You are on page 1of 3

OLOGI C

GE A
D

DE
SOC I E DA

C HILE
un

F
da 6

2
la serena octubre 2015 d a e n 19

Massive Iron Oxide-Apatite (IOA) Deposit Formation by


Efficient Flotation of Magmatic Magnetite Suspensions: A
Genetic Link Between IOA and IOCG Deposits?
1, * 2 2 1 2 3
Martin Reich , Jaayke Knipping , Adam Simon , Fernando Barra , Laura Bilenker , Rodrigo Munizaga
1
Department of Geology and Andean Geothermal Center of Excellence (CEGA), Universidad de Chile, Santiago
2
Department of Earth and Environmental Sciences, University of Michigan, Ann Arbor, MI, US
3
Compañía Minera del Pacífico (CMP), Brasil N 1050, Vallenar, Región de Atacama 1610000, Chile

*Contact email: mreich@ing.uchile.cl

Abstract. The link between iron oxide-apatite (IOA, or from IOCG deposits sensu stricto. A major challenge to
“Kiruna-type”) and iron oxide-copper-gold (IOCG) deposits testing these hypotheses seems to be the need to peel back
remains controversial. Based on geochemical and isotopic layers of metasomatism that possibly obscure the original
data of magnetite from the Los Colorados IOA deposit in chemistry of the deposits. In this study, we combine high-
northern Chile, we report compelling evidence for a change precision Fe and O isotope data with chemical analyses of
from purely magmatic to magmatic-hydrothermal conditions magnetite grains to peel back these layers in the world-
that agrees with the proposed model in which IOAs may
class Los Colorados IOA deposit. Based on these data, we
represent the deeper roots of some Andean IOCG systems.
propose a new model (Knipping et al., in press) that
Our proposed model invokes efficient concentration of
magnetite that takes place by buoyant segregation of early-
explains the observed coexisting purely magmatic and
formed magmatic magnetite-bubble pairs, followed by high- hydrothermal-magmatic features at Los Colorados and
temperature hydrothermal magnetite formation, driven by allows a genetic connection between Kiruna-type IOA and
decompression. These results suggest a link to mafic- IOCG deposits.
intermediate magmas and are indicative that IOA deposits
can source Fe-Cu-Au-rich fluids that according to
experimental studies, can migrate and cool to form IOCG 2 Samples
mineralization at shallower levels in the crust.
Samples from two drill cores were collected from the Los
Keywords: IOA, Kiruna-type, IOCG, magnetite, flotation Colorados (LC) IOA deposit, in the Chilean Iron Belt,
which hosts more than 40 IOA deposits in the Coastal
Range of northern Chile. All the deposits occur within the
1 Introduction southern segment of the Atacama Fault System. The LC
deposit (350 Mt Fe) hosts two parallel, magnetite-rich
Iron oxide-apatite (IOA) and iron oxide-copper-gold (>90 modal %) dikes that measure 1500 m strike, 150 m
(IOCG) deposits are of economic interest due to their wide, and ~500 m deep (total). The West dike contains
mineable amounts of magnetite and/or hematite and/or ~63% total iron and the East dike ~55% total iron. The
variable amounts of Cu, Au, REE, P, U, Ag and Co dikes are bounded on the west by a fault and the east by a
(Barton, 2014). While IOCG deposits are mostly thought brecciated zone, which contains ~25% total iron. LC is
to be formed by hydrothermal processes mainly due to a Cretaceous in age, and hosted in igneous rocks of the
lack of clear igneous correlation, the origin of IOA or Punta del Cobre Formation.
Kiruna-type deposits remains controversial. Some authors
invoke a hydrothermal origin, which can be either a non-
magmatic surface derived fluid that scavenges Fe from 3 Results
surrounding dioritic plutons and metasomatically replaces
volcanic rocks (Menard, 1995; Barton and Johnson, 2004;
Sillitoe and Burrows, 2002), or a magmatic-hydrothermal 3.1 High precision Fe and O isotopes
fluid that sources Fe directly from magmas (Pollard, 2006).
We report stable Fe and O isotope pairs for magnetite
A third hypothesis invokes liquid immiscibility between a
samples from two drill cores of the LC deposit. The
Fe-, P-rich oxide melt and a conjugate Si-rich melt, with
coalescence, separation and crystallization of the Fe-, P- resulting δ56Femgt values range from 0.09‰ to 0.24‰
rich oxide melt forming IOA deposits (Nyström and (average δ56Femgt = 0.17‰ ± 0.05‰ [2 σ]) and
Henríquez, 1994; Naslund et al., 2002). The first two δ18Omgt values range from 1.92‰ to 3.17‰ (average
hypotheses allow the possibility for a genetic connection δ18Omgt = 2.60‰ ± 0.04‰ [2 σ]. The Fe and O isotope
between IOA and IOCG deposits, which have been compositions of magnetite at LC overlap those of
observed within the same district (Sillitoe, 2003), whereas magnetite precipitated from a silicate melt or magmatic-
the third hypothesis distinguishes completely IOA deposits hydrothermal aqueous fluid, reported in the literature (i.e.,
394
AT 2 geología económica y recursos naturales

