You are on page 1of 17

Miner Deposita (2012) 47:483499

DOI 10.1007/s00126-011-0383-2

ARTICLE

The boron isotope geochemistry of tourmaline-rich alteration


in the IOCG systems of northern Chile: implications
for a magmatic-hydrothermal origin
Fernando Tornos & Michael Wiedenbeck &
Francisco Velasco

Received: 8 July 2011 / Accepted: 18 August 2011 / Published online: 27 September 2011
# Springer-Verlag 2011

Abstract Hydrothermal tourmaline is common in the iron


oxide-copper-gold (IOCG) deposits of the Coastal Cordillera
of Chile where it occurs as large crystals in the groundmass of
magmatic-hydrothermal breccias, such as in the Silvita or
Tropezn ore bodies, or as small grains in replacive bodies or
breccia cement in the ore-bearing andesite, as seen at the
Candelaria or Carola deposits. Tourmaline shows strong
chemical zoning and has a composition of schorldravite
with significant povondraite and uvite components. The
observed boron isotope composition is fairly variable,
between 10.4 and +6.0 with no major differences
among the different deposits, suggesting a common
genetic mechanism. The 11B values are significantly
lower than those of seawater or marine evaporites and very
similar to those of younger porphyry copper deposits and
volcanic rocks in the region, indicating that the boron has
a common, likely magmatic, origin. The predominant
boron source was ultimately dewatering of the subducting
slab with a significant contribution derived from the

overlying continental basement. The range of 11B values


is between those of the porphyry copper deposits and the
porphyry tin deposits of the Andes, suggesting that the
IOCG mineralization might be genetically related to fluids
having more crustal contamination than the porphyry
copper deposits; such an interpretation is at odds with
current models that propose that the Andean IOCG
deposits are related to juvenile melts or to the circulation
of basinal brines. Furthermore, the obtained 11B data are
markedly different from those of the tourmaline in the
Carajs IOCG district (Brazil), suggesting that IOCGs do
not form by a unique mechanism involving only one type
of fluids.

Editorial handling: B. Lehmann

One of the major controversies concerning the origin of


iron oxide-copper-gold (IOCG) deposits relates to the
source of the ore-forming fluids. Many IOCG districts are
within belts rich in both plutonic rocks (which are broadly
contemporaneous with the mineralizing event) as well as
shallow marine sequences that locally contain evaporites.
Neither stable isotope nor radiogenic isotope geochemistry
has been able to unravel this conundrum because most can
be interpreted as fluids having been equilibrated either with
magmatic or basinal reservoirs. This is the case for the
Cloncurry district in Australia and the Skellefte district in
Sweden (Mark et al. 2006; Williams et al. 2003, 2005). Thus,
current models for IOCG systems fluctuate between wholly
magmatic (Pollard 2006; Tornos et al. 2010), those that
invoke circulation of basinal brines with the igneous rocks

F. Tornos (*)
Instituto Geolgico y Minero de Espaa,
Ros Rosas, 23,
28003 Madrid, Spain
e-mail: f.tornos@igme.es
M. Wiedenbeck
Helmholtz Centre Potsdam,
GFZ German Research Centre for Geosciences,
Telegrafenberg C161,
14473 Potsdam, Germany
F. Velasco
Departamento de Mineraloga y Petrologa,
Universidad del Pas Vasco,
Campus de Leioa, Sarriena s/n,
48940 Leioa Vizcaya, Spain

Keywords Boron isotopes . Tourmaline . IOCG . Chile .


Andes

Introduction

484

acting exclusively as heat engines (Barton and Johnson


1996, 2000) or intermediate models in which a magmatichydrothermal event is followed by the invasion of the system
by basinal brines (Marschik and Fontbot 2001a).
In surficial environments, boron can be contained in
clays and borates, but in deep settings, it is only hosted in
small amounts in white micas and sillimanite and in more
significant amounts in axinite, vonsenite, fluoborite, datolite, dumortierite, and tourmaline. Tourmaline is the most
common B-bearing mineral and is found in magmatic,
metamorphic, and hydrothermal rocks. It is very stable and
resistant to diffusion (Nakano and Nakamura 2001; Bebout
and Nakamura 2003), thus archiving information about
original conditions of crystallization. Tourmaline is a
particularly useful mineral for studies of ore deposits
because its chemical composition is sensitive to changes
occurring in the geochemical environment during ore
formation, providing a petrogenetic indicator able to
discriminate among mineralizing regimes (Palmer and
Slack 1989; Slack et al. 1993). The tourmaline crystal
structure can accommodate a wide variety of cations with
different size and valence, typically producing a complex
typology and variable compositions. It is a mineral widespread in magmatic-hydrothermal mineralized systems occurring in breccias and veins related to copper, molybdenum, tintungsten, and gold deposits (Jiang et al. 1998; Lehmann et al.
2000; Krienitz et al. 2008; Wittenbrink et al. 2009).
Boron isotope geochemistry is a powerful tool for
constraining the source of boron and, by extension, the
origin of fluids responsible for B-rich alteration. Boron is
not bound directly to cations, and its isotope ratio is not
influenced by redox state, salinity of the fluid, or biogenic
activity; only temperature, pH, and related aqueous speciation seem to have a significant control over 11B values
(Palmer and Slack 1989). Boron is incompatible in most
magmatic processes not involving the crystallization of
tourmaline or axinite (Bebout et al. 1999).
Boron isotopes show a wide range in nature and the
11 10
B/ B ratios of different reservoirs are markedly different
(Barth 1993) making them an excellent tracer for distinguishing between continental and marine boron reservoirs.
Continental rocks and their derived evaporites have
negative 11B values, whereas the large positive value of
seawater (+39.61, 0.04; Foster et al. 2010) causes
marine sediments to be strongly enriched in 11B (Chaussidon
and Albarde 1992; Palmer and Swihart 1996; Kasemann et
al. 2000). Boron isotopes have been used for tracking mass
and fluid transfer during plate convergence and subduction,
mostly because the different reservoirs participating in
magma generationcontinental crust, mantle, sediments,
and altered upper oceanic crusthave distinct B isotope
signatures (Benton et al. 2001; Tonarini et al. 2007). Another
strength of B isotope geochemistry is that the fractionation

Miner Deposita (2012) 47:483499

factors between tourmaline and water at high temperatures


that dominate in magmatic-hydrothermal systems (>400C)
are low, ca. 1.81.0 -units (Meyer et al. 2008), with only a
minimal pressure dependence.
Boron isotopes can be a valuable tool for discerning the
origin of fluids responsible of IOCG formation, since the
current genetic discussion fluctuates between fluids derived
from evaporitic basins or magmatic-hydrothermal systems.
Some studies have documented the occurrence of tourmaline in porphyry CuMo and WSn deposits (Frikken et al.
2005; Klemm et al. 2007; Wagner et al. 2009; Wittenbrink
et al. 2009), but there is limited information on tourmaline
related to IOCG deposits. The lack of tourmaline in most
IOCG systems has precluded its systematic utilization.
Until now, the studies dealing with B isotopes in IOCG
deposits are those of Dreher et al. (2008) and Xavier et al.
(2008) in the Carajs Province, Brazil. Here, high 11B
values are indicative of the ore-forming brines having been
derived from marine evaporites; lower 11B values are
interpreted as mixed signals which some of the B came
from felsic crust. These deposits, as is the case in most
IOCG districts, occur in highly deformed Precambrian belts
where superimposed events of hydrothermal activity related
to magmatism or metamorphism can modify the original
isotope composition.
Tourmaline-rich rocks are widespread in magmatichydrothermal systems of the Andes, including porphyry
copper and epithermal deposits (Sillitoe and Sawkins
1971), but only recently has tourmaline been described as
an abundant hydrothermal mineral in IOCG-like systems
(Tornos et al. 2010); previous works (Benavides et al. 2007;
Marschik and Fontbot 2001a) have quoted it as an
accessory phase in some zones of the IOCG deposits of
the Copiap area. This study reports the first tourmaline
compositional and 11B data for several IOCG deposits of
the Coastal Cordillera of the Andes. Tourmalinite is present
in both the ore-related hydrothermal alteration zones, and
associated igneous rocks and has been found in the
Candelaria, Carola, Tropezn, and Silvita deposits; these
ore bodies are Early Cretaceous in age, and the primary
features are virtually unmodified since the time of their
formation. Tourmaline has been also described in other
IOCG systems (e.g., Mumin et al. 2007).
Geologic setting and description of the tourmaline-rich
rocks
The IOCG and magnetite-apatite deposits of the Coastal
Cordillera of northern Chile and Peru (ca. 2,000 Mt of ore)
are within the so-called Cretaceous Iron Belt (Henriquez et
al. 1994; Vivallo et al. 2000; Sillitoe 2003) that forms a
narrow (ca. 2540 km), 1,000 km long corridor from near
Santiago, Chile, to near Lima, Peru. This belt shows a close

Miner Deposita (2012) 47:483499

spatial relationship with the Atacama Fault System (AFS).


