulveda (Llahuin pluton in the north-central area and Illapel–Caimanes
1991). The Espino Member consists of limestone, siltstone, pluton to the south; Fig. 3).
sandstone, and conglomerate with local gypsum lenses. Vertical
and lateral facies variations of the Espino Member made it Stratigraphy
difficult to define a representative stratigraphic column. The
Espino member is interpreted to have been deposited in The Arqueros Formation in the El Espino district has a min-
relatively small marine to transitional basins with sea level imum thickness of 1,500 m and consists of gray-greenish
variations (Rivano and Sepulveda 1991). The Quelen Member, andesitic lava, andesitic breccia, and andesitic lapilli tuff with
with estimated thickness of 1,200 m, consists of red colored lenses up to 100 m thick of andesitic conglomerate, sandstone,
volcaniclastic and sedimentary rocks including lava, pyroclastic siltstone, and limestone (Fig. 4). The unit displays a high
breccia, and volcanically derived sandstone and conglomerate. degree of lateral facies variability.
The Quelen member represents deposition in a continental The Quebrada Marquesa Formation (Figs. 3 and 4) dis-
environment accompanied by intense volcanism. plays a conformable contact with the underlying Arqueros
The Early Cretaceous Espino and Quelen members in the Formation in the southern portion of the El Espino district
El Espino district are covered by the Late Cretaceous Sala- but is in fault contact with the Arqueros Formation in the
manca Formation (initially defined as the Viñita Formation by north. The Quebrada Marquesa Formation consists of a mixed
Aguirre and Egert 1965, but later included into Salamanca sequence of volcanic and sedimentary rocks.
Formation by Rivano and Sepulveda 1991). The Salamanca The Espino Member of the Quebrada Marquesa Formation
Formation is preserved from the Coastal Range up to the Main is composed of continental to marginal marine volcaniclastic
Cordillera and ranges in thickness from approximately 3,500– and sedimentary rocks. The marine sedimentary rocks of
4,000 m. The formation is subdivided into two units: lower Espino member were deposited in a relatively small, structur-
Santa Virginia Member and upper Rio Manque Member ally controlled basin that is approximately 13 km long in a
(Rivano and Sepulveda 1991). The Santa Virginia Member north–south direction and approximately 5 km wide. The base
consists of conglomerate and red, hematitic sandstone with of the Espino Member consists of a 5–10 m thick, well-
minor siltstone and lacustrine limestone intercalations. These stratified red andesitic conglomerate that grades upward into
sedimentary rocks are interpreted to have been deposited as red graywacke beds, then siltstone, and finally limestone. This
alluvial fans and debris flows with finer-grained sediments basal unit is overlain by two additional, normally graded
deposited in the basin center. The overlying Rio Manque sequences each with a base of sandstone that grades upward
Member consists of tuff and volcanic breccia; it represents to siltstone, limestone, and local gypsum beds. Graywacke is
deposition during a period of intense and explosive latest dominantly gray in color and is composed of feldspar, quartz,
Cretaceous volcanism (Rivano and Sepulveda 1991). and lithic clasts of andesite. Sandstone is yellow to pink and
Miner Deposita (2014) 49:235–259 239

Fig. 3 Geological map of El Espino mining district, showing the location of the three mineralized areas that were examined in detail: Espino, Romero,
and Llahuin

fine to coarse grained. There is complete gradation from with pyroxene and hornblende phenocrysts, lapilli tuff, and
volcanic-derived arkosic graywacke to feldspathic sandstone. breccia. Major element chemistry of volcanic rocks indi-
Siltstone beds range from dark gray to yellow with the former cates these rocks are medium- to high-K basaltic andesite
being locally organic rich with relatively abundant diagenetic to andesite (Fig. 5). The uppermost sequence in the Espino
pyrite. Limestone is thick bedded and commonly contains Member consists of inversely graded sandstone beds, with
chert nodules. The largest gypsum bed in the district is subsidiary limestone and siltstone. Sandstone is yellow to
100 m thick and has been mined over a strike length of gray and fine to medium grained. Limestone beds are a few
800 m (San Enrique mine). In the central portion of the meters thick and commonly contain chert nodules. The
district, the sedimentary rocks within these sequences thickness of the Espino Member is estimated to be approxi-
are intercalated with volcanic lavas, pyroclastic beds, and mately 900–1,000 m in the southern portion of the district and
volcaniclastic sedimentary rocks. Volcanic and pyroclastic approximately 600 m in the northwest where it is truncated by
rocks include gray to green colored porphyritic lava flows the Llahuin pluton.
240 Miner Deposita (2014) 49:235–259

Fig. 4 Cretaceous stratigraphy in


the El Espino mining district
showing the stratigraphic
intervals of the Espino, Romero,
and Llahuin areas

The Quelen Member of the Quebrada Marquesa Formation


is characterized by several centimeter thick medium- to
coarse-grained sandstone base that grades upward to a region-
ally extensive coarse conglomerate with clasts of well-
rounded porphyritic andesite, andesitic breccia, and phaneritic
dioritic intrusive rocks. The conglomerate is overlain by a
thick greenish lapilli tuff of the Late Cretaceous Salamanca
Formation.
The Late Cretaceous Salamanca Formation (Figs. 3 and 4)
is in fault contact with the Quebrada Marquesa Formation in
the El Espino mining district, although a few outcrops at
higher elevations show an erosional unconformity between
red conglomeratic sandstone of the Quelen Member and
Fig. 5 Plot of SiO2 vs. K2O (Tatsumi and Eggins 1995) showing
lapilli tuff of the Santa Virginia member of Salamanca chemical compositions of volcanic rocks in the Arqueros and
Formation. In the study area, the Salamanca Formation Quebrada formations. Rocks are classified as medium- to high-potassium
contains a basal unit that comprises volcanic breccia and andesite and basaltic andesite
Miner Deposita (2014) 49:235–259 241

locally gray thin-bedded siltstone. Higher in the sequence, Structure


this formation contains red to gray polymictic conglomerate
and red sandstone. A fault architecture model of the El Espino mining district was
developed by mapping topographic lineaments on Landsat
and Aster imagery that were then field checked during geo-
Intrusive rocks logical mapping (Fig. 6). Fault orientation data from outcrop
show two principal strike sets: NNW–NS, and NE (Fig. 7).
Approximately 30 % of the El Espino mining district is Fault dips are mostly >45° (Fig. 7d). Fault cores are mostly
underlain by intrusive rocks (Fig. 2). The northern portion of about 0.5–1 m thick, but ranging up to 20 m thick and consist
the study area exposes the large, multiphase Llahuin pluton. of matrix-supported breccia, gouge, and locally a penetrative
The central portion of the area exposes a number of isolated fracture cleavage. Damage zones of fractured and locally
diorite plutons, and the southern portion of the area exposes folded rocks can be up to 150 m thick. Analysis of slickenline
the large multiphase Illapel–Caimanes pluton (Fig. 3). rake (Fig. 7d) shows that, although the majority of faults are
The Llahuin pluton (or Llahuin plutonic suite) comprises at strike-slip faults, oblique-slip and dip-slip fault also are pres-
least nine individual intrusions, which are distinguished by ent. Multiple slickenlines on some major faults indicate a
texture and mineralogy, and geographical location. Few protracted faulting history. This is supported by the presence
intrusive contacts between intrusions were observed due of hydrothermal alteration (specular hematite) and copper
to soil, debris, and/or vegetation cover, so some textural mines along NNW-strike faults that were probably active
and mineralogical variations observed in the district might previously during basin formation.
be due to zonation within a single stock. Some of the dioritic The Quebrada Marquesa formation has a general NS trend
intrusions have caused local contact metamorphism of adjacent and is bounded to the west by the Espino–Illapel fault that
volcano-sedimentary rocks. Biotite-pyroxene hornfels occurs extends for ∼20 km from near the city of Illapel to the El
adjacent to sills in drill cores in the El Espino deposit. Locally Espino area. Although the fault trends generally N to NNE,
andesitic breccias in the Llahuin area display a meter-wide about 8 km south of the area this structure changes trend to
metamorphic aureole composed of fine-grained recrystallized NNW and becomes a set of right-stepping en echelon fault
quartz with lesser magnetite and epidote adjacent to a quartz segments along the west boundary of the basin (Fig. 3). With-
monzonite intrusion. in the basin, strata generally have low homoclinal dip to the
Three intrusions present in the El Espino deposit area NE, but are broadly folded in the southern portion. Boudinage
include a hornblende-bearing porphyritic dioritic stock and in limestone and fine-grained sandstone units of the El Espino
sills; an hornblende-bearing, fine-grained monzodiorite; and member suggests bedding-parallel extension during diagenet-
an equigranular, fine-grained hornblende and pyroxene- ic compaction or during subsequent folding. Folds have EW
bearing diorite. A sharp contact was observed between the trend in the east and NE trend in the west (Fig. 6). Near the
amphibole-bearing monzodiorite and the hornblende and southern edge of the basin, a broad open NE plunging anti-
pyroxene-bearing diorite. A medium- to coarse-grained, cline is present and strata dip to the E and SE. In addition, dips
equigranular quartz-dioritic to tonalitic intrusion is exposed are locally steep where strata are drag folded within about
at higher elevations north of the El Espino deposit. East of 10 m of some major faults. A π-analysis, which modeled the
the El Espino deposit at low to medium elevations, a average axis of rotation of bedding in the Quebrada Marquesa
hornblende and pyroxene-bearing microdiorite, NS-strike Formation (in this case caused by fault-block rotation), sug-
gray dioritic dikes, and a NS-trending tabular body of leuco- gests that the eastern margin of the basin may be bounded by a
monzodiorite are present. Dikes display a sharp contact with NNW-strike fault (∼350°) (Fig. 7a). A fault with this approx-
the leuco-monzodiorite. Northeast of the El Espino deposit an imate strike, although not the basin margin fault, is present
equigranular, medium-grained, olivine and pyroxene-bearing east of the Romero area (Fig. 6).
gabbro diorite, an equigranular medium-grained quartz mon- The Quebrada Marquesa strata are also cut by a set of NE-
zonite, and a monzodiorite occur in topographically high strike, left-stepping en echelon faults (Fig. 3). Most mineral-
areas. The monzodiorite unit displays a transitional contact ized veins in the district, including those in the El Espino,
between a dark gray idiomorphic medium-grained unit and a Romero, and Llahuin areas, strike NS to ENE, and many are
pink xenomorphic finer-grained unit. related to these NE-strike faults (Fig. 7). Veins are defined
U–Pb geochronology on zircons from a weakly altered here as fractures with infill of more of 5 cm width, and veins
diorite from the Llahuin pluton at the El Espino deposit that display a fault plane or fault zone with gouge and/or
yielded an age of 88.5±1.7 Ma. A quartz diorite from the breccia are termed fault-veins. Several types of veins were
Llahuin area in the northern portion of the district yielded an identified based on their ore and/or gangue mineralogy in the
age of 88.1±1.1 Ma. More detailed descriptions of samples district and include iron oxide sulfide (with or without-quartz)
analyzed, and age results are reported in Lopez (2012). veins, quartz sulfide veins, and calcite- and/or barite sulfide veins.
242 Miner Deposita (2014) 49:235–259

