29.10.2014 Views

Iron oxide-copper-gold deposits in the Kolari region, Finland (PDF)

Iron oxide-copper-gold deposits in the Kolari region, Finland (PDF)

Iron oxide-copper-gold deposits in the Kolari region, Finland (PDF)

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Ore Geology Reviews 30 (2007) 75–105<br />

www.elsevier.com/locate/oregeorev<br />

Geology, geochemistry, fluid <strong>in</strong>clusion characteristics, and U–Pb age<br />

studies on iron <strong>oxide</strong>–Cu–Au <strong>deposits</strong> <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong>,<br />

nor<strong>the</strong>rn F<strong>in</strong>land<br />

Tero Niiranen a,b, *, Matti Poutia<strong>in</strong>en a , Irmeli Mänttäri c<br />

a Department of Geology, P.O. Box 64, FIN-00014 University of Hels<strong>in</strong>ki, F<strong>in</strong>land<br />

b Northland Resources Inc., Teknotie 14-16, 11, FIN-96930 Rovaniemi, F<strong>in</strong>land<br />

c Geological Survey of F<strong>in</strong>land, P.O. Box 96, FIN-02151 Espoo, F<strong>in</strong>land<br />

Received 21 February 2005; accepted 23 November 2005<br />

Available onl<strong>in</strong>e 19 January 2006<br />

Abstract<br />

Several iron <strong>oxide</strong>–<strong>copper</strong>–<strong>gold</strong> <strong>deposits</strong> are known <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong>, <strong>in</strong> <strong>the</strong> western part of <strong>the</strong> Central Lapland<br />

greenstone belt, nor<strong>the</strong>rn F<strong>in</strong>land. They are hosted by cl<strong>in</strong>opyroxene-dom<strong>in</strong>ated skarns that were formed near to contact zone<br />

between ca. 1860 Ma Haparanda Suite <strong>in</strong>trusions and N2050 Ma Savukoski Group supracrustal rocks. All <strong>deposits</strong> are located<br />

with<strong>in</strong> or next to shear and fault zones form<strong>in</strong>g parts of <strong>the</strong> major, NNE-trend<strong>in</strong>g, <strong>Kolari</strong> shear zone. Three of <strong>the</strong> Fe–Cu–Au<br />

<strong>deposits</strong>, Kuervitikko, Cu-Rautuvaara, and Laur<strong>in</strong>oja were studied; all conta<strong>in</strong> significant amounts of Cu (0.1% to 4.5%) and Au<br />

(0.1 to 6.6 ppm). At Laur<strong>in</strong>oja and Kuervitikko, Cu and Au are hosted by ironstone and skarn. At Cu-Rautuvaara, <strong>the</strong> host rock<br />

is a magnetite-dissem<strong>in</strong>ated albitite. The <strong>deposits</strong> have a dist<strong>in</strong>ct metal association of Fe–Cu–AuFAg, Bi, Ba, Co, Mo, Sb, Se,<br />

Te, Th, U, LREE. The wall and host rocks are <strong>in</strong>tensely altered and display a deposit-scale zonal pattern. The distal alteration<br />

zone is characterised by albiteFbiotite, K-feldspar, and scapolite and <strong>the</strong> proximal zone by cl<strong>in</strong>opyroxene–magnetiteFamphibole,<br />

scapolite, calcite, and sulphides. Mass balance calculations <strong>in</strong>dicate that Al 2 O 3 ,TiO 2 , and Zr were immobile<br />

dur<strong>in</strong>g alteration. The calculations also <strong>in</strong>dicate that significant quantities of Fe 2 O 3 , CaO, CO 2 , S, Cu, Au, Bi, and Te were<br />

added to <strong>the</strong> proximal altered rocks. The ma<strong>in</strong> ga<strong>in</strong>s <strong>in</strong> <strong>the</strong> distal altered rocks are <strong>in</strong> Na 2 O, K 2 O, and Ba. Fluid <strong>in</strong>clusion data<br />

suggest that <strong>the</strong> fluids which circulated <strong>in</strong> <strong>the</strong> rocks dur<strong>in</strong>g <strong>the</strong> ma<strong>in</strong> m<strong>in</strong>eralisation event and subsequent brittle fractur<strong>in</strong>g were<br />

highly sal<strong>in</strong>e (V56 wt.% NaCl) H 2 OFCO 2 fluids. The temperature dur<strong>in</strong>g <strong>the</strong> ma<strong>in</strong> m<strong>in</strong>eralisation event was between 450 and<br />

550 8C and <strong>the</strong> pressure was 1.5 to 3.5 kbar.<br />

Based on U–Pb age data of magmatic zircon from altered hang<strong>in</strong>g wall diorite and granite that brecciates <strong>the</strong> ore, <strong>the</strong> age of <strong>the</strong><br />

<strong>deposits</strong> is between 1864F5 and 1766F5 Ma. The 1797F5 Ma age of zircon <strong>in</strong> skarn, comb<strong>in</strong>ed with <strong>the</strong> 1810 to 1780 Ma ages<br />

of <strong>the</strong> metamorphic titanite <strong>in</strong> altered wall rocks and skarn, suggest that <strong>the</strong> <strong>deposits</strong> were most likely formed ca. 1800 Ma. This age<br />

has been <strong>in</strong>terpreted to be broadly contemporaneous with <strong>the</strong> D 3 thrust<strong>in</strong>g event <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> dur<strong>in</strong>g which <strong>the</strong> <strong>Kolari</strong> shear<br />

zone was activated.<br />

The data presented <strong>in</strong> this work are <strong>in</strong>consistent with previously proposed models for <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> that <strong>the</strong>y are<br />

metamorphosed syngenetic iron formations, or skarns related to ca. 1860 Ma monzonite <strong>in</strong>trusions. Instead, we propose that <strong>the</strong><br />

<strong>Kolari</strong> Fe–Cu–Au <strong>deposits</strong> are of metasomatic replacement-type, and are controlled by <strong>the</strong> <strong>Kolari</strong> shear zone structures related to<br />

* Correspond<strong>in</strong>g author. Northland Resources Inc., Teknotie 14-16, 11, FIN-96930 Rovaniemi, F<strong>in</strong>land. Fax: +358 16 362 620.<br />

E-mail address: tero.n@nrm<strong>in</strong>e.com (T. Niiranen).<br />

0169-1368/$ - see front matter D 2005 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.oregeorev.2005.11.002


76<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

post-peak metamorphic D 3 thrust<strong>in</strong>g event <strong>in</strong> nor<strong>the</strong>rn F<strong>in</strong>land. Our data suggest that <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> best fit <strong>in</strong>to <strong>the</strong> category of<br />

epigenetic iron <strong>oxide</strong>–<strong>copper</strong>–<strong>gold</strong> m<strong>in</strong>eralisation.<br />

D 2005 Elsevier B.V. All rights reserved.<br />

Keywords: IOCG <strong>deposits</strong>; Hydro<strong>the</strong>rmal alteration; Proterozoic; Fluid <strong>in</strong>clusions; <strong>Kolari</strong>; F<strong>in</strong>land<br />

1. Introduction<br />

Several Fe FCu, Au <strong>deposits</strong> are known <strong>in</strong> <strong>the</strong><br />

<strong>Kolari</strong> <strong>region</strong> <strong>in</strong> <strong>the</strong> western part of <strong>the</strong> Proterozoic<br />

Central Lapland greenstone belt (CLGB), F<strong>in</strong>land<br />

(Fig. 1). Deposits at Hannuka<strong>in</strong>en and Rautuvaara<br />

were m<strong>in</strong>ed for iron between 1974 and 1990 with a<br />

total of 16 Mt iron ore produced (Puust<strong>in</strong>en, 2003). The<br />

total <strong>in</strong>ferred amount of iron ore <strong>in</strong> <strong>the</strong> known <strong>deposits</strong><br />

is several tens of Mt (Hiltunen, 1982). Besides iron,<br />

four <strong>deposits</strong> <strong>in</strong> <strong>the</strong> <strong>region</strong> conta<strong>in</strong> significant amounts<br />

of <strong>copper</strong> and <strong>gold</strong>. However, up until now, Au and Cu<br />

have only been extracted from <strong>the</strong> Laur<strong>in</strong>oja ore body<br />

of <strong>the</strong> Hannuka<strong>in</strong>en m<strong>in</strong>e (Fig. 1; Hiltunen, 1982;<br />

Puust<strong>in</strong>en, 2003).<br />

Almost all of <strong>the</strong> known Fe–Cu–Au <strong>deposits</strong> are<br />

located near to contact zones between ca. 1860 Ma<br />

monzonite–diorite <strong>in</strong>trusions and N2050 Ma Savukoski<br />

Fig. 1. General geological features of nor<strong>the</strong>rn F<strong>in</strong>land (a), location and general geological features of <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> (b), and location of<br />

FeFCu–Au <strong>deposits</strong> (c). Location of <strong>the</strong> <strong>Kolari</strong> shear zone (KSZ) (c), modified after Korsman et al. (1997), l<strong>in</strong>eation and fold axis data after<br />

Hiltunen (1982) and Väisänen (2002); stratigraphy groups after Lehtonen et al. (1998). The coord<strong>in</strong>ate system is <strong>the</strong> F<strong>in</strong>nish national system (YKJ).<br />

CLGB=Central Lapland greenstone belt.


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 77<br />

Group supracrustal rocks (Fig. 1; Hiltunen, 1982; Lehtonen<br />

et al., 1998). Nearly all of <strong>the</strong> <strong>deposits</strong> are located<br />

<strong>in</strong>, or next, to thrust or o<strong>the</strong>r fault zones that form <strong>the</strong><br />

complex N- to NNE-trend<strong>in</strong>g <strong>Kolari</strong> Shear System (e.g.,<br />

Väisänen, 2002). Hiltunen (1982) suggests that <strong>the</strong><br />

<strong>Kolari</strong> <strong>deposits</strong> are stratabound skarns that were formed<br />

by contact metasomatic events related to <strong>in</strong>trusion of<br />

monzonites. Mäkelä and Tammenmaa (1978) and<br />

Frietsch et al. (1995) proposed, based on <strong>the</strong> sulphur<br />

isotope composition of sulphides, that <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong><br />

are metamorphosed syngenetic iron formations. However,<br />

<strong>the</strong> <strong>deposits</strong> display characteristics that are<br />

<strong>in</strong>consistent with both <strong>the</strong>se models. These <strong>in</strong>clude a<br />

ca. 65 Ma time difference between <strong>the</strong> U–Pb zircon<br />

ages of <strong>the</strong> monzonite and skarn that hosts <strong>the</strong> ore<br />

(Hiltunen, 1982) and <strong>the</strong> lack of any textural evidence<br />

that would imply a sedimentary orig<strong>in</strong> for <strong>the</strong> iron.<br />

Instead, many features of <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> are typical<br />

of iron <strong>oxide</strong>–<strong>copper</strong>–<strong>gold</strong> (IOCG) <strong>deposits</strong> that are now<br />

widely recognised as a global class of ore <strong>deposits</strong> (e.g.,<br />

Hitzman et al., 1992; Barton and Johnson, 1996; Hitzman,<br />

2000).<br />

The purpose of this work is to present U–Pb, fluid<br />

<strong>in</strong>clusion, and new geochemical data on <strong>the</strong> <strong>Kolari</strong><br />

<strong>deposits</strong> <strong>in</strong> order to establish whe<strong>the</strong>r <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong><br />

are skarns sensu stricto, metamorphosed syngenetic<br />

iron formations, or if <strong>the</strong> IOCG model better expla<strong>in</strong>s<br />

<strong>the</strong>ir characteristics.<br />

2. General geology<br />

The bedrock of nor<strong>the</strong>rn F<strong>in</strong>land (Fig. 1) consists of<br />

an Archean basement, Palaeoproterozoic greenstone<br />

and schist belts, ca. 1900 Ma granulite belt, and Svecofennian<br />

1930 to 1770 Ma granitoids (e.g., Hanski et<br />

al., 1997). The CLGB was formed dur<strong>in</strong>g prolonged<br />

stages of rift<strong>in</strong>g of <strong>the</strong> Archean craton, with sedimentation,<br />

and magmatism <strong>in</strong> <strong>in</strong>tracratonic and cratonic<br />

marg<strong>in</strong> rift sett<strong>in</strong>gs between 2500 and 1900 Ma (e.g.,<br />

Hanski et al., 1997; Lehtonen et al., 1998), and was<br />

subjected to multiphase deformation and metamorphism<br />

dur<strong>in</strong>g orogenic events between 1920 and 1770<br />

Ma (Ward et al., 1989; Lehtonen et al., 1998; Väisänen,<br />

2002; Laht<strong>in</strong>en et al., 2003; Sorjonen-Ward et al., 2003;<br />

Hölttä et al., <strong>in</strong> press).<br />

Supracrustal rocks of <strong>the</strong> CLGB consist of terrestrial<br />

to deep mar<strong>in</strong>e sediments with thick <strong>in</strong>tercalations of<br />

ultramafic to mafic volcanic rocks deposited on <strong>the</strong><br />

metamorphosed Archean basement. The supracrustal<br />

sequence has been divided <strong>in</strong>to seven lithostratigraphic<br />

units which are, from oldest to youngest: Salla,<br />

Onkamo, Sodankylä, Savukoski, Kittilä, Kumpu and<br />

La<strong>in</strong>io Groups (Fig. 2; Lehtonen et al., 1998). Intrusives<br />

<strong>in</strong> <strong>the</strong> CLGB consist of ca. 2440 Ma layered<br />

<strong>in</strong>trusions, ca. 2220 Ma mafic sills and dykes, 2050<br />

Ma layered <strong>in</strong>trusions and diabases, 1930 to 1860 Ma<br />

dom<strong>in</strong>antly <strong>in</strong>termediate to felsic <strong>in</strong>trusions, and ca.<br />

1820 to 1770 Ma felsic <strong>in</strong>trusions (Lehtonen et al.,<br />

1998; Mutanen and Huhma, 2001; Hanski et al.,<br />

2001; Rastas et al., 2001).<br />

In <strong>the</strong> <strong>Kolari</strong> <strong>region</strong>, <strong>the</strong> supracrustal rocks consist of<br />

Onkamo Group tholeiites and komatiites, Sodankylä<br />

Group quartzites, mica schists and gneisses, and conglomerates,<br />

Savukoski Group Fe–tholeiites, tuffites,<br />

dolomitic marbles and black schists, and La<strong>in</strong>io<br />

Group quartzites, siltstones, conglomerates and m<strong>in</strong>or<br />

volcanic rocks (Fig. 1; Hiltunen, 1982; Lehtonen et al.,<br />

1998). The ca. 1930 to 1860 Ma diorites, tonalites, and<br />

monzonites constitute <strong>the</strong> ma<strong>in</strong> <strong>in</strong>trusives <strong>in</strong> <strong>the</strong> <strong>region</strong><br />

(Hiltunen, 1982; Väänänen and Lehtonen, 2001). In<br />

addition to <strong>the</strong>se, variably albitised 2200 to 2000 Ma<br />

diabases as well as ca. 1820 to 1770 Ma granitic rocks<br />

occur <strong>in</strong> <strong>the</strong> <strong>region</strong> (Hiltunen, 1982; Väänänen and<br />

Lehtonen, 2001; this work).<br />

The structural evolution of <strong>the</strong> CLGB dur<strong>in</strong>g <strong>the</strong><br />

Svecofennian orogeny is currently under revision, lead<strong>in</strong>g<br />

to somewhat confus<strong>in</strong>g nomenclature between <strong>the</strong><br />

older and <strong>the</strong> more recent reports. Lehtonen et al.<br />

(1998) divide <strong>the</strong> ma<strong>in</strong> deformation stages <strong>in</strong> <strong>the</strong><br />

CLGB <strong>in</strong>to five different events, whereas Väisänen<br />

(2002) and Hölttä et al. (<strong>in</strong> press) dist<strong>in</strong>guish three<br />

ductile deformation stages and subsequent brittle stages<br />

which all took place with<strong>in</strong> <strong>the</strong> period 1920 to 1770<br />

Ma. The deformation stages discussed <strong>in</strong> this work are<br />

after Väisänen (2002) and Hölttä et al. (<strong>in</strong> press), except<br />

that <strong>the</strong> completely brittle stages follow<strong>in</strong>g <strong>the</strong> D 3 are<br />

referred here as D 4 .<br />

The D 1 structures comprise bedd<strong>in</strong>g parallel S 1 foliation<br />

that is only microscopically observable (Väisänen,<br />

2002; Hölttä et al., <strong>in</strong> press). Lehtonen et al. (1998)<br />

suggest that gentle fold<strong>in</strong>g was associated with D 1 .<br />

They also proposed that a static stage occurred between<br />

D 1 and D 2 dur<strong>in</strong>g which <strong>the</strong> metamorphic conditions<br />

already reached greenschist to amphibolite facies conditions.<br />

The most prom<strong>in</strong>ent structural features <strong>in</strong> <strong>the</strong><br />

CLGB were developed dur<strong>in</strong>g <strong>the</strong> D 2 stage between<br />

1890 and 1860 Ma (Hiltunen, 1982; Sorjonen-Ward et<br />

al., 1992; Lehtonen et al., 1998; Väisänen, 2002; Hölttä<br />

et al., <strong>in</strong> press).<br />

The D 2 structures are characterised by tight to isocl<strong>in</strong>al<br />

F 2 folds and S 2 foliation. Generally <strong>the</strong> S 2 foliation<br />

is subparallel to bedd<strong>in</strong>g, gently dipp<strong>in</strong>g to<br />

horizontal and <strong>the</strong> F 2 folds are recumbed to recl<strong>in</strong>ed,<br />

although locally <strong>the</strong> foliation and axial planes are steep-


78<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Fig. 2. Stratigraphy of <strong>the</strong> Central Lapland greenstone belt toge<strong>the</strong>r with age data for different igneous rocks <strong>in</strong> <strong>the</strong> area. After Lehtonen et al. (1998)<br />

and Hanski et al. (2001) and references <strong>the</strong>re<strong>in</strong>.<br />

ly dipp<strong>in</strong>g or vertical. K<strong>in</strong>ematic <strong>in</strong>dicators suggest that<br />