The AFS includes multiple crustal-scale, deeply plunging,
NS anastomosing, sub-parallel shear zones and faults
accompanied by NE and SW subsidiary structures that have
been active since at least the Early Jurassic (Brown et al.
1993; Cembrano et al. 2005). The AFS is parallel to the
modern subduction zone to the west and has accommodated
significant deformation induced by oblique subduction; it
also channelized regional magmatic-hydrothermal activity
and mineralization.
The ore bodies are hosted by shallow marine to subaerial
calc-alkaline andesite of Late Jurassic to Early Cretaceous
age that was erupted in an arc-back arc environment during
early stages of plate convergence. This andesite is earlier to
coeval with plutonism of intermediate to felsic composition. Dated magnetite-apatite and IOCG deposits formed in
the 13298 Ma range (Mathur et al. 2002; Diaz et al. 2003;
Gelcich et al. 2005; Benavides et al. 2007; Tornos et al.
2010). As a whole, the Cretaceous Iron Belt seems to have
formed in response to a major tectonic breakup that took
place in the Early Cretaceous, when there was a major
change in the convergence vector of the Nazca Plate with
the South American Plate, with orthogonal subduction
changing to an oblique one with maximum compressive
stress-oriented NW-SE (Scheuber and Andriessen 1990).
This shift produced a change in the style of magmatism,
which gradually evolved from volumetrically large, fissuredominated, flood-like juvenile andesite of Late Jurassic to
Early Cretaceous age that was extruded in an extensional
setting, to intrusion of the Coastal Batholith that was
emplaced as several large laccoliths in pull-apart structures
within a transpressive setting (Grocott and Taylor 2002).
The onset of left lateral, strike-slip deformation was synchronous with eastward migration of the magmatic arc (Taylor et
al. 1998). The Cretaceous Iron Belt is the westernmost ore
province of the South American Cordillera; eastward, it is
bound by younger sub-parallel, NS trending Cu, Ag, and
Sn belts (e.g., Sillitoe 1976).
The Iron Belt hosts hundreds of small copper prospects,
most of which are hematite-rich bodies that comprise stockwork zones and massive replacements related to faults and
breccia bodies. The prospects are associated with a sericite
chloritecarbonate hydrothermal alteration assemblage that
postdates an earlier and deeper amphiboleK feldspar
magnetite assemblage (Benavides et al. 2007; Rieger et al.
2011), of which the Mantoverde (140 Mt at 0.63% Cu),
Carmen de Cobre (71 Mt at 0.65% Cu), and Teresa de Colmo
(92 Mt at 0.53% Cu) deposits are the most significant.
The magnetite-rich deposits replace volcanic and plutonic
rocks and are associated with a Kfeldsparactinolitebiotite
alteration assemblage; they probably formed at higher depths
and temperatures than the hematite-rich ones. The major
deposits near Copiap (Fig. 1) include the world-class

485

Candelaria deposit (470 Mt at 0.95% Cu), Carola (20 Mt at


1.4% Cu), Alcaparrosa, and Santos. These deposits are
hosted by the Punta del Cobre Formation of Early
Cretaceous age, that consists of ca. 500 m of shallow marine
volcaniclastic andesite deposited in a back arc basin
(Marschik and Fontbot 2001b). The Punta del Cobre
Formation is overlain by the Late Berriasian to Early Albian
Chaarcillo Group (7002,000 m), which includes both
volcanic and sedimentary rocks; the ore is predominantly
hosted by the andesite and to a much lesser extent by the
Chaarcillo Group. Both the Punta del Cobre Formation and
the Chaarcillo Group formed in a NS trending, elongate,
ensialic back arc shallow marine pull-apart basin that formed
during onset of the oblique subduction, being intimately
related to the deformation that accommodated the curvature
of the Atacama Fault System (Taylor et al. 1998). Rocks
belonging to the Chaarcillo Group interfinger with the
subaerial andesite of the Bandurrias Formation, representing
the volcanic facies of the magmatic arc to the east.
Mineralization is associated with several steeply dipping,
NNWSSE structures and a large aureole of hydrothermal
alteration that has replaced the andesite. The alteration
assemblages are biotitemagnetiteactinoliteK feldspar
epidotequartz in the deep zones and biotiteK feldspar
quartz in the shallow ones; the sulfide assemblage consists of
chalcopyrite and pyrite with magnetite commonly replacing
earlier hematite. The overlying limestone is irregularly
replaced by a nearly barren calcic skarn (Marschik and
Fontbot 2001a). Marschik and Fontbot (2001a) reported
minor tourmaline in the earliest hydrothermal assemblage,
but we found in our study that it is fairly common within the
deep parts of the deposit.
Near Taltal, about 200 km north of the Punta del Cobre
district, some IOCG-like deposits share many features with
those of the Copiap area, but differ in being related to
large tourmaline breccia pipes. Tropezn is a small CuMo
mine that works a zone of massive CaFeK alteration
replacing diorite dated at 110.02.1 Ma (Tornos et al.
2010). The host diorite intrusion is pervasively altered with
irregular replacement of the primary assemblage by a
potassiccalcic alteration of K-feldspar, biotite, actinolite,
and magnetite. A more intense CaFeK alteration occurs
as a dome-shaped body in the core of the intrusion,
dominated by actinolite with accessory magnetite, tourmaline, quartz, and sulfides (mostly pyrite, chalcopyrite, and
molybdenite). The uppermost part grades into an epidote
quartzK-feldspar rock with locally abundant anhydrite.
The CaFeK alteration postdates the emplacement of two
tourmaline breccia pipes; these pipes contain monolithic
fragments of the host diorite with a potassiccalcic
alteration assemblage supported and replaced by tourmaline
and quartz. Groundmass minerals include coarse-grained
tourmaline, magnetite, molybdenite, pyrite, chalcopyrite,

486

Miner Deposita (2012) 47:483499

Fig. 1 Geologic map of the


Coastal Cordillera of Chile
showing the location of the
studied IOCG mines. Modified
from Tornos et al. (2010)

calcite, and local anhydrite. Similar to Tropezn, the Silvita


prospect has at least two monolithic tourmaline breccia
pipes but lacks visible actinolite-rich alteration and related
mineralization. Tourmaline is intergrown with quartz and
chalcopyrite and locally supports highly heterometric fragments of the host diorite. Late fluids that invaded the
system seem to have had minimal influence in all these
deposits because retrograde alteration assemblages are
accessory and volumetrically minor.
The age of the epizonal Coastal Batholith (ca. 131
104 Ma; Berg et al. 1983; Dallmeyer et al. 1996; Gelcich et
al. 2005) is similar to that of the ore deposits. This
polyphase batholith is characterized by rather primitive

isotopic signatures, with initial 87Sr/86Sr ratios close to


0.703 and Ndi values of +1.8 to +2.8 (Marschik et al.
2003a). A clear genetic relationship between intrusive rocks
and mineralization is found in the Tropezn and Silvita
mines, where the ore body is directly related to tourmaline
breccia pipes, and hydrothermal alteration is restricted to
discrete diorite intrusions. In the Punta del Cobre district,
the situation is not so straightforward, since the ore deposits
are a few kilometers from the intrusive rocks and probably
replace a roof pendant of volcanic rocks. However, in this
district, no report exists of a relationship of the deposits and
a specific intrusive unit or a direct link with a known
magmatic-hydrothermal system.

Miner Deposita (2012) 47:483499

487

Fluid composition and stable isotope data are somewhat


ambiguous. At Candelaria, fluid inclusions indicate the
circulation of CO2-bearing fluids having salinities between
25 and 47 wt.% (Marschik and Fontbot 2001a) and
suggest the presence of a first pulse of magmatic brines
that was gradually replaced by Ca-rich fluids of basinal
derivation. However, both metals and sulfur are interpreted
to be of magmatic origin. Sulfur isotope data show highly
variable signatures: while Tropezn and Candelaria have
close to 0 values (Marschik and Fontbot 2001a)
compatible with a magmatic derivation, Mantoverde displays highly variable data (6.8 to +11.2) which
suggest derivation of most of the sulfur from the adjacent
sedimentary rocks. Initial Os isotope signatures support a
basinal derivation of the fluids (Mathur et al. 2002), whereas
37Cl values (0.2004.6) of fluid inclusions are also
consistent with mixing of magmatic brines with a dominant
fluid equilibrated with evaporitic sources (Chiaradia et al.
2006). Lead isotopes from Candelaria are consistent with the
lead being derived from mixed crustal and mantle sources
(Marschik et al. 2003b). At Tropezn, ore-forming fluids
are both CO2-poor and high saline (5070 wt.% NaCl eq.)
NaKCa brines that are interpreted as having exsolved
from the underlying intrusion above the two phase curve
(Tornos et al. 2010). Thus, the geology and geochemistry
of these deposits is consistent with their derivation from
one of two contrasting fluid sources: either the intrusive
rocks, the Coastal Batholith or marine strataperhaps
including evaporitesof the Chaarcillo Formation. Both
systems could have coexisted in time and space with the
ore-forming processes.

complex irregular zonation with evidence of multiple


reabsorption events; a patchy zoning is found locally. Also
present there are embayments with cleaner, more transparent and colorless outer zones. The occurrence of abundant
inclusions of chalcopyrite and magnetite indicates that these
external zones grew during the ore-forming stage. This
feature differs from that in the crystals from the Silvita
mine, which display more homogeneous cores with only
weak oscillatory zoning surrounded by narrow, dark-brown
rims. Tentatively, it could be suggested that cores of the
large prismatic crystals at TropeznSilvita and the disseminated fine-grained crystals at Carola-Candelaria are equivalent and belong to a first pulse of tourmalinization,
whereas the pale brown regrowths, patches, and rims
correlate with a second event of tourmaline growth related
to ore formation (Fig. 2).
Tourmaline is also abundant in several sub-volcanic
monzodioritetonalite plugs that crosscut andesite of the
Punta del Cobre Formation. These plugs are related to
NWSE trending dikes and show widespread albitization
with replacement of the groundmass and most phenocrysts by albite and lesser quartz, actinolite, magnetite,
and minor scapolite and zircon. The most altered zones
host irregular breccia bodies with 210 mm angular
fragments of texturally variable plutonic rock having
similar matrix, but a fine-grained assemblage containing
abundant quartz and tourmaline. This tourmaline is
petrographically and chemically identical to that in the
nearby mineralized zones.

Tourmaline occurrences

We used the Cameca ims 6f SIMS instrument in Potsdam,


Germany, to obtain in situ 11B determinations of selected
domains of tourmaline crystals identified in polished thin
sections. Prior to analysis, the samples were cleaned
ultrasonically for 5 min in high-purity ethanol, sputtercoated with a 35 nm-thick film of high-purity Au, and
placed in an in-house designed high-vacuum storage
chamber (Wiedenbeck et al. 2004) at least 1 day in advance
of analysis. Calibration of the SIMS analyses was based on
three natural tourmaline samples whose 11B values had
been characterized as parts of reference material development projects: Harvard Mineralogical Museum #112566
schorl and #108796 dravite (Dyar et al. 2001; Leeman and
Tonarini 2001) and B4 schorl (Gonfiantini et al. 2003;
Tonarini et al. 2003). These three reference materials were
analyzed a total of 42 times during the 4 days during which
sample analyses were obtained. These data yield a
repeatability of 1.2 (1), for overall instrumental mass
fractionation, which is the predominant source of uncertainty in our analytical set-up; this is our best estimate of
the robustness of the isotopic data.