Fig. 6 Structural map of El


Espino mining district showing
faults in blue, lineaments in
black, folds in black, and veins in
red. Blue triangles indicate
inverse fault, dots indicate normal
faults, and blue arrows indicate
sense of movement along faults.
The district is bounded on the
west and east by major fault zones

Iron oxide sulfide veins are the most common followed by quartz dispersion in orientation than iron oxide sulfide veins (Fig. 8).
sulfide veins. Iron oxide sulfide veins range from 5 to 2 m in Veins frequently terminate at contacts with overlying limestone
width, whereas quartz sulfide veins range from 5 cm to 1 m. units where the calcareous rock above is mostly silicified and
In the El Espino deposit area, iron oxide and quartz sulfide the rock beneath shows stratabound metal disseminations, in-
veins occur in the Espino Member and intrusive units dicating that locally iron oxide and sulfide mineralization was
of Llahuin pluton. Orientations are variable but average focused immediately beneath these contacts. The wall rocks
∼240°/75° (right-hand rule); quartz sulfide veins have a wider surrounding iron oxide sulfide veins are commonly altered to
Miner Deposita (2014) 49:235–259 243

calcic and hydrolytic associations depending on their structural Kinematic data indicate that the NNW- and NE-striking
level. The true width of the alteration zones range from 20 cm faults in the district had both dextral and sinistral strike-slip
to 5 m. Wall rocks surrounding quartz sulfide veins are altered and normal and reverse dip-slip motion (Fig. 7c). Fault-vein
to hydrolytic to argillic associations with widths ranging from kinematic data display dextral and sinistral strike-slip compo-
20 cm to 1 m. nents, but all dip-slip components show normal fault motion
In the Llahuin area iron oxide and quartz sulfide veins (Fig. 8c, d). The orientations of the instantaneous principal
occur in the Quelen Member of Quebrada Marquesa Forma- strain axes were modeled from fault kinematic data using the
tion. Veins have an average strike of ∼020°; quartz sulfide Faultkin program (Allmendinger et al. 2001). The results
veins have an average dip of ∼55° (towards the east), while (Fig. 9) show near vertical shortening and NNW–SSE exten-
iron oxide sulfide veins are steeper with average dip of sion for the major faults and near vertical shortening and
∼90° (Fig. 8). Barite sulfide veins occur in the Arqueros WNW–ESE extension for the fault veins. Although the results
Formation and in the Espino Member of Quebrada Marquesa are only an approximation due to the likelihood of multiple
Formation. They are distal to the El Espino deposit area and faulting events, the models suggest that the fault architecture
have scattered orientations. was established prior to mineralization during NNW–SSE
In the Romero area, iron oxide and quartz sulfide veins crustal transtension with a dextral component and that the
occur mostly in the Quelen Member and have strikes ranging fault veins formed during extensional faulting and WNW–
from NNW to NNE and averaging ∼185°. Dips of the veins ESE oriented crustal extension.
average 65° (Fig. 8). Vein density is higher in the Romero area The overall geometry and kinematics of the faults and veins
(up to ∼6/km) than in the Espino area (∼3/km), where suggests the district contains a negative flower structure de-
stockworks of specular hematite are a common feature related veloped in a right-stepping jog locally along a major N-strike
to major veins. dextral fault system. This architecture also controlled the

Fig. 7 Lower hemisphere, equal


area stereonets showing bedding
a b
and fault orientation data. a
Bedding poles in Quebrada
Marquesa Formation excluding
areas of drag folds near exposed 2%

major faults. The average bedding


orientation is ∼000°/20°. Fault 4%
rotation axis (π-axis
perpendicular to best fit great
circle) is nearly horizontal and 6%

trends 350–170°. b Pole density


contours for 66 faults with gouge
width >5 cm. There are three
main fault set orientations:
∼135°/90°, 205°/80°, 350°/70°.
c Major faults orientations shown
as great circles. Slickenlines in
faults are depicted as dots with
arrows showing the relative
motion of the hanging wall. Black
squares labeled 1, 2, and 3 are the c d
instantaneous strain axes
calculated by Faultkin program
(Allmendinger et al. 2001) that
correspond to σ 3, σ 2, and σ 1
principal stress axes, respectively,
if minimal rotational strain is
assumed. d Histograms showing
dip of major faults (top) and the
rake of major fault slickenlines
(bottom)
244 Miner Deposita (2014) 49:235–259

Fig. 8 Lower hemisphere, equal


area stereonets showing vein a b
orientation data. a Poles to veins;
Espino IOCG veins (black
circle), Espino quartz-sulfide and 2%
calcite-quartz-sulfide veins (open
circle), Romero IOCG (black 4%
square), Romero quartz-sulfide
veins (open square), Llahuin 6%

IOCG veins (black triangle),


Llahuin quartz-sulfide veins 8%

(open triangle), and barite-sulfide


veins (star). b Pole density
contours for 44 IOCG veins in the
El Espino mining district; average
IOCG vein orientation is
∼210°/85°. c Major veins
orientations plotted as great
circles with dots indicating
slickenlines and arrows on the c d 6
dots showing relative motion of

Frequency
the hanging wall. Numbers 1–3
4
are the instantaneous strain axes
calculated by Faultkin program
(Allmendinger et al. 2001) that 2
correspond to σ 3, σ 2, and σ 1
principal stress axes, respectively, 0
if minimal rotational strain is 10 20 30 40 50 60 70 80 90
assumed. d Histograms showing Fault-vein dip
4
dip of major fault veins (top) and
the rake of major fault-vein
Frequency
slickenlines (bottom)
2

0
10 20 30 40 50 60 70 80 90
Slickenline rake

localization of later IOCG alteration and mineralization sulfide mineralization. The district displays a complex series
around the Llahuin pluton (Fig. 10). The Espino vein system of late hydrolytic alteration styles that also have associated
formed mostly as NE-strike extension fractures and normal copper and gold mineralization.
faults within the jog, whereas the Romero and Llahuin veins
formed along the eastern N- to NNE-strike basin margin fault Structural controls on alteration and mineralization
system. The Romero area contains NS-striking dioritic dikes
that locally controlled the location of IOCG fault veins. The distribution of hydrothermally altered and mineralized
zones in the El Espino district indicates a strong structural
control. The NNE-striking Lagarrigue fault forms a fundamen-
Alteration and mineralization tal eastern boundary to the district with altered rocks occurring
to the west but not to the east of this fault. Sodic hydrothermal
The El Espino mining district displays a variety of different alteration appears to have occurred between Lagarrigue and
alteration types and styles (Figs. 11 and 12). The paragenetic Illapel–Espino faults. Major alteration zones are distributed
sequence of alteration (Fig. 11) was built based on crosscut- along, or subparallel to, the NNW- to NE-striking faults in the
ting relationships between alteration minerals observed either district. Although the NNW-strike faults may contain iron
in the field or petrographically with the microscope. An early oxides and sulfides, these faults rarely contain ore bodies.
sodic alteration event affected much of the district. It was Many of the iron oxide and sulfide-bearing veins in the
followed by sodic–calcic and potassic alteration events. These district have a generally N to NE strike. At the El Espino
alteration events are associated with iron oxide mineralization. deposit steeply dipping sulfide-bearing veins strike NE to
A later calcic alteration event was synchronous with copper NNE. Similar veins in the Romero area also have an N- to
Miner Deposita (2014) 49:235–259 245