δ56Femgt: 0.06‰ to 0.5‰, and δ18Omgt: from 1.0‰ to 4.0‰, 4 A new model for IOA deposits: flotation
Heimann et al., 2008). The data also overlap the Fe and O of magmatic magnetite suspensions
isotope signature of magnetite from the Kiruna district
(Jonsson et al., 2013), and eliminate a purely low- The LC magnetite records a transition from purely
temperature hydrothermal origin for the Fe ore at LC. magmatic conditions (Type 1) to high-T magmatic-
hydrothermal conditions (Type 2) with decreasing T (Type
3.2 High resolution magnetite geochemistry 3). This compositional change suggests that the formation
of the LC magnetite ore resulted from a sequence of events
To distinguish between purely igneous and magmatic- involving a melt and a magmatic-hydrothermal fluid.
hydrothermal signatures that are merged as “magmatic” in
the previous section, high-resolution trace element Our model invokes hydrous, oxidized arc-magmas, where
analyses were performed on individual magnetite grains. magnetite is often the liquidus phase at 200 MPa (Fig. 2A).
This magmatic magnetite is enriched in elements such as
Electron microprobe analyzer (EMPA) data of most Ti, V, Mn, Al and Ga, consistent with Type 1 magnetite
magnetite grains from the center of the western dike and cores. The exsolving magmatic-hydrothermal fluid prefers
its border zone indicate a high-T magmatic origin to nucleate bubbles initially on mineral surfaces, and thus
(porphyry type) according to discrimination diagrams crystallizing magnetite promotes water supersaturation.
([Ti+V] vs. [Al+Mn]) of Dupuis and Beaudoin (2011) and The attachment of bubbles is energetically favored on
Nadoll et al. (2014) (Fig. 1). However, some magnetite magnetite microlites, and magnetite-bubble pairs form.
grains are zoned with euhedral cores rich in silicate The total density of these pairs is less than the surrounding
inclusions (Type 1) within a less porous magnetite matrix melt, allowing magnetite-bubble pairs to ascend (Fig. 2B).
(Type 2), which can be surrounded by a third generation of
porous magnetite (Type 3). The compositions of the During ascent, the magnetite-bubble pairs are able to
magnetite cores (Type 1) are consistent with Ti-rich “sweep up” other magnetite microlites becoming a rising
magnetite in nelsonites, which are thought to form by suspension rich in primary magnetite (Fig. 3C). Since H2O
purely magmatic processes, while Type 2 magnetite has a saturation is followed by significant partitioning of Cl into
high-T magmatic-hydrothermal fluid signature (porphyry the fluid phase, the exsolving fluid will become Cl-rich
field) (Fig. 1). Only samples distal from the dike center or which in turn has the ability to scavenge from silicate melt
distal from the grain cores (i.e., late growth zones) have up to several wt% Fe as FeCl2 (Simon et al., 2004) (Fig.
(Ti+V) and (Al+Mn) as low as expected for magnetite of 3C). The originally igneous magnetite can continue to
the Kiruna field (Type 3 magnetite). grow by sourcing Fe from the magnetite-fluid suspension,
and this magnetite is expected to have a chemical signature
The chemical patterns are therefore best interpreted to consistent with magmatic-hydrothermal magnetite.
reflect a change from purely magmatic to magmatic-
hydrothermal conditions during crystallization of the LC In fact, the lack of Na and K alteration that is common in
magnetite. magmatic-hydrothermal IOCG deposits (Barton, 2014) can
also be explained at LC by magnetite growth from a highly
saline brine instead from a low salinity vapor. It has been
shown experimentally that with decompression the
solubility of Na and K increase in the brine phase at 800°C
while iron solubility slightly decreases (Simon et al.,
2004). Consequently, Fe precipitation from brine would be
possible in the pressure range of the estimated paleodepth
of Los Colorados (3-4 km ~145-110 MPa) without the
formation of simultaneous K- and Na-rich minerals during
adiabatic decompression, in contrast to low salinity vapor.