The petrographic and compositional features of the tourmaline vary depending on location and stage of deposition.
In detail, two groups of tourmaline are distinguished, one
from Tropezn and Silvita and another one from Candelaria
and Carola.
Fine grained, uniformly submicroscopic tourmaline from
the Candelaria and Carola deposits is euhedral and shows a
poorly defined chemical zonation traceable by a very thin,
dark rim. Typically, it occurs as small disseminated euhedral
grains (ca. 1530 m), only locally forming true tourmalinite (>30% tourmaline). These tourmaline crystals have a
pleochroism ranging from greenish blue to pale yellow or
colorless. In contrast, the tourmaline in Tropezn and
Silvita is coarse grained, up to several millimeters in
length, prismatic to acicular, and shows a well-defined
chemical zonation. Tourmaline from the Tropezn breccia
displays an intense pleochroism from brown-yellow to dark
green in the core, to pale brown in the regrowths and rims.
The cores have abundant fluid inclusions and show a

Analytical methods

488

Miner Deposita (2012) 47:483499

Fig. 2 Photographs of tourmaline-bearing rocks from the IOCG


deposits of the Coastal Cordillera. a Acicular fine-grained greenbrown tourmaline, with a matrix of quartz, plagioclase, and chlorite
from the Candelaria deposit (sample CAND-1); b breccia with
fragments of altered andesite supported by fine-grained tourmaline
and quartz and cut by late veins of sulfides. Width of view 12 cm
Candelaria; c late tourmaline intergrown with K-feldspar, quartz,
actinolite, epidote, and sulfides and magnetite (black). Tropezn Mine.
Sample TR-145; d breccia from the Tropezn deposit with fragments
of diorite affected by potassic alteration and supported by a cement of
tourmaline, quartz, pyrite, magnetite and, molybdenite; e tourmaline
intergrown with sulfides (chalcopyrite and pyrite) and quartz.

Tropezn Mine. Sample TR4; f andesite replaced by tourmaline from


the Carola mine. Tourmaline grains locally contain minute crystals of
magnetite. Sample 330.1; g tourmaline breccia from the Silvita mine
including large fragments of diorite with potassic alteration and
supported by quartz and tourmaline. Widespread supergene alteration
highlight the fragments; h prismatic tourmaline from the Silvita
breccias exhibiting fine-scale chemical zonation with greenish to
brown cores and pale-brown rims, in a cement of actinolite, quartz,
and magnetite. Microveins and patches correspond to late colorless to
green tourmaline. All microphotographs are 1 cm across and with
plane polarized transmitted light

The SIMS analyses employed an 800-pA, nominally


12.5-kV 16O primary beam that was focused to a ca. 5 m
diameter spot on the sample surface. Peak stepping
sequence for our mono-collection instrument consisted of

5.95 m/e (0.1 s per cycle; needed for the 180 s duration
presputtering used to clean the sample surface and to
establish equilibrium sputtering conditions), 6Li (1 s), 10B
(4 s), and 11B (2 s). A total of 40 cycles of data were

Miner Deposita (2012) 47:483499

collected, leading to a total analysis time, including


presputtering, of 9 min per spot. The count rate on the
11
B peak was typically around 350 counts/s, leading to an
observed uncertainty on single measurements typically of
0.5 (1).
Electron microprobe (EPMA) analyses were performed
using an automatized Camebax MBX (Cameca) instrument at the Country Basque University. The analytical
conditions were 15 kV accelerating voltage, 10 nA beam
current, 12 m beam size, and 30 s peak counting time.
Data were managed using a conventional ZAF routine,
and data were reduced using a conventional ZAF
procedure. Natural silicate minerals and synthetic reference material (Spi Structure Probe Inc.) were used for
calibration.
Composition of the tourmaline
The general formula of the tourmaline may be expressed as
XY3Z6(T6O18)(BO3)3V3W with X containing Na, K, Ca, or
vacant; Y representing Fe2+, Mg, Mn2+, or Al, Fe3+, Cr3+,
V3+, Ti4+, and Li; Z for Al, Mg, Fe3+, Cr3+, V3+; T=Si, Al;
B=B; V=OH, O and W=OH, F, O (Hawthorne and Henry
1999). For the normalization of the structural formula, the
analyses were recalculated on the basis of 15 (T+Z+Y)
cations (Henry and Dutrow 1996), assuming the theoretical
amount of 3 a.p.f.u. for boron despite that some authors
have suggested that excess of B may exist in T positions
(Tagg et al. 1999). For the calculation of the general
formula and the corresponding end-member components,
we have used the procedure of Pesquera et al. (2008).
Representative compositions of tourmaline from ca. 300
electron microprobe analyses from Tropezn (n=231),
Silvita (n=32), Candelaria (n=16), and Carola (n=20)
mines are shown in Table 1. The most significant variations
are observed for FeO, MgO, CaO, and Na2O, although the
TiO2 and V2O3 variations are also noteworthy. Fluorine
contents are up to 0.2 wt.%. The K2O, MnO, and Cr2O3 are
low and close to the detection limit in all analyses. Oxide
totals for most of the analyses are between 84 and 87 wt.%.
Most tourmaline analyses fall within the field of tourmaline
from Fe3+-rich quartz-tourmaline rocks, calc-silicate rocks,
and metapelite according to Henry and Guidotti (1985)
(Fig. 3a), and classify in the alkali group with significant
Ca contents and low X-site vacancies (Fig. 3b). In general,
most compositions of tourmaline seem to be influenced by
the chemistry of the host rock, defining separate fields in
terms of major elements with rims showing wider compositional variation than the cores. Probably, this variability
could reflect the changing nature of the mineralizing fluids
able to cause the observed oscillatory zoning, the embayments by dissolution and the later regrowths of the
tourmaline crystals.

489

All tourmaline studied from the Coastal Cordillera


belongs to Fe- and Mg-rich varieties (schorldravite series),
but the incorporation of other cations (Ca, V, Ti, Cr, or Mn)
produces numerous substitutions in the form of mineral
solid solutions. Tourmaline from the Silvita mine has the
highest concentrations of Fe and the lowest Mg contents,
suggesting a major povondraite component. At the edge of
the nearly optically opaque outer rims, the iron content is
up to 21.2 wt.%; however, it is relatively poor in Al.
Tourmaline from the Tropezn and Silvita mines also
belongs to the schorldraviteuvite series but approaches
the Mg-rich end-member with Mg/(Fe+Mg) ratios from 0.5
to 0.9. EPMA data indicate that much of the chemical
variability of the Tropezn tourmaline is in the atomic
ratios of Mg# (Mg/(Fe + Mg)), Ca/(Ca + Na), and Na
(Fig. 3ce); however, they have an important uvite
component with Ca/(Na+Ca) ratios between 0.2 and 0.7.
This tourmaline can have noteworthy high V2O3 contents,
which can reach values up to 0.5 wt.% (average 0.15 wt.%)
while the TiO2 and F contents yield averages of 1.1 wt.%
and 0.2 wt.%, respectively.
Tourmaline from the Candelaria and Carola mines also
belongs to the draviteschorl series and exhibits an
intermediate composition with depletion in Ca and a
relative enrichment in Al that is more pronounced in the
tourmaline from Carola (Table 1 and Fig. 3a). It is
important to note that the ferromagnesian a.p.f.u. contents
are significantly less than three in the Candelaria and
Carola tourmaline, probably revealing the presence of Al,
Fe3+, vacancies, and/or others cations in the Y site.
Normalization of the tourmaline to a stoichiometric
value of 3 B of a.p.f.u. generally gives a significant excess
in the cation charge, something that suggests that a
significant part of the Fe is in the form of Fe3+. This agrees
with the observed positive correlation between Al (mainly
below 6 a.p.f.u) and the (Fe+Mg) content and is consistent
with some involvement of the FeAl-1 and/or AlO(Fe(OH))1 exchange vectors (Fig. 4e). The projection of the data
along the povondraite-oxy-dravite join also suggests an
important influence of the MgFe-1 exchange, mainly in the
outer rims of the tourmaline. This is easy to observe on a
RX site vacancy Ca+Ti vs. AlX site vacancy +Ca plot.
Here, R represents the sum of Mg and Fe and the
tourmaline analyses fall roughly parallel to the FeAl-1 and
AlO(R(OH))-1 exchange vectors. This plot enables to
evaluate the significant proportion of the povondraite
component because the dravite and schorl end-members
fall projected in a same point along the oxy-dravite
povondraite join (Henry et al. 2008). This confirms that the
Fe enrichment observed in the crystal rims could be related
with an increase of the aFe3+ during ore deposition or a
shift in the redox state of the fluids. The abundance of
magnetite and the presence of anhydrite in the hydrother-

29.07

6.79

0.01

10.07

0.04

0.13

1.30

1.87

0.07

0.24

Al2O3

MgO

MnO

FeO(t)

Cr2O3

V2O3

CaO

Na2O

K2O

86.53

0.23

0.06

1.88

1.25

0.18

0.01

11.05

0.04

6.59

85.63

0.12

0.08

1.66

1.77

0.12

0.00

10.01

0.00

6.88

28.17

0.79

36.15

CAND1
Core

85.73

0.04

0.05

1.60

1.51

0.12

0.00

10.04

0.03

6.51

29.27

0.65

35.82

CAND1
Core

0.000

5.686

0.000

0.061

1.678

0.001

1.397

0.005

0.017

0.230

0.601

0.016

0.124

0.153

Al (total)

Al (T)