Fig. 9 Lower hemisphere, equal


area stereonets showing fault 3 2
plane solutions for a major faults
and b IOCG fault veins in El
Espino mining district. Labels σ 3, 3
σ 2, and σ 1 correspond to
2
principal stress axes

1
1

NNE strike that extend upward into the Llahuin area. Vein The NW-strike Tamara fault (Correa 2003) apparently
density is higher at Romero and veins stockworks, commonly postdates both alteration and mineralization. While it locally
containing abundant specular hematite, are more prominent contains sericite, it is never mineralized. Instead, the fault
here than in other prospects. The width and intensity of appears to have acted as a conduit for meteoric waters that
alteration around NE-striking faults and veins are commonly resulted in supergene alteration of sulfide-rich zones that it
comparable to, or better developed than, alteration around the transects.
NW striking faults.
The N- to NE-striking fault veins in the district are related Sodic alteration
to a transtensional stress regime during the prolonged hydro-
thermal event. The strike of these structures suggests WNW– Sodic alteration affected a large portion of the El Espino
ESE extension with an associated dextral component. Calcic district. It is best developed along a north-striking corridor
and hydrolytic alteration together with the majority of iron from the Caimanes–Illapel pluton to the El Espino deposit.
oxide and sulfide mineralization appear to be synchronous The style and mineralogy of sodic alteration associations
with this extensional structural event. vary by rock type but generally resulted in the formation of

Fig. 10 Proposed structural


model for El Espino mining
district. a Satellite image
(Landsat 741) with lineament
interpretation. b Fault jog model
showing general orientation and
location of veins (red) in El
Espino (e), Romero (r), and
Llahuin (l) areas. Plutons shown
as pink ellipses: d Llahuin diorite
and equivalents, g dioritic pluton
along southern margin of basin
246 Miner Deposita (2014) 49:235–259

Fig. 11 Paragenetic sequence of alteration minerals. Continuous lines represent a major phase, segmented lines represent a minor phase, and a dotted
line represents a trace mineral

albite (Fig. 13a). Igneous rocks generally display pervasive replaces the cores of plagioclase, pyroxene, and hornblende
alteration with complete replacement of plagioclase by albite crystals. Actinolite replaces hornblende and pyroxene. Volcanic
and commonly wholesale destruction of igneous mafic min- rocks display pods of epidote grading outward to albite.
erals by chlorite. Veinlets of massive albite up to 1 cm in width Breccias and conglomerates show clasts replaced by intergrown
are common in intrusive rocks. Sedimentary rocks generally epidote and albite, or surrounded by an albite rim. Thin-bedded
display selective replacement of fine-grained beds by albite or siltstone and sandstone are altered to albite and epidote zones
albite and chlorite, where albite replaces plagioclase and chlo- along bedding (Fig. 14b). Some siltstone samples show
rite replaces amphibole and pyroxene. In conglomerate and irregular, <1-mm-wide epidote veinlets that cut across
volcanic breccias, albite replaces the matrix and/or forms rims albite-altered beds to connect parallel epidote-altered beds.
around clasts. Intrusions contain epidote veinlets with albite selvages or
pseudo-breccia textures formed due to irregular replacement
Sodic–calcic alteration of the rock by albite, actinolite, and chlorite. Subhedral dis-
seminated magnetite (1–5 %) is commonly present in sodic–
Sodic alteration associations were overprinted by sodic–calcic calcic alteration associations, and it is often partially to totally
alteration associations of albite, epidote, actinolite, titanite, martitized. Intense sodic–calcic alteration with scapolite is
scapolite, and minor apatite. The area affected by sodic–calcic restricted to zones within the leucodiorite intrusion located
alteration is more restricted than that affected by sodic alter- in the Romero area.
ation. As with sodic alteration associations, sodic–calcic alter-
ation formed textural and mineralogical differences related to Potassic alteration
rock type.
The most common alteration minerals are epidote, actino- Potassic alteration assemblages are present in the El Espino,
lite, and albite. The most common mineral associations are Romero, and Llahuin areas. These assemblages are locally
epidote–albite, actinolite–albite, epidote–actinolite–chlorite– and weakly developed and/or highly affected by subsequent
albite–(−titanite), and actinolite–scapolite–albite (Fig. 11). Al- calcic and hydrolytic alteration. Subsequent alteration of po-
bite partially or totally replaces plagioclase. Epidote partially tassic assemblages has commonly resulted in biotite altering
Miner Deposita (2014) 49:235–259 247

Fig. 12 Alteration map of El Espino mining district

Fig. 13 Outcrop photographs


showing examples of alteration.
a Sodic alterated rock as albite
stockwork veinlets in quartz
diorite from Caimanes–Illapel
pluton. b Sodic–calcic altered red
sandstone from Romero area with
some beds partially altered to
albite and epidote
248 Miner Deposita (2014) 49:235–259

Fig. 14 Photographs showing


examples of potassic alteration.
a b
K-spar potassium feldspar, bt
biotite, chl chlorite, mt magnetite.
a Hand sample of sandstone from
El Espino prospect with bt + mt
disseminated secondary biotite K-spar
and magnetite alteration
assemblage. b Hand sample of
andesite from El Espino prospect
with K-feldspar and chlorite 2 cm 2 cm
cal
alteration association.
c Photomicrograph (crossed c d
polars) showing leucodiorite
from Romero area replaced
mostly by K-feldspar and lesser K-spar
epidote and biotite. Primary
plagioclase is replaced by
bt bt
K-feldspar and biotite. chl
d Photomicrograph (crossed
polars) showing secondary
biotite replaced by chlorite 200 m 200 m

to chlorite and potassium feldspar generally being replaced by pods. Some pods have specular hematite in their centers. The
sericite, clays, and chlorite; potassium feldspar is more com- least altered leuco-monzonite sample located at medium ele-
monly preserved than biotite. Potassium feldspar either re- vations in this area displays interstitial potassium feldspar
places original plagioclase or replaces albite formed during (10 % of the rock), disseminated magnetite, and ferromagne-
earlier sodic or sodic–calcic alteration. sian minerals that are replaced by actinolite and chlorite. A
Secondary biotite accompanied by disseminated subhedral more altered sample displays secondary biotite and potassium
magnetite was found in several deep drill holes at the El feldspar. Potassium feldspar replaces plagioclase phenocrysts
Espino deposit (Fig. 14a). The style of alteration is typical of and is commonly intergrown with minor biotite, actinolite,
that observed in other Chilean IOCG deposits such as epidote, and/or chlorite (Fig. 14c, d). Secondary biotite
Candelaria (Marschik and Fontbote 2001). However, its ex- accompanied by disseminated subhedral magnetite was
tent is poorly constrained. Potassic alteration assemblages found deep at El Espino deposit (few meters of core) and
containing potassium feldspar, commonly with hematite its distribution is unclear.
(Fig. 14b), are most common in the El Espino district. The
age relationships between the biotite-bearing and potassium Calcic alteration
feldspar-bearing potassic assemblages are currently unclear.
While it is possible that the different potassic mineral assem- Calcic alteration affected a northwest-trending area of approx-
blages simply represent originally different protolitths (Barton imately 25 km2 extending from the El Espino deposit to the
and Johnson 1996), they may also represent fundamentally Romero prospect (Fig. 12). Calcic alteration mineral associa-
different assemblages formed at different times and in differ- tions contain epidote, actinolite, and chlorite with lesser calcite,
ent structural levels. titanite, apatite, and quartz and minor garnet (andradite) in
Volcanic and intrusive rocks affected by potassic alteration calcareous sedimentary rocks. The most common calcic alter-
contain interstitial metasomatic potassium feldspar. In pyro- ation associations are epidote–actinolite and epidote–chlorite
clastic and sedimentary rocks, potassium feldspar partially or (Fig. 15a). A first stage of calcic alteration is characterized by
totally replaces albite either in the matrix or in previously partial replacement in volcanic rocks and pervasive replace-
altered fragments. Potassium feldspar alteration is selective, ment in sedimentary rocks that locally obliterates their primary
forming bands in thin bedded siltstone or rims around clasts. texture. Calcic alteration epidote replaces igneous feldspar
Locally, potassium feldspar has pervasively replaced all the and mafic minerals and cuts hydrothermal albite and potassi-
primary minerals in clastic sedimentary rocks. um feldspar. Epidote occurs as both fine-grained anhedral
The eastern side of Romero area displays extensive evi- aggregates and well-developed crystals up to 10 mm in length.
dence of potassic and potassic–calcic alteration. A conglom- It occurs as veinlets, disseminations, or massive replacements
erate of the Quelen Member in the Romero area contains of the rock. Actinolite occurs as acicular aggregates with
rounded pods (3–5 cm diameter) of epidote grading outward crystal lengths averaging 0.5–1 cm, but ranging up to 5 cm;
to interstitial pink potassium feldspar as haloes around the it is commonly disseminated but locally can pervasively
Miner Deposita (2014) 49:235–259 249