The tectonic stress change in the back-arc setting, which


was responsible for generating the Atacama Fault System
(AFS) during the late Lower Cretaceous, may have created
hydraulic fractures that served as conduits for the ascent of
suspensions into the overlying crust. This explains the
efficient fluid transport and precipitation of massive
Figure 1. [Ti+V] versus [Al+Mn] diagram show the distribution magnetite as dikes (Fig. 2D) instead of a less efficient
of LC magnetite samples, as measured by EMPA (dark grey: segregation such as magnetite-rich enclaves observed in
center of the dike; light grey: distal from the dike center). The andesitic volcanoes (e.g., Soufrière Hills, Edmonds et al.,
green star shows the average composition of magnetite. Fields
2014). According to model calculations of Knipping et al.
after Dupuis and Beaudoin (2011) and Nadoll et al. (2014).
(in press) a magma chamber size with 50 km3 would be
sufficient to supply enough water and iron to create a
395
SIM 3 DEPÓSITOS DE TIPO IOCG DE LOS ANDES

deposit such as Los Colorados (~350 Mt Fe) even with a References


depositional efficiency of only 50% iron. This volume is in
the range of typical arc volcano magma chambers (~4-60 Barton, M.D. 2014. Iron oxide(–Cu–Au–REE–P–Ag–U–Co)
km3) and similar to estimated caldera sizes of extrusive systems. In Scott, S.D., ed., Treatise on Geochemistry (2nd
edition, volume 13): Amsterdam, Elsevier, p. 515–541.
IOA deposits (~30 km2; El Laco, Chile).
Barton, M.D.; Johnson, D.A., 1996. Evaporitic-source model for
igneous-related Fe-oxide (REE-Cu-Au-U) mineralization.
Geology 24: 259-262.

Dupuis, C.; Beaudoin, G. 2011. Discriminant diagrams for iron oxide


trace element fingerprinting of mineral deposit types. Mineralium
Deposita 46: 319–335.

Edmonds, M.; Brett, A.; Herd, R.A.; Humphreys, M.C.S.; Woods, A.


2014. Magnetite-bubble aggregates at mixing interfaces in
andesite magma bodies. In Zellmer, G.F., et al., eds., Geological
Society of London Special Publication 410: 95–121.

Heimann, A.; Beard, B.L.; Johnson, C.M. 2008. The role of volatile
exsolution and subsolidus fluid/rock interactions in producing
high 56Fe/54Fe ratios in siliceous igneous rocks. Geochimica et
Cosmochimica Acta 72: 4379–4396.