<strong>the</strong>se structures were caused by horizontal movement<br />

related to thrust tectonics and that <strong>the</strong> movement direction<br />

was probably from SSW to NNE (Väisänen, 2002;<br />

Hölttä et al., <strong>in</strong> press).<br />

The thrusts, which were <strong>in</strong>itiated <strong>in</strong> earlier stages,<br />

were re-activated dur<strong>in</strong>g <strong>the</strong> D 3 stage when <strong>the</strong> D 1–2<br />

structures were overpr<strong>in</strong>ted by F 3 folds, S 3 foliation,<br />

and L 3 l<strong>in</strong>eation with an extreme variation <strong>in</strong> direction,<br />

dip, plunge, and <strong>in</strong>tensity (Lehtonen et al., 1998;<br />

Väisänen, 2002; Hölttä et al., <strong>in</strong> press). The variation<br />

<strong>in</strong> <strong>the</strong> direction of <strong>the</strong> D 3 structures has been<br />

expla<strong>in</strong>ed by ei<strong>the</strong>r simultaneous thrust<strong>in</strong>g from oppos<strong>in</strong>g<br />

direction or with progressive thrust<strong>in</strong>g that was<br />

accompanied with rotation (Sorjonen-Ward et al.,<br />

1992; Lehtonen et al., 1998; Väisänen, 2002; Hölttä<br />

et al., <strong>in</strong> press). The absolute age of <strong>the</strong> D 3 probably<br />

varies among <strong>the</strong> different parts of <strong>the</strong> CLGB but it is<br />

between 1.86 and 1770 Ma (Väisänen, 2002; Hölttä et<br />

al., <strong>in</strong> press).<br />

The peak metamorphic conditions <strong>in</strong> <strong>the</strong> CLGB<br />

were probably reached dur<strong>in</strong>g <strong>the</strong> D 2 stage (Lehtonen<br />

et al., 1998; Hölttä et al., <strong>in</strong> press). The metamorphic<br />

grade varies from greenschist to granulite<br />

facies conditions <strong>in</strong> <strong>the</strong> metamorphic zones that are<br />

separated by <strong>the</strong> D 3 thrust planes (Hölttä et al., <strong>in</strong><br />

press).<br />

Structural features of <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> are dom<strong>in</strong>ated<br />

by a number of thrust, shear, and fault zones that<br />

form <strong>the</strong> <strong>Kolari</strong> Shear System which is part of <strong>the</strong> N–S<br />

oriented, up to 50 to 100 km wide, Baltic-Bothnian<br />

Mega-shear (Fig. 1; Ber<strong>the</strong>lsen and Marker, 1986; Väisänen,<br />

2002). The thrust and shear zones lie parallel or<br />

subparallel to <strong>the</strong> F 2 fold axis of <strong>the</strong> syncl<strong>in</strong>e–anticl<strong>in</strong>e<br />

structures developed dur<strong>in</strong>g <strong>the</strong> NNE directed D 2<br />

thrust<strong>in</strong>g (Hiltunen, 1982; Väisänen, 2002; Hölttä et<br />

al., <strong>in</strong> press). Thrust<strong>in</strong>g dur<strong>in</strong>g <strong>the</strong> D 3 stage led to<br />

reactivation of <strong>the</strong> D 2 thrust and fault zones, refold<strong>in</strong>g<br />

of <strong>the</strong> D 2 structures, and development of strong L 3<br />

l<strong>in</strong>eation <strong>in</strong> <strong>the</strong> vic<strong>in</strong>ity of <strong>the</strong> thrust zones (Hiltunen,<br />

1982; Väisänen, 2002). The k<strong>in</strong>ematic <strong>in</strong>dicators suggest<br />

that, <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong>, <strong>the</strong> D 3 structures were<br />

caused by W to E directed thrust<strong>in</strong>g (Hiltunen, 1982;<br />

Väisänen, 2002; Hölttä et al., <strong>in</strong> press). As a result, <strong>the</strong><br />

areas around <strong>the</strong> thrust zones are characterised by welldeveloped,<br />

gently to moderately W to SW plung<strong>in</strong>g L 3<br />

l<strong>in</strong>eation, and F 3 folds with axes plung<strong>in</strong>g parallel to <strong>the</strong><br />

l<strong>in</strong>eation (Fig. 1).


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 79<br />

3. Analytical procedures and m<strong>in</strong>eral abbreviations<br />

Geochemical analyses were carried out at <strong>the</strong> laboratories<br />

of <strong>the</strong> Geological Survey of F<strong>in</strong>land (GSF).<br />

Whole-rock analyses were made us<strong>in</strong>g standard wavelength<br />

dispersive X-ray fluorescence spectrometry<br />

(XRF) from pressed powder pellets, and rare earth<br />

elements (REE) were analysed us<strong>in</strong>g <strong>in</strong>ductively coupled<br />

plasma mass spectrometry (ICP-MS) after HF–<br />

HClO 4 digestion and lithium metaborate–sodium perborate<br />

fusion. Ag and Co were analysed with <strong>in</strong>ductively<br />

coupled plasma atomic emission spectrometry<br />

(ICP-AES) after aqua regia digestion. Au, Bi, Sb, Se,<br />

and Te were analysed after aqua regia digestion us<strong>in</strong>g<br />

atomic absorption spectrometry and electro<strong>the</strong>rmal atomisation<br />

<strong>in</strong> a graphite furnace (GFAAS).<br />

Fig. 3. Geology of <strong>the</strong> Hannuka<strong>in</strong>en deposit. Note that <strong>the</strong> Kivivuopio ore body is not <strong>in</strong>cluded <strong>in</strong>to <strong>the</strong> surface geology map because its geometry<br />

and size is poorly known. After Hiltunen (1982).


80<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Fluid-<strong>in</strong>clusion micro<strong>the</strong>rmometry was carried out<br />

us<strong>in</strong>g a L<strong>in</strong>kam THMSG 600 heat<strong>in</strong>g-freez<strong>in</strong>g stage at<br />

<strong>the</strong> University of Hels<strong>in</strong>ki. Accuracy levels of F0.1 and<br />

F2.0 8C were achieved for subzero and higher temperatures,<br />

respectively, by calibration us<strong>in</strong>g syn<strong>the</strong>tic fluid<strong>in</strong>clusion<br />

standards (SynFl<strong>in</strong>c). FLINCOR (Brown,<br />

1989) and L<strong>in</strong>kam PVTX software were used to analyse<br />

laboratory micro<strong>the</strong>rmometric data.<br />

For conventional U–Pb age determ<strong>in</strong>ations, <strong>the</strong> decomposition<br />

of zircons and titanites and extraction of U<br />

and Pb follows ma<strong>in</strong>ly <strong>the</strong> procedure described by<br />

Krogh (1973, 1982). 235 U– 208 Pb-spiked and unspiked<br />

isotopic ratios were measured us<strong>in</strong>g a VG Sector 54<br />

<strong>the</strong>rmal ionization multicollector mass spectrometer.<br />

The measured Pb and U isotopic ratios were normalised<br />

to <strong>the</strong> accepted ratios of SRM 981 and U500 standards.<br />

The Pb/U ratios were calculated us<strong>in</strong>g <strong>the</strong> PbDat-program<br />

(Ludwig, 1991) and fitt<strong>in</strong>g of <strong>the</strong> discordia l<strong>in</strong>es as<br />

well as calculation of <strong>the</strong> ages were completed us<strong>in</strong>g <strong>the</strong><br />

Isoplot/Ex 3 program (Ludwig, 2003).<br />

M<strong>in</strong>eral abbreviations used <strong>in</strong> figures, diagrams and<br />

tables <strong>in</strong> this work are: ab =albite, act=act<strong>in</strong>olite,<br />

am=amphibole, atp=anthophyllite, Au=native <strong>gold</strong>,<br />

bt=biotite, cc=calcite, cpx=cl<strong>in</strong>opyroxene, cpy=chalcopyrite,<br />

ep=epidote, gr=garnet, hbl=hornblende,<br />

kfs=K-feldspar, mo=molybdenite, mgt =magnetite,<br />

po=pyrrhotite, py=pyrite, sca=scapolite, te=tellurides,<br />

tit=titanite, ura=uran<strong>in</strong>ite.<br />

Fig. 4. Vertical sections across <strong>the</strong> Kuervitikko (a) and Cu-Rautuvaara (b) <strong>deposits</strong>. After Hiltunen (1982).


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 81<br />

4. Deposit description<br />

Three Cu- and Au-rich iron <strong>oxide</strong> <strong>deposits</strong> were<br />

selected for this study: Cu-Rautuvaara, Kuervitikko,<br />

and Laur<strong>in</strong>oja (Figs. 1, 3, and 4); Table 1). Of<br />

<strong>the</strong>se, Laur<strong>in</strong>oja is one of <strong>the</strong> five dist<strong>in</strong>ct ore bodies<br />

of <strong>the</strong> Hannuka<strong>in</strong>en m<strong>in</strong>e. All of <strong>the</strong> selected <strong>deposits</strong><br />

are located <strong>in</strong> or next to <strong>the</strong> same thrust zone (Fig. 1).<br />

The host rocks and alteration sequence of <strong>the</strong> Laur<strong>in</strong>oja<br />

and Kuervitikko are very similar with each<br />

o<strong>the</strong>r (Figs. 2 and 3). In <strong>the</strong> Cu-Rautuvaara deposit,<br />

located ca. 7 km SW of <strong>the</strong> Hannuka<strong>in</strong>en m<strong>in</strong>e, <strong>the</strong><br />

lithology is relatively similar to <strong>the</strong> Laur<strong>in</strong>oja and<br />

Kuervitikko <strong>deposits</strong>, but <strong>in</strong>stead of ironstone and<br />

skarn, <strong>the</strong> Cu–Au m<strong>in</strong>eralisation is hosted by magnetite<br />

dissem<strong>in</strong>ated albitite (Fig. 4). Never<strong>the</strong>less, <strong>the</strong><br />

structural sett<strong>in</strong>g and general features of <strong>the</strong> Cu-Rautuvaara<br />

are so similar to <strong>the</strong> Laur<strong>in</strong>oja and Kuervitikko<br />

that it can be considered to have been formed by<br />

<strong>the</strong> same k<strong>in</strong>d of processes that led to <strong>the</strong> formation of<br />

<strong>the</strong> latter two <strong>deposits</strong>.<br />

4.1. Laur<strong>in</strong>oja and Kuervitikko<br />

The Laur<strong>in</strong>oja and Kuervitikko <strong>deposits</strong> consist of<br />

massive to silicate-banded magnetite-rich lenses (ironstones)<br />

(Fig. 5), that are hosted by skarns formed near to<br />

<strong>the</strong> contact zone between <strong>the</strong> Savukoski Group supracrustal<br />

sequence and ca. 1860 Ma monzonite–diorite<br />

<strong>in</strong>trusions (Hiltunen, 1982). The footwall of <strong>the</strong>se<br />

<strong>deposits</strong> consists of variably altered mafic metavolcanic<br />

rocks (Fig. 6d–g), quartzite quartz–feldspar schist and<br />

mica gneiss. Monzonite (Fig. 6a) and variably altered<br />

diorite (Fig. 6b–c) comprise <strong>the</strong> hang<strong>in</strong>g wall (Figs. 3<br />

and 4). The skarns consist ma<strong>in</strong>ly of cl<strong>in</strong>opyroxene<br />

(diopside–hedenbergite) and amphibole (act<strong>in</strong>olite–<br />

hornblende) dom<strong>in</strong>ated rocks. Also, cl<strong>in</strong>opyroxenebear<strong>in</strong>g<br />

scapolite (marialite) skarns (Fig. 6i) are abundant,<br />

and th<strong>in</strong> (b0.5 m) garnet-rich (andradite) horizons<br />

occur locally <strong>in</strong> <strong>the</strong> sequence. In addition, rocks consist<strong>in</strong>g<br />

ma<strong>in</strong>ly of albite with vary<strong>in</strong>g amounts of biotite,<br />

amphibole, K-feldspar and quartz (Fig. 6h), and referred<br />

to as albitites <strong>in</strong> this work, are common <strong>in</strong> <strong>the</strong> sequence.<br />

A small lens of carbonate rock occurs with<strong>in</strong> <strong>the</strong> cl<strong>in</strong>opyroxene–amphibole<br />

skarn at Kuervitikko (Fig. 4),<br />

whereas no carbonate rocks have been found at Laur<strong>in</strong>oja.<br />

Pegmatitic granite dykes, b1 to several m <strong>in</strong><br />

thickness, are abundant <strong>in</strong> both areas and, <strong>in</strong> places,<br />

ca. 1780 Ma medium-gra<strong>in</strong>ed granite dykes crosscut<br />

and brecciate <strong>the</strong> ore and wall rocks (Fig. 5c–d).<br />

The ma<strong>in</strong> <strong>oxide</strong> m<strong>in</strong>eral <strong>in</strong> both <strong>deposits</strong> is magnetite;<br />

pyrite, pyrrhotite, and chalcopyrite are <strong>the</strong> dom<strong>in</strong>ant<br />

sulphides (Fig. 7). Locally, m<strong>in</strong>or amounts of<br />

molybdenite occur <strong>in</strong> association with chalcopyrite<br />

(Fig. 7g). Native <strong>gold</strong> has been detected <strong>in</strong> silicate<br />

gangue, chalcopyrite, and magnetite <strong>in</strong> both <strong>the</strong> Laur<strong>in</strong>oja<br />

and Kuervitikko occurrences (Fig. 7i–l). Cl<strong>in</strong>opyroxene,<br />

amphibole, albite, scapolite, biotite, calcite,<br />

and quartz comprise <strong>the</strong> gangue <strong>in</strong> <strong>the</strong> ironstones.<br />

Sulphides ma<strong>in</strong>ly occur as dissem<strong>in</strong>ations, but locally<br />

also as ve<strong>in</strong>s <strong>in</strong> <strong>the</strong> ironstones, skarns and altered<br />

wall rocks. In places, pyrrhotite and chalcopyrite form<br />

narrow (b30 cm wide), massive lenses or ve<strong>in</strong>s <strong>in</strong> <strong>the</strong><br />

ironstone, and locally chalcopyrite and pyrite, toge<strong>the</strong>r<br />

with variable amounts of quartz, scapolite, K-feldspar<br />

and calcite occur as fracture <strong>in</strong>fill <strong>in</strong> <strong>the</strong> ironstone (Fig.<br />

Table 1<br />

Grades, tonnages and general characteristics of <strong>the</strong> Laur<strong>in</strong>oja, Kuervitikko, and Cu-Rautuvaara <strong>deposits</strong><br />

Deposit/prospect Laur<strong>in</strong>oja Kuervitikko Cu-Rautavaara<br />

Size and grade<br />

4.56 Mt* at 43% Fe,<br />

1 ppm Au, 0.88% Cu;<br />

33 Mt** at 36–53% Fe,<br />

b0.1–11.0% Cu,<br />

b0.1–6.6 ppm Au,<br />

b0.1–17.7 ppm Ag,<br />

20–1000 ppm Co<br />

Monzonite, diorite,<br />

mafic metavolcanic rock<br />

1.2 Mt** at 36–53% Fe,<br />

b0.1–8.3% Cu, b0.1–6.0 ppm Au,<br />

b0.1–2.1 ppm Ag, 50–830 ppm Co<br />

N4 Mt** weakly to moderately<br />

Cu–Au m<strong>in</strong>eralised rock,<br />

b0.1–1.5% Cu, b0.1–2.6 ppm Au,<br />

b0.5–1.2 ppm Ag<br />

Wall rocks<br />

Monzonite, diorite,<br />

Monzonite, diorite, mafic<br />

mafic metavolcanic rock, albitite metavolcanic rock<br />

Ma<strong>in</strong> host rock(s) <strong>Iron</strong>stone, cpx–afb skarn <strong>Iron</strong>stone, cpx–afb skarn Albitite<br />

Opaques mgt, py, po, cpyFmo, Au, te mgt, py, cpy, poFAu, te mgt, po, cpyFpy, ura<br />

Gangue<br />

cpx, hbl–act, ab, bt, Fsca, cpx, hbl–act, ab, Fsca, bt,<br />

ab, atf, bt, Fcpx, qtz, tit<br />

qtz, cc, kfs, gr, ep<br />

cc, kfs, qtz, gr, ep<br />

Metal association Fe, Cu, Au, SFAg, Bi, Ba, Fe, Cu, Au, SFAg, Ba, Bi,<br />

Fe, Cu, Au, SFAg, Ba, Bi,<br />

Co, Mo, Sb, Te, LREE<br />

Co, Mo, Se, Te, LREE<br />

Mo, Se, Te, Th, U, LREE<br />

Data from K<strong>in</strong>nunen (1980), Hiltunen (1982), Ke<strong>in</strong>änen (1995), Puust<strong>in</strong>en (2003) and this work. See Analytical procedures and m<strong>in</strong>eral<br />

abbreviations for m<strong>in</strong>eral abbreviations. *Production 1978–1990. **Inferred.


82<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Fig. 5. (a) Typical barren massive ironstone. Note <strong>the</strong> S 3 foliation <strong>in</strong> <strong>the</strong> rock (dotted white l<strong>in</strong>e). The S 3 is crosscut by chalcopyrite and scapolite–<br />

quartz–chalcopyrite–pyrite filled fractures along <strong>the</strong> direction of <strong>the</strong> white dashed l<strong>in</strong>e. (b) Close up photograph of <strong>the</strong> fractures from <strong>the</strong> area<br />

marked with white rectangle <strong>in</strong> (a). (c) Massive barren ironstone that is brecciated by ca. 1780 Ma granite. (d) Cu–Au-m<strong>in</strong>eralised ironstone that is<br />

<strong>in</strong>truded by ca. 1780 Ma granite. The co<strong>in</strong> <strong>in</strong> <strong>the</strong> figures is 26 mm <strong>in</strong> diameter. All photos from Laur<strong>in</strong>oja.<br />

5a–b). Magnetite <strong>in</strong> <strong>the</strong> sulphide-bear<strong>in</strong>g ironstones is<br />

typically coarser and more euhedral <strong>in</strong> form than <strong>in</strong> <strong>the</strong><br />

barren ironstones (Fig. 7a–c).<br />

4.2. Cu-Rautuvaara<br />

In <strong>the</strong> Cu-Rautuvaara deposit, <strong>the</strong> bulk of <strong>the</strong> Cu–Au<br />

ore is hosted by magnetite-dissem<strong>in</strong>ated, biotite- and/or<br />

anthophyllite-bear<strong>in</strong>g albitite which forms a ca. 45 m<br />

thick, elongated lens (Fig. 4). <strong>Iron</strong>stone lenses and cl<strong>in</strong>opyroxene–amphibole<br />

skarns similar to those at Laur<strong>in</strong>oja<br />

and Kuervitikko are also present, but <strong>the</strong>y are<br />

significantly smaller <strong>in</strong> size. As <strong>in</strong> Kuervitikko and<br />

Laur<strong>in</strong>oja, <strong>the</strong> hang<strong>in</strong>g wall consists of diorite and monzonite,<br />

and altered mafic metavolcanic rock comprises<br />

<strong>the</strong> immediate footwall. Chalcopyrite and pyrrhotite are<br />

<strong>the</strong> dom<strong>in</strong>ant sulphides, although pyrite is locally abundant.<br />

Magnetite is <strong>the</strong> dom<strong>in</strong>ant <strong>oxide</strong> m<strong>in</strong>eral, and small<br />

clusters of uran<strong>in</strong>ite have been detected <strong>in</strong> <strong>the</strong> cl<strong>in</strong>opyroxene–amphibole<br />

skarn (Fig. 7h). Albite, anthophyllite,<br />

biotite, and quartz are <strong>the</strong> ma<strong>in</strong> gangue m<strong>in</strong>erals.<br />