replace host rocks. Chlorite replaces igneous ferromagnesian deposition. Many of the veins display actinolite selvages that
minerals as well as hydrothermal actinolite and biotite. Epi- are commonly 5–10 mm wide. Some actinolite selvages grade
dote and actinolite associations, sometimes with calcite, are outwards to epidote with minor intergrown chlorite.
accompanied by abundant magnetite or less frequently hema- Calcic alteration associations also occur in specular
tite (Fig. 15b). These mineral associations may contain mag- hematite-bearing breccias found at low elevations within
netite, ilmenite, pyrite, and chalcopyrite with minor the leuco-monzonite unit in the Romero prospect area.
bornite and covellite (Fig. 15c, d). Sulfides replace iron These breccias contain fragments replaced by albite, scap-
oxide minerals or form minute inclusions in quartz olite, potassium feldspar, or quartz in a matrix of actino-
(Fig. 15c, d). Minor chalcopyrite, bornite, pyrite, and lite with specular hematite and disseminated chalcopyrite
native gold, together with acicular actinolite, were found and pyrite.
in quartz grains around or within replacive epidote–actinolite
pods. Chalcopyrite either formed later as inclusions in idio- Hydrolytic alteration
blastic pyrite (Fig. 15e, f) or as partial replacement of
pyrite. Hydrolytic alteration associations occur within hydrothermal
A second stage of calcic alteration is characterized by iron breccia zones, around a mineralized manto body at the El
oxide- and sulfide-bearing veinlets with actinolite selvages Espino deposit, and in veins (Fig. 16a, b). Hydrolytic
that cut earlier calcic, sodic–calcic, and sodic alteration asso- alteration associations contain variable amounts of quartz,
ciations. These veinlets range in width from 1 to 10 mm and calcite, sericite (illite and phengite), chlorite, iron oxide
contain actinolite, epidote, and quartz with either magnetite or minerals, dominantly hematite, and sulfides (Fig. 16a–d);
specular hematite and pyrite and chalcopyrite. Specular he- some hydrolytic associations contain minor rutile and Na–
matite in some veinlets has been replaced by mushketovite Fe tourmaline. Hydrolytic alteration associations contain the
(magnetite pseudomorphs of hematite) related to carbonate vast majority of sulfides in the El Espino mining district.

Fig. 15 Photomicrographs a b
showing examples of calcic
alteration associations. ab albite, hm
act actinolite, cal calcite, epi act ab cal
epidote, qz quartz, hm specular
hematite, mt magnetite, bn hm
bornite, cpy chalcopyrite, py epi
pyrite. a Conglomerate from epi
Romero area replaced by act
actinolite and epidote
overprinting earlier albite. 1 cm 1 mm
b Same conglomerate as A
showing actinolite, epidote, c d
cal
calcite, and hematite (crossed epi cal
polars). c Conglomerate from bn epi
bn
Romero area showing quartz with
hm hm
inclusions of acicular actinolite
cpy
crystals and opaques (crossed
polars). d Same as f, but with
reflected light showing that
opaques are bornite and qz qz
chalcopyrite replacing hematite.
e Mineralized sandstone from El act 100 m hm act 100 m
Espino area showing pyrite and
chalcopyrite (reflected light). e f mt
f Pyrite and chalcopyrite
replacing magnetite in siltstone
from El Espino deposit
pervasively altered to actinolite qz
and minor epidote (reflected light)
py
cpy
py cpy
act
mt
100 m 100 m
250 Miner Deposita (2014) 49:235–259

Hydrolytic stage hydrothermal breccias are found at the El (Fig. 16e, f). Minor amounts of bornite, galena, and
Espino, Romero, and Llahuin areas. These hydrothermal brec- hypogene chalcocite are present locally as replacements
cias contain fragments with muscovite replacing earlier of chalcopyrite. Trace native gold is present as inclusions
formed alteration minerals such as albite and potassium feld- in chalcopyrite.
spar in a matrix of chlorite and specular hematite.
Alteration zones adjacent to El Espino manto commonly Calcite veins and argillic alteration
contain a white mica–chlorite–specular hematite assemblage.
Both hydrothermal breccias and mantos are commonly The final hypogene alteration and mineralization event in
overprinted by 1–2 cm wide veinlets of specular hematite, the Espino district resulted in the formation of calcite–
minor quartz, and chalcopyrite. barite–(quartz) veins that may contain chalcopyrite, pyrite,
Three stages of hydrolytic alteration/mineralization were rec- and/or galena. These veins have selvages composed of various
ognized from cross-cutting vein relationships. The earliest veins combinations of chlorite, quartz, calcite, illite, clinochlore,
contain quartz and sulfides. These are cut by veins containing montmorillonite, and laumontite. Thick veins of this have
hematite and sulfides with variable amounts of quartz (Fig. 16d). been exploited by small miners for copper and locally silver,
The hematite-sulfide veins have well to poorly developed sel- presumably hosted in the galena. These veins are distal to the
vages of white mica and chlorite. The latest veins are composed main zones of hydrolytic alteration and mineralization and
of quartz, calcite, sulfide and minor iron oxide (Fig. 16c). occur found at higher elevations in the Llahuin and north of El
Hydrolytic alteration assemblages commonly contain Espino deposit areas.
chalcopyrite, pyrite, and specular hematite. Sulfides and
hematite at the El Espino deposit occur in veins and within Supergene alteration and mineralization
a manto within a favorable fine-grained volcaniclastic
unit of the Espino Member. Pyrite and chalcopyrite are Weathering of sulfide-rich associations in the El Espino
coexisting with chalcopyrite formed later replacing pyrite district has resulted in the formation of supergene copper

Fig. 16 Photographs showing a b


hydrolytic alteration associations.
cal calcite, chl chlorite, cpy chl
chalcopyrite, hm specular cpy cpy
hematite, qz quartz, py pyrite, ser py
sericite. a Hematite-sulfide hm
hm py ser
veinlet with chlorite selvages in
sandstone from Romero area.
b Breccia within 1 m wide vein in
chl
Romero area, showing massive
chalcopyrite accompanied by 2 cm 1 cm
chlorite and sericite and cut by
hematite veinlet. c Calcite– c ser d py
chalcopyrite veinlet in diorite
ser+chl
from El Espino deposit hm1
overprinting specular hematite– chl cal
sericite–chlorite–quartz veinlet. qz
d Veinlet in sandstone from El ser
Espino deposit showing two hm2
hm qz
stages of hematite, the first shows chl
actinolite altered to chlorite and cpy
the second shows sericite, quartz 2 cm
hm 2 cm
and minor chlorite (reflected
light). e Mineralized sandstone
from Romero area showing e f
chalcopyrite inclusions in pyrite cpy
cpy hm
(reflected light). f Mineralized
sandstone from El Espino deposit py
showing chalcopyrite after pyrite cpy
(reflected light)
py

py
Miner Deposita (2014) 49:235–259 251

oxides and limited amounts of supergene chalcocite. Supergene elevations (1,000–800 m depth from Llahuin area) in the El
alteration also resulted in the formation of clays. Supergene Espino deposit area and at medium to low elevations (1,000–
alteration has been observed along fault zones and is particu- 500 m depth) in the Romero area. While surface exposures of
larly well developed along the Tamara fault zone (Fig. 6). calcic-altered rocks contain minor specular hematite or mag-
netite, pyrite, and traces of chalcopyrite, deep exposures of
this alteration type at the El Espino deposit contain zones of
Vertical and lateral distribution of alteration massive magnetite cut by specular hematite veins with locally
and mineralization significant amounts of pyrite and chalcopyrite.
Potassic-altered rocks occur as small zones both within and
Alteration assemblages in the El Espino mining district outside zones of sodic alteration (Fig. 11). Potassically altered
display a distinct vertical and lateral zonation (Fig. 17). Both rocks are found at both low (El Espino) and high (Llahuin)
the El Espino and Romero prospects contain areas at low to elevations. Extensive zones with hydrothermal potassium
intermediate elevations (1,000–500 m deep from Llahuin feldspar are common present within intrusions in the Llahuin
area) that contain sodic alteration assemblages that have been area. Hydrothermal biotite accompanied by K-feldspar occurs
overprinted by sodic–calcic, potassic, and calcic alteration deep at the eastern margin in the Romero area. Hydrothermal
assemblages. The generally high elevation Llahuin area biotite at the Romero prospect area appears to have been more
contains well-developed hydrolytic alteration assemblages. widespread but was almost completely replaced by later calcic
Hydrothermal alteration was centered on high angle faults and hydrolytic alteration assemblages.
with alteration generally decreasing in intensity and complexity Hydrolitic alteration assemblages are most common at
outward from these structures. medium to high elevations throughout the district. Hydrolytic
Sodic alteration affected over 100 km2 in the district. Sodic alteration is concentrated in veins that served as feeders for El
alteration was most intense along north–south- and NE- Espino manto deposit. Hydrolytic assemblages appear to
trending regional lineaments (Fig. 12). Sodic alteration is best change mineralogy with depth. Iron oxide-rich hydrolytic
developed at low elevations and only minor zones of sodic- veins at El Espino and Romero grade upward into quartz veins
altered rocks are found at higher elevations, primarily in the with little to no hematite at Llahuin. At the highest elevations
Llahuin area. Sodic–calcic altered zones are spatially more in the district, veins are dominated by argillic assemblages and
restricted than sodic-altered rocks and cover an area of contain dominantly calcite and barite with minor argentiferous
approximately 50 km2. Sodic–calcic altered rocks surround sulfides..
dioritic intrusions located in the southern end of the Llahuin
plutonic suite and are present in structural zones within and
between the Espino and Romero areas. Sodic–calcic altered Age of alteration and mineralization
zones are best developed at low to medium elevations (1,000–
500 m deep from Llahuin area) but may extend to higher The age of hydrothermal alteration and mineralization in the
elevations in the Llahuin area and in the very far north of El Espino district was investigated by Lopez (2012) through
40
the district. Ar- 39 Ar isotopic analysis of hydrothermal minerals
Calcic altered rocks occur primarily in an area 5 km long× (Table 1). Analysis of potassium feldspar, separated from an
2 km wide from El Espino deposit on the west to the Romero irregular calcite vein developed within a potassically altered
area on the east (Fig. 12). Calcic altered rocks occur at low biotite–magnetite assemblage from deep within the Espino