Jonsson, E.; Troll, V.R.; Hoegdahl, K.; Harris, C.; Weis, F.; Nilsson,
Figure 2. (A) Bubble nucleation on primary igneous magnetite; K.P.; Skelton, A. 2013. Magmatic origin of giant ‘Kiruna-type’
(B) ascent of bubble-magnetite pairs due to positive buoyancy; apatite-iron-oxide ores in Central Sweden. Nature Scientific
(C) further ascent and coalescence of primary magnetite as well Reports 3: 1644–1652.
as scavenging of Fe into the high-salinity fluids; (D) formation of
hydraulic fractures (due to tectonic stress changes) allowing Knipping, J.L.; Bilenker, L.D.; Simon, A.C.; Reich, M.; Barra, F.;
efficient segregation growth of hydrothermal magnetite. Deditius, A.P.; Lundstrom, C.; Bindeman, I.; Munizaga, R. in
press. Giant Kiruna-type deposits form by efficient flotation of
magmatic magnetite suspensions. Geology.
5 A genetic link between IOAs and IOCGs? Menard, J.J. 1995. Relationship between altered pyroxene diorite and
the magnetite mineralization in the Chilean Iron Belt, with
Our new model accounts for the observed geochemical emphasis on the El Algarrobo iron deposits (Atacama region,
Chile). Mineralium Deposita 30: 268-274.
features of magnetites at LC and the lack of Na and K
alteration, and can potentially support a genetic link Nadoll, P.; Angerer, T.; Mauk, J.L.; French, D.; Walshe, J. 2014. The
between IOA and IOCG deposits. Iron concentration of a chemistry of hydrothermal magnetite: A review. Ore Geology
Cl-rich aqueous fluid decreases slightly during Reviews 61: 1–32.
decompression (Simon et al., 2004), while concentrations
of Na and K strongly increase, allowing for magnetite Naslund, H.R.; Henriquez, F.; Nystroem, J.O.; Vivallo, W.; Dobbs,
F.M. 2002. Magmatic iron ores and associated mineralisation:
precipitation without simultaneous Na and K Examples from the Chilean high Andes and coastal Cordillera. In
mineralization. However, owing to retrograde solubility of Porter, T.M., ed., Adelaide, Australia, PGC Publishing 2: 207–
metals such as Fe, Cu, and Au, the magmatic-hydrothermal 226.
fluid that precipitates magnetite will continue transporting
significant amounts of dissolved Fe (plus Cu, Au) after Nyström, J.O.; Henríquez, F. 1994. Magmatic features of iron ores of
IOA deposition. Further ascent and cooling promotes the the Kiruna type in Chile and Sweden: Ore textures and magnetite
geochemistry. Economic Geology 89: 820–839.
precipitation of Cu-sulfides at T<420°C and at shallow
levels within the crust, as observed for IOCG deposits. Pollard, P.J. 2006. An intrusion-related origin for Cu-Au
This is consistent with the proposed model in which IOA mineralization in iron oxide-copper-gold (IOCG) provinces.
deposits represent the deeper roots of IOCG systems (e.g., Mineralium Deposita 41: 179–187.
Sillitoe, 2003), and may therefore be a step toward a
Sillitoe, R.H.; Burrows, D.R. 2002. New field evidence bearing on
systematic formation model for IOCG deposits.
the origin of the El Laco magnetite deposit, northern Chile.
Economic Geology 97: 1101–1109.
Acknowledgements
Sillitoe, R.H. 2003. Iron oxide-copper-gold deposits: An Andean
Martin Reich and Fernando Barra acknowledge funding view. Mineralium Deposita 38: 787–812.
from the MSI “Millennium Nucleus for Metal Tracing
Simon, A.C.; Pettke, T.; Candela, P.A.; Piccoli, P.M.; Heinrich, A.H.
Along Subduction” NC130065, FONDECYT grant 2004. Magnetite solubility and iron transport in magmatic-
#1140780 and FONDAP grant 15090013. We thank CAP hydrothermal environments. Geochimica et Cosmochimica Acta
geologist Mario Lagos for his help during fieldwork. 68: 4905–4914.

396

You might also like