4.3. Ma<strong>in</strong> structural features of <strong>the</strong> <strong>deposits</strong><br />

Both Laur<strong>in</strong>oja and Kuervitikko are located <strong>in</strong> <strong>the</strong><br />

easternmost thrust zone of <strong>the</strong> <strong>Kolari</strong> Shear System<br />

next to <strong>the</strong> Äkäsjoki fault (Fig. 1). The strong foliation<br />

and l<strong>in</strong>eation developed <strong>in</strong>to rocks <strong>in</strong> <strong>the</strong> vic<strong>in</strong>ity<br />

of <strong>the</strong> thrust zone dur<strong>in</strong>g <strong>the</strong> D 3 stage which<br />

overpr<strong>in</strong>ts <strong>the</strong> structural features of <strong>the</strong> D 1 and D 2<br />

stages so that, for example at Hannuka<strong>in</strong>en and<br />

Kuervitikko, <strong>the</strong> older structural features cannot be<br />

recognised.<br />

The S 3 foliation and <strong>the</strong> L 3 l<strong>in</strong>eation are best developed<br />

<strong>in</strong> altered wall rocks and silicate-banded ironstones,<br />

and locally can be dist<strong>in</strong>guished <strong>in</strong> barren<br />

massive ironstones (Figs. 5 and 6). However, <strong>the</strong> S 3<br />

foliation is less developed <strong>in</strong> sulphide-bear<strong>in</strong>g ironstones<br />

which locally appear unfoliated. The S 3 foliation<br />

at Laur<strong>in</strong>oja and Kuervitikko dips at ca. 308 to <strong>the</strong> west<br />

and <strong>the</strong> L 3 l<strong>in</strong>eation plunges gently to <strong>the</strong> southwest. In<br />

both areas, <strong>the</strong> elongated lens-shaped ironstone units<br />

plunge gently to <strong>the</strong> SW follow<strong>in</strong>g <strong>the</strong> direction of <strong>the</strong><br />

L 3 l<strong>in</strong>eation (Figs. 3 and 4). The youngest structural<br />

features detected at Laur<strong>in</strong>oja and Kuervitikko are th<strong>in</strong><br />

mylonite seams and brittle fractures that crosscut <strong>the</strong> S 3<br />

foliation at a steep angle (Fig. 5a, b).<br />

The Cu-Rautuvaara deposit is structurally located<br />

<strong>in</strong> <strong>the</strong> NW limb of a major SW open<strong>in</strong>g syncl<strong>in</strong>e next<br />

to <strong>the</strong> same thrust zone <strong>in</strong> which Laur<strong>in</strong>oja and<br />

Kuervitikko are located (Figs. 1 and 4; Hiltunen,<br />

1982). Although <strong>the</strong> S 3 foliation and <strong>the</strong> L 3 l<strong>in</strong>eation


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 83<br />

Fig. 6. Photomicrographs of <strong>the</strong> wall rocks from Laur<strong>in</strong>oja, Kuervitikko and Cu-Rautuvaara. (a) Typical hang<strong>in</strong>g wall monzonite from DDH79-<br />

76.70. (b) Typical distal altered diorite from DDH75-33.60. Hornblende with m<strong>in</strong>or biotite comprises <strong>the</strong> dark m<strong>in</strong>erals and <strong>the</strong> pale areas consist of<br />

albite with m<strong>in</strong>or scapolite. (c) Albite–biotite–K-feldspar-altered diorite from DDH75-43.75. Biotite is <strong>the</strong> dom<strong>in</strong>ant mafic m<strong>in</strong>eral, and <strong>the</strong> pale<br />

areas consist of albite and K-feldspar. In addition, <strong>the</strong>re is a f<strong>in</strong>e-gra<strong>in</strong>ed magnetite dissem<strong>in</strong>ation <strong>in</strong> <strong>the</strong> rock (black spots). (d) Least-altered footwall<br />

metavolcanic rock from DDH79-217.69. Ma<strong>in</strong> m<strong>in</strong>erals are hornblende (dark) and plagioclase (pale). (e) Intensely biotite-altered metavolcanic rock<br />

from DDH79-205.90. Dark m<strong>in</strong>erals are biotite with small amounts of hornblende and act<strong>in</strong>olite. Pale areas consist of albite with traces of quartz. (f)<br />

Biotite–K-feldspar-altered metavolcanic rock from DDH79-190.85. Dark m<strong>in</strong>erals consist of biotite with traces of act<strong>in</strong>olite. Moderate magnetite<br />

dissem<strong>in</strong>ation (black spots). K-feldspar, albite and quartz comprise <strong>the</strong> pale m<strong>in</strong>erals. (g) K-feldspar–biotite–cl<strong>in</strong>opyroxene-altered footwall<br />

metavolcanic rock from DDH79-177.70. Pale m<strong>in</strong>erals consist of K-feldspar porphyroblasts and m<strong>in</strong>or amounts of quartz and albite. Biotite,<br />

act<strong>in</strong>olite and m<strong>in</strong>or amounts of cl<strong>in</strong>opyroxene and magnetite comprise <strong>the</strong> dark m<strong>in</strong>erals. (h) Magnetite-dissem<strong>in</strong>ated albitite from DDH109-<br />

217.25. (i) Scapolite skarn with dissem<strong>in</strong>ated magnetite, chalcopyrite, and pyrrhotite (dark) from DDH109-236.00. Younger biotite–pyrite ve<strong>in</strong>s<br />

crosscut <strong>the</strong> skarn.<br />

are well developed <strong>in</strong> <strong>the</strong> wall and host rocks of <strong>the</strong><br />

Cu-Rautuvaara deposit, <strong>the</strong>y are less pronounced than<br />

those at Laur<strong>in</strong>oja and Kuervitikko. The dom<strong>in</strong>ant<br />

direction of <strong>the</strong> ma<strong>in</strong> structural trend is SE and <strong>the</strong><br />

dip is steep (Fig. 4). As at Laur<strong>in</strong>oja and Kuervitikko,<br />

th<strong>in</strong> mylonite seams and brittle fractures crosscut <strong>the</strong><br />

ma<strong>in</strong> structures and are <strong>the</strong> youngest deformational<br />

feature.


84<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Fig. 7. Reflected light photomicrographs of <strong>the</strong> ores and skarns from <strong>Kolari</strong>. (a) Typical f<strong>in</strong>e-gra<strong>in</strong>ed magnetite (pale grey) ore with cl<strong>in</strong>opyroxene<br />

(dark grey) gangue, Laur<strong>in</strong>oja. (b, c) Typical chalcopyrite-bear<strong>in</strong>g magnetite ore from Laur<strong>in</strong>oja. Note <strong>the</strong> significantly coarser gra<strong>in</strong> size of <strong>the</strong><br />

magnetite compared to <strong>the</strong> magnetite without sulphides <strong>in</strong> (a), and <strong>the</strong> well-developed crystal form of <strong>the</strong> magnetite gra<strong>in</strong>s <strong>in</strong> (c). (d) Cu–Aum<strong>in</strong>eralised<br />

albitite from Cu-Rautuvaara. (e) Pyrite, magnetite, and chalcopyrite <strong>in</strong> cl<strong>in</strong>opyroxene skarn, Kuervitikko. (f) Chalcopyrite and<br />

magnetite <strong>in</strong> cl<strong>in</strong>opyroxene skarn from Laur<strong>in</strong>oja. Th<strong>in</strong> chalcopyrite ve<strong>in</strong>s or fracture <strong>in</strong>fills crosscut <strong>the</strong> cl<strong>in</strong>opyroxene and magnetite. (g)<br />

Molybdenite flakes <strong>in</strong> association with chalcopyrite and pyrite <strong>in</strong> cl<strong>in</strong>opyroxene skarn at Laur<strong>in</strong>oja. (h) Uran<strong>in</strong>ite and small pyrite gra<strong>in</strong>s <strong>in</strong><br />

cl<strong>in</strong>opyroxene, Cu-Rautuvaara. (i) A small <strong>gold</strong> gra<strong>in</strong> <strong>in</strong> chalcopyrite next to magnetite, Laur<strong>in</strong>oja. (j) Native <strong>gold</strong> <strong>in</strong> amphibole, Kuervitikko. (k)<br />

Native <strong>gold</strong> <strong>in</strong> coarse-gra<strong>in</strong>ed magnetite, Laur<strong>in</strong>oja; oil immersion. (l) Gold gra<strong>in</strong>s <strong>in</strong> a fracture <strong>in</strong> cl<strong>in</strong>opyroxene, Laur<strong>in</strong>oja; oil immersion.<br />

4.4. Size, grade, and distribution of <strong>copper</strong> and <strong>gold</strong><br />

The total <strong>in</strong>ferred amount of iron ore <strong>in</strong> <strong>the</strong> Kuervaara,<br />

Laur<strong>in</strong>oja, Vuopio, and Lauku ore bodies of <strong>the</strong><br />

Hannuka<strong>in</strong>en deposit is ca. 68 Mt (Hiltunen, 1982).<br />

Accord<strong>in</strong>g to Hiltunen (1982), <strong>the</strong> Laur<strong>in</strong>oja ore body<br />

conta<strong>in</strong>ed ca. 33 Mt of iron ore with Fe content vary<strong>in</strong>g<br />

between 36 and 53 wt.% and is <strong>the</strong> largest ore body at<br />

Hannuka<strong>in</strong>en (Fig. 1; Table 1). The Cu and Au contents<br />

at Laur<strong>in</strong>oja are poorly constra<strong>in</strong>ed, but Hiltunen (1982)<br />

estimates that <strong>the</strong> average concentration of <strong>the</strong> whole<br />

ironstone lens is 0.36 wt.% and 0.15 ppm for Cu and Au,<br />

respectively. Copper and <strong>gold</strong> are, however, not evenly<br />

distributed throughout <strong>the</strong> ironstone. The richest parts of<br />

<strong>the</strong> ore are located <strong>in</strong> <strong>the</strong> nor<strong>the</strong>astern part of <strong>the</strong> ironstone<br />

lens whereas <strong>the</strong> grades are lower <strong>in</strong> <strong>the</strong> marg<strong>in</strong>s<br />

and southwestern part (Hiltunen, 1982). For <strong>the</strong> 4.6 Mt<br />

of ore m<strong>in</strong>ed from Laur<strong>in</strong>oja, <strong>the</strong> grades were 43 wt.%<br />

Fe, 0.88 wt.% Cu, and 1 ppm Au (Puust<strong>in</strong>en, 2003).<br />

The distribution of Cu and Au <strong>in</strong> drill hole 170<br />

drilled across <strong>the</strong> Cu–Au rich part of <strong>the</strong> Laur<strong>in</strong>oja<br />

ore body shows that Cu–Au m<strong>in</strong>eralisation chiefly


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 85<br />

occurs with<strong>in</strong> <strong>the</strong> ironstone, and that <strong>the</strong>re is a positive<br />

correlation between Cu and Au (Fig. 8). Very little<br />

analytical data is available for rocks outside <strong>the</strong> ironstones,<br />

but <strong>the</strong> new data presented <strong>in</strong> this work show<br />

that elevated Cu and Au concentrations occur also <strong>in</strong><br />

<strong>the</strong> altered wall rocks and skarns (Table 2).<br />

Hiltunen (1982) estimated that <strong>the</strong> size and grade of<br />

<strong>the</strong> Kuervitikko deposit is 1.2 Mt with Fe grades similar<br />

Fig. 8. Distribution of Cu and Au <strong>in</strong> selected cross sections from Laur<strong>in</strong>oja, Kuervitikko, and Cu-Rautuvaara. See Figs. 3 and 4 for drill hole<br />

locations. Data after K<strong>in</strong>nunen (1980) and Ke<strong>in</strong>änen (1995).


86<br />

Table 2<br />

Chemical composition of <strong>the</strong> wall rocks, skarns, and ores from <strong>the</strong> Laur<strong>in</strong>oja, Cu-Rautuvaara, and Kuervitikko<br />

Sample i.d. 1 2A 2B 2C 3A 3B 3C 3D 4A 4B 4C 4D 5 6A 6B 6C 7A 7B 8A 8B<br />

DDH 79 75 75 79 78 170 75 170 170 78 129 109 109 109 127 109 75 170 170 127<br />

Depth 37.50 33.60 43.75 148.30 212.45 113.40 136.25 99.00 65.85 147.80 90.40 237.30 231.70 219.30 65.20 274.00 108.70 48.45 71.65 42.80<br />

Rock<br />

Monzonite Diorite Diorite Diorite m-volc m-volc m-volc m-volc Skarn Skarn Skarn Skarn sca skarn Albitite Albitite Albitite+<br />

Cu–Au<br />

<strong>Iron</strong>stone <strong>Iron</strong>stone <strong>Iron</strong>stone+<br />

Cu–Au<br />

XRF (wt.%)<br />

SiO 2 58.78 56.69 55.79 51.23 48.44 54.50 49.20 33.31 48.64 41.61 27.14 43.22 44.70 67.08 54.77 45.15 2.73 26.48 17.56 18.36<br />

TiO 2 0.87 0.30 0.29 0.23 1.05 0.89 0.86 0.42 0.05 0.06 0.29 0.07 0.38 0.63 0.79 0.72 0.03 0.10 0.04 0.21<br />

Al 2 O 3 16.31 20.69 20.74 17.31 13.32 13.80 12.89 6.68 0.92 2.43 4.32 7.28 17.19 15.40 13.75 12.87 0.57 1.71 1.14 4.36<br />

Fe 2 O 3 (t) 6.83 4.85 8.23 13.42 15.21 13.79 13.41 32.99 12.22 29.03 14.28 12.06 12.79 4.49 10.35 21.51 86.49 59.39 59.33 63.46<br />

MnO 0.10 0.05 0.05 0.05 0.21 0.12 0.17 0.12 0.29 0.24 0.43 0.52 0.11 0.05 0.10 0.03 0.46 0.09 0.12 0.11<br />

MgO 2.67 2.11 1.24 1.33 6.49 5.11 7.06 5.43 11.06 5.20 6.27 12.29 0.62 1.00 3.25 1.65 4.65 7.38 4.31 2.66<br />

CaO 4.57 5.80 3.43 5.26 9.35 3.36 4.07 10.47 20.83 16.52 25.85 16.17 8.75 1.98 4.38 0.69 3.64 3.53 5.22 2.13<br />

Na 2 O 5.12 7.05 6.54 6.66 3.08 5.95 1.54 1.80 0.68 1.34 0.50 1.00 6.08 6.22 5.52 7.36 b0.07 0.59 0.14 1.69<br />

K 2 O 3.88 1.12 3.03 0.70 0.66 1.95 6.17 0.45 0.01 0.05 0.39 0.95 0.69 1.63 2.59 0.06 b0.01 0.26 0.12 0.24<br />

P 2 O 5 0.36 0.24 0.24 0.20 0.09 0.09 0.06 0.13 0.03 0.21 0.12 0.10 0.08 0.19 0.24 0.05 0.05 0.07 0.10 0.13<br />

C 0.01 0.02 b0.01 0.03 0.03 0.01 0.02 0.03 0.01 0.03 3.40 0.06 0.38 0.08 0.07 0.04 0.95 0.02 b0.01 0.02<br />

Cl 0.08 0.10 0.05 0.05 0.13 0.05 0.16 0.07 0.03 0.02 0.10 0.37 1.69 0.11 0.02 0.01 0.02 0.02 0.01 0.03<br />

S 0.01 0.01 0.05 4.60 0.12 0.27 0.02 6.07 0.25 0.04 2.51 0.01 4.85 0.08 3.11 6.43 0.20 0.06 7.25 5.66<br />

Total 99.61 99.03 99.67 101.07 98.15 99.89 95.63 97.97 95.03 96.76 85.60 94.10 98.30 98.94 98.95 96.58 99.80 99.69 95.32 99.06<br />

XRF (ppm)<br />

As b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 147 b30 b30 b30<br />

Ba 924 405 1447 190 121 211 821 78 b20 25 34 78 156 1948 736 44 24 31 46 336<br />

Ce 115 45 62 42 b30 b30 50 68 b30 b30 54 55 51 365 81 b30 b30 854 b30 42<br />

Cr 31 b30 b30 b30 140 117 125 101 b30 b30 68 b30 b30 b30 189 100 30 b30 b30 54<br />

Cu 32 b20 28 955 140 59 b20 24,990 1744 29 5282 b20 2177 26 2133 51,360 995 44 65,130 27,910<br />

La 47 b30 b30 b30 b30 b30 b30 61 b30 b30 b30 b30 b30 185 32 b30 b30 729 b30 b30<br />

Mo b10 b10 b10 b10 b10 b10 b10 10 b10 b10 b10 b10 b10 b10 60 b10 b10 b10 b10 b10<br />

Nb 11 b10 b10 b10 b10 b10 b10 11 b10 b10 b10 46 50 b10 12 b10 b10 b10 b10 b10<br />

Ni b20 20 b20 194 85 70 79 64 21 67 50 b20 117 21 225 137 b20 137 66 b20<br />

<strong>Iron</strong>stone+<br />

Cu–Au<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105


Pb b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 b30 154 b30 b30 b30 b30 b30 b30 b30 41<br />

Rb 148 42 67 18 38 73 128 19 b10 b10 b10 b10 11 48 43 b10 b10 b10 b10 20<br />

Sr 707 667 354 394 113 79 59 191 14 26 83 16 143 378 221 54 18 b10 12 61<br />

Th 11 b10 b10 b10 b10 b10 b10 b10 b10 b10 b10 180 40 19 b10 b10 b10 b10 b10 b10<br />

U b10 b10 b10 b10 b10 b10 b10 b10 b10 b10 b10 330 20 b10 b10 b10 b10 b10 b10 b10<br />

V 141 59 51 39 349 233 250 92 b3 95 95 b3 33 60 545 138 b3 156 b3 61<br />

Y 20 b10 b10 b10 31 27 29 18 b10 21 15 66 48 12 35 17 b10 b10 b10 b10<br />

Zn 69 b20 b20 b20 63 b20 58 156 44 34 105 70 24 26 25 54 36 22 123 33<br />

Zr 268 38 41 31 54 131 66 77 11 18 59 181 379 523 206 33 b10 17 13 44<br />

TiO 2 /Al 2 O 3 10 4 53.3 14.5 14.0 13.3 78.8 64.5 66.7 62.9 54.3 24.7 67.1 9.62 22.1 40.9 57.5 55.9 52.6 58.5 35.1 48.2<br />

Zr/Al 2 O 3 16.4 1.84 1.98 1.81 4.05 9.49 5.12 11.5 12.0 7.41 13.7 24.9 22.0 34.0 15.0 2.56 9.94 11.4 10.1<br />

Zr/TiO 2 308 127 141 135 51.4 147 76.7 183 220 300 203 2590 997 830 261 45.8 170 325 210<br />

ICP-AES (ppm)<br />

Ag b1 b1 b1 b1 b1 b1 b1 3 1 b1 n.a. n.a. n.a. n.a. n.a. n.a. 1 b1 6 n.a.<br />

Co 16 6 13 418 22 47 43 660 6 19 n.a. n.a. n.a. n.a. n.a. n.a. 383 24 154 n.a.<br />