Al (Z)

Al (Y)

Ti

Mg

Mn

Fe (t)

Cr

Ca

Na

Vac

6.033

0.158

0.121

0.012

0.608

0.222

0.023

0.001

1.540

0.005

1.638

0.126

0.000

5.533

0.000

5.533

6.066

0.125

0.065

0.018

0.539

0.318

0.016

0.000

1.404

0.000

1.721

0.100

0.000

5.572

0.000

5.572

6.005

0.198

0.023

0.010

0.521

0.270

0.016

0.000

1.407

0.004

1.626

0.082

0.000

5.785

0.000

5.785

0.271

0.007

0.005

0.535

0.188

0.015

0.000

1.279

0.000

1.561

0.043

0.023

6.000

0.000

6.023

6.016

85.97

0.01

0.02

1.67

1.06

0.11

0.00

9.23

0.00

6.32

30.85

0.35

36.33

CAR
Core

0.322

0.007

0.009

0.512

0.157

0.002

0.000

1.175

0.001

1.375

0.007

0.285

6.000

0.000

6.285

6.031

86.37

0.01

0.04

1.61

0.90

0.01

0.00

8.58

0.01

5.63

32.55

0.05

36.83

CAR
Core

0.246

0.052

0.010

0.568

0.176

0.007

0.000

1.167

0.000

1.480

0.016

0.099

6.000

0.000

6.099

6.062

85.28

0.10

0.05

1.77

0.99

0.05

0.00

8.41

0.00

5.98

31.18

0.13

36.54

CAR
Core

0.195

0.088

0.010

0.622

0.174

0.008

0.000

1.460

0.000

1.580

0.016

0.000

5.730

0.000

5.730

6.114

85.43

0.17

0.04

1.91

0.97

0.06

0.00

10.40

0.00

6.32

28.95

0.13

36.42

CAR
Core

0.067

0.043

0.009

0.722

0.201

0.002

0.000

2.456

0.014

1.183

0.064

0.000

5.412

0.035

5.447

5.965

86.46

0.08

0.04

2.15

1.08

0.01

0.00

16.92

0.09

4.57

26.62

0.49

34.36

SILV2
Core

0.146

0.000

0.014

0.622

0.217

0.003

0.002

2.117

0.018

1.196

0.062

0.000

5.616

0.000

5.616

6.036

86.28

0.00

0.07

1.87

1.18

0.02

0.01

14.77

0.12

4.68

27.81

0.48

35.24

SILV2
Core

0.115

0.029

0.008

0.622

0.256

0.013

0.008

2.086

0.010

1.324

0.140

0.000

5.436

0.000

5.436

6.005

86.47

0.05

0.03

1.88

1.39

0.10

0.06

14.59

0.07

5.20

26.97

1.09

35.12

SILV2
Core

0.068

0.068

0.016

0.563

0.352

0.015

0.000

3.040

0.013

0.770

0.244

0.000

4.700

0.000

4.700

6.200

85.81

0.12

0.07

1.63

1.84

0.10

0.00

20.34

0.09

2.89

22.31

1.82

34.69

SILV2
Rim

0.302

0.028

0.007

0.564

0.126

0.000

0.000

1.144

0.008

1.851

0.010

0.000

5.866

0.000

5.866

6.112

87.13

0.06

0.04

1.79

0.73

0.00

0.00

8.41

0.06

7.64

30.60

0.08

37.59

TROP4
Core

0.007

0.126

0.015

0.476

0.502

0.032

0.000

2.054

0.015

2.061

0.040

0.000

4.926

0.133

5.059

5.867

86.22

0.23

0.07

1.42

2.71

0.23

0.00

14.21

0.10

8.00

24.83

0.31

33.95

TR56
Core

0.010

0.000

0.009

0.400

0.582

0.034

0.007

1.744

0.003

2.107

0.116

0.000

5.071

0.153

5.224

5.847

87.21

0.00

0.04

1.22

3.21

0.25

0.05

12.33

0.02

8.36

26.20

0.91

34.57

TR56
Core

0.179

0.000

0.002

0.559

0.260

0.046

0.012

1.024

0.006

1.869

0.048

0.000

5.784

0.000

5.784

6.096

85.63

0.00

0.01

1.75

1.47

0.35

0.09

7.43

0.04

7.61

29.77

0.39

36.99

TR76
Core

0.218

0.163

0.016

0.185

0.581

0.056

0.006

1.139

0.004

2.424

0.141

0.000

5.107

0.000

5.107

6.019

85.77

0.31

0.07

0.57

3.26

0.42

0.05

8.19

0.03

9.78

26.05

1.13

36.19

TROP4
Rim

0.138

0.179

0.008

0.436

0.418

0.033

0.000

0.812

0.001

2.286

0.094

0.000

5.566

0.012

5.579

5.988

84.23

0.34

0.04

1.35

2.34

0.25

0.00

5.83

0.01

9.21

28.41

0.75

35.95

TR56
Rim

0.123

0.193

0.021

0.323

0.532

0.026

0.001

0.738

0.005

2.360

0.189

0.000

5.425

0.155

5.580

5.845

86.50

0.38

0.10

1.02

3.05

0.20

0.01

5.42

0.04

9.73

29.09

1.55

35.92

TROP34
Rim

Structural formula on the basis of T+Z+Y=15 according to Henry and Dutrow (1996). Total Fe measured as FeO. Samples CAN Candelaria, CAR Carola, SILV Silvita, TR Tropezn

6.034

5.686

Si

Cations on the basis of 24.5 O and 3 B per formula unit

86.38

1.00

0.49

Total

36.19

36.37

SiO2

TiO2

28.15

CAND1
Core

CAND1
Core

Sample

Table 1 Selected electron microprobe analyses of tourmaline from IOCG deposits of the Coastal Cordillera of Chile

0.121

0.101

0.007

0.405

0.467

0.029

0.003

0.824

0.001

2.383

0.109

0.000

5.501

0.005

5.506

5.995

86.17

0.20

0.03

1.28

2.67

0.22

0.02

6.03

0.00

9.78

28.58

0.89

36.68

TR56
Rim

490
Miner Deposita (2012) 47:483499

Miner Deposita (2012) 47:483499

491

Fig. 3 Chemical composition of tourmaline from the studied IOCG


deposits expressed in atoms per formula unit (a.p.f.u.) and atomic ratios. a
Al-Fe-Mg ternary diagram showing the field defined by Henry and
Guidotti (1985) with the tourmaline data occupying mainly the fields 3,
5 and 6. Labeled field are: 1 Li-rich granitic rocks, pegmatite, and aplite;
2 Li-poor granitic rocks, pegmatite, and aplite; 3 hydrothermally altered

granitic rocks; 4 metapelite and meta-arenite (aluminous); 5 metapelite


and meta-arenite (Al-poor); 6 Fe3+-rich quartz-tourmaline rocks, calcsilicate hornfels, and metapelite; 7 low-Ca meta-ultramafic rocks; and
8 meta-carbonate rocks and meta-pyroxenite. b X-site-vacancy-Na-Ca
ternary diagram (Hawthorne and Henry 1999) showing that most
analyses plot in the alkalic and calcic groups

mal assemblage is consistent with a relatively oxidizing


environment.
An evaluation of the different exchange vectors that
control substitutions in the tourmaline structure indicate
that both cores and rims of the crystals grew in open
hydrothermal systems. According to Taylor and Slack
(1984), the oscillatory zoning observed in the rims of the
tourmaline crystals likely reflect mineral growth in a
presumably open system with high fluid/rock ratios.

rims can be either enriched or depleted in 11B relative to


corresponding cores (Table 2). The database of 11B data
from magmatic hydrothermal systems of the Andes is
scarce and limited to the data of Wittenbrink et al. (2009)
from the El Salvador porphyry copper deposit. Thus, we
have also obtained eight 11B analyses of tourmaline from
the ca. 60 Ma Copucha porphyry copper deposit in the
Sierra Gorda district (Iriarte 1993). The obtained values
range between 3.1 and 1.4 (Table 2).
Within a repeatability of ca. 1, the observed variations,
even in single crystals, are larger than the analytical
uncertainty, indicating that they are not an analytical
artifact, but instead reflect real variations in the 11B of
the tourmaline. Some 11B values of tourmaline from the
Carola Mine fall in the +3.5 to +8.3 range and are
significantly heavier than those measured on other spots of
the same sample and from the other mines (Table 2).
Duplicate analyses in adjacent spots, only a few micrometers from the original crater, match the average of
regional values. Furthermore, these high 11B values are
systematically accompanied by low B contents, suggesting
that the ion beam sampled inclusions that affected the
signal. Thus, these results are excluded from the overall
interpretation.