Fig. 17 Schematic EW cross-


section showing vertical
distribution of alteration types in
the El Espino mining district.
Note that intense calcic alteration
was confined largely to mixed
volcanic and sedimentary rocks
adjacent to the mostly sodically
and lesser potasically altered
Llahuin pluton. The location of
the section is illustrated in
figure 11. Mineral abbreviations:
act actinolite, chl chlorite,
hm specular hematite, ser sericite,
tm tourmaline
252 Miner Deposita (2014) 49:235–259

deposit, yielded an age of 86.09±0.45 Ma (Table 1). Analysis around 0 ‰ (Figs. 18 and 19). Copper sulfide-bearing
of actinolite, from a calcic alteration veinlet containing chal- veins lacking iron oxides yielded δ34S values between
copyrite, actinolite, epidote, and minor magnetite at the El −4.3 and +1.0 ‰, whereas sulfides from veins containing
Espino deposit, yielded an age of 88.4±1.2 Ma (Table 1). The iron oxide minerals yielded slightly higher δ34S values of
age of the hydrolytic alteration was determined from between −4.3 and +6.1 ‰. Sulfides within late stage
40
Ar–39Ar analyses of white micas (Table 1). Analysis of calcite veins are distinctively lower with δ34S values between
white mica from a core sample with a hydrolytic quartz- −4.8 and −9.7 ‰ (Fig. 18).
sulfide veinlet at El Espino yielded an inverse isochron age Although the sulfur isotopic values of pyrite and chalco-
of 87.9±0.6 Ma (Table 1). White mica from a specular hema- pyrite in the calcic and hydrolytic associations overlap, there
tite veinlet at Espino yielded a similar plateau age of 87.9± are slight differences in isotopic compositions. Pyrite values
0.6 Ma (Table 1). are generally higher and range from −2.2 to +6.1, whereas
Geochronological U–Pb and 40Ar/39Ar data indicate that chalcopyrite values range from −4.4 to +4.5 ‰. Some samples
hydrothermal alteration was coeval with magmatic intrusive with coexisting chalcopyrite and pyrite have significantly
activity. One particular dioritic intrusion (88.5 Ma) preceded different isotopic values, suggesting that chalcopyrite formed
the calcic stage (88.4 Ma), which was accompanied by by utilizing reduced sulfur in the hydrothermal fluid rather
iron oxide copper and gold mineralization. Hydrolytic than incorporating sulfur from pre-existing pyrite (Ohmoto
alteration, related to economic iron oxide copper and gold and Goldhaber 1997).
mineralization, came immediately after at 87.9 Ma. It also The wide range of sulfur isotope compositions are compat-
appears that hydrothermal alteration and mineralization ible with two sources variably mixed, a sulfur source of δ34S
may have been accomplished in a short period of two to close to 0 ‰ and a sulfate source with variable degrees of
three million years. reduction. The sulfur isotope composition range related to
calcic alteration is relatively more restricted (range of 8 ‰)
with a mean near 0 ‰. That alteration surrounds a dioritic
Sulfur isotopes of sulfides from calcic and hydrolytic intrusion. The relatively smaller range around 0 ‰ and
alteration associations pattern of the alteration suggest a magmatic source for the
former of the fluids. Hydrolytic stage sulfur isotope ratios
Sulfur isotopic analyses were performed on sulfide minerals show a wider range (11 ‰), generally higher than in the
collected from calcic, hydrolytic, and late calcite vein associ- calcic stage, suggesting incorporation of sulfur derived
ations across the district (Table 2). Samples were acquired from gypsum. High temperature reduction of sulfate
from a range of locations to investigate potential changes in may have occurred during the convective circulation of
sulfur isotopic composition vertically and laterally within the reduced magmatic hydrothermal fluids through the gyp-
hydrothermal system. Single analysis of gypsum from the sum beds of the Quebrada Marquesa Formation. Sulfur
Espino Member of the Quebrada Marquesa Formation yielded isotope ratios in the late calcite veins suggest incorporation of
a δ34S value of 12.4 ‰. Single analysis of diagenetic pyrite in biogenic sulfur.
sedimentary rocks of Espino Member yielded a δ34S value of Hydrothermal sulfides in the district show a trend toward
−7.3 ‰. higher sulfur isotope values with stratigraphic height in the
Sulfides extracted from veins and rock replacements Quebrada Marquesa Formation (Fig. 20). The greatest
related to calcic alteration yielded δ34S values between variability of sulfur isotopic values occurs within the lower
−4.0 and +2.4 ‰, with a mean value around −1 ‰ (Figs. 18 marine, locally gypsiferous sedimentary rocks of the Espino
and 19). Sulfides within hydrolytic altered rock yielded Member with δ34S values ranging from −9.7 to +2.4 ‰,
δ34S values between −4.4 and +6.2 ‰, with a mean value whereas the sulfides in the overlying continental conglomerate

Table 1 40Ar/39Ar geochronology


data for alteration ages Sample Locality Material Age±2σ Ma Type Alteration type

ERD-6 247.4 Espino Actinolite 88.4±1.2 Inverse isochron Calcic


ERD-6 128 Espino Sericite 87.9±0.6 Inverse isochron Hydrolytic
ERD-6 142.3 Espino Sericite 87.9±0.6 Plateau Hydrolytic
EDH-3 246 Espino K-Feldspar 86.21±0.41 Plateau Calcite vein
Rab-01 Espino Muscovite 98.9±2.2 Inverse isochron Hydrolytic
179682 Romero Muscovite 94.17±0.49 Total fusion Hydrolytic
172261 Llahuin Illite 91.71±0.47 Total fusion Hydrolytic
Miner Deposita (2014) 49:235–259 253

Table 2 Sulfur isotope data of mineralized rocks from El Espino mining district

Rock description Alteration and mineralization S isotopes

Sample Locality Host rock Morphology Mineralogy Alteration type Sulfide δ34S

EDH4-83 Espino Siltstone Siltstoneiron oxide–sulfide vein alb-act-chl-hm-mushk-cal-py-cpy Sodic–calcic py −1.51


ERD6-242 Espino Andesite Iron oxide–sulfide vein act-epi-calc-cpy-hm Calcic cpy −3.66
ERD9-137 Espino Sandstone Rock replacement epi-act-chl-py-cpy Calcic py −3.51
ERD4-174 Espino Sandstone Rock replacement epi-chl-mgt-cpy-py Calcic cpy −2.94
ERD4-174 Espino Sandstone Rock replacement epi-chl-mgt-cpy-py Calcic py −1.36
179694 Espino Siltstone Rock replacement epi-act-cpy-py Calcic cpy −1.26
ERD6-236 Espino Siltstone Rock replacement chl-cal-mgt-py-cpy Calcic py −0.89
ERD6-236 Espino Siltstone Iron oxide–sulfide vein act-qtz-cal-mushk-cpy-py Calcic py −0.83
ERD6-236 Espino Siltstone Iron oxide–sulfide vein act-qtz-cal-mushk-cpy-py Calcic cpy −0.06
ERD4-136 Espino Siltstone Rock replacement alb, epi-act-mgt-py-cpy Calcic py 0.27
179695 Espino Diorite Iron oxide–sulfide vein act-qtz-mgt-calc-py Calcic py 2.58
ERD5-165 Espino Calc. siltstone Calcite vein chl-ser-cal-cpy Calcite vein cpy −9.69
ERD7-179 Espino Siltstone Calcite vein calc-hm-cpy Calcite vein cpy −9.14
ERD6-83 Espino Diorite Calcite vein alb-act-chl-cpy (rock), calc (vein) Calcite vein cpy −4.73
172271 Llahuin Diorite Quartz sulfide–iron oxide vein chl-ser-qtz-calc-hm-cpy Hydrolytic cpy −4.30
ERD6-128 Espino Sandstone Quartz sulfide–iron oxide vein chl-ser-hm-mushk-qtz-calc-cpy Hydrolytic cpy −4.20
179553 Romero Leucodiorite Quartz vein chl-ser-hm-cpy-py Hydrolytic cpy −2.65
172252A Romero Conglomerate Quartz–iron oxide vein calc-qtz-chl-ser-hm-mushk-cpy Hydrolytic cpy −2.26
179682 Romero Conglomerate Iron oxide–sulfide breccia chl-calc-(ser)-cpy-py Hydrolytic py −2.00
ERD5-176 Espino Sandstone Rock replacement chl, calc-ser-mushk-cpy Hydrolytic cpy −1.79
179553 Romero Leucodiorite Quartz vein chl-ser-hm-cpy-py Hydrolytic py −1.59
179534A Romero Conglomerate Rock replacement chl-ser-mgt-hm-cpy Hydrolytic cpy −1.66
179534B Romero Conglomerate Iron oxide vein hm-chl-cpy Hydrolytic cpy −1.46
ERD5-176 Espino Sandstone Iron oxide–quartz vein chl-ser-cal-mushk-cpy Hydrolytic cpy −0.78
179550 Romero Diorite Rock replacement chl-ser-(hm)-py-cpy Hydrolytic py −0.29
179689 Espino Andesite Quartz vein chl-ser-hm-calc-py-cpy Hydrolytic py 0.59
179682 Romero Conglomerate Iron oxide–sulfide breccia chl-calc-(ser)-cpy-py Hydrolytic cpy 1.13
179515 Llahuin Diorite Quartz vein chl-ser-(hm)-cpy Hydrolytic cpy 1.04
EDH4-157 Espino Sandstone Iron oxide–sulfide vein chl-ser-mgt-py, hm-mgt-cpy Hydrolytic cpy 1.87
179696 Romero Conglomerate Iron oxide–sulfide vein chl-ser-hm-mgt-calc-cpy-py Hydrolytic cpy 4.60
179696 Romero Conglomerate Iron oxide–sulfide vein chl-ser-hm-mgt-calc-cpy-py Hydrolytic py 6.38
179534B Romero Conglomerate Iron oxide breccia hm-chl-cpy Hydrolytic cpy 0.87
179683 Romero Conglomerate Disseminated ser-clays-py, cpy veinlets Hydrolitic/argillic cpy 0.36
172255 Romero Diorite Quartz vein qtz-chl-py Hydrolytic py 4.22