GFAAS (ppb)<br />

Au b0.5 b0.5 b0.5 41 5 b0.5 3 818 191 b0.5 70 1 84 2 44 387 16 2 1020 3780<br />

Bi 5 b2 b2 172 8 8 17 389 73 b2 125 91 475 67 221 b20 2460 b2 396 853<br />

Sb 29 b25 b25 82 44 31 b25 b25 b25 b25 n.a. n.a. n.a. b25 n.a. n.a. 72 b25 b25 n.a.<br />

Se n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. n.a. 1960 b50 4140 b50 4130 470 n.a. n.a. n.a. 1560<br />

Te b2 2 3 812 127 43 11 1510 132 4 830 3 1710 10 1000 970 123 12 4070 892<br />

(1) Typical hang<strong>in</strong>g wall monzonite, (2A) albite- and scapolite-altered diorite from upper distal alteration zone, (2B) biotite–K-feldspar-altered diorite from upper distal alteration zone, (2C)<br />

cl<strong>in</strong>opyroxene–magnetite-altered diorite from upper proximal alteration zone, (3A) least altered footwall metavolcanic rock, (3B) albite–biotite-altered mafic metavolcanic rock from lower distal<br />

alteration zone, (3C) K-feldspar–biotite-altered mafic metavolcanic rock from lower proximal alteration zone, (3D) cl<strong>in</strong>opyroxene–magnetite-altered and m<strong>in</strong>eralised mafic metavolcanic rock, lower<br />

proximal alteration zone, (4A) cl<strong>in</strong>opyroxene skarn, upper proximal alteration zone, (4B) magnetite-dissem<strong>in</strong>ated cl<strong>in</strong>opyroxene dom<strong>in</strong>ated skarn, lower proximal alteration zone, (4C) weakly<br />

m<strong>in</strong>eralised cl<strong>in</strong>opyroxene skarn, upper proximal alteration zone, (4D) cl<strong>in</strong>opyroxene–amphibole skarn, Cu-Rautuvaara, (5) cl<strong>in</strong>opyroxene-bear<strong>in</strong>g , magnetite-dissem<strong>in</strong>ated and weakly m<strong>in</strong>eralised<br />

scapolite skarn, Cu-Rautuvaara, (6A) albitite, Cu-Rautuvaara. (6B) Albitite, Kuervitikko. Upper proximal alteration zone. (6C) M<strong>in</strong>eralised albitite, Cu-Rautuvaara. (7A) Barren, allanite–monatzitebear<strong>in</strong>g<br />

ironstone, Laur<strong>in</strong>oja, (7B) barren ironstone, Laur<strong>in</strong>oja. (8A) M<strong>in</strong>eralised ironstone, Laur<strong>in</strong>oja, (8B) m<strong>in</strong>eralised ironstone, Kuervitikko. DDH=diamond drill hole. The depth refers to <strong>the</strong><br />

down-hole depth and is given as meters. n.a.=not analysed. Zr, TiO 2 , and Al 2 O 3 ratios calculated as Zr (ppm)/Al 2 O 3 (wt.%), and Zr (ppm)/TiO 2 (wt.%).<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 87


88<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

to those at Laur<strong>in</strong>oja. Typical concentrations <strong>in</strong> <strong>the</strong><br />

Kuervitikko deposit are 0.7 ppm and 0.5 wt.% for Au<br />

and Cu, respectively (Fig. 8; Ke<strong>in</strong>änen, 1995). Besides<br />

ironstone, <strong>the</strong> Cu–Au m<strong>in</strong>eralisation is hosted by a<br />

magnetite-dissem<strong>in</strong>ated albite–biotite-altered diorite<br />

that appears similar to <strong>the</strong> albitite that hosts <strong>the</strong> Cu–<br />

Au m<strong>in</strong>eralisation at Cu-Rautuvaara (Fig. 8). The spatial<br />

correlation between Cu and Au is stronger at Kuervitikko<br />

than at Laur<strong>in</strong>oja.<br />

There is no previously published estimate of <strong>the</strong> size<br />

and grade of <strong>the</strong> Cu-Rautuvaara. Based on drill<strong>in</strong>g <strong>the</strong><br />

lens shaped magnetite-dissem<strong>in</strong>ated albitite unit host<strong>in</strong>g<br />

<strong>the</strong> bulk of <strong>the</strong> Cu–Au m<strong>in</strong>eralisation (Fig. 8) is at least<br />

200 m <strong>in</strong> <strong>the</strong> longest dimension, up to 45 m thick, and<br />

is open at <strong>the</strong> depth of about 250 m (Fig. 4; Hiltunen,<br />

1982). Accord<strong>in</strong>g to our calculations, <strong>the</strong> size of <strong>the</strong><br />

weakly to moderately m<strong>in</strong>eralised albitite <strong>in</strong> <strong>the</strong> Cu-<br />

Rautuvaara deposit is 4.2 Mt. The Au concentration at<br />

Cu-Rautuvaara is generally 1 ppm or less, except for a<br />

few th<strong>in</strong>, V1 m, sections where it is up to 3 ppm.<br />

Copper concentrations up to 1.3 wt.% have been detected,<br />

but typically <strong>the</strong> Cu concentrations are 0.7 wt.%<br />

or less (Fig. 8; Ke<strong>in</strong>änen, 1995). The correlation between<br />

Cu and Au is similar to that at Kuervitikko.<br />

5. Alteration<br />

Metasomatic alteration styles related to <strong>the</strong> ironstones<br />

and Cu–Au m<strong>in</strong>eralisation consist of comb<strong>in</strong>ations of<br />

albite, amphiboles (act<strong>in</strong>olite–hornblende–anthophyllite),<br />

biotite, cl<strong>in</strong>opyroxenes (diopside–hedenbergite),<br />

K-feldspar, magnetite, scapolite (marialite), calcite, and<br />

sulphides (pyrite–pyrrhotite–chalcopyrite) alteration<br />

(Fig. 9). At <strong>the</strong> deposit scale, zon<strong>in</strong>g can be dist<strong>in</strong>guished<br />

around <strong>the</strong> Laur<strong>in</strong>oja and Kuervitikko ironstones whereas<br />

clear zon<strong>in</strong>g is difficult to dist<strong>in</strong>guish around <strong>the</strong> Cu-<br />

Rautuvaara deposit.<br />

At Laur<strong>in</strong>oja, where <strong>the</strong> zon<strong>in</strong>g is most dist<strong>in</strong>ct, <strong>the</strong><br />

alteration halo can be divided <strong>in</strong>to distal and proximal<br />

zones <strong>in</strong> both hang<strong>in</strong>g wall and footwall host rocks<br />

(Fig. 9). The alteration assemblages vary somewhat<br />

depend<strong>in</strong>g on <strong>the</strong> primary rock type, but <strong>the</strong> general<br />

pattern is that albite Fscapolite, biotite, and K-feldspar<br />

are <strong>the</strong> dom<strong>in</strong>ant alteration m<strong>in</strong>erals <strong>in</strong> <strong>the</strong> distal zones,<br />

and cl<strong>in</strong>opyroxenes, amphiboles, and magnetiteF<br />

calcite are <strong>the</strong> dom<strong>in</strong>ant m<strong>in</strong>erals <strong>in</strong> <strong>the</strong> proximal alteration<br />

zones. The change from distal to proximal alteration<br />

zones is def<strong>in</strong>ed by <strong>the</strong> appearance of<br />

cl<strong>in</strong>opyroxene and <strong>the</strong> outer boundary of <strong>the</strong> distal<br />

alteration zone is def<strong>in</strong>ed by <strong>the</strong> appearance of albite.<br />

Distal alteration zones display a m<strong>in</strong>eral variation with<br />

<strong>the</strong> amount of biotite and K-feldspar <strong>in</strong>creas<strong>in</strong>g toward<br />

Fig. 9. Schematic m<strong>in</strong>eral paragenetic diagrams across <strong>the</strong> Laur<strong>in</strong>oja,<br />

Kuervitikko, and Cu-Rautuvaara <strong>deposits</strong>. Thickness of <strong>the</strong> l<strong>in</strong>es<br />

corresponds to <strong>the</strong> abundance of <strong>the</strong> m<strong>in</strong>eral. The proximal and distal<br />

zones cannot be dist<strong>in</strong>guished at Cu-Rautuvaara due to poorly developed<br />

zonation.


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 89<br />

proximal alteration zones. Next to <strong>the</strong> proximal zone,<br />

<strong>the</strong> distal alteration zone is biotite- and K-feldspar-dom<strong>in</strong>ated.<br />

This feature, and <strong>the</strong> change from distal to<br />

proximal alteration, are best developed <strong>in</strong> footwall<br />

mafic metavolcanic rocks, where <strong>the</strong> m<strong>in</strong>eral assemblage<br />

changes from plagioclase–hornblende <strong>in</strong> <strong>the</strong><br />

least altered rock (Fig. 6d), via biotite–albite (Fig. 6e)<br />

and K-feldspar–biotite–albite (Fig. 6f) <strong>in</strong> <strong>the</strong> distal zone,<br />

to K-feldspar–biotite–amphibole–cl<strong>in</strong>opyroxene–magnetite<br />

<strong>in</strong> <strong>the</strong> proximal alteration zone (Fig. 6g). In <strong>the</strong><br />

hang<strong>in</strong>g wall distal alteration zone, <strong>the</strong> assemblages are<br />

similar, except that <strong>the</strong> albite alteration is more extensive<br />

and <strong>in</strong>tense, and scapolite is more abundant than <strong>in</strong> <strong>the</strong><br />

footwall.<br />

The dom<strong>in</strong>ant alteration m<strong>in</strong>eral <strong>in</strong> proximal zones<br />

is cl<strong>in</strong>opyroxene (Fig. 9). However, calcic amphibole<br />

commonly occurs toge<strong>the</strong>r or <strong>in</strong>stead of cl<strong>in</strong>opyroxene<br />

as <strong>the</strong> ma<strong>in</strong> alteration m<strong>in</strong>eral. Variable amounts<br />

of albite, biotite, K-feldspar, scapolite, calcite, and<br />

quartz occur <strong>in</strong> <strong>the</strong> proximal zones and locally <strong>the</strong>ir<br />

amount can exceed that of cl<strong>in</strong>opyroxene and amphibole.<br />

Besides abundant magnetite, <strong>the</strong> alteration<br />

assemblages <strong>in</strong> ironstone are similar to <strong>the</strong> proximal<br />

alteration zones (Fig. 9). Scapolite skarn units (Fig.<br />

6i), dm to several m <strong>in</strong> thickness, occur locally with<strong>in</strong><br />

<strong>the</strong> proximal alteration zones, and <strong>in</strong> <strong>the</strong> ironstones<br />

scapolite is locally <strong>the</strong> ma<strong>in</strong> gangue m<strong>in</strong>eral. Outside<br />

<strong>the</strong> ironstone, magnetite is most abundant <strong>in</strong> <strong>the</strong><br />

proximal alteration zones, but can locally account<br />

for up to 15 vol.% <strong>in</strong> <strong>the</strong> distal alteration zones. In<br />

<strong>the</strong> Laur<strong>in</strong>oja ore body, most of <strong>the</strong> sulphides occur<br />

<strong>in</strong> <strong>the</strong> ironstone and <strong>in</strong> <strong>the</strong> proximal alteration zones,<br />

but both Fe- and Cu-sulphides are locally abundant (1<br />

to 10 vol.%) <strong>in</strong> all altered rocks <strong>in</strong>dependent of <strong>the</strong><br />

alteration assemblages.<br />

In <strong>the</strong> Kuervitikko deposit, <strong>the</strong> alteration assemblages<br />

def<strong>in</strong>e a similar zon<strong>in</strong>g as described for Laur<strong>in</strong>oja<br />

(Fig. 9). The most significant differences are that<br />

at Kuervitikko, K-feldspar and biotite are less common<br />

<strong>in</strong> <strong>the</strong> hang<strong>in</strong>g wall distal alteration zone, and <strong>the</strong>re is<br />

an albitite unit between <strong>the</strong> diorite and cl<strong>in</strong>opyroxene–<br />

amphibole skarn. In addition, amphibole is a slightly<br />

more abundant proximal alteration m<strong>in</strong>eral than at<br />

Laur<strong>in</strong>oja. Also, compared to Laur<strong>in</strong>oja, sulphides appear<br />

to be more abundant <strong>in</strong> <strong>the</strong> distal alteration zones,<br />

especially <strong>in</strong> <strong>the</strong> albitite (Fig. 9).<br />

At Cu-Rautuvaara, <strong>the</strong> alteration zon<strong>in</strong>g is poorly<br />

developed, although <strong>the</strong> alteration assemblages are similar<br />

to Laur<strong>in</strong>oja and Kuervitikko (Fig. 9). A th<strong>in</strong><br />

sequence of cl<strong>in</strong>opyroxene skarns with alteration assemblage<br />

similar to <strong>the</strong> proximal alteration zones at<br />

Laur<strong>in</strong>oja and Kuervitikko occurs <strong>in</strong> <strong>the</strong> hang<strong>in</strong>g wall.<br />

However, as stated above, <strong>the</strong> ma<strong>in</strong> Cu–Au m<strong>in</strong>eralisation<br />

is with<strong>in</strong> a magnetite-dissem<strong>in</strong>ated anthophylliteand<br />

biotite-bear<strong>in</strong>g albitite (Figs. 6h and 9).<br />

6. Geochemistry<br />

6.1. Monzonite<br />

The chemical composition of <strong>the</strong> monzonite <strong>in</strong>dicates<br />

that it is metalum<strong>in</strong>ous, low <strong>in</strong> Nb and Y, and<br />

shows a calc-alkal<strong>in</strong>e aff<strong>in</strong>ity (Table 2). These features<br />

Fig. 10. Chondrite-normalised REE patterns for <strong>the</strong> samples presented<br />

<strong>in</strong> Table 3. Normalis<strong>in</strong>g values for C1 chondrite after McDonough<br />

and Sun (1995).


90<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

are typical for <strong>the</strong> 1900 to 1860 Ma Haparanda Suite<br />

<strong>in</strong>trusions abundant throughout northwestern F<strong>in</strong>land<br />

and nor<strong>the</strong>rn Sweden (e.g., Mellquist et al., 2003).<br />

The chondrite-normalised REE pattern of <strong>the</strong> monzonite<br />

shows enrichment <strong>in</strong> LREE <strong>in</strong> respect to HREE and<br />

a weak negative Eu anomaly (Fig. 10a).<br />

6.2. Diorite and mafic metavolcanic rocks<br />

Due to <strong>the</strong> highly altered nature of <strong>the</strong> wall and<br />

host rocks of <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong>, <strong>the</strong> primary classification<br />

of <strong>the</strong> variably altered rocks are dist<strong>in</strong>guished<br />

us<strong>in</strong>g TiO 2 –Al 2 O 3 , Zr–Al 2 O 3 , and Zr–TiO 2 ratios;<br />

elements that usually display immobile behaviour <strong>in</strong><br />

most hydro<strong>the</strong>rmal systems. The variably altered diorites<br />

display remarkably constant TiO 2 –Al 2 O 3 , Zr–<br />

Al 2 O 3 , and Zr–TiO 2 ratios (Fig. 11). The chondritenormalised<br />

REE pattern of <strong>the</strong> diorite is identical to<br />

monzonite, except for <strong>the</strong> slightly positive Eu anomaly<br />

and <strong>the</strong> lower general REE content (Fig. 10a). This,<br />

comb<strong>in</strong>ed with <strong>the</strong> age data (see below), suggests that<br />

<strong>the</strong> monzonite and diorite were derived from <strong>the</strong> same<br />

magma source.<br />

The metavolcanic rocks form two groups based on<br />

<strong>the</strong>ir TiO 2 –Al 2 O 3 , Zr–Al 2 O 3 , and Zr–TiO 2 ratios (Fig.<br />

11). One group has a similar Zr–TiO 2 ratio to least<br />

altered mafic metavolcanic rock (average Zr/<br />

TiO 2 =65F15 ppm/wt.%, n =6), while <strong>the</strong> o<strong>the</strong>r<br />

group shows a considerably higher (average Zr/<br />

TiO 2 =174F50 ppm/wt.%, n = 6) ratio (Fig. 11). This<br />

<strong>in</strong>dicates a primary geochemical variation. Based on <strong>the</strong><br />

difference <strong>in</strong> Zr/TiO 2 ratios, <strong>the</strong> metavolcanic rocks are<br />

divided <strong>in</strong>to type-1 (lower Zr–TiO 2 ) and type-2 (higher<br />

Zr–TiO 2 ). The Zr–TiO 2 ratio of <strong>the</strong> type-1 is typical for<br />

basalts while <strong>the</strong> ratio of <strong>the</strong> type-2 rocks straddles <strong>the</strong><br />

border of basalts and andesites <strong>in</strong> <strong>the</strong> classification<br />

diagram of W<strong>in</strong>chester and Floyd (1977). Both types<br />

are present at Laur<strong>in</strong>oja and Kuervitikko whereas at Cu-<br />

Rautuvaara, only type-1 has been detected. However,<br />

<strong>the</strong> data from Cu-Rautuvaara are too limited to rule out<br />

<strong>the</strong> possibility that both types of mafic metavolcanic<br />

rocks occur <strong>the</strong>re, too.<br />

At Laur<strong>in</strong>oja and Kuervitikko, <strong>the</strong> chemical composition<br />

of diorite and metavolcanic rocks displays a<br />

zon<strong>in</strong>g that corresponds to zon<strong>in</strong>g <strong>in</strong> alteration assemblages.<br />

The rocks <strong>in</strong> <strong>the</strong> distal alteration zones are rich<br />

<strong>in</strong> Na 2 OFK 2 O, Cl, Ba, and <strong>the</strong> proximally altered<br />

rocks are enriched <strong>in</strong> CaO FFe 2 O 3t , S, Ag, Au, Bi, C,<br />

Ce, Co, Cu, La, and Te compared to distal altered rocks<br />

(Table 2; Fig. 9). For <strong>the</strong> evaluation of <strong>the</strong> effect of<br />

alteration on <strong>the</strong> chemical composition of rocks, major<br />

element <strong>oxide</strong>s and selected m<strong>in</strong>or elements are plotted<br />

Fig. 11. Immobile element ratios <strong>in</strong> variably altered diorite, metavolcanic rocks, skarns, and ironstones from Laur<strong>in</strong>oja, Kuervitikko, and Cu-<br />

Rautuvaara. C8=slope for <strong>the</strong> l<strong>in</strong>es def<strong>in</strong>ed by average TiO 2 –Al 2 O 3 , Zr–Al 2 O 3 , and Zr–TiO 2 ratios for diorite and type-1 and -2 metavolcanic rocks.