Results of the isotopic study


11B values of analyzed tourmaline range from 10.4
to +6.0, with an average value of 2.43.3 (n=60;
Table 2 and Fig. 5). These values are only slightly more
negative than those obtained by Wittenbrink et al. (2009)
for boron in melt inclusions in porphyry deposits of
northern Chile. These signatures show differences among
different outcrops and generations of tourmaline. In the
Tropezn deposit, tourmaline in the breccia pipe has 11B
values of 4.6 to +0.5, whereas later ore-associated
tourmaline has more negative signatures of 7.6 and
2.1. The nearby Silvita mine has 11B between 4.9
and +6.0, which is the largest variation with the most
positive values observed. 11B values in the Candelaria and
Carola deposits are very similar, 6.1 to +1.2 and
10.4 to +1.5, respectively, with the 10.4 value being
the most negative measured. At the crystal scale, there are
no marked differences. 11B variation is uniformly less than
4. No systematic trend was found in the analyzed crystals
from the two Tropezn samples TR-34 and TR-56; wherein

Discussion: the source of B in the IOCG deposits


Due to the steadily growing database for boron isotopes
from many different rock types, the overall signatures of
major reservoirs are now well established (Fig. 6; Chaussidon
and Albarde 1992; Palmer and Swihart 1996; Kasemann et

492

Miner Deposita (2012) 47:483499

Fig. 4 Chemical composition of


tourmalines expressed in terms
of atomic ratio and a.p.f.u. a Ca/
(Ca+Na) vs. Mg#; b Fe(total)
vs. Mg; c Na vs. Ca; d vacancy
vs. Ca; e Fe vs. Al+Ca; and f (R
+X site vacancyCa+Ti) vs.
(AlX-site vacancy +Ca). R is
the sum of Mg and Fe for
achieving the projection down
to the MgFe-1 exchange vector;
thus, dravite and schorl are
projected in the same point.
Plots B, C, D, E, and F show the
direction of the main exchange
vectors. Note that in all plots,
the tourmaline composition
defines a trend towards the
schorl, dravite, uvite, and
povondraite end-members

al. 2000). In the studied ore deposits, precipitation of


tourmaline in an otherwise B-poor igneous rock indicates
that the boron isotope composition is controlled by the oreforming magmatic fluids. The original fluid signature could
have been modified by fractionation during separation, fluid
mixing, or interaction with B-bearing host rocks. These
results may be also influenced by depositional temperature or
equilibration with late fluids (e.g., Palmer and Slack 1989).
However, the influence of most of these variables is likely
quite small. Variations in the temperature alone cannot
account for the range observed in the data. Fluid inclusion

microthermometry shows that tourmaline in the Tropezn


and Silvita mines precipitated between 400C and more than
600C (Tornos et al. 2010). This temperature range is similar
to that estimated by Marschik and Fontbot (2001a) for the
early mineralizing stage at Candelaria (500C to 600C).
Boron isotope fractionation between tourmaline and an
aqueous fluid is only 2.7 at 400C and 1.3 at
600C (Meyer et al. 2008). Thus, changes in the temperature
can only explain a relatively small variation of perhaps 1,
a range close to our analytical uncertainty. Mixing with B
inherited from the host rocks was probably negligible

Miner Deposita (2012) 47:483499

493

Table 2 11B analysis and Li contents of tourmaline from the IOCG


deposits of the Coastal Cordillera of Chile

35

0.5

26

Edge

4.1

0.6

13

pt1 north

+0.7

0.5

31

16

SILV-2

16

CAND-1

g1 outer rim

7.6

0.5

g1 inner rim

7.6

0.5

TR-56

0.5

+2.0

Edge

13

TR-56

+0.5

Fibrous

SILV-2

0.5

g1 core

Li g/g

SILV-2
5.9

TR-56

Error

Point

Li g/g

Point

11B

Sample

Error

Sample

Table 2 (continued)
11B

TR-56

g1 outer core

5.6

0.5

19

CAND-1

pt2 north

6.1

0.6

29

TR-56

g2 interior

5.1

0.5

15

CAND-1

pt3 north tip

6.1

0.8

26

TR-56

g2 rim

6.4

0.5

18

CAND-1

pt4 east

+1.2

0.7

30

TR-56

g3 interior

4.1

0.5

22

CAND-1

pt5 mid

+0.7

0.8

29

TR-56

g3 rim

3.8

0.5

20

CAND-1

pt6

2.8

0.6

41

TR-56

g4 core clear

4.9

0.6

14

CAND-1

pt7 south

0.6

0.5

32

14

TR-56

g4 core dark

3.3

0.6

TR-56

g4 rim

5.1

0.6

COPU-1

3.1

0.4

15

COPU-1

2.1

0.4

187

TR-56

g5 core

5.4

0.5

14

COPU-1

2.6

0.5

27

TR-56

g5 rim

2.1

0.4

15

COPU-1

1.9

0.4

16

COPU-1

1.4

0.4

243

TR-34

g1 core

1.8

0.6

TR-34

g1 rim

1.1

0.5

11

TR-34

g2 core

+0.5

0.8

12

TR-34

g2 rim

+0.0

0.4

10

TR-34

g3 core

1.8

0.7

14

383-1 Carola, COPU Copucha, CAND Candelaria, SILV Silvita, TR


Tropezn

TR-34

g3 rim

5.9

0.6

15

TR-34

g4 core

2.1

0.6

18

TR-34

g4 rim

+0.5

0.6

18

TR-34

g5 core

3.3

0.6

17

Error indicates the 1 observed uncertainty of the individual


analysis based on the scatter of 40 isotope ratios recorded during the
run. The overall uncertainty for a given datum is estimated to be
around 1.2 (1 ) as based on the observed repeatability of the
instrumental mass fractionation as measured using 42 analyses of the
suite of three tourmaline reference materials

TR-34

g5 rim

4.4

0.5

15

Analysis GFZ Potsdam, SIMS laboratory

TR-34

g6 core

0.6

0.6

12

TR-34

g6 rim

0.3

0.6

13

TR-34

g7 core

0.6

0.7

12

TR-34

g7 rim

4.6

0.4

16

383.1

G1-1

+8.3a

0.6

292

383.1

G2-1

+3.5a

0.8

688

383.1

G3-1

5.1

0.5

117

383.1

G4-1

+1.5

0.6

59

383.1

G5-1

+6.8a

0.5

121

383.1

G6-1

2.8

0.4

46

383.1

G6-2

10.4

0.6

41

383.1

G7-1 lower circle

3.6

0.5

39

383.1

G8-1 upper circle

+1.0

0.5

16

383.1

g9-9

4.1

0.5

383.1

g10-1

1.6

0.5

31

383.1

g5-2

3.8

0.5

33

383.1

g1-2

7.4

0.5

33

383.1

g2-3

1.8

0.5

45

SILV-2

G1-1 tip

3.1

0.6

40

SILV-2

G1-2 massive

+2.5

0.5

87

SILV-2

G1-3 massive

+6.0

0.5

21

SILV-2

G1-4 massive

+2.0

0.6

49

SILV-2

G1-5 core

+3.0

0.5

20

SILV-2

Fibrous

0.6

0.5

17

SILV-2

Fibrous

4.9

0.4

19

SILV-2

Fibrous

+0.7

0.5

18

SILV-2

Massive edge

+0.2

0.5

27

SILV-2

Massive edge

2.1

0.6

41

SILV-2

Massive edge

+4.8

0.5

53

Samples having inclusions of other minerals

because of the likely high fluid/rock ratios expected during


magmatic and hydrothermal tourmaline growth. Fluid-melt
separation induces an increase in the 11Bfluid values due to
the preferential fractionation of 11B to the fluid (Hervig et al.
2002), whereas boiling produces the reverse effect (Smith
and Yardley 1996). Boiling has been recorded in both the
Tropezn and Silvita mines, but not in the Candelaria or
Carola deposits. However, because the 11B values are
similar in all the deposits, it seems that the effect of boiling
in the boron isotope composition is minor compared with
other possible sources of variation.
With a limited B magmatic reservoir, the isotopic
composition of different batches of the releasing fluid

Fig. 5 Histogram showing the boron isotope composition of


tourmaline from the IOCG deposits of the Coastal Cordillera

494

would be governed by Rayleigh fractionation and producing progressively 11B enriched fluids. Preliminary LA ICPMS fluid inclusion analyses suggest that hydrothermal
fluids at Tropezn carried 102103 g/g B, a content similar
to that reported by Audetat et al. (2000) for the Mole
Granite, in Australia. Precipitation of tourmaline having ca.
3 wt.% B should gradually deplete the system in B. At
500C, the calculated 11Bfluid values at Tropezn are
between 5.7 and +2.4; a fluid equilibrated with the
isotopically heavier tourmaline (11Btourmaline =+0.5)
would imply f values (residual B in the system) as low as
0.01 if the original 11Bfluid was in equilibrium with a
tourmaline having a signature to 7.6. Thus, the
observed 11B range could reflect tourmaline growing from
a system with major variation in the degree of distillation.
These results show that within individual deposits, the
observed 11B variations cannot be due to the precipitation
of tourmaline at different temperatures and degrees of
distillation unless extreme fractionation is assumed; thus,
the observed variations must be interpreted as inherited
from the boron source. The overall isotopic data and
geologic evidence are consistent with a model in which
the variation of 11B values has a regional significance and
is inherited from the B sources.
Arc melts usually have 11B isotopic signatures that are
different from those of the underlying crust and mantle
wedge, suggesting that material from the subducted slab is
incorporated in the melts. Boron in both the slab and the arc
trench is mainly hosted by clay within the pelagic sediments, which is rapidly converted into phengite. It is also
contained in entrained seawater and in serpentine in the
altered oceanic crust. Thus, the 11B values can vary
significantly depending on the thickness and nature of the
sedimentary cover and the degree of alteration of the
seafloor. Pelagic sediments have 11B signatures of 9.2
to 6.6; boron predominantly derived from the pelagic
sediments is accompanied by smaller contributions from
continental detritus (11B, 13 to 2.8), biogenic silica
(+2.1 to +4.5) and limestone (+8 to +35)
(Spivack et al. 1987; Ishikawa and Nakamura 1993). In
the present-day ocean floor east of the Pacific trench, the
proportion of continental sediments and carbonates is low,
with the sedimentary cover dominated by clay-rich pelagic
sediments (Plank and Langmuir 1998). The 11B values of
the altered oceanic crust are between 5 and +21
(Spivack and Edmond 1987; Leeman and Sisson 2002;
Rosner et al. 2003). Ishikawa and Nakamura (1994) stated
that 11B values are also influenced by the thickness of the
mantle wedge but its low B content, similar to that of
unaltered MORB-like crust (ca. 0.05 g/g), suggest that
this contribution is usually negligible (Chaussidon and
Marty 1995). Thus, 11B values of the subducted slab
should be controlled by proportions of pelagic sediments