Quelen Member have heavier values of −2.7 to +6.2 ‰. Fluid inclusion studies
These compositions suggest that isotopically heavy sulfur
was contributed to the hydrothermal fluids during upward Eight samples of quartz from veins with iron oxides and sulfides
fluid migration. were surveyed fluid inclusions suitable for microthermometric
Calculation of the precipitation temperature of coexisting study. Homogenization temperatures were collected only
(i.e., textural equilibrium) chalcopyrite and pyrite in calcic from assemblages of inclusions that showed consistent
alteration associations using the method of Ohmoto and Rye liquid to vapor volumetric proportions and that yielded
(1979) and Campbell and Larson (1998) yields temperatures consistent results (Goldstein and Reynolds 1994). Such
between 529 and 280 °C. Hydrolytic alteration sulfide yielded assemblages of inclusions were very rare, perhaps because
temperatures of 216 °C (Table 3). of overprinting of hydrothermal events with concomitant
254 Miner Deposita (2014) 49:235–259

Fig. 18 Sulfur isotope 7

compositions of chalcopyrite and


pyrite organized by alteration
Calcite veining
stages for calcic, hydrolytic, and 6
calcite veins. Calcite veins sulfide
correspond to chalcopyrite. Both Hydrolytic alteration
sedimentary gypsum and
5
diagenetic pyrite sulfur isotope Calcic alteration
compositions are also included
for comparison

Frequency
4
Diagenetic pyrite

Gypsum
3

0
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10 11 12
34S (o/ )
oo

recrystallization of quartz and post-entrapment changes of association of quartz, pyrite, chalcopyrite, specular hematite,
the fluid inclusions. epidote, actinolite, chlorite after actinolite, and sericite. Most
A core sample from Espino deposit (ERD-6, 79.2 m FIAs hosted by quartz show inconsistent liquid/vapor ratios,
depth), is a late calcic alteration stage quartz veinlet with but one FIAwas found that yielded consistent homogenization
intergrown hematite and chalcopyrite in an amphibole temperatures at 340–350° C.
diorite host rock. The veinlet displays euhedral quartz As there is no evidence of brine-vapor immiscibility in all
with no evidence of quartz annealing (Fig. 21), though samples surveyed, T h only provides a minimum temperature
the sample did contain many fluid inclusion assemblages of formation conditions. A pressure correction must be ap-
(FIAs) with inconsistent liquid to vapor ratios. The FIAs plied to get the true formation conditions. Although the depth
measured were intimately associated with tiny specular of formation is unknown, sample 179689 displays textures
hematite and acicular actinolite solid inclusions within such as wispy inclusion texture and recrystallized quartz in-
quartz and thus are deemed primary in origin. These dicative of formation under lithostatic conditions (Reynolds,
primary inclusions contain transluscent halite crystals. Three 2012, personal communication). If a 350 °C temperature is
FIAs yielded consistent homogenization temperatures (T h) pressure corrected to 425 °C (assuming that ductile deforma-
between 280 and 295 °C. Salinities were between 32 and tion styles would occur at such temperatures under lithostatic
34 wt% NaCl equivalent. No other FIAs at higher T h were load), then the pressure required would be roughly 700–1,000
found in this vein sample. bars for a saline brine, corresponding to a minimum depth of
A second core sample from a calcically altered siltstone formation of roughly 3–4 km, consistent with the estimated
deeper in the El Espino deposit (ERD-6, 236.8 m depth) thickness of the Late Cretaceous lithostatic load (Salamanca
contains an actinolite and magnetite assemblage that is Formation) in the area.
overprinted by a second alteration association composed of
calcite, quartz, mushketovite, chalcopyrite, sericite, and
chlorite. Euhedral quartz associated with the second calcic Discussion
alteration association is intergrown with acicular fibers of
actinolite and contains growth zones along the margins of Early sodic alteration (albite) was the most aerially extensive
the vein. Liquid/vapor ratios were not consistent in these and external alteration event in the district and is spatially
inclusions due to necking; however, salinities from these related to the Llahuin and Illapel–Caimanes plutons especially
primary inclusions yield ∼15 wt% NaCl equivalent. along major fault zones and lineaments within and around
Sample number 179689 is from an intermediate elevation these plutons. Although best developed at stratigraphically
north of the El Espino deposit. This sample contains a mineral deep levels, i.e., ∼1,000 m deep from the top of the alteration
Miner Deposita (2014) 49:235–259 255

exposed higher in the Llahuin area, the sodic alteration as- hydrothermal system, as proposed by Barton and Johnson
semblage extends upward in the El Espino mining district to (2000) for a number of IOCG systems.
the near surface in Espino and Romero areas where it gener- Sodic–calcic alteration (albite–epidote–actinolite) is
ally seems to transition into a sodic–calcic assemblage com- relatively spatially more restricted than sodic alteration and
posed of albite and epidote. Although some units of Llahuin grades into a center of calcic alteration assemblages (actino-
suite are affected by pervasive sodic alteration, other units lite–epidote–iron oxide) displaying a zoned pattern around the
such as a diorite pluton and sills in the Espino deposit are not southern end of the Llahuin Pluton.
affected by this alteration, whereas other units higher in the Potassic alteration in the El Espino district is relatively
Llahuin area show albite veining which might have developed restricted in extent. Deep, biotite–magnetite assemblages,
later. The distribution of intense sodic alteration is compatible similar to those at Candelaria, are present but are poorly
with it being developed within inflow zones to the broad known due to a lack of deep drilling. Higher level potassium

Fig. 19 Sulfur isotope a


compositions for pyrite and 4
chalcopyrite in a calcic and b Calcic
hydrolytic alteration associations

pyrite
3

chalcopyrite
Frecuency

0
-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8
34S (o/
oo)

b 7
Hydrolytic
6

pyrite
5
chalcopyrite

4
Frecuency

0
-8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8
34S (o/
oo)
256 Miner Deposita (2014) 49:235–259

Fig. 20 Sulfur isotope 9


compositions organized by
location and therefore, 8
stratigraphic level. El Espino
sulfides occur in the Espino
7 Llahuin (n=2)
Member, whereas Romero
sulfides occur stratigraphically Romero (n=13)
higher in the Quelen Member 6
Espino (n=20)
5

Frecuency
4

0
-10 -9 -8 -7 -6 -5 -4 -3 -2 -1 0 1 2 3 4 5 6 7 8 9 10
34S ( o/
oo)