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 91<br />

aga<strong>in</strong>st <strong>the</strong> immobile Al 2 O 3 (Fig. 12). The variation <strong>in</strong><br />

mobile elements shows apparent loss or ga<strong>in</strong> of elements<br />

dur<strong>in</strong>g <strong>the</strong> alteration. Correlation of mobile element<br />

with immobile Al 2 O 3 suggests that <strong>the</strong> change of<br />

<strong>the</strong> former was coupled with a net mass change of <strong>the</strong><br />

rock dur<strong>in</strong>g alteration.<br />

Fig. 12. Major element <strong>oxide</strong> variation aga<strong>in</strong>st <strong>the</strong> immobile Al 2 O 3 <strong>in</strong> altered diorite, metavolcanic rock, cl<strong>in</strong>opyroxene–amphibole-dom<strong>in</strong>ated<br />

skarns, and ironstones from <strong>the</strong> Laur<strong>in</strong>oja, Kuervitikko, and Cu-Rautuvaara <strong>deposits</strong> with <strong>in</strong>terpreted alteration trends.


92<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Both <strong>the</strong> distal altered diorite and metavolcanic<br />

rocks show considerable variation <strong>in</strong> K 2 O and<br />

Na 2 O concentrations reflect<strong>in</strong>g <strong>the</strong>ir proportions of<br />

albite, biotite, and K-feldspar. Barium displays very<br />

high (up to 1500 ppm; Table 2) local values <strong>in</strong> distal<br />

altered rocks. S<strong>in</strong>ce no barite has been detected, it is<br />

very likely that Ba is bound with<strong>in</strong> biotite and/or K-<br />

feldspar <strong>in</strong> <strong>the</strong>se rocks. The concentration of Si 2 O,<br />

Fe 2 O 3t , CaO and MgO show negative correlation<br />

with Al 2 O 3 <strong>in</strong> <strong>the</strong> proximal altered rocks <strong>in</strong>dicat<strong>in</strong>g<br />

that proximal alteration was accompanied with significant<br />

net mass ga<strong>in</strong> (Fig. 12). The <strong>in</strong>crease <strong>in</strong> <strong>the</strong><br />

CaO, Fe 2 O 3t , and MgO reflects <strong>the</strong> <strong>in</strong>crease <strong>in</strong><br />

amount of cl<strong>in</strong>opyroxene, magnetite, and amphiboles<br />

<strong>in</strong> proximal altered rocks compared to distal altered<br />

rocks. In addition, high CaO concentrations coupled<br />

with elevated C concentrations <strong>in</strong> <strong>the</strong> proximal altered<br />

rocks reflect <strong>the</strong> presence of calcite. Sulphur<br />

shows elevated (N1.5 wt.%) values only <strong>in</strong> proximal<br />

altered diorites and metavolcanic rocks (Fig. 12). The<br />

elevated sulphur concentrations display positive correlation<br />

with concentrations of Ag, Au, Bi, Co, Cu,<br />

and Te (Table 2).<br />

6.3. Cl<strong>in</strong>opyroxene–amphibole skarns and ironstones<br />

The skarns and ironstones, toge<strong>the</strong>r with albitites,<br />

are <strong>the</strong> most <strong>in</strong>tensely altered rocks <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong>.<br />

The TiO 2 –Al 2 O 3 , Zr–Al 2 O 3 , and Zr–TiO 2 ratios <strong>in</strong><br />

<strong>the</strong> cl<strong>in</strong>opyroxene skarns and ironstones are similar for<br />

both rock types suggest<strong>in</strong>g common orig<strong>in</strong>s for <strong>the</strong>se<br />

rocks (Fig. 11). However, <strong>the</strong>re is some variation <strong>in</strong><br />

<strong>the</strong> immobile element ratios, imply<strong>in</strong>g that <strong>the</strong> protolith<br />

for <strong>the</strong>se rocks was at least locally heterogeneous<br />

and/or consisted of several rock types. Never<strong>the</strong>less,<br />

<strong>the</strong> ratios <strong>in</strong> <strong>the</strong> skarns and ironstones show considerable<br />

similarities with o<strong>the</strong>r rock types, especially with<br />

type-2 metavolcanic rocks, and a cont<strong>in</strong>uum from<br />

altered metavolcanic rocks and diorites to skarns and<br />

ironstones can be envisioned <strong>in</strong> a number of <strong>the</strong> major<br />

element <strong>oxide</strong> bivariant plots (Figs. 11 and 12). The<br />

elements that display high concentrations <strong>in</strong> proximal<br />

altered diorites and metavolcanic rocks also have high<br />

concentrations <strong>in</strong> <strong>the</strong> skarns and ironstones (Fig. 12;<br />

Table 2). The ma<strong>in</strong> difference between <strong>the</strong> chemical<br />

composition <strong>in</strong> <strong>the</strong> skarns and ironstones is <strong>the</strong> drastically<br />

higher Fe 2 O 3t concentrations <strong>in</strong> <strong>the</strong> latter (Fig.<br />

12). In addition, concentration of CaO is significantly<br />

higher <strong>in</strong> skarns than ironstones, yet this can be<br />

expla<strong>in</strong>ed with <strong>the</strong> higher amount of cl<strong>in</strong>opyroxene<br />

and calcic amphibole <strong>in</strong> <strong>the</strong> former compared to <strong>the</strong><br />

latter. The very high (N20 wt.%) CaO concentrations<br />

<strong>in</strong> <strong>the</strong> skarns are coupled with elevated C values and<br />

this is due to significant amount of calcite <strong>in</strong> <strong>the</strong> rock<br />

(Fig. 12; Table 2). The behaviour of Ag, Au, Bi, Co,<br />

Cu, S, and Te <strong>in</strong> <strong>the</strong> skarns and ironstones is similar to<br />

<strong>the</strong> proximal altered diorites and metavolcanic rocks,<br />

but <strong>the</strong> concentration of <strong>the</strong>se elements is typically<br />

higher <strong>in</strong> <strong>the</strong> former rock types <strong>in</strong>dicat<strong>in</strong>g higher<br />

sulphide contents (Table 2). Concentrations of As,<br />

Pb, and Zn are typically low <strong>in</strong> all rocks. The 180<br />

ppm Th and 330 ppm U concentration <strong>in</strong> <strong>the</strong> uran<strong>in</strong>ite-bear<strong>in</strong>g<br />

cl<strong>in</strong>opyroxene–amphibole skarn suggest<br />

that <strong>the</strong>se elements have also been mobile dur<strong>in</strong>g <strong>the</strong><br />

skarn alteration.<br />

The P REE <strong>in</strong> <strong>the</strong> skarns and ironstones is typically<br />

low, although locally <strong>the</strong> ironstones display considerably<br />

high REE concentrations (Table 3). The ma<strong>in</strong><br />

REE-bear<strong>in</strong>g m<strong>in</strong>erals <strong>in</strong> <strong>the</strong>se ironstones are allanite<br />

and monazite. The chondrite-normalised REE patterns<br />

for both skarns and ironstones are relatively flat with<br />

only slight enrichment <strong>in</strong> LREE <strong>in</strong> respect to HREE<br />

(Fig. 10b, c). However, <strong>the</strong> allanite- and monazite-rich<br />

ironstone displays a strong LREE enrichment (Fig.<br />

10c). This suggests that at least LREE, and possibly<br />

also some of <strong>the</strong> HREE, have been mobile dur<strong>in</strong>g <strong>the</strong><br />

alteration, and fur<strong>the</strong>r expla<strong>in</strong>s <strong>the</strong> locally elevated La<br />

and Ce values <strong>in</strong> all altered rocks (Table 2). Never<strong>the</strong>less,<br />

<strong>the</strong> REE patterns of <strong>the</strong> skarns and ironstones<br />

are relatively similar to distal altered type-1 metavolcanic<br />

rock (Fig. 10). This fur<strong>the</strong>r supports <strong>the</strong> idea that<br />

<strong>the</strong> metavolcanic rocks are <strong>the</strong> precursors for <strong>the</strong><br />

skarns and ironstones.<br />

6.4. Albitites and scapolite skarns<br />

Geochemically <strong>the</strong> most peculiar rock types at <strong>Kolari</strong><br />

are <strong>the</strong> scapolite skarns and albitites. Typical for both<br />

rock types is <strong>the</strong> high Na 2 O concentration (up to 8.6<br />

wt.%) and <strong>the</strong> highly variable concentrations <strong>in</strong> Al 2 O 3 ,<br />

TiO 2 ,Zr(Table 2). These rocks can be found throughout<br />

<strong>the</strong> ore-bear<strong>in</strong>g sequence around all <strong>the</strong> <strong>deposits</strong>. Albitites<br />

are most commonly found <strong>in</strong> <strong>the</strong> hang<strong>in</strong>g and<br />

footwalls whereas <strong>the</strong> scapolite skarns are typically<br />

located <strong>in</strong> <strong>the</strong> proximal alteration zones. Thus, it is<br />

likely that <strong>the</strong> protolith for <strong>the</strong>se rocks <strong>in</strong>cludes several<br />

rock types. In places, especially at Cu-Rautuvaara, <strong>the</strong>se<br />

rocks are m<strong>in</strong>eralised with a metal association that is<br />

similar to m<strong>in</strong>eralised ironstones and skarns (Table 2).<br />

Like <strong>the</strong> o<strong>the</strong>r altered rocks, <strong>the</strong> albitites and scapolite<br />

skarns conta<strong>in</strong> locally elevated concentrations of Ba<br />

(1950 ppm), Ce (365 ppm), and La (185 ppm). Generally<br />

<strong>the</strong> concentrations of Cr, Mo, Nb, Ni, Th, U, and V<br />

are <strong>in</strong> places higher <strong>in</strong> <strong>the</strong>se rocks than anywhere <strong>in</strong>


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 93<br />

Table 3<br />

Rare earth element composition of unaltered and altered rocks and ironstones<br />

Sample i.d. 1 2A 3E 4E 4F 6C 7A 7B 7C 7D<br />

DDH 79 75 75 75 75 109 75 170 170 130<br />

Depth (m) 37.50 33.60 144.15 53.80 71.50 274.00 108.70 48.45 107.15 107.40<br />

La 37.0 9.00 7.32 4.32 2.68 10.0 2.12 698 6.50 1.93<br />

Ce 75.1 18.5 17.0 12.0 5.25 16.8 4.39 841 6.16 2.92<br />

Pr 8.36 2.29 2.23 1.91 0.77 1.74 0.57 58.4 0.63 0.32<br />

Nd 33.4 9.43 10.7 8.95 3.94 6.92 2.26 145 2.49 1.57<br />

Sm 5.35 1.77 2.79 2.14 1.14 1.20 0.52 7.75 0.69 0.45<br />

Eu 1.32 0.66 0.94 0.66 0.53 0.39 0.10 1.83 0.47 0.18<br />

Gd 5.16 1.44 3.77 1.82 0.96 1.67 0.69 7.71 1.13 0.41<br />

Tb 0.71 0.20 0.61 0.32 0.16 0.26 0.10 0.86 0.22 b.d.<br />

Dy 3.28 1.04 3.35 1.87 0.98 1.76 0.44 1.25 1.24 0.23<br />

Ho 0.6 0.15 0.74 0.34 0.17 0.38 b.d. 0.18 0.25 b.d.<br />

Er 1.70 0.50 2.33 0.97 0.56 0.98 0.30 0.45 0.72 b.d.<br />

Yb 1.69 0.44 1.85 1.17 1.00 1.15 0.21 0.54 0.68 b.d.<br />

Lu<br />

P<br />

0.23 b.d. 0.29 0.16 0.19 0.16 b.d. 0.12 0.11 b.d.<br />

REE 173.90 45.42 53.92 36.63 18.33 43.41 11.70 1763.09 21.29 8.01<br />

Sample number<strong>in</strong>g as <strong>in</strong> Table 2. 1=unaltered monzonite, Laur<strong>in</strong>oja, 2A=albitised diorite, Laur<strong>in</strong>oja, 3E=weakly albitised mafic metavolcanic<br />

rock, Laur<strong>in</strong>oja, 4E–F=cl<strong>in</strong>opyroxene skarn from upper proximal alteration zone, Laur<strong>in</strong>oja, 6C=Cu–Au m<strong>in</strong>eralised albitite, Cu-Rautuvaara,<br />

7A=weakly m<strong>in</strong>eralised ironstone, Laur<strong>in</strong>oja, 7B=barren, allanite and monazite bear<strong>in</strong>g ironstone, Laur<strong>in</strong>oja, 7C=Cu–Au m<strong>in</strong>eralised ironstone,<br />

Laur<strong>in</strong>oja, 7D=barren ironstone, Kuervitikko. All values given <strong>in</strong> ppm. b.d.=below detection limit. DDH=diamond drill hole. Depth as down-hole<br />

depth (meters).<br />

o<strong>the</strong>r altered rock types (Table 2). The chondrite-normalised<br />

REE pattern of <strong>the</strong> Cu–Au-m<strong>in</strong>eralised, magnetite-dissem<strong>in</strong>ated<br />

albitite at Cu-Rautuvaara shows<br />

LREE enrichment relative to HREE, but <strong>the</strong> pattern<br />

does not seem to correlate to any o<strong>the</strong>r rock type<br />

(Fig. 10).<br />

6.5. Mass balance calculations<br />

There probably are several protoliths for skarns and<br />

ironstones at <strong>Kolari</strong>, but <strong>the</strong> immobile-element ratios<br />

suggest that <strong>the</strong> dom<strong>in</strong>ant protolith is <strong>the</strong> type-2 metavolcanic<br />

rock (Fig. 11). It is evident that significant net<br />

mass changes did occur dur<strong>in</strong>g <strong>the</strong> proximal alteration<br />

(Figs. 11 and 12). This h<strong>in</strong>ders <strong>the</strong> estimation of which<br />

rocks were protoliths to which parts of skarn and<br />

ironstone units. However, <strong>the</strong> effect of <strong>the</strong> net mass<br />

changes is elim<strong>in</strong>ated by plott<strong>in</strong>g <strong>the</strong> data presented <strong>in</strong><br />

Figs. 11 and 12 <strong>in</strong>to <strong>the</strong> TiO 2 /Al 2 O 3 vs. Zr/Al 2 O 3<br />

b<strong>in</strong>ary plot (Fig. 13). For comparison, <strong>the</strong> albitites,<br />

scapolite skarn and monzonite presented <strong>in</strong> Table 2<br />

are also plotted on <strong>the</strong> diagram. In <strong>the</strong> ratio to ratio<br />

plot <strong>the</strong> rocks form dist<strong>in</strong>ct groups of <strong>the</strong>ir own. Variably<br />

altered diorites form group I, type-1 metavolcanic<br />

rocks, group II, and type-2 metavolcanic rocks, group<br />

III (Fig. 13). In addition, three skarn, and three ironstone<br />

samples plot near to each o<strong>the</strong>r, form<strong>in</strong>g a dist<strong>in</strong>ct<br />

group (IV). The majority of skarn and ironstone sam-<br />

Fig. 13. TiO 2 –Al 2 O 3 vs. Zr–Al 2 O 3 plot of <strong>the</strong> samples shown <strong>in</strong> Figs. 11 and 12. Monzonite, albitites, and scapolite skarn presented <strong>in</strong> Table 2<br />

added for comparison. Encircled areas def<strong>in</strong>ed as follows: (I) distal and proximal altered diorites, (II) distal and proximal altered type-1<br />

metavolcanic rocks, (III) distal and proximal altered type-2 metavolcanic rocks and overlapp<strong>in</strong>g skarns and ironstones, (IV) skarn and ironstone<br />

samples that appear to correlate with each o<strong>the</strong>r but not with any o<strong>the</strong>r rock types.


94<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

ples overlap <strong>the</strong> area def<strong>in</strong>ed by group III <strong>in</strong>dicat<strong>in</strong>g<br />

that <strong>the</strong> protolith for <strong>the</strong>se is <strong>the</strong> type-2 metavolcanic<br />

rock. The skarn and ironstone samples form<strong>in</strong>g group<br />

IV as well as o<strong>the</strong>r samples that scatter around <strong>the</strong><br />

diagram do not seem to correlate with any known<br />

rock type, and thus <strong>the</strong> protolith for <strong>the</strong>se rema<strong>in</strong>s<br />

unknown. Possible candidates are <strong>the</strong> sedimentary<br />

rocks <strong>in</strong> <strong>the</strong> sequence (Figs. 3 and 4).<br />

All samples with<strong>in</strong> <strong>the</strong> group III (Fig. 13) were<br />

selected for <strong>the</strong> mass balance calculations. Due to<br />

heterogeneity of <strong>the</strong> immobile element ratios <strong>in</strong> <strong>the</strong><br />

type-2 metavolcanic rocks average concentrations<br />

were used for calculations <strong>in</strong>stead of <strong>in</strong>dividual samples.<br />

The average concentrations were calculated for<br />

distal and proximal altered type-2 metavolcanic rocks,<br />

skarns, and ironstones represent<strong>in</strong>g moderate proximal<br />

alteration, <strong>in</strong>tense proximal alteration, and <strong>in</strong>tense Fealteration,<br />

respectively (Table 4). The mass balance<br />

calculations were carried out us<strong>in</strong>g isocon method described<br />

by Grant (1986). S<strong>in</strong>ce <strong>the</strong>re is no unaltered<br />

type-2 metavolcanic rock available, <strong>the</strong> calculated<br />

averages of <strong>the</strong> distal altered varieties were used as<br />

reference samples for <strong>the</strong> calculations <strong>in</strong>stead of unaltered<br />

rocks. Alum<strong>in</strong>ium, titanium and zirconium were<br />

used as immobile elements <strong>in</strong> <strong>the</strong> calculations. Calculations<br />

are based on constant volume assumption that is<br />

probably not true for <strong>the</strong> most <strong>in</strong>tense alteration. Never<strong>the</strong>less,<br />

<strong>the</strong> results (Fig. 14) are <strong>in</strong> accordance with<br />

<strong>the</strong> changes <strong>in</strong> m<strong>in</strong>eral assemblage, and we thus feel<br />

that <strong>the</strong>y reflect <strong>the</strong> true chemical changes that occurred<br />

dur<strong>in</strong>g <strong>the</strong> alteration.<br />

The isocons <strong>in</strong>dicate constant and considerable net<br />

mass ga<strong>in</strong>s between both proximally altered samples<br />

(moderately and <strong>in</strong>tensely altered) and <strong>in</strong>tensely Fealtered<br />

samples (Fig. 14). The Zr, TiO 2 , and Al 2 O 3<br />

composition plot very close to <strong>the</strong> isocon <strong>in</strong> all diagrams<br />

<strong>in</strong>dicat<strong>in</strong>g that <strong>the</strong>y were conserved dur<strong>in</strong>g <strong>the</strong><br />

alteration. In addition, SiO 2 , MgO, MnO, and Cr<br />

appear to have been conserved, for <strong>the</strong> most part, <strong>in</strong><br />

moderate proximal alteration. All o<strong>the</strong>r major elements,<br />

except alkalis, show ga<strong>in</strong>s <strong>in</strong> all three groups.<br />

Of <strong>the</strong>se <strong>the</strong> ga<strong>in</strong>s <strong>in</strong> Fe 2 O 3t , CaO, and S are highest<br />