Miner Deposita (2012) 47:483499

and altered oceanic crust, varying between extreme


values of 13 to +21, with average compositions
between 5 and +20 (Marschall et al. 2009). Values
close to the lower limit have been proposed for the 11B
composition of continental crust, between 13 and 7,
with an average of 8.9 (Chaussidon and Albarde 1992;
Palmer and Swihart 1996; Kasemann et al. 2000; Marschall
and Ludwig 2006).
The architecture of the Andean convergent margin and
its associated magmatic arc is relatively well known (see
Coira et al. 1982; Mpodozis and Ramos 1990), and the
geochemistry of boron in this system is relatively well
constrained (e.g., Rosner et al. 2003). Volcanic andesite to
dacite of Cenozoic age have 11B values between 0 and
6, interpreted as reflecting the composition of the
subducting slab (Rosner et al. 2003); Neogene ignimbrite
has more negative values of 11B of 6.6 to 1.0; this
boron isotope composition is interpreted as reflecting
mixing of boron derived from the subducted slab and the
underlying continental crust since the early Paleozoic
basement has an average 11B signature of 11 to 5
(Kasemann et al. 2000). The 11B values of melt inclusions
in mineralized porphyry systems in the Andes show that,
whereas the tin porphyries have negative signatures
(14.1 to 8.7), the porphyry copper deposit of El
Salvador has a broad range between 7 to +12
(Wittenbrink et al. 2009). Our data for the Copucha
porphyry copper deposit (3.1 to 1.4) range within
these latter values.
Current tectonic models call for the formation of IOCG
deposits of the Andes during the early stages of subduction
(Sillitoe 2003) when the magmatic arc was in its incipient
stage. The 11B signatures shown in Table 2 are similar to
those of the nearby Cenozoic volcanic rocks, both mafic
and felsic, and somewhat depleted in 11B compared with
the wide range observed in the porphyry copper deposits.
The uppermost observed 11B value, +6, is only slightly
higher than the +4 signature predicted for fluids derived
directly from a subducting slab. More negative values could
also be explained by B sourced from dewatering of the slab
at much greater depths, but it seems unlikely that this
occurred during the early stages of subduction with magmas
generated near the trench. Assuming that the 11B composition of the slab and basement has not changed since the
Cretaceous, the simplest interpretation is that these values
track boron derived from two or more reservoirs hosted by
the continental crust and the underlying subducted slab
(Fig. 6).
Negative Ndi values (11 to 5.5) in the tin porphyries
and +2 to +4 Nd values in the porphyry copper deposits
(Wittenbrink et al. 2009) support that in the Andes, not all
the observed variation in boron isotopes is due to
progressive dehydration of the subducting slab but reflects

Miner Deposita (2012) 47:483499

495

Sossego
Igarap-Baha (high grade)

Carajs district

Igarap-Bahia & Salobo


Tropezon
Silvita
Carola
Candelaria
Copper porphyries (Copucha)

Ore deposits

Copper porphyries (EL Salvador)

Tin porphyries
Salars

Present day volcanoes


Cenozoic andesite-dacite
Cretaceous sediments

Rocks
Central Andes

Jurassic limestone
Paleozoic basement
avg. continental crust
metapelite
granitic rocks
MORB

progressive
dehydration

altered oceanic crust


subducting slab

marine detritic sediments


pelagic sediments

Global
Reservoirs

biogenic silica

biogenic carbonates

marine evaporites
non marine evaporites

-40

-30

-20

-10

10

11B

20

30

seawater

40

Fig. 6 11B values of different boron reservoirs in the Andes and


elsewhere. Data of the volcanic and metasedimentary rocks and the
basement of the Andes are from Kasemann et al. (2000), Rosner et al.
(2003), and Schmitt et al. (2002), and those of the global reservoirs are

from Chaussidon and Albarde (1992), Ishikawa and Nakamura


(1993), Marschall and Ludwig (2006) and Marschall et al. (2009),
Palmer and Swihart (1996), Spivack and Edmond (1987), and Spivack
et al. (1987)

two distinct sources of B, crustal melts generated from


pelitic rocks and slab-derived fluids. The data of Stern and
Skewes (1995) show rather primitive initial values for
porphyry copper deposits, with 87Sr/86Sri lower than 0.7049
and Ndi higher than 1.0 (e.g., Stern and Skewes 1995).
However, preliminary radiogenic isotope data for the
Tropezn deposit with 87Sr/86Srinitial and Ndinitial values
of 0.7075 and 2.9, respectively (Tornos et al. 2010),
indicate that melts related to the IOCG deposits had
intermediate degree of contamination between those of the
porphyry copper and the tin-related intrusions. The boron
isotopes are consistent with this model, and the most
coherent interpretation is that the observed 10.4 to
+6.0 values reflect mixing between boron derived from
an underlying continental basement and fluids released
from a shallow subducting slab. If that holds true, primitive
melts were not the sole source of the IOCG deposits in the
Andes, and the interaction of juvenile melts with the
overlying continental crust was a prerequisite for its
formation. Our model differs from the interpretations of
Marschik et al. (2003a) and Sillitoe (2003), who suggested

that IOCG systems are related to primitive melts having


minor interaction with crustal rocks or metamorphic,
basinal, or surficial fluids.
The geologic environment of formation of the studied
IOCG deposits excludes predominant derivation of boron
from continental evaporites, unmodified seawater, or
marine carbonates. It can be argued that the observed
11B signature reflects the interaction of seawater derived
from the coeval back arc basin where the Chaarcillo
Formation was deposited with an underlying continental
basement. Percolation of evaporitic brines into the crust and
further equilibration might produce the observed 11B
signature. However, a numerical approach using the
equations presented by Faur (1986) shows that this
mechanism is unlikely. The interaction of a fluid derived
from seawater having a 11B value of +39 and 4.5 g/g
B (Ishikawa and Nakamura 1993) with a basement having a
11B of 8.9, and 43 g/g B (Kasemann et al. 2000)
implies that only equilibration at fluid/rock mass ratios lower
than 0.20.3assuming equilibration in a dominantly open
systemcan produce the observed 11B values. Despite that

496

this is numerically plausible, such low fluid/rock ratios seem


unrealistic for these systems because metals would never
have been leached in sufficient amounts to form the observed
ore deposit. Also, it is inconsistent with the geologic setting
of the Tropezn depositlocated in the apical zone of an
intrusion and showing many of the characteristics of
magmatic-hydrothermal systemsand the similitude of the
measured 11B values with those of porphyry Cu and Sn
deposits and volcanic rocks.
Finally, our results are somewhat different from those of
Xavier et al. (2008) for the Carajs district, Brazil, who
reported a broad range of 11B values between 8 and
+26.5 and with significant variations among deposits.
Their data were interpreted to suggest that the ore-forming
fluids are derived from marine evaporites, which are mixed
with fluids from or equilibrated with felsic igneous rocks.
However, the data for the Sossego deposit (11B, 8
to +11) are very similar to those for the magmatichydrothermal deposits of the Andes. The isotopically
heavier boron of Igarap-Baha and Salobo (+5.8
to +26.1) suggest a clear evaporitic influence (Figs. 6
and 7). As a whole, these results reinforce the concept that,
viewed worldwide, IOCG deposits do not reflect a single

Fig. 7 11B of tourmaline from the IOCG deposits of the Coastal


Cordillera, nearby rocks, and other ore deposits of the Andes, showing
the similarity of the boron isotope compositions of the IOCG deposits
with those of the nearby but younger porphyry copper and tin deposits
and volcanic rocks. Data from the El Salvador porphyry copper
deposit and the tin porphyry deposits of Bolivia are from Wittenbrink
et al. (2009). For comparison, 11B signatures of tourmaline from the
IOCG deposits of the Carajas district (Brazil) are shown (Xavier et al.
2008). Data for the plutonic and metasedimentary rocks and the

Miner Deposita (2012) 47:483499

uniform evolutionary system and that otherwise similar ore


deposits differ markedly in their evolution.

Conclusions
Tourmaline from the Tropezn, Silvita, Candelaria, and
Carola IOCG deposits of the Coastal Cordillera of Chile has
11B values ranging from 10.4 to +6. Both the boron
isotope data and the composition and zonation of tourmaline are consistent with a common origin for boron in all
these deposits, probably representing different zones in a
large magmatic-hydrothermal system. This system would
include subvertical tourmaline breccia pipes in the lowermost zones (Tropezn and Silvita) and more diffuse
tourmaline-rich alteration that is found only in the deeper
zones of Carola and Candelaria.
As a whole, these 11B values are similar to those of the
nearby Cenozoic volcanic rocks and are intermediate
between those of the porphyry copper deposits and
intrusive rocks related to SnAg deposits, suggesting a
common origin for the boron. The simplest interpretation is
that the 11B values track derivation of boron extracted

basement of the Andes are from Kasemann et al. (2000, 2004),


Schmitt et al. (2002), and Rosner et al. (2003). The 11B values of
recent plutonic rocks are interpreted as reflecting the composition of
deep melts with the lower values being closer to those of the
subducting slab that are gradually modified by interaction with the
basement. The low 11B values of the porphyry tin deposits suggest
that part of the basement has even more negative signatures than those
predicted by Kasemann et al. (2000)

Miner Deposita (2012) 47:483499

during devolatilization of altered oceanic crust and overlying sediments of the subducting slab, which subsequently
interacted with the Andean continental basement in variable
proportions. These results are consistent with Sr and Nd
isotope data, all of which indicate appreciable crustal
contamination of igneous rocks associated with the IOCG
deposits. Local processes such as changes in the temperature of precipitation or distillation seem to have played only
a minor role.
The 11B values are, thus, consistent with a magmatichydrothermal origin for the IOCG deposits of the Coastal
Cordillera of Chile and exclude basinal brines related to
marine evaporites as a major source of the hydrothermal
boron, although non-marine evaporites cannot be excluded;
however, the similitude of the obtained d11B values with
those of porphyry deposits, and volcanic rocks indicate that
the boron is of magmatic/basement derivation. These
conclusions are at odds with those for other settings, where
positive B isotopes values suggest a marine evaporitic
origin of the boron.
Acknowledgements This study was done under the framework of
project DGI-FEDER CGL2006-0378 of the Spanish Government and
by internal funding of the SIMS Laboratory in Potsdam. It would not
have been possible without the collaboration of Nicolae Pop (Minera
Carola) and Manuel Erazo and Walter Gil (Minera Cenizas) who
granted access to the mine site and assisted with petrologic
interpretation of the samples. Thanks are also extended to Fernando
Barra and Diego Morata (Universidad de Chile) for help on the study
of the IOCG deposits. Our acknowledgement to Horst Marschall and
John Slack for reviewing earlier versions of this manuscript and Bernd
Lehmann for final editing.