feldspar-bearing potassic assemblages are better developed at medium levels (i.e., 1,000–500 m deep from the top of the
the current level of exposure of the system. Cross-cutting exposed alteration in the Llahuin area). The calcic alteration
relationships suggest that potassium feldspar-bearing as- assemblage extends upward and outward transitioning into a
semblages cut early sodic alteration assemblages but were sodic–calcic assemblage dominated by albite–actinolite–
locally formed synchronous with potassic–calcic altered epidote(−scapolite). Significant amounts of iron oxides are
rocks. In the Romero area, spatially restricted potassic– associated with calcic alteration assemblages. Geochronolog-
calcic zones developed within the calcic area that are related ical data indicate that calcic alteration was synchronous with
to iron oxide mineralization. However, potassic alteration intrusive magmatic activity in the district.
zones do not always display a close spatial relationship to Hydrolytic alteration assemblages (chlorite–sericite–
sodic, sodic–calcic, and calcic alteration zones. The spatial quartz–hematite) in the El Espino district occur along faults
and temporal distribution of potassic alteration zones could and veins. These asemblages are best developed at medium to
be evidence that they developed as outflow zones as is high stratigraphic levels, i.e., 500 to 0 m deep from the top of
hypothesized for several other IOCG districts (Barton and the exposed alteration in the Llahuin area. Although many
Johnson 2000). hydrolytic zones cut rocks that were previously sodic, sodic-
The calcic alteration zone forms the core of the large-scale calcic, or calcic altered, some hydrolytic alteration zones
alteration system in the district. It postdates sodic alteration, occur in relatively unaltered rocks along mineralized veins.
but is probably gradational in time and space with sodic– Hydrolytic veins are feeders to a zone of manto-type alteration
calcic and locally potassic–calcic alterations. The calcic and mineralization within favorable volcano-sedimentary
alteration zone is best developed at stratigraphically deep to lithologies in the El Espino deposit. Sulfides (pyrite and
chalcopyrite) are largely confined to calcic and hydrolytic
alteration assemblages.
Table 3 Estimates of the sulfur isotope fractionation temperature derived The paragenetic sequences in the El Espino and Candelaria–
from sulfur isotope compositions of coexisting chalcopyrite and pyrite for Punta del Cobre IOCG districts are distinct. The economic iron
each alteration type {T(°K)=[(0.45×106)/(δ34Spy −δ34Scpy)]1/2
oxide and sulfide mineralization are related to calcic–potassic
Location Sample δ34Spy (‰) δ34Scpy (‰) Alteration T (°C) alteration zones in Candelaria (Marschik and Fontbote 2001),
stage while in El Espino, the bulk of the iron oxide and copper
mineralization occurs with hydrolytic alteration. Alteration
Espino ERD6-236 −0.31 −1.01 Calcic 1st 523
stage
distribution and paragenesis for the early and late alteration
Romero 179553 −1.58 −2.36 Calcic 1st 486 stages in the El Espino district is similar to that seen in the
stage Manto Verde district in northern Chile (Benavides et al. 2007).
Espino ERD4-174 −1.43 −3.01 Calcic 2nd 261 Both districts show regional albitization and then local hydro-
stage
Romero 179696 6.37 4.50 Hydrolytic 218
lytic alteration related to economic iron oxide and sulfide
mineralization.
Miner Deposita (2014) 49:235–259 257

Fig. 21 Photomicrographs
showing quartz growth zones
with primary inclusions (a) and
primary inclusions coexisting
with acicular actinolite crystals
within quartz growth zones (b)

Extensive potassium and iron metasomatism was not seen probably sulfur. This magmatic source might have progres-
in El Espino mining district as it has been described for other sively mixed with a sulfate-derived source. In contrast, in
deposits such as Candelaria and Raul Condestable (Marschik hydrolytic assemblages, sulfur isotopic compositions of sul-
and Fontbote 2001; de Haller 2006). In the El Espino mining fides show a much wider spread, more clearly suggesting the
district, potassic metasomatism occurred only locally in the involvement of a sulfate-derived source. Under the moderate
Espino, Romero, and Llahuin areas, within or outside sodic, temperature, relatively oxidizing conditions indicated by the
sodic–calcic, and calcic altered areas. Isolated small zones of mineralogy and fluid inclusions, marine sulfate is likely to be
potassium feldspar and magnetite were even identified in the an important sulfur source in the hydrothermal system. The
Arqueros formation to the west of the district sometimes individual data sets of sulfur isotopic compositions show wide
spatially related to dioritic intrusions. spreads, and the coexisting minerals (as indicated in the para-
The host rocks, paragenetic sequence, and alteration pat- genesis) show that most formed with iron oxide, especially
terns at El Espino and the Raul-Condestable district in Peru hematite. All this evidence is consistent with partial reduction
(de Haller 2006) are very similar. At Raul-Condestable, the of heavy marine sulfate. This is what is observed in other areas
calcic alteration zone is developed in a volcanic and sedimen- such as Raul-Condestable (de Haller 2006) and modern Salton
tary sequence and grades upward into a hydrolytic zone and Sea geothermal system in Southern California. The close
outward into a sodic–calcic zone. In both districts iron oxide spatial and temporal association of well-altered zones with
and sulfide mineralization are related to calcic and hydrolytic generally intermediate composition intrusions suggests that
stages of hydrothermal alteration (de Haller et al. 2006). The part of Llahuin plutonic suite (including possible unexposed
potassic core (biotite) described at Raul-Condestable is either plutons) probably provided heat and contributed fluid and
weakly developed or not exposed in the El Espino district. sulfur early to the hydrothermal system, but the bulk of iron
Extensive biotite alteration might have taken place in the oxide, copper, and gold mineralization occurred relatively late
Romero area, but it has been masked by the later pervasive in the system by mixing with a fluid containing marine-
and extensive chloritization. There is surface evidence of derived sulfate.
secondary biotite in that area which is overprinted by either Estimates of the sulfur isotope fractionation temperature
calcic or hydrolytic assemblages. The core of the alteration show a wide range. These estimates assume that sulfur isotope
pattern at El Espino district is given by the calcic alteration fractionation factors are correct and that samples analyzed did
zone, dominated by actinolite, epidote, iron oxide, and sulfide. not have microscopic inclusions of another sulfide. The esti-
This type of alteration is mineralized and pervasive in Raul- mation of fluid temperature indicated by the high quality fluid
Condestable deposit where it surrounds a potassic core inclusion study performed seems to better represent the con-
(secondary biotite) related to tonalite intrusions (de Haller ditions of the fluids responsible for iron oxide, copper, and
2006), suggesting that a potassic core in El Espino district gold precipitation. The fluid responsible for specular hematite
might be still unexposed beneath the Llahuin area and that and sulfide mineralization in El Espino deposit was saline and
significant undiscovered deposits may exist in calcic-altered moderate temperature, which is consistent with results
zones deeper within the system. obtained by Rieger et al. (2012) for the Manto Verde district.
Sulfur isotopic compositions of sulfides in calcic assem- Homogenization temperatures (T h) measured in El Espino
blages show a wide range of values with a mean near 0 per quartz hosted inclusions within veins containing hematite
mile (Fig. 18) that may be interpreted as consistent with a were between 280 and 350 °C, consistent with quartz hosted
mixture of two sources, one magmatic and one external with inclusions related to hematite formation in Manto Verde,
variable mixing. The alteration pattern and timing of alteration which were in the range between 208 and 470 °C (Rieger
and magmatism suggests that dioritic intrusions played a role et al. 2012). Early magnetite formation in quartz fluid inclu-
in the formation of the system providing heat, fluid, and sions at Manto Verde deposit occured around 435 °C and
258 Miner Deposita (2014) 49:235–259

reached up to 530 °C (Rieger et al. 2012), suggesting that minor copper and silver are interpreted to represent the
calcic alteration related to massive magnetite formation highest-level expression of the hydrothermal system.
may have occurred at that similar range of temperature at Fluid inclusion provides evidence of a minimum fluid
El Espino district. temperature of 425 °C (pressure corrected) and minimum
depth of formation of 3–4 km for the formation of quartz–
hematite–actinolite veins related to the second stage of calcic
alteration. Sulfur isotopes, alteration mineralogy, alteration
Conclusions patterns, and timing of magmatism and alteration suggest that
early high temperature magmatically derived fluids were
The paragenetic sequence and structural/stratigraphic controls mixed with cooler, saline, oxidized fluids that included
of alteration and mineralization are similar to those observed marine-derived sulfate. The close spatial and temporal associ-
at a number of other Andean IOCG systems. As is typical of ation of well-altered calcic zones with generally intermediate
many Andean systems (Sillitoe 2003), an early large-scale composition ∼88 Ma intrusions suggests that these plutons
sodic alteration event was followed by a complex series of could have provided both heat and fluid in the early stage of
sodic–calcic, potassic, and calcic alteration events, which in the IOCG system.
turn were followed by a late, and generally high-level hydro- The Andean IOCG systems in Chile generally become
lytic alteration stage. As at a number of other Andean systems, younger to the south. Manto Verde’s age range between 123
sulfide precipitation in the El Espino mining district took place and 117 Ma, Candelaria–Punta del Cobre has been constrained
with calcic and hydrolytic associations. Unlike the Candelaria to 116–110 Ma (Sillitoe 2003), whereas El Espino deposit, with
(Marschik and Fontbote 2001) and Raul-Condestable deposits an apparent age of approximately 88 Ma, represents one of the
(de Haller et al. 2006), where the bulk of sulfides formed with youngest and southernmost systems yet defined. The relatively
calcic–potassic and calcic alteration associations, respectively, shallow level of erosion of the El Espino system compared to
the majority of the sulfides currently observed at El Espino older systems farther north in Chile suggests that significant
district deposits occur in hydrolytic alteration associations undiscovered deposits may exist in calcic-altered zones deeper
overprinting calcic alteration. within the system.
Iron oxides in the El Espino district display a general tem-
poral trend from early magnetite associated with sodic, sodic– Acknowledgements This work was supported by the National Science
calcic, potassic alteration, and early calcic to later hematite Foundation project EAR-0207217. Assistance from Teckcominco for
associated with calcic and hydrolytic alteration. This overall field work logistics and chemical rock analysis and assistance from John
change in mineralogy may reflect declining temperature of Humphrey and Jim Reynolds (Fluidinc) on the sulfur isotope and fluid
inclusion studies, respectively, are gratefully acknowledged. GPL also
hydrothermal fluids through time or a progressive increase of acknowledges invaluable help from the Society of Economic Geologists
the oxidation state of the fluids, perhaps due to fluid mixing with through Terrones and McKinstry student research grants, from a
a more oxidized, meteorically derived fluid. The presence of Newcrest Resources Economic Geology Fellowship, and from Sarah
mushketovite with some calcic alteration mineral associations Gleeson (University of Alberta) for discussions and access to facilities.
may reflect a late temperature increase during an overall cooling
trend or a change to slightly more reducing conditions during an References
overall higher oxidation trend.
The El Espino mining district appears to lack bodies of
Åberg G, Aguirre L, Levi B, Nyström JO (1984) Spreading-subsidence
massive magnetite–apatite that are present at a number of and generation of ensialic marginal basins: an example from Early
Andean deposits including Manto Verde (Rieger et al. 2010) Cretaceous of Central Chile. In: Kokellar BP, Howells MF (eds)
and the Marcona-Mina Justa deposits in Peru (Chen et al. Volcanic and associated sedimentary and tectonic processes in mod-
2011). Such deposits may be absent or may be present below ern and ancient marginal basins, vol 16. Geological Society of
London, Special Publication, London, pp 185–193
the current level of erosion. Aguirre L, Egert E (1965) Cuadrángulo Quebrada Marquesa, Provincia
Hydrothermal fluid flow in the El Espino mining district de Coquimbo, vol 15. Instituto de Investigaciones Geológicas, Carta
was channeled along high angle, brittle faults and related Geológica de Chile, p 92
extension fractures that developed in a right-stepping jog Allmendinger R, Marrett R, Cladouhos T (2001) FaultKin Version
1.2.2 (for Windows), computer program and user’s manual (for
locally along a major north-striking dextral fault system. vol 1.1), 29 p. http://www.geo.cornell.edu/geology/faculty/RWA/
Mantos were developed within the mixed volcano-sedimentary programs.html
sequence of the district, particularly beneath relatively imper- Arancibia G (2004) Mid-cretaceous crustal shortening: evidence from a
meable limestone beds. Mineralized breccias along structures regional-scale ductile shear zone in the coastal range of central Chile
(32 degrees S). J S Am Earth Sci 17:209–226
and particular beds are locally present in the El Espino district Barton M, Johnson DA (1996) Evaporitic source model for igneous
but are not well developed at the current level of erosion. related Fe oxide-(REE–Cu–Au–U) mineralization. Geology 24:
Distal, hydrolytic, and argillic stage calcite–barite veins with 259–262
Miner Deposita (2014) 49:235–259 259