<strong>in</strong> all three groups, and MgO display significant ga<strong>in</strong><br />

<strong>in</strong> <strong>in</strong>tensely altered groups. Of <strong>the</strong> trace elements <strong>the</strong><br />

highest ga<strong>in</strong>s <strong>in</strong> all groups are <strong>in</strong> light rare earth<br />

elements and chalcophile elements. Only V shows<br />

constant depletion <strong>in</strong> all groups. S<strong>in</strong>ce <strong>the</strong> isocons<br />

were constructed us<strong>in</strong>g distal altered rocks as reference<br />

samples (i.e., Na–K altered rocks) <strong>in</strong>stead of<br />

truly unaltered rocks, it is possible that some Na 2 O<br />

and/or K 2 O was also added <strong>in</strong>to <strong>the</strong> skarns and ironstones<br />

dur<strong>in</strong>g <strong>the</strong> alteration event though <strong>the</strong> isocon<br />

Table 4<br />

Average chemical composition values <strong>in</strong> <strong>the</strong> rocks used <strong>in</strong> mass<br />

balance calculations<br />

A B C D<br />

n 3 3 5 7<br />

SiO 2 54.75 42.51 32.32 23.58<br />

TiO 2 0.83 0.56 0.27 0.18<br />

Al 2 O 3 14.19 10.09 4.31 3.28<br />

Fe 2 O 3 (t) 11.97 22.22 12.15 54.76<br />

MnO 0.08 0.10 0.43 0.20<br />

MgO 5.00 4.24 7.48 5.01<br />

CaO 2.59 8.47 25.34 6.67<br />

Na 2 O 4.53 4.10 0.46 0.97<br />

K 2 O 3.90 0.61 1.21 0.37<br />

P 2 O 5 0.11 0.12 0.17 0.10<br />

C 0.01 0.02 2.74 0.08<br />

Cl 0.08 0.03 0.06 0.05<br />

S 0.95 5.95 1.86 3.71<br />

(ppm)<br />

Ba 390 176 182 90<br />

Ce 39 68 49 164<br />

Cr 130 88 55 46<br />

Cu 122 10,584 1979 20,550<br />

La 22 30 27 152<br />

Mo 29 57 9 44<br />

Ni 119 122 63 80<br />

Sr 63 153 67 45<br />

V 347 110 139 66<br />

Zn 22 62 47 91<br />

Zr 134 100 47 36<br />

(ppb)<br />

Au 7 303 61 849<br />

Bi 74 279 105 333<br />

Te 221 1803 519 828<br />

TiO 2 /Al 2 O 3 0.0584 0.0550 0.0630 0.0542<br />

Zr/Al 2 O 3 * 9.46 9.94 10.96 10.89<br />

Zr/TiO 2 * 162 181 174 201<br />

A=distal altered type 2 metavolcanic rocks, B=proximal altered type<br />

2 metavolcanic rocks, C=skarns, D=ironstones, barren and Cu–Au<br />

m<strong>in</strong>eralised. n =number of samples. *(ppm/wt.%). 0.5detection<br />

limit is used for values below detection limits <strong>in</strong> calculations. See<br />

Table 2 for <strong>the</strong> detection limits.<br />

diagrams suggest depletion of <strong>the</strong>se elements <strong>in</strong> some<br />

of <strong>the</strong> groups.<br />

No suitable sample pairs were found for mass balance<br />

calculations <strong>in</strong> distal alteration. However, s<strong>in</strong>ce <strong>the</strong><br />

distal alteration (albite–scapolite–biotite–K-feldspar alteration)<br />

caused no significant mass changes (Figs. 11<br />

and 12), <strong>the</strong> change <strong>in</strong> chemistry between least altered<br />

type-1 metavolcanic rocks and distal altered varieties of<br />

<strong>the</strong>m reflects <strong>the</strong> absolute ga<strong>in</strong>s or losses that occurred<br />

dur<strong>in</strong>g <strong>the</strong> alteration (Table 2). Accord<strong>in</strong>g to this, <strong>the</strong><br />

ma<strong>in</strong> ga<strong>in</strong>s <strong>in</strong> <strong>the</strong> distal altered rocks are <strong>in</strong> Na 2 O, and<br />

K 2 O accompanied with loss <strong>in</strong> CaO. The Ba concentrations<br />

<strong>in</strong> biotite–K-feldspar altered rocks are typically


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 95<br />

Fig. 14. Isocon diagrams for type-2 metavolcanic rocks show<strong>in</strong>g elemental changes dur<strong>in</strong>g alteration. Average values of <strong>the</strong> groups plott<strong>in</strong>g with<strong>in</strong><br />

field III <strong>in</strong> Fig. 13 are used. The calculated averages for all three groups and reference samples are shown <strong>in</strong> Table 4. Major <strong>oxide</strong>s, C, Cl, and S<br />

plotted <strong>in</strong> wt.%, trace elements <strong>in</strong> ppm. C8=slope of <strong>the</strong> isocon, R 2 =correlation coefficient between Zr, TiO 2 , and Al 2 O 3 .


96<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

high, thus it is evident that <strong>the</strong> distal altered rocks also<br />

ga<strong>in</strong>ed Ba dur<strong>in</strong>g <strong>the</strong> alteration.<br />

7. Fluid <strong>in</strong>clusion data<br />

7.1. Sample description<br />

Fluid <strong>in</strong>clusion work was carried out on five samples,<br />

three from <strong>the</strong> Laur<strong>in</strong>oja deposit (DDH78-148.80,<br />

DDH170-51.40, 1-LAU-01), and two from <strong>the</strong> Kuervitikko<br />

deposit (DDH129-135.40, DDH130-115.10). The<br />

DDH78-148.80 is a quartz–magnetite ve<strong>in</strong> from cl<strong>in</strong>opyroxene<br />

skarn next to <strong>the</strong> massive ironstone and<br />

DDH170-51.40 is a barren ironstone with quartz, K-<br />

feldspar, act<strong>in</strong>olite, and albite gangue. Sample 1-LAU-<br />

01 was taken from sulphide-bear<strong>in</strong>g quartz–scapolite<br />

ve<strong>in</strong> <strong>in</strong> ironstone (Fig. 5a, b). Sample DDH129-<br />

135.40 is a pyrrhotite–chalcopyrite–pyrite–magnetitebear<strong>in</strong>g<br />

quartz ve<strong>in</strong> from footwall proximal altered<br />

metavolcanic rock and DDH130-115.10 is a chalcopy-<br />

Fig. 15. (a, b) Typical samples from which fluid <strong>in</strong>clusion work were carried out. (a) Pyrrhotite–pyrite–chalcopyrite–magnetite-bear<strong>in</strong>g quartz ve<strong>in</strong><br />

from Kuervitikko. (b). Chalcopyrite- and carbonate-bear<strong>in</strong>g quartz ve<strong>in</strong> from <strong>the</strong> Kuervitikko. Pyrite and chlorite filled fractures brecciate <strong>the</strong> ve<strong>in</strong>.<br />

(c) Photomicrograph of a pyrrhotite gra<strong>in</strong> from sample (a) with a dense, sphere-like cloud of type-1 and -3 <strong>in</strong>clusions around <strong>the</strong> gra<strong>in</strong>. (d) Close up<br />

of <strong>the</strong> <strong>in</strong>clusions <strong>in</strong> quartz around <strong>the</strong> pyrrhotite gra<strong>in</strong> shown <strong>in</strong> (c). Secondary, lower sal<strong>in</strong>ity <strong>in</strong>clusions occur <strong>in</strong> a th<strong>in</strong> fracture that crosscut <strong>the</strong><br />

type-1 and -3 <strong>in</strong>clusions. (e) Type-1a <strong>in</strong>clusions <strong>in</strong> calcite of <strong>the</strong> sample shown <strong>in</strong> (b). The type-1a <strong>in</strong>clusions <strong>in</strong> this sample conta<strong>in</strong> ubiquitously an<br />

opaque solid. (f) A typical type-1a <strong>in</strong>clusion that conta<strong>in</strong>s up to five different solids <strong>in</strong> addition to halite, Kuervitikko. (g) Co-exist<strong>in</strong>g type-1b and -3<br />

<strong>in</strong>clusions, Laur<strong>in</strong>oja. (h) Type-1b <strong>in</strong>clusions from <strong>the</strong> brittle scapolite–quartz–carbonate ve<strong>in</strong> shown <strong>in</strong> Fig. 5a–b. Halite is <strong>the</strong> only solid phase <strong>in</strong><br />

this case. (i). Type-2 <strong>in</strong>clusion <strong>in</strong> quartz.


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 97<br />

rite-bear<strong>in</strong>g quartz–calcite ve<strong>in</strong>, brecciated by pyrite<br />

and chlorite filled fractures, from hang<strong>in</strong>g wall proximal<br />

altered diorite (Fig. 15).<br />

7.2. Fluid <strong>in</strong>clusion types and micro<strong>the</strong>rmometry<br />

Three ma<strong>in</strong> types of fluid <strong>in</strong>clusions were recognised<br />

<strong>in</strong> <strong>the</strong> <strong>Kolari</strong> samples. These are identified here us<strong>in</strong>g a<br />

comb<strong>in</strong>ation of textural, phase proportional, and micro<strong>the</strong>rmometric<br />

observations: (1) H 2 OFCO 2 +halite F<br />

solids, (2) H 2 OFCO 2 , and (3) CO 2 (Table 5; Fig. 15).<br />

Type-1 <strong>in</strong>clusions are divided <strong>in</strong>to two subgroups<br />

based on <strong>the</strong>ir mode of occurrence and solid content.<br />

Type-1a <strong>in</strong>clusions are complex aqueous- or aqueouscarbonic<br />

liquid-vapour multisolid (up to 6 phases)<br />

<strong>in</strong>clusions that were found as solitary <strong>in</strong>clusions or<br />

primary clusters <strong>in</strong> quartz paragenetically associated<br />

with sulphides (Fig. 15). Type-1b <strong>in</strong>clusions were<br />

found as a network of brecciat<strong>in</strong>g <strong>in</strong>clusion trails <strong>in</strong><br />

quartz and calcite associated with sulphides and as<br />

solitary <strong>in</strong>clusions toge<strong>the</strong>r with type-2 <strong>in</strong>clusions<br />

(Fig. 15). The type-1a <strong>in</strong>clusions occur between <strong>the</strong><br />

type-1b <strong>in</strong>clusion trails. The longest dimensions of<br />

both <strong>in</strong>clusion types range from about 10 to 50 Am<br />

and <strong>the</strong> vapour bubble occupies only a small volume of<br />

<strong>the</strong>m (less than 30 vol.%). Vapour bubbles of <strong>the</strong><br />

<strong>in</strong>clusions conta<strong>in</strong> enough CO 2 to form clathrate but<br />

not a visible phase of liquid CO 2 dur<strong>in</strong>g cool<strong>in</strong>g. Sal<strong>in</strong>ity<br />

estimates based on halite dissolution temperatures<br />

are <strong>in</strong> <strong>the</strong> range of 45 to 48 and 32 to 56 wt.%<br />

NaCl equiv. for <strong>the</strong> type-1a and -1b <strong>in</strong>clusions, respectively.<br />

In type-1a <strong>in</strong>clusions, halite disappears prior to<br />

vapour bubble and <strong>in</strong> <strong>the</strong> type-1b <strong>in</strong>clusions vapour<br />

bubble disappears simultaneously or more typically<br />

prior to halite dissolution. Some daughter m<strong>in</strong>erals<br />

rema<strong>in</strong>ed undissolved up to halite dissolution temperatures.<br />

Density of <strong>the</strong> type-1a and -1b <strong>in</strong>clusions varies<br />

from 1.08 to 1.09 g/cm 3 and from 1.14 to 1.28 g/cm 3 ,<br />

respectively.<br />

Type-2 <strong>in</strong>clusions are simple aqueous Fcarbonic liquid-vapour<br />

<strong>in</strong>clusions <strong>in</strong> which <strong>the</strong> vapour bubbles do<br />

not conta<strong>in</strong> enough CO 2 to form ei<strong>the</strong>r clathrate or a<br />

visible phase of liquid CO 2 dur<strong>in</strong>g cool<strong>in</strong>g. This suggests<br />

that <strong>the</strong> CO 2 content of <strong>the</strong> bubble is less than<br />

0.85 mol.% (Hedenquist and Henley, 1985). The vapour<br />

bubble occupies less than 15 vol.% and <strong>the</strong> longest<br />

dimensions of <strong>the</strong> <strong>in</strong>clusions range from about 10 to<br />

50 Am. They occur as <strong>in</strong>ter- and <strong>in</strong>tragranular trails <strong>in</strong><br />

quartz and calcite associated with sulphides (Fig. 15).<br />

In calcite, some <strong>in</strong>clusions conta<strong>in</strong> a s<strong>in</strong>gle daughter<br />

m<strong>in</strong>eral that dissolves at temperatures around 120 8C<br />

dur<strong>in</strong>g heat<strong>in</strong>g. The <strong>in</strong>clusions have homogenisation<br />

temperatures to <strong>the</strong> liquid phase rang<strong>in</strong>g from 126 to<br />

315 8C. Sal<strong>in</strong>ity estimates based on f<strong>in</strong>al ice melt<strong>in</strong>g<br />

temperatures are <strong>in</strong> <strong>the</strong> range from 9 to 22 wt.% NaCl<br />

equivalent. First melt<strong>in</strong>g temperatures of about 35 8C<br />

Table 5<br />

Summary of fluid <strong>in</strong>clusion micro<strong>the</strong>rmometric data (8C)<br />

Deposit M<strong>in</strong> FI type ThCO 2 (L) TmCO 2 TmH 2 O ThH 2 OFCO 2 (L) TdNaCl Sal<strong>in</strong>ity<br />

Kuervitikko Qtz Type 1a 386 to 463 375 to 402 45 to 48<br />

Qtz Type 1b 132 to 266 200 to 447 32 to 56<br />

Qtz Type 2 7.3 to 5.1 155 to 174 8 to 11<br />

Qtz Type 3 20.3 to 25.7 56.7 to 56.8 Dist<strong>in</strong>ct<br />

co-occurr<strong>in</strong>g<br />

type-1b and type-3<br />

<strong>in</strong>clusion trails <strong>in</strong><br />

<strong>in</strong>clusion bcloudsQ<br />

or halos surround<strong>in</strong>g<br />

sulphides (Fig. 15)<br />

10.2 to 24.5 57.4 to 58.4<br />

0.6 to 24.6 61.1 to 60.2 Dist<strong>in</strong>ct<br />

<strong>in</strong>tragranular trails<br />

Cc Type 1b 146 to 209 198 to 244 32 to 35<br />

Cc Type 2 5.6 to 4.5 142 to 157 7 to 9<br />

Laur<strong>in</strong>oja Qtz Type 1b 109 to 216 209 to 251 32 to 35<br />

Qtz Type 2 19 to 6 126 to 315 9 to 22<br />

Qtz Type 3 20 to 2.7 56.6 Inter- and<br />

<strong>in</strong>tragranular trails<br />

M<strong>in</strong>=m<strong>in</strong>eral, FI=fluid <strong>in</strong>clusion type-1a=H 2 OFCO 2 -halite-solids, type-1b=H 2 OFCO 2 -halite, type-2=H 2 OFCO 2 , type-3=CO 2 FCH 4 ,<br />

ThCO 2 =temperature of homogenisation of CO 2 to liquid (L), TmCO 2 =f<strong>in</strong>al melt<strong>in</strong>g temperature of CO 2 , TmH 2 O=temperature of melt<strong>in</strong>g of<br />

ice, ThH 2 O/CO 2 =temperature of homogenisation of H 2 O to liquid (L) and temperature of total homogenisation of H 2 O and CO 2 to liquid (L),<br />

TdNaCl=dissolution temperature of halite, Sal<strong>in</strong>ity=equivalent weight % NaCl (TmH 2 O) or weight % NaCl (TdNaCl).


98<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

for <strong>the</strong> <strong>in</strong>clusions match <strong>the</strong> eutectic for <strong>the</strong> H 2 O–<br />

MgCl 2 –NaCl system (Crawford, 1981). Their density<br />

varies from 0.92 to 1.00 g/cm 3 .<br />

Type-3 <strong>in</strong>clusions are s<strong>in</strong>gle phase liquid CO 2 -rich<br />

<strong>in</strong>clusions which were found <strong>in</strong> quartz (not <strong>in</strong> calcite) as<br />

<strong>in</strong>tra- and <strong>in</strong>tergranular trails, as halos around sulphide<br />

gra<strong>in</strong>s, or <strong>the</strong>y occur toge<strong>the</strong>r with <strong>the</strong> type-1b <strong>in</strong>clusions<br />

<strong>in</strong> short <strong>in</strong>tragranular trails (Fig. 15). Their longest<br />

dimensions vary between 5 and 25 Am. Some<br />

<strong>in</strong>clusions conta<strong>in</strong> a s<strong>in</strong>gle solid phase. Based on<br />

CO 2 -freez<strong>in</strong>g po<strong>in</strong>t depression, <strong>in</strong> <strong>in</strong>tergranular and<br />

<strong>in</strong>tragranular trails with <strong>the</strong> type-1b <strong>in</strong>clusions <strong>the</strong>y<br />

conta<strong>in</strong> pure CO 2 , whereas <strong>in</strong> halos and <strong>in</strong> <strong>in</strong>tragranular<br />

trails <strong>the</strong>y conta<strong>in</strong> 7 to 10 and 15 to 20 mol.% CH 4 /N 2 ,<br />

respectively. Density of <strong>the</strong> type-3 <strong>in</strong>clusions range<br />

from 0.72 to 1.03 g/cm 3 <strong>in</strong> <strong>in</strong>ter- and <strong>in</strong>tragranular<br />

trails, 0.70 to 0.77 g/cm 3 <strong>in</strong> trails toge<strong>the</strong>r with <strong>the</strong><br />

type-1b <strong>in</strong>clusions, and from 0.72 to 0.98 g/cm 3 <strong>in</strong><br />

halos surround<strong>in</strong>g sulphides.<br />

8. U–Pb geochronology<br />

8.1. Previous age determ<strong>in</strong>ations<br />

Previous U–Pb age data from <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong><br />

have been published by Hiltunen (1982) and were<br />

obta<strong>in</strong>ed from a monzonite sample from <strong>the</strong> Hannuka<strong>in</strong>en<br />

deposit (A840), albitite <strong>in</strong> <strong>the</strong> hang<strong>in</strong>g wall of <strong>the</strong><br />

Rautuvaara deposit (A958), a coarse-gra<strong>in</strong>ed bmafic<br />

pegmatoidQ (A959) from <strong>the</strong> Rautuvaara deposit, and<br />

cl<strong>in</strong>opyroxene skarn from <strong>the</strong> Hannuka<strong>in</strong>en deposit<br />