References
Audetat A, Gunther D, Heinrich CA (2000) Magmatic hydrothermal
evolution in a fractionating granite: a microchemical study of the
Sn-W-F-mineralized Mole Granite (Australia). Geochimica et
Cosmochimica Acta 64:33733393
Barth S (1993) Boron isotope variations in nature: a synthesis. Geol
Rundsch 82:640651
Barton MD, Johnson DA (1996) Evaporitic source model for igneousrelated Fe oxide-(REE-Cu-Au-U) mineralization. Geology
24:259262
Barton MD, Johnson DA (2000) Alternative brine sources for Fe oxide(Cu-Au) systems: implications for hydrothermal alteration and
metals. In: Porter TM (ed), Hydrothermal Iron Oxide Copper-Gold
& Related Deposits: A Global Perspective, Australian Mineral
Foundation, Adelaide, v.I., pp 4360
Bebout GE, Nakamura E (2003) Record in metamorphic tourmalines
of subduction-zone devolatilization and boron cycling. Geology
31:407410
Bebout GE, Ryan JG, Leeman WP, Bebout AE (1999) Fractionation of
trace elements by subduction-zone metamorphismeffect of
convergent-margin thermal evolution. Earth and Planetary Science
Letters 171:6381
Benavides J, Kyser TK, Clark AH, Oates CJ, Zamora R, Tarnovschi
R, Castillo B (2007) The Mantoverde iron oxide-copper-gold

497
district, III region, Chile: the role of regionally derived, nonmagmatic fluids in chalcopyrite mineralization. Econ Geol
102:415440
Benton LD, Ryan JG, Tera F (2001) Boron isotope systematics of slab
fluids as inferred from a serpentine seamount, Mariana forearc.
Earth and Planetary Science Letters 187:273282
Berg K, Breitkreuz C, Damm KW, Pichowiak S, Zeil W (1983) The
North-Chilean Coast Rangean example for the development of
an active continental margin. Geol Rundsch 72:715731
Brown M, Diaz F, Grocott J (1993) Displacement history of the
Atacama Fault System 2500S-27-00S, Northern Chile. Geol
Soc Am Bull 105:11651174
Cembrano J, Gonzalez G, Arancibia G, Ahumada I, Olivares V,
Herrera V (2005) Fault zone development and strain partitioning
in an extensional strike-slip duplex: a case study from the
Mesozoic Atacama fault system, Northern Chile. Tectonophysics
40:105125
Chaussidon M, Albarde F (1992) Secular boron isotope variations in
the continental crustan ion microprobe study. Earth and
Planetary Science Letters 108:229241
Chaussidon M, Marty B (1995) Primitive boron isotope composition
of the mantle. Science 269:383386
Chiaradia M, Banks D, Cliff R, Marschik R, de Haller A (2006)
Origin of fluids in iron oxide-copper-gold deposits: constraints
from 37Cl, 87S/86Sri and Cl/Br. Mineralium Deposita 41:565
573
Coira B, Davidson J, Mpodozis C, Ramos V (1982) Tectonic and
magmatic evolution of the Andes of northern Argentina and
Chile. Earth-Science Reviews 18:303332
Dallmeyer RD, Brown M, Grocott J, Taylor GK, Treloar PJ (1996)
Mesozoic magmatic and tectonic events within the Andean plate
boundary zone, 262730S, North Chile: constraints from
40
Ar/39Ar mineral ages. J Geol Soc 104:1940
Diaz A, Vivallo W, Jorquera R, Pizarra N (2003) Depsitos de Fe,
xidos de Fe-Cu-Au y su relacin con el magmatismo del
Cretcico Inferior, III Regin de Atacama, Chile 10Congreso
Geolgico Chileno. Concepcin
Dreher AM, Xavier RP, Taylor BE, Martini SL (2008) New geologic,
fluid inclusion and stable isotope studies on the controversial
Igarape Bahia Cu-Au deposit, Carajs Province, Brazil. Mineralium Deposita 43:161184
Dyar MD, Wiedenbeck M, Robertson D, Cross LR, Delaney JS,
Ferguson K, Francis CA, Grew ES, Guidotti CV, Hervig RL,
Hughes JM, Husler J, Leeman W, McGuire AV, Rhede D, Rothe
H, Paul RL, Richards I, Yates M (2001) Reference minerals for
microanalysis of light elements. Geostandarts Newsletter 25:441
463
Faur G (1986) Principles of isotope geology. Wiley & Sons, New
York, p 589
Foster GL, Pogge von Strandmann PAE, Rae JWB (2010) Boron and
magnesium isotopic composition of seawater. Geochemistry
Geophysics Geosystems 11, Q08015, 10 pp. doi:10.1029/
2010GC003201
Frikken PH, Cooke DR, Walshe JL, Archibald D, Skarmeta J, Serrano
L, Vargas R (2005) Mineralogical and isotopic zonation in the
Sur-Sur tourmaline breccia, Rio Blanco-Los Bronces Cu-Mo
deposit, Chile; implications for ore genesis. Econ Geol 100:935
961
Gelcich S, Davis DW, Spooner ETC (2005) Testing the apatitemagnetite geochronometer: U-Pb and 40Ar/39Ar geochronology
of plutonic rocks, massive magnetite-apatite tabular bodies and
IOCG mineralization in Northern Chile. Geochimica et Cosmochimica Acta 69:33673384
Gonfiantini R, Tonarini S, Grning M, Adorni-Braccesi A, Al-Ammar
AS, Astner M, Bchler S, Barnes RM, Bassett RL, Cocherie A,
Deyhle A, Dini A, Ferrara G, Gaillardet J, Grimm J, Guerrot C,

498
Krhenbhl U, Layne G, Lemarchand D, Meixner A, Northington DJ, Pennisi M, Reitznerov E, Rodushkin I, Sugiura N,
Surberg R, Tonn S, Wiedenbeck M, Wunderli S, Xiao Y, Zack T
(2003) Intercomparison of boron isotope and concentration
measurements. Part II: evaluation of results. Geostandards
Newsletter 27:4157
Grocott J, Taylor GK (2002) Magmatic arc fault systems, deformation
partitioning and emplacement of granitic complexes in the
Coastal Cordillera, north Chilean Andes (2530S to 2700S). J
Geol Soc 159:425442
Hawthorne FC, Henry DJ (1999) Classification of the minerals of the
tourmaline group. Eur J Mineral 11:201215
Henriquez F, Dobbs FM, Espinoza S, Nystrom J, Travisany V, Vivallo
W (1994) Origin of chilean magnetite-apatite ore deposits. In: 7.
Congreso Geologico Chileno 2:822824
Henry DJ, Dutrow BL (1996) Metamorphic tourmaline and its
petrologic applications. Rev Mineral Geochem 33:503557
Henry DJ, Guidotti CV (1985) Tourmaline as a petrogenetic indicator
mineralan example from the staurolitegrade metapelites of
NW Maine. Am Mineral 70:115
Henry DJ, Sun H, Slack JF, Dutrow BL (2008) Tourmaline in metaevaporites and highly magnesian rocks: perspectives from
Namibian tourmalines. Eur J Mineral 20:889904
Hervig RL, Moore GM, Williams LB, Peacock SM, Holloway JR,
Roggensack K (2002) Isotopic and elemental partitioning of
boron between hydrous fluid and silicate melt. Am Mineral
87:769774
Iriarte S (1993) Control estructural de la mineralizacin vetiforme en
Mina Faride y su relacin con el distrito de Sierra Gorda, regin
de Antofagasta. Thesis, Universidad de Chile, Santiago, Chile, p
200
Ishikawa T, Nakamura E (1993) Boron isotope systematics of marine
sediments. Earth and Planetary Science Letters 117:567580
Ishikawa T, Nakamura E (1994) Origin of the slab component in arc
lavas from across arc variation of B and Pb isotopes. Nature
370:205208
Jiang SY, Palmer MR, Slack JF, Shaw DR (1998) Paragenesis and
chemistry of multistage tourmaline formation in the Sullivan PbZn-Ag deposit, British Columbia. Econ Geol 93:4767
Kasemann S, Erzinger J, Franz G (2000) Boron recycling in the
continental crust of the central Andes from the Palaeozoic to
Mesozoic, NW Argentina. Contrib Mineral Petrol 140:328343
Kasemann SA, Meixner A, Erzinger J, Viramonte JG, Alonso RN,
Franz G (2004) Boron isotope composition of geothermal fluids
and borate minerals from salar deposits (central Andes/NW
Argentina). Journal of the South American Earth Sciences
16:685697
Klemm LM, Pettke T, Heinrich CA, Campos E (2007) Hydrothermal
evolution of the El Teniente deposit (Chile): porphyry Cu-Mo ore
deposition from low salinity magmatic fluids. Econ Geol
102:10211045
Krienitz MS, Trumbull RB, Hellmann A, Kolb J, Meyer FM,
Wiedenbeck M (2008) Hydrothermal gold mineralization at the
Hira Buddini gold mine, India: constraints on fluid evolution and
fluid sources from boron isotopic compositions of tourmaline.
Mineralium Deposita 43:421434
Leeman WP, Sisson VB (2002) Geochemistry of boron and its
implications for crustal and mantle processes. Rev Mineral
Geochem 33:645708
Leeman W, Tonarini S (2001) Boron isotopic analysis of proposed
borosilicate mineral reference samples. Geostandards Newsletter
25:399403
Lehmann B, Dietrich A, Heinhorst J, Metrich N, Mosbahm M,
Palacios C, Schneider H, Wallianos A, Webster J, Winkelmann L
(2000) Boron in the Bolivian tin belt. Mineralium Deposita
35:223232