Barton M, Johnson D (2000) Alternative brine sources for Fe oxide Morata D, Aguirre L (2003) Extensional lower cretaceous volcanism in
(–Cu–Au) systems: implication for hydrothermal, alteration and the Coastal Range (29 degrees 20′–30 degrees S), Chile: geochemistry-
metal. In: Porter TM (ed) hydrothermal iron-oxide cooper-gold and petrogenesis. J S Am Earth Sci 16:459–476
and related deposits: A global perspective. Australian Mineral Mpodozis C, Ramos VA (1990) The Andes of Chile and Argentina In:
Foundation, Adelaide, pp 43–60 Ericksen GE, Cañas Pinochet MT, Reinemud JA (eds) Geology of
Benavides J, Kyser TK, Clark A, Oates C, Zamora R, Tarnovschi C, the Andes and its relation to hydrocarbon and mineral resources,
Castillo B (2007) The Mantoverde Iron-Oxide-Copper-Gold district, Circumpacific Council for Energy and Mineral Resources, Earth
III Region, Chile: The role of regionally derived nonmagmatic fluids Sciences Series, 11:59–90
in chalcopyrite mineralization. Econ Geol 102:415–440 Ohmoto H, Goldhaber M (1997) Sulfur and carbon isotopes. In: Barnes
Campbell AR, Larson PB (1998) Introduction to stable isotope applica- HD (ed) Geochemistry of hydrothermal ore deposits (3rd edn).
tions in hydrothermal systems. Rev Econ Geol 10:173–193 Wiley, New York, pp 517–612
Chen H, Clark A, Kyser TK, Ullrich T, Baxter R, Chen Y, Moody T Ohmoto H, Rye RO (1979) Isotopes of sulfur and carbon. In: Barnes HL
(2011) Evolution of the giant Marcona–Mina Justa iron oxide- (ed) Geochemistry of hydrothermal ore deposits. Wiley, New York,
copper-gold district. South-Central Peru Econ Geol 105:155–185 pp 509–567
Chiaradia M, Banks D, Cliff R, Marschik R, de Haller A (2006) Origin of Parada MA, Nystrom J, Levi B (1999) Multiple sources for the Coastal
fluids in iron oxide–copper–gold deposits: constraints from delta Cl- Batholith of central Chile (31–34 degrees S): geochemical and
37, Sr-87/Sr-86(i) and Cl/Br: Mineralium Deposita 41: 565–573 Sr–Nd isotopic evidence and tectonic implications. Lithos 46:
Correa A (2003) El Espino, un nuevo depósito del tipo Fe Cu (Au) en 505–521
Chile. Comuna de Illapel, Region de Coquimbo. MS thesis, Parada MA, Féraud G, Fuentes F, Aguirre L, Morata D, Larrondo
Universidad Catolica del Norte, Antofagasta, Chile, 69 pp P (2005) Ages and cooling history of the Early Cretaceous
de Haller A (2006) The Raul-Condestable iron oxide copper–gold de- Caleu pluton: testimony of a switch from a rifted to a
posit, Central Coast of Peru. PhD thesis, Geneva, Switzerland, compressional continental margin in central Chile. J Geol
University of Geneva, Terre et Environment, vol 58, 123 p Soc Lond 162:273–287
de Haller A, Corfu F, Fontbote L, Barra F, Chiaradia M, Frank M, Zuniga Pollard PJ (2006) An intrusion-related origin for Cu–Au mineralization in
Alvarado J (2006) Geology, geochronology, and Hf and Pb isotope iron oxide–copper–gold (IOCG) provinces. Miner Deposita 41:
data of the Raúl-Condestable iron oxide–copper–gold deposit, 179–187
Central Coast of Peru. Econ Geol 101:281–310 Rieger AA, Marschik R, Diaz M, Hölzl S, Chiaradia M, Akker B,
Explorator Resources Report (2011) Preliminary Assesment on El Espino- Spangenberg J (2010) The hypogene iron oxide copper–gold min-
Venus Property, Region IV, Chile. NI 43–101 Technical Report eralization in the Mantoverde district. North Chile Econ Geol 105:
Gana P, Zentilli M (2000) Historia termal y exhumación de intrusivos de 1271–1299
la Cordillera de la Costa de Chile Central. Congreso Geológico Rieger A, Marschik R, Diaz D (2012) The evolution of the hydrothermal
Chileno IX, Puerto Varas, Chile, vol 2, pp. 664–668 IOCG system in the Mantoverde district, northern Chile: new evi-
Goldstein R, Reynolds T (1994) Systematics of fluid inclusions in diage- dence from microthermometry and stable isotope geochemistry.
netic minerals: SEPM Short Course, 31 p Miner Deposita 47:359–369
Groves D, Bierlein F, Meinert L, Hitzman M (2010) Iron oxide copper– Rivano S, Sepulveda H (1991) Hoja Illapel, Region de Coquimbo In:
gold (IOCG) deposits through Earth history: implications for origin, Mpodozis C (ed) Carta Geologica de Chile, scale 1:250,000
lithospheric setting, and distinction from other epigenetic iron oxide Sillitoe RM (2003) Iron oxide–copper–gold deposits: an Andean view.
deposits. Econ Geol 105:641–654 Miner Deposita 38:787–812
Haynes DW (2000) Iron oxide copper (−gold) deposits: their position in Tatsumi Y, Eggins S (1995) Subduction zone magmatism. Blackwell
the ore deposit spectrum and modes of origin. In: Potter TM (ed) Science, Boston, p 211
Hydrothermal iron oxide copper–gold & related deposits: a global Vergara M, Levi B, Nystrom J, Cancino A (1995) Jurassic and
perspective. Australian Mineral Foundation, Adelaide, pp 71–90 Early Cretaceous island arc volcanism, extension and subsi-
Hitzman MW (2000) Iron oxide–Cu–Au deposits: what, where, when, dence in the coast range of Central Chile. Geol Soc Am Bull
and why. In: Porter TM (ed) Hydrothermal iron oxide–copper–gold 107:1427–1440
and related deposits: a global perspective. Australian Mineral Williams PJ, Barton MD, Johnson DA, Fontboté L, de Haller A,
Foundation, Adelaide, pp 9–25 Mark G, Oliver, NHS, Marschik R (2005) Iron oxide copper–
Levi B, Aguirre L (1981) Ensialic spreading-subsidence in the Mesozoic gold deposits: geology, space–time distribution, and possible
and Paleogene Andes of Central Chile. J Geol Soc Lond 138:75–81 modes of origin: Economic Geology 100th Anniversary Vol-
Lopez G (2012) The El Espino iron oxide copper gold (IOCG) district, ume, pp. 371–405
Coastal Cordillera of North-Central Chile, PhD thesis, Colorado Wilson M (1992) Magmatism and continental rifting during the opening
School of Mines, 120 p of the South Atlantic Ocean: a consequence of Lower Cretaceous
Marschik R, Fontbote L (2001) The Candelaria–Punta del Cobre Superplume activity? In: Storey B, Alabaster T, Pankhurst R (eds)
iron oxide Cu–Au (–Zn–Ag) deposits. Chile Econ Geol 96: Magmatism and the causes of continental break-up. Geol Soc Spec
1799–1826 Pub 68: 241–255

Potrebbero piacerti anche