(A963). The ages for samples A840, A958, A959,<br />

and A963 given below are from Väänänen and Lehtonen<br />

(2001) <strong>in</strong> which Hiltunen’s (1982) recalculated U–<br />

Pb data are presented. All <strong>the</strong>se previously dated samples<br />

were re-checked from <strong>the</strong> GSF archives, and brief<br />

updated sample descriptions are given below.<br />

Sample A840 is a typical monzonite that occurs <strong>in</strong><br />

<strong>the</strong> hang<strong>in</strong>g wall for <strong>the</strong> Kuervitikko and Cu-Rautuvaara<br />

and Hannuka<strong>in</strong>en <strong>deposits</strong>. In macroscopic <strong>in</strong>vestigation,<br />

it appears as slightly altered, possibly albitised.<br />

It displays a well developed foliation and l<strong>in</strong>eation,<br />

probably S 3 and L 3 . The 1862F4 Ma U–Pb zircon<br />

age of <strong>the</strong> monzonite is based on four fractions that<br />

are slightly discordant. The age for titanite is 1784 Ma.<br />

Hiltunen (1982) proposed that A958 is an albitic<br />

variety of <strong>the</strong> monzonite derived via magmatic fractionation.<br />

In contrast, our op<strong>in</strong>ion is that albitites are metasomatic<br />

ra<strong>the</strong>r than orthomagmatic rocks. Macroscopic<br />

<strong>in</strong>vestigation of A958 revealed that it is similar to <strong>the</strong><br />

distal altered diorite at Laur<strong>in</strong>oja and Kuervitikko. The<br />

A958 albitite is strongly foliated and <strong>the</strong> l<strong>in</strong>eation is<br />

well developed. The ages for zircon and titanite are<br />

1849F19 Ma and 1783 Ma, respectively.<br />

Accord<strong>in</strong>g to Hiltunen (1982), sample A959 was<br />

taken from a narrow, coarse-gra<strong>in</strong>ed mafic pegmatoid<br />

ve<strong>in</strong> that crosscuts <strong>the</strong> skarn and skarn hosted ore <strong>in</strong> <strong>the</strong><br />

Rautuvaara deposit. The macroscopic review of A959<br />

showed that it is unoriented, coarse-gra<strong>in</strong>ed (V15 mm),<br />

and consists of plagioclase (albite?), amphibole, cl<strong>in</strong>opyroxene,<br />

quartz, and possibly scapolite. The orig<strong>in</strong> of<br />

<strong>the</strong> ve<strong>in</strong> is ambiguous but it could be l<strong>in</strong>ked to <strong>the</strong> late<br />

(D 4 ) brittle deformation (e.g., Fig. 5a, b). The 1748F7<br />

Ma zircon age of <strong>the</strong> ve<strong>in</strong> from six highly discordant<br />

zircon fractions is <strong>the</strong> youngest age obta<strong>in</strong>ed from <strong>the</strong><br />

whole of nor<strong>the</strong>rn F<strong>in</strong>land (Hanski et al., 2001).<br />

Sample A963 is a typical cl<strong>in</strong>opyroxene skarn that<br />

occurs as <strong>the</strong> ma<strong>in</strong> host rock for <strong>the</strong> ironstones for <strong>the</strong><br />

Kuervitikko and Hannuka<strong>in</strong>en <strong>deposits</strong>. It consists almost<br />

solely of cl<strong>in</strong>opyroxene and is <strong>in</strong> very prist<strong>in</strong>e<br />

condition show<strong>in</strong>g no signs of retrograde alteration.<br />

The 1797 F5 Ma age for <strong>the</strong> skarn is constructed by<br />

three slightly discordant zircon fractions and titanite is<br />

coeval.<br />

8.2. U–Pb age determ<strong>in</strong>ations of A1697 granite and<br />

A1698 diorite<br />

Two new rock samples (A1697 and A1698) were<br />

taken for U–Pb dat<strong>in</strong>g. U–Pb data on zircon and titanite<br />

are shown <strong>in</strong> Table 6 and <strong>the</strong> concordia plots presented<br />

<strong>in</strong> Fig. 16. All U–Pb ages from <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> are<br />

summarised <strong>in</strong> Fig. 17.<br />

The A1697 granite was taken from <strong>the</strong> waste rock<br />

piles at Hannuka<strong>in</strong>en. It is even gra<strong>in</strong>ed, unoriented and<br />

brecciates and crosscuts Cu–Au-m<strong>in</strong>eralised ironstone<br />

(Fig. 5c–d). Four fractions of prismatic, magmatic zircon<br />

and one fraction of dark brown, translucent titanite<br />

were dated (Table 6). On <strong>the</strong> concordia diagram, <strong>the</strong><br />

prismatic zircons give an upper <strong>in</strong>tercept age of<br />

1766F5 Ma(Fig. 16). The concordia age for titanite<br />

is 1759F3 Ma. Accord<strong>in</strong>gly, <strong>the</strong> ages of <strong>the</strong> magmatic<br />

zircon and titanite are coeval with<strong>in</strong> <strong>the</strong> error limits<br />

(Fig. 17).<br />

The granite also conta<strong>in</strong>s a small amount of bright,<br />

colourless and rounded zircon without crystal faces.<br />

Consequently, <strong>the</strong>se gra<strong>in</strong>s have characteristics typical<br />

of metamorphic zircon. The Pb–U ratios of <strong>the</strong>se zircon<br />

gra<strong>in</strong>s plot quite close to <strong>the</strong> concordia curve (Fig. 16).<br />

The age of <strong>the</strong> only slightly discordant data can be<br />

approximated us<strong>in</strong>g 207 Pb/ 206 Pb ratio to be ca.<br />

1800 Ma. As <strong>the</strong> magmatic age of <strong>the</strong> Hannuka<strong>in</strong>en<br />

granite is significantly younger, <strong>the</strong> 1800 Ma zircon<br />

must be <strong>in</strong>herited <strong>in</strong> orig<strong>in</strong>. Interest<strong>in</strong>gly, this age is


Table 6<br />

Conventional U–Pb age data on zircon and titanite, samples A1697 granite, and A1698 diorite<br />

Sample <strong>in</strong>formation, analysed<br />

m<strong>in</strong>eral and fraction<br />

Sample<br />

(weight/mg)<br />

U<br />

(ppm)<br />

Pb<br />

(ppm)<br />

206 Pb/ 204 Pb<br />

(measured)<br />

208 Pb/ 206 Pb<br />

(radiogenic)<br />

Isotopic ratios a<br />

Apparent ages/MaF2r<br />

206 Pb/ 238 U F2r 207 Pb/ 235 U F2r 207 Pb/ 206 Pb F2r Rho 206 Pb/ 238 U 207 Pb/ 235 U 207 Pb/ 206 Pb<br />

A1697 granite<br />

(A) Zircon d: N4.2 g/cm 3 , N75 Am,<br />

bright, colourless, rounded,<br />

abraded 4 h<br />

(B) Zircon d: N4.2 g/cm 3 , b75 Am,<br />

long prismatic, translucent,<br />

abraded 10 h<br />

(C) Zircon d: N4.2 g/cm 3 , b75 Am,<br />

long prismatic, translucent,<br />

abraded 1 h<br />

(D) Zircon d: 4.2–4.0 g/cm 3 ,<br />

long prismatic,<br />

translucent, brownish, abraded 10 h<br />

(E) Zircon d: 4.2–4.0 g/cm 3 ,<br />

long prismatic,<br />

translucent, brownish, abraded 1 h<br />

(F) Titanite, dark brown, transparent,<br />

abraded 1/3 h<br />

A1698 diorite<br />

(A) Zircon d: N4.2 g/cm 3 , N75 Am,<br />

prismatic, transparent, colourless,<br />

abraded 4 h<br />

(B) Zircon d: N4.2 g/cm 3 , b75 Am,<br />

prismatic, transparent, colourless,<br />

abraded 5 h<br />

(C) Zircon d: N4.2 g/cm 3 , b75 Am,<br />

prismatic, transparent, colourless,<br />

abraded 3 h<br />

(D) Zircon d: N4.2 g/cm 3 , b75 Am,<br />

prismatic, transparent, colourless<br />

(E) Titanite, dark brown, translucent,<br />

abraded 1/2 h<br />

0.59 152 52 2078 0.16 0.3058 0.65 4.619 0.65 0.1096 0.15 0.97 1720 1753 1792F2<br />

0.32 1425 361 3244 0.14 0.2318 0.65 3.373 0.65 0.1055 0.15 0.97 1344 1498 1724F2<br />

0.26 1120 229 3061 0.13 0.1895 0.65 2.710 0.65 0.1037 0.15 0.97 1119 1331 1691F2<br />

0.26 1100 297 2978 0.14 0.2472 0.65 3.623 0.65 0.1063 0.15 0.97 1424 1555 1737F2<br />

0.48 1317 317 2554 0.13 0.2212 0.65 3.212 0.65 0.1053 0.15 0.97 1288 1460 1720F2<br />

1.97 173 96 654 0.83 0.3143 0.40 4.662 0.42 0.1076 0.20 0.88 1762 1761 1759F4<br />

0.24 579 221 1659 0.25 0.3149 0.40 4.939 0.42 0.1137 0.13 0.95 1765 1809 1860F3<br />

0.50 456 183 1828 0.30 0.3205 0.40 5.027 0.42 0.1137 0.07 0.98 1792 1824 1860<br />

0.31 432 170 966 0.28 0.3132 0.40 4.908 0.42 0.1136 0.10 0.96 1757 1804 1858F2<br />

0.16 537 202 1014 0.23 0.3115 0.40 4.872 0.42 0.1135 0.17 0.89 1748 1798 1855F3<br />

1.85 127 63 759 0.56 0.3241 0.40 4.924 0.60 0.1102 0.46 0.62 1810 1806 1803F9<br />

a Isotopic ratios corrected for fractionation, blank (50 pg), and age related common lead (Stacey and Kramers, 1975; 206 Pb/ 204 PbF0.2, 207 Pb/ 204 PbF0.1, 208 Pb/ 204 PbF0.2).<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 99


100<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Fig. 16. Concordia diagrams show<strong>in</strong>g <strong>the</strong> conventional U–Pb age data on zircons and titanites, samples A1697 granite (a) and A1698 diorite (b)<br />

from <strong>the</strong> Laur<strong>in</strong>oja deposit. See Table 5 for <strong>the</strong> U–Pb data. Error ellipses are 2r. Concordia ages are calculated at 2r confidence level and decayconstant<br />

errors ignored.<br />

equal with <strong>the</strong> age of zircon and titanite from <strong>the</strong> A963<br />

cl<strong>in</strong>opyroxene skarn (Fig. 17). Fur<strong>the</strong>rmore, <strong>the</strong> age<br />

from <strong>the</strong> titanite <strong>in</strong> A840 monzonite is also ca. 1800<br />

Ma.<br />

The A1698 diorite was also taken from <strong>the</strong> waste<br />

rock piles at Hannuka<strong>in</strong>en. It is a typical distal altered<br />

diorite that occurs as a wall rock for all studied <strong>deposits</strong><br />

display<strong>in</strong>g well developed S 3 foliation and L 3 l<strong>in</strong>eation<br />

(e.g., Fig. 6b). Four fractions consist<strong>in</strong>g of prismatic,<br />

colourless, transparent zircon were dated from <strong>the</strong> sample<br />

(Table 6). The data plot on a discordia l<strong>in</strong>e which<br />

<strong>in</strong>tercepts <strong>the</strong> concordia curve at 1864F5 Ma and<br />

170F160 Ma (Fig. 16). With<strong>in</strong> <strong>the</strong> error limits, <strong>the</strong><br />

A1698 diorite is contemporaneous with <strong>the</strong> A840 monzonite<br />

and A958 albitite (Fig. 17).<br />

A titanite fraction from <strong>the</strong> diorite gives a concordia<br />

age of 1805 F5 Ma(Table 6 and Fig. 16). The significant<br />

age difference with <strong>the</strong> zircon and titanite <strong>in</strong> <strong>the</strong><br />

A1698 diorite is consistent with <strong>the</strong> age difference<br />

between <strong>the</strong>se m<strong>in</strong>erals <strong>in</strong> <strong>the</strong> A840 monzonite and<br />

A958 albitite (Fig. 17). The titanite ages thus must<br />

reflect <strong>the</strong> time of a later metamorphism.<br />

9. Discussion<br />

The data presented <strong>in</strong> this work suggest that <strong>the</strong><br />

skarns and ironstones <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> overpr<strong>in</strong>t<br />

several different rock types, chiefly type-2 metavolcanic<br />

rocks (Fig. 13). This strongly suggests that <strong>the</strong><br />

<strong>deposits</strong> are of metasomatic replacement type ra<strong>the</strong>r<br />

than metamorphosed syngenetic iron formations as<br />

was previously suggested (Mäkelä and Tammenmaa,<br />

1978; Frietsch et al., 1995). Although <strong>the</strong> calcic alteration<br />

assemblages <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> are quite similar<br />

to <strong>the</strong> ones described <strong>in</strong> association with calcic-iron<br />

skarns (sensu stricto), <strong>the</strong> zon<strong>in</strong>g at Laur<strong>in</strong>oja and<br />

Kuervitikko is atypical for a classic skarn deposit. In<br />

<strong>the</strong> skarn <strong>deposits</strong>, <strong>the</strong> zon<strong>in</strong>g is typically characterised<br />

Fig. 17. Summary of U–Pb ages from <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong>. A840, A958, A959, and A963 after Hiltunen (1982) and Väänänen and Lehtonen (2001).


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 101<br />

by variations <strong>in</strong> proportions and compositions of pyroxenes<br />

and garnet (e.g., E<strong>in</strong>audi et al., 1981). In <strong>the</strong><br />

<strong>Kolari</strong> <strong>deposits</strong>, garnet-rich skarns are practically absent,<br />

and no clear variation can be seen <strong>in</strong> <strong>the</strong> colour or<br />

optical properties of cl<strong>in</strong>opyroxene that would <strong>in</strong>dicate<br />

significant variation <strong>in</strong> its composition. Instead, <strong>the</strong><br />

zon<strong>in</strong>g <strong>in</strong> <strong>Kolari</strong> <strong>deposits</strong> consists of moderately to<br />

<strong>in</strong>tensely alkali-altered distal, and calcic-iron-altered<br />

proximal zones. The zon<strong>in</strong>g is roughly symmetrical<br />

around <strong>the</strong> ironstones, although <strong>the</strong> distal alteration<br />

appears to be more extensive and <strong>in</strong>tense <strong>in</strong> <strong>the</strong> hang<strong>in</strong>g<br />

wall compared to footwall. The characteristics of <strong>the</strong><br />

zon<strong>in</strong>g, toge<strong>the</strong>r with <strong>the</strong> <strong>in</strong>timate relationship of <strong>the</strong><br />

<strong>deposits</strong> with thrust and shear zones suggest a hydro<strong>the</strong>rmal<br />

system where <strong>the</strong> ironstones represent <strong>the</strong> structural<br />

base of <strong>the</strong> shear system and <strong>the</strong> zon<strong>in</strong>g is<br />

controlled by fluid flux, fluid evolution, and fluidwall<br />

rock reactions <strong>in</strong> and around <strong>the</strong> fault and thrust<br />

zones. The lack of suitable structures that would focus<br />

<strong>the</strong> fluid flow could expla<strong>in</strong> <strong>the</strong> poorly developed zon<strong>in</strong>g<br />

<strong>in</strong> <strong>the</strong> Cu-Rautuvaara deposit.<br />

9.1. Temporal relationship between ironstones, Cu–Au<br />

m<strong>in</strong>eralisation, and deformation stages<br />

Two generations of sulphides exist <strong>in</strong> <strong>the</strong> <strong>Kolari</strong><br />

<strong>deposits</strong>, and both of <strong>the</strong>se conta<strong>in</strong> chalcopyrite. The<br />

majority of sulphides pre-date <strong>the</strong> brittle deformation.<br />

In <strong>the</strong> magnetite-dissem<strong>in</strong>ated and Cu–Au-m<strong>in</strong>eralised<br />

albitites <strong>the</strong> textural features suggest that magnetite and<br />

sulphides were contemporaneous (Fig. 7). The gangue<br />

m<strong>in</strong>eralogy is similar between <strong>the</strong> barren and sulphiderich<br />

parts of <strong>the</strong> ironstones. No systematic dist<strong>in</strong>ction<br />

can be made <strong>in</strong> chemical composition between <strong>the</strong> Cu–<br />

Au m<strong>in</strong>eralised and unm<strong>in</strong>eralised ironstone units, except<br />

for concentrations of sulphur and elements correlated<br />

with that. These facts suggest that precipitation of<br />

<strong>the</strong> magnetite and sulphides was contemporaneous. The<br />

S 3 foliation <strong>in</strong> <strong>the</strong> ironstones and <strong>the</strong> orientation of <strong>the</strong><br />

ironstone lenses parallel to L 3 l<strong>in</strong>eation at Laur<strong>in</strong>oja and<br />

Kuervitikko (Figs. 3 and 4), and <strong>the</strong> U–Pb age data (see<br />

below) suggests that <strong>the</strong> formation of ironstones and<br />

Cu–Au m<strong>in</strong>eralisation were contemporaneous to <strong>the</strong> D 3<br />

thrust<strong>in</strong>g event. The good correlation between Cu and<br />

Au <strong>in</strong> all <strong>deposits</strong> suggests that <strong>the</strong>y most likely were<br />

precipitated <strong>in</strong> <strong>the</strong> same event.<br />

The youngest generation are related to brittle fractures<br />

that crosscut <strong>the</strong> S 3 foliation and <strong>the</strong>refore postdate<br />

D 3 be<strong>in</strong>g probably l<strong>in</strong>ked to D 4 stage (Figs. 5 and<br />

15b). However, it is unknown whe<strong>the</strong>r <strong>the</strong> sulphides <strong>in</strong><br />

this stage are just remobilised varieties of <strong>the</strong> pre-exist<strong>in</strong>g<br />

ones, or if <strong>the</strong>y are derived from a more distant<br />

source. Never<strong>the</strong>less, consider<strong>in</strong>g <strong>the</strong> low abundance of<br />

late fractures <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> deposit, <strong>the</strong>ir contribution to<br />

<strong>the</strong> Cu–Au budget is low.<br />

9.2. Implications of <strong>the</strong> U–Pb age data<br />

The 1864F5 Ma U–Pb zircon age of <strong>the</strong> Hannuka<strong>in</strong>en<br />

diorite is, with<strong>in</strong> <strong>the</strong> error limits, <strong>the</strong> same as <strong>the</strong> U–<br />

Pb zircon age of <strong>the</strong> Hannuka<strong>in</strong>en monzonite suggest<strong>in</strong>g<br />

that <strong>the</strong>y are coeval (Fig. 17). The age of <strong>the</strong> diorite<br />

also def<strong>in</strong>es a maximum age for <strong>the</strong> ores. Based on <strong>the</strong><br />

contact relations between <strong>the</strong> <strong>in</strong>trusives and <strong>the</strong> supracrustal<br />

rocks, Hiltunen (1982) proposed that monzonite<br />

and diorite <strong>in</strong>trusions <strong>in</strong>truded conformably <strong>in</strong>to <strong>the</strong><br />

supracrustal rocks dur<strong>in</strong>g <strong>the</strong> early stages of <strong>the</strong> ma<strong>in</strong><br />

fold<strong>in</strong>g. This would mean that <strong>the</strong>y <strong>in</strong>truded dur<strong>in</strong>g <strong>the</strong><br />