Miner Deposita (2012) 47:483499


Mark G, Oliver NHS, Williams PJ (2006) Mineralogical and chemical
evolution of the Ernest Henry Fe oxide-Cu-Au ore system,
Cloncurry district, northwest Queensland, Australia. Mineralium
Deposita 40:769801
Marschall HR, Ludwig T (2006) Re-examination of the boron isotopic
composition of tourmaline from the Lavicky Granite, Czech
Republic, by secondary ion mass spectrometry: back to normal.
Critical comment on Chemical and boron isotopic compositions
of tourmaline from the Lavicky leucogranite, Czech Republic
by S.-Y. Jiang et al., Geochemical Journal, 37, 545556, 2003.
Geochemical Journal 40: 631638
Marschall HR, Korsakov AY, Luvizotto GL, Nasdala L, Ludwig T
(2009) On the occurrence and boron isotopic composition of
tourmaline in (ultra)high-pressure metamorphic rocks. J Geol Soc
166:801823
Marschik R, Fontbot L (2001a) The Candelaria-Punta del Cobre iron
oxide Cu-Au(Zn-Ag) deposits, Chile. Econ Geol 96:17991828
Marschik R, Fontbot L (2001b) The Punta del Cobre Formation,
Punta del Cobre-Candelaria area, Northern Chile. Journal of
South American Earth Sciences 14:401433
Marschik R, Fontignie D, Chiaradia M, Voldet P (2003a) Geochemical
and Sr-Nd-Pb-O isotope composition of granitoids of the Early
Cretaceous Copiap Plutonic Complex (2730S), Chile. Journal
of South American Earth Sciences 16:281398
Marschik R, Chiaradia M, Fontbot L (2003b) Implications of Pb
isotope signatures of rocks and iron oxide Cu-Au ores in the
Candelaria-Punta del Cobre district, Chile. Mineralium Deposita
38:900912
Mathur R, Marschik R, Ruiz J, Munizaga F, Leveille RA, Martin W
(2002) Age of mineralization of the Candelaria Fe Oxide Cu-Au
deposit and the origin of the Chilean Iron belt, based on Re-OS
isotopes. Econ Geol 97:5971
Meyer C, Wunder B, Meixner A, Romer RL, Heinrich W (2008)
Boron-isotope fractionation between tourmaline and fluid: an
experimental re-investigation. Contrib Mineral Petrol 156:259267
Mpodozis C, Ramos V (1990) The Andes of Chile and Argentina. In:
Erikssen GE, Caas MT, Reinemund JA (eds) Geology of the
Andes and its relationship to hydrocarbon and mineral resources,
vol 11. Circum Pacific Council Energy and Mineral Resources
Earth Science Series, pp 5990
Mumin AH, Corriveau L, Somarin AK, Ootes L (2007) Iron oxide
copper-gold-type polymetallic mineralization in the Contact Lake
Belt, Great Bear Magmatic Zone, Northwest Territories, Canada.
Explor Min Geol 16:187208
Nakano T, Nakamura E (2001) Boron isotope geochemistry of
metasedimentary rocks and tourmalines in a subduction zone
metamorphic suite. Physics of the Earth and Planetary Interiors
127:233252
Palmer MR, Slack JF (1989) Boron isotopic composition of
tourmaline from massive sulfide deposits and tourmalinites.
Contrib Mineral Petrol 103:434451
Palmer MR, Swihart GH (1996) Boron isotope geochemistry: an
overview. Rev Mineral Geochem 33:709744
Pesquera A, Torres F, Gil-Crespo P, Torres-Ruiz J (2008) TOURCOMP: a program for estimating end-member proportions in
tourmalines. Mineral Mag 72:10211034
Plank T, Langmuir CH (1998) The chemical composition of subducting
sediment and its consequences for the crust and mantle. Chem Geol
145:325394
Pollard PJ (2006) An intrusion-related origin for Cu-Au mineralization
in iron oxide-copper-gold (IOGG) provinces. Mineralium Deposita
41:179187
Rieger AA, Marschik R, Diaz M, Holzl S, Chiaradia M, Akker B,
Spangenberg JE (2011) The hypogene iron oxide copper-gold
mineralization in the Mantoverde District, Northern Chile. Econ
Geol 105:12711299

Miner Deposita (2012) 47:483499


Rosner M, Erzinger J, Franz G, Trumbull RB (2003) Slab-derived
boron isotope signatures in arc volcanic rocks from the Central
Andes and evidence for boron isotope fractionation during
progressive slab dehydration. Geochemistry, Geophysics, Geosystems 4:125
Scheuber E, Andriessen PAM (1990) The kinematic and geodynamic
significance of the Atacama fault zone, northern Chile. J Struct
Geol 12:243257
Schmitt AK, Kasemann S, Meixner A, Rhede D (2002) Boron in
central Andean ignimbrites: implications for crustal boron cycles
in an active continental margin. Chem Geol 183:333347
Sillitoe RH (1976) Andean mineralization: a model for the metallogeny of convergent plate margins. In: Strong DF (ed) Metallogeny and Plate Tectonics. Geological Association of Canada
Special Paper 14:59100
Sillitoe RH (2003) Iron oxide-copper-gold deposits: an Andean view.
Mineralium Deposita 38:787812
Sillitoe RH, Sawkins FJ (1971) Geologic, mineralogic and fluid
inclusion studies relating to the origin of copper-bearing
tourmaline breccia pipes, Chile. Econ Geol 66:10281041
Slack JF, Palmer MR, Stevens BPJ, Barnes RG (1993) Origin and
significance of tourmaline-rich rocks in the Broken Hill district,
Australia. Econ Geol 88:505541
Smith MP, Yardley BWD (1996) The boron isotopic composition of
tourmaline as a guide to fluid processes in the southwestern
England orefield: an ion microprobe study. Geochimica et
Cosmochimica Acta 60:14151427
Spivack AJ, Edmond JM (1987) Boron isotope exchange between
seawater and the oceanic crust. Geochimica et Cosmochimica
Acta 51:10331043
Spivack AJ, Palmer MR, Edmond JM (1987) The sedimentary cycle
of boron isotopes. Geochimica et Cosmochimica Acta 51:1939
1949
Stern CR, Skewes ME (1995) Miocene to present magmatic evolution
at the northern end of the Andean Southern Volcanic Zone,
Central Chile. Revista Sociedad Geologica Chile 22:261272
Tagg SL, Cho H, Dyar MD, Grew ES (1999) Tetrahedral boron in
naturally occurring tourmaline. Am Mineral 84:14511455
Taylor BE, Slack JF (1984) Tourmalines from AppalachianCaledonian
massive sulfide deposits: textural, chemical, and isotopic relationships. Econ Geol 79:17031726
Taylor GK, Grocott J, Popec A, Randall DE (1998) Mesozoic fault
systems, deformation and fault block rotation in the Andean
forearc: a crustal scale strike-slip duplex in the Coastal Cordillera
of northern Chile. Tectonophysics 229:93109

499
Tonarini S, Pennisi M, Adorni-Braccesi A, Dini A, Ferrara G, Gonfiantini
R, Wiedenbeck M, Grning M (2003) Intercomparison of boron
isotope concentration measurements. Part I: selection, preparation
and homogeneity tests of the intercomparison materials. Geostandards Newsletter 27:2139
Tonarini S, Agostini S, Doglioni C, Innocenti F, Manetti P (2007)
Evidence for serpentinite fluid in convergent margin systems: The
example of El Salvador (Central America) arc lavas. Geochemistry
Geophysics Geosystems 8. doi:10.1029/2006GC001508
Tornos F, Velasco F, Barra F, Morata D (2010) The Tropezn CuMo-(Au) deposit, Northern Chile: the missing link between
IOCG and porphyry copper systems? Mineralium Deposita
45:313321
Vivallo W, Daz A, Gelcich S, Lled H (2000) Estilos y tipos de
mineralizacin del Jursico y Cretcico Inferior en la Cordillera
de la Costa de la regin de Copiap, Chile. In: IX Congreso
Geolgico Chileno, vol 2, Puerto Varas, pp. 179182
Wagner T, Mlynarczyk MSJ, Williams-Jones AE, Boyce AJ (2009)
Stable isotope constraints on ore formation at the San Rafael tincopper deposit, Southeast Peru. Econ Geol 104:223248
Wiedenbeck M, Rhede D, Lieckefett R, Witzki H (2004) Cryogenic
SIMS and its applications in the earth sciences. Appl Surf Sci
231(232):888892
Williams P, Guoyi D, Pollard P, Broman C, Martinsson O,
Wanhainen C, Mark G, Ryan CG, Mernagh T (2003) The
nature of iron oxide-copper-gold ore fluids: Fluid inclusion
evidence from Norbotten (Sweden) and the Cloncurry district
(Australia). In: Eliopoulos,D.G., et al. (eds.), Mineral Exploration and Sustainable Development, Millpress Rotterdam,
pp. 11271130
Williams P, Barton MD, Johnson DA, Fontbot L, Haller A, Mark G,
Oliver NHS, Marschik R (2005) Iron oxide copper-gold deposits:
geology, space-time distribution, and possible modes of origin.
In: Hedenquist JW, Thompson JFH, Goldfarb RJ, Richards JP
(eds) Economic geologyone hundredth anniversary volume.
Society of Economic Geologists, Littleton, pp 371406
Wittenbrink J, Lehmann B, Wiedenbeck M, Wallianos A, Dietrich A,
Palacios A (2009) Boron isotope composition of melt inclusions
from porphyry systems of the Central Andes: a reconnaissance
study. Terra Nova 21:111118
Xavier RP, Wiedenbeck M, Trumbull RB, Dreher AM, Monteiro LVS,
Rhede D, Arajo CEG, Torresi I (2008) Tourmaline B-isotopes
fingerprint marine evaporites as the source of high-salinity ore
fluids in iron oxide copper-gold deposits, Carajs Mineral
Province (Brazil). Geology 36:743746

You might also like