D 2 stage of Väisänen (2002) and Hölttä et al. (<strong>in</strong> press).<br />

Thus, <strong>the</strong> D 2 stage <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> is 1865 Ma or<br />

older. The m<strong>in</strong>imum age for <strong>the</strong> ores is constra<strong>in</strong>ed by<br />

<strong>the</strong> undeformed, 1766F5 Ma granite that brecciates <strong>the</strong><br />

ore. Coeval undeformed granites occur throughout<br />

nor<strong>the</strong>rn F<strong>in</strong>land (e.g., Hanski et al., 2001), and <strong>the</strong>y<br />

also constra<strong>in</strong> <strong>the</strong> m<strong>in</strong>imum age for <strong>the</strong> D 3 thrust<strong>in</strong>g<br />

event (e.g., Väisänen, 2002).<br />

The 1797 F5 Ma U–Pb zircon age of <strong>the</strong> cl<strong>in</strong>opyroxene<br />

skarn, which is ca. 65 Ma younger than <strong>the</strong> age<br />

of <strong>the</strong> hang<strong>in</strong>g wall <strong>in</strong>trusions, makes <strong>the</strong> genetic model<br />

presented by Hiltunen (1982) dubious. A number of<br />

analyses of titanite display metamorphic ages that are<br />

similar to <strong>the</strong> zircon age of <strong>the</strong> skarn (Fig. 17). Fur<strong>the</strong>rmore,<br />

<strong>the</strong> <strong>in</strong>herited zircon from <strong>the</strong> granite that brecciates<br />

<strong>the</strong> ore is ca. 1800 Ma <strong>in</strong> age. Hiltunen (1982) does<br />

not give any explanation to age difference between <strong>the</strong><br />

skarn and <strong>the</strong> monzonite <strong>in</strong>trusions, which is clearly too<br />

large if <strong>the</strong> cl<strong>in</strong>opyroxene skarn is related to <strong>the</strong> hang<strong>in</strong>g<br />

wall <strong>in</strong>trusions. The U–Pb age of 1797F5 Ma for<br />

<strong>the</strong> cl<strong>in</strong>opyroxene skarn is constructed by only three<br />

bulk zircon fractions which actually form a two po<strong>in</strong>t<br />

discordia l<strong>in</strong>e (Hiltunen, 1982). The titanite is apparently<br />

slightly younger than <strong>the</strong> upper <strong>in</strong>tercept age for<br />

<strong>the</strong> zircon fractions, and <strong>the</strong> U–Pb data for A840 monzonite<br />

plot on <strong>the</strong> bolder sideQ of <strong>the</strong> A963 data. Thus,<br />

although <strong>the</strong> skarn zircon can also conta<strong>in</strong> some <strong>in</strong>herited<br />

zircon components, <strong>the</strong> transparent, brown, euhedral<br />

zircon most likely crystallized at <strong>the</strong> time of skarn<br />

formation. The age of <strong>the</strong> skarn cannot be younger than<br />

<strong>the</strong> 1780 Ma, titanite age and <strong>the</strong> age for <strong>the</strong> new zircon<br />

must be younger than <strong>the</strong> age of <strong>the</strong> monzonite especially<br />

when <strong>in</strong>herited, altered zircon also are mixed with<br />

<strong>the</strong> dated fractions. However, fur<strong>the</strong>r ion microprobe<br />

dat<strong>in</strong>g would be required to thoroughly understand <strong>the</strong><br />

ages and systematics of <strong>the</strong> skarn zircon. Regardless of


102<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

all <strong>the</strong> problems deal<strong>in</strong>g with <strong>the</strong> skarn zircon age, we<br />

believe that <strong>the</strong> skarn alteration <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong><br />

took place at ca. 1800 Ma, and that <strong>the</strong> titanite ages of<br />

<strong>the</strong> altered monzonite, diorite, and albitite reflect <strong>the</strong><br />

age of <strong>the</strong> distal alteration that was contemporaneous<br />

with <strong>the</strong> proximal, skarn alteration. Hence, <strong>the</strong> formation<br />

of <strong>the</strong> ironstones and Cu–Au occurrences at <strong>Kolari</strong><br />

most likely took place ca. 1800 Ma. This is co<strong>in</strong>cident<br />

with <strong>the</strong> <strong>the</strong>rmal event that led to <strong>the</strong> formation of ca.<br />

1800 to 1770 Ma granites that are abundant throughout<br />

nor<strong>the</strong>rn F<strong>in</strong>land and nor<strong>the</strong>rn Sweden. Fur<strong>the</strong>rmore,<br />

<strong>the</strong> <strong>in</strong>timate relation of <strong>the</strong> <strong>Kolari</strong> ores to <strong>the</strong> D 3 thrust<br />

zones suggests that <strong>the</strong> D 3 stage <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong><br />

took place around 1800 Ma, if our <strong>in</strong>terpretation of <strong>the</strong><br />

zircon age of <strong>the</strong> skarn is correct. This <strong>in</strong>terpretation is<br />

supported by <strong>the</strong> U–Pb data from <strong>the</strong> Pajala Shear Zone<br />

<strong>in</strong> nor<strong>the</strong>rn Sweden, which forms <strong>the</strong> sou<strong>the</strong>rn extension<br />

to <strong>the</strong> <strong>Kolari</strong> Shear Zone, where ages between<br />

1810 and 1774 Ma were def<strong>in</strong>ed from metamorphic<br />

zircon <strong>in</strong> mica gneiss (Bergman and Skiöld, 1998). This<br />

also <strong>in</strong>dicates that a thrust event correspond<strong>in</strong>g to D 3 <strong>in</strong><br />

<strong>Kolari</strong> <strong>region</strong> took place between 1810 and 1774 Ma.<br />

The 1748F7 Ma age of <strong>the</strong> bmafic pegmatoidQ is<br />

ambiguous. No <strong>in</strong>trusives with a similar age have been<br />

dated from nor<strong>the</strong>rn F<strong>in</strong>land (cf. Hanski et al., 2001). If<br />

<strong>the</strong> pegmatoid ve<strong>in</strong> is related to brittle fractures, <strong>the</strong> age<br />

of <strong>the</strong> D 4 deformation <strong>in</strong> <strong>the</strong> nor<strong>the</strong>rn F<strong>in</strong>land is <strong>in</strong>deed<br />

significantly younger than previously considered as<br />

proposed by Väisänen (2002).<br />

9.3. Pressure and temperature considerations,<br />

evolution of <strong>the</strong> fluid<br />

The early type-1a H 2 OFCO 2 –haliteFsolids <strong>in</strong>clusions<br />

<strong>in</strong> quartz with abundant sulphides <strong>in</strong>dicate conditions<br />

under which <strong>the</strong> enclos<strong>in</strong>g host were formed. The<br />

fluids were trapped homogeneously (constant phase<br />

ratios) <strong>in</strong> a s<strong>in</strong>gle-phase <strong>region</strong> of <strong>the</strong> H 2 O–CO 2 –halite<br />

system (Fig. 18). However, <strong>the</strong> orig<strong>in</strong>al P–T conditions<br />

of entrapment cannot be reconstructed from fluid <strong>in</strong>clusion<br />

evidence alone. Never<strong>the</strong>less, <strong>the</strong> total homogenisation<br />

temperature (386 to 463 8C) and <strong>the</strong> total<br />

homogenisation pressure (ca. 1.5 kbar) serve as m<strong>in</strong>imum<br />

constra<strong>in</strong>ts on <strong>the</strong> entrapment conditions. The<br />

pressure dur<strong>in</strong>g <strong>the</strong> peak of <strong>the</strong> <strong>region</strong>al metamorphism<br />

<strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> was between 3.3 and 3.8 kbar<br />

(Hölttä et al., <strong>in</strong> press). S<strong>in</strong>ce <strong>the</strong> D 3 stage post-dates<br />

<strong>the</strong> metamorphic peak, <strong>the</strong>se pressures can be considered<br />

to represent <strong>the</strong> upper limits dur<strong>in</strong>g <strong>the</strong> alteration<br />

and <strong>the</strong> Fe and Cu–Au m<strong>in</strong>eralisation. Based on this, <strong>the</strong><br />

pressure dur<strong>in</strong>g <strong>the</strong> entrapment of <strong>the</strong> type-1a <strong>in</strong>clusions<br />

was between 1.5 and 3.5 kbar and <strong>the</strong> temperature was<br />

Fig. 18. Pressure–temperature diagram show<strong>in</strong>g <strong>the</strong> location of <strong>the</strong><br />

isochors for <strong>the</strong> relevant fluid <strong>in</strong>clusion types detected at <strong>Kolari</strong>.<br />

Solvus for <strong>the</strong> H 2 O–40 wt.% NaCl system with 5 mol.% CO 2 is<br />

from Schmidt and Bodnar (1994) and liquidus for <strong>the</strong> 30 and 40 wt.%<br />

NaCl are from Cl<strong>in</strong>e and Bodnar (1994) and Bodnar (1994). Pressure<br />

dur<strong>in</strong>g <strong>the</strong> peak of <strong>the</strong> <strong>region</strong>al metamorphism after Hölttä et al. (<strong>in</strong><br />

press).<br />

between 450 and 550 8C (Fig. 18). Subsequent cool<strong>in</strong>g<br />

and brittle fractur<strong>in</strong>g resulted <strong>in</strong> trapp<strong>in</strong>g of <strong>the</strong> type-1b<br />

<strong>in</strong>clusions. The co-exist<strong>in</strong>g type-1b and -3 <strong>in</strong>clusions<br />

suggest that dur<strong>in</strong>g <strong>the</strong> evolution of <strong>the</strong> hydro<strong>the</strong>rmal<br />

system at Laur<strong>in</strong>oja and Kuervitikko, a phase separation<br />

occurred between <strong>the</strong> aqueous and carbonic components<br />

of <strong>the</strong> fluid. This could have occurred ei<strong>the</strong>r due to<br />

cool<strong>in</strong>g, or due to pressure fluctuations or both as<br />

<strong>in</strong>dicated by <strong>the</strong> considerable variations between <strong>the</strong><br />

densities of <strong>the</strong> type-3 <strong>in</strong>clusions.<br />

9.4. IOCG m<strong>in</strong>eralisation at <strong>Kolari</strong>?<br />

The ca. 65 Ma age difference between <strong>the</strong> monzonite<br />

and diorite <strong>in</strong>trusions and skarn <strong>in</strong>dicates that <strong>the</strong> genesis<br />

of <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> cannot be related to <strong>the</strong><br />

<strong>in</strong>trusions as suggested by Hiltunen (1982), and <strong>the</strong>refore<br />

an alternative orig<strong>in</strong> should be considered. The<br />

general features of <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> display characteristics<br />

typical for IOCG occurrences that form a controversial<br />

spectrum of <strong>deposits</strong> <strong>in</strong>to which several<br />

diverse FeFCu, Au, REE, U <strong>deposits</strong> have been <strong>in</strong>cluded<br />

(Hitzman et al., 1992; Barton and Johnson,<br />

1996, 2000; Hitzman, 2000; Williams and Pollard,<br />

2003). Common characteristics for both <strong>the</strong> <strong>Kolari</strong><br />

and <strong>the</strong> IOCG <strong>deposits</strong> elsewhere <strong>in</strong>clude (1) high<br />

bulk Fe–S ratio, (2) metal association Fe–Cu–<br />

AuFAg, Co, Mo, REE, U, (3) relatively low Cu and<br />

Au concentrations, (4) relation to highly sal<strong>in</strong>e aqueous


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 103<br />

to carbonic br<strong>in</strong>es, (5) distal Na FK, Cl and proximal<br />

Ca–FeFCO 2 , S alteration (e.g., Hitzman et al., 1992;<br />

Barton and Johnson, 2000; Perr<strong>in</strong>g et al., 2000; Pollard,<br />

2001; Oliver et al., 2004). Generally <strong>in</strong> IOCG occurrences,<br />

<strong>the</strong> proximal alteration is dom<strong>in</strong>ated by K- and<br />

Fe-rich phases, i.e., biotite, K-feldspar, sericite, and Fe<strong>oxide</strong>s<br />

(e.g., Ernest Henry; Mark et al., 2000). Never<strong>the</strong>less,<br />

<strong>in</strong> a number of occurrences (e.g., Candelaria,<br />

Mt Elliott; cf. Pollard, 2000) <strong>the</strong> Cu–Au m<strong>in</strong>eralisations<br />

are hosted by magnetite-rich dskarnsT. Of <strong>the</strong> known<br />

IOCG occurrences, <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> are perhaps<br />

most similar to <strong>the</strong> Mt Elliott deposit (Cloncurry district,<br />

Australia), where <strong>the</strong> Cu–Au m<strong>in</strong>eralisation is<br />

hosted by cl<strong>in</strong>opyroxene–magnetite skarn comparable<br />

to <strong>the</strong> one at Laur<strong>in</strong>oja and Kuervitikko (Wang and<br />

Williams, 2001).<br />

We can only speculate about <strong>the</strong> ultimate source of<br />

<strong>the</strong> metals enriched <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong>. S<strong>in</strong>ce <strong>the</strong><br />

general features of <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> are so apparently<br />

similar with <strong>deposits</strong> classified <strong>in</strong>to <strong>the</strong> IOCG type, we<br />

feel that it is reasonable to try to apply <strong>the</strong> IOCG<br />

genetic model. The IOCG models emphasise <strong>the</strong> role<br />

of distal sodic alteration for <strong>the</strong> source of <strong>the</strong> elements<br />

enriched <strong>in</strong> <strong>the</strong> ironstones, barren or Cu–Au m<strong>in</strong>eralised<br />

(e.g., Williams, 1994; Barton and Johnson, 1996;<br />

Hitzman, 2000; Oliver et al., 2004). The true extent of<br />

<strong>the</strong> sodic alteration around <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> is not<br />

known, but it probably covers several km 2 of <strong>the</strong> country<br />

rocks along <strong>the</strong> strike of <strong>the</strong> thrust zones. Also, <strong>the</strong><br />

alteration can be vertically extensive consider<strong>in</strong>g that<br />

<strong>the</strong> <strong>deposits</strong> are located <strong>in</strong> major thrust structures. Thus,<br />

similar to IOCG models, <strong>the</strong> <strong>region</strong>ally albite-altered<br />

rocks could be <strong>the</strong> source for <strong>the</strong> bulk of <strong>the</strong> elements<br />

enriched <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong>. It is, however, possible<br />

that part of <strong>the</strong> CaO and CO 2 <strong>in</strong> <strong>the</strong> proximal alteration<br />

zones are derived from a local source, i.e., from a<br />

carbonaceous unit, s<strong>in</strong>ce such rocks are known to<br />

occur <strong>in</strong> <strong>the</strong> supracrustal sequence <strong>in</strong> which <strong>the</strong> <strong>deposits</strong><br />

are located (Hiltunen, 1982).<br />

Recent work by Oliver et al. (2004) shows that Cu is<br />

not systematically stripped from rocks subjected to<br />

albite alteration, <strong>in</strong> contrast to Fe, K, Ba, Rb FCa, Sr,<br />

Co, V, Mn, Pb, and Zn that are more consistently<br />

leached out of <strong>the</strong> albitised rocks. They propose a<br />

model where Cu <strong>in</strong> <strong>the</strong> IOCG <strong>deposits</strong> is derived<br />

from <strong>the</strong> source of <strong>the</strong> br<strong>in</strong>es, and precipitation of <strong>the</strong><br />

Cu was a result of mix<strong>in</strong>g of Cu-bear<strong>in</strong>g br<strong>in</strong>e that is<br />

evolved extensively via albitisation with sulphur bear<strong>in</strong>g<br />

fluid, or reaction of <strong>the</strong> br<strong>in</strong>e with sulphur-bear<strong>in</strong>g<br />

country rocks. In <strong>the</strong>ir model <strong>the</strong> ultimate source for <strong>the</strong><br />

albitis<strong>in</strong>g br<strong>in</strong>es and Cu is a deep seated granitoid. In<br />

<strong>the</strong> cases where <strong>the</strong> fluid lacks Cu, or sulphur source is<br />

absent, or both, br<strong>in</strong>es enriched <strong>in</strong> elements released by<br />

albitisation produce barren ironstones. Apply<strong>in</strong>g this<br />

model to <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> could expla<strong>in</strong> <strong>the</strong> large<br />

variation <strong>in</strong> <strong>the</strong> sulphur isotope compositions <strong>in</strong> <strong>the</strong>m<br />

(Mäkelä and Tammenmaa, 1978; Hiltunen, 1982), i.e.,<br />

that <strong>the</strong> sulphur <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> was derived from<br />

<strong>the</strong> Savukoski Group metavolcanic and sedimentary<br />

rocks, and also possibly from <strong>the</strong> abundant <strong>in</strong>trusives.<br />

Similarly, <strong>the</strong> availability of Cu and/or S could expla<strong>in</strong><br />

why some of <strong>the</strong> ironstones <strong>in</strong> <strong>the</strong> <strong>Kolari</strong> <strong>region</strong> are Cu<br />

m<strong>in</strong>eralised and o<strong>the</strong>rs are not.<br />

10. Conclusions<br />

Current data on <strong>the</strong> <strong>Kolari</strong> Fe–Cu–Au <strong>deposits</strong> show<br />

that <strong>the</strong>y are hydro<strong>the</strong>rmal replacement-type <strong>deposits</strong><br />

and <strong>the</strong>y overpr<strong>in</strong>t <strong>the</strong> ca. 1860 Ma Haparanda-type<br />

<strong>in</strong>trusives and <strong>the</strong> metavolcanic rocks and metasediments<br />

of <strong>the</strong> Savukoski Group. U–Pb age data presented<br />

<strong>in</strong> this work and by previous authors <strong>in</strong>dicate<br />

that <strong>the</strong> ores were formed between 1864F5 Ma and<br />

1766F5 Ma, most likely at ca. 1800 Ma as suggested<br />

by <strong>the</strong> age of <strong>the</strong> cl<strong>in</strong>opyroxene skarn at Hannuka<strong>in</strong>en.<br />

The temperature dur<strong>in</strong>g <strong>the</strong> ma<strong>in</strong> alteration and m<strong>in</strong>eralisation<br />

event was between 450 and 550 8C and <strong>the</strong><br />

pressure between 1.5 and 3.5 kbar. The key features of<br />

<strong>the</strong> <strong>Kolari</strong> <strong>deposits</strong> that should be considered <strong>in</strong> any<br />

discussion are:<br />

1. Metal association Fe–Cu–Au FAg, Bi, Ba, Co, Mo,<br />

Sb, Se, Te, Th, U, LREE.<br />

2. Alteration styles that, <strong>in</strong> <strong>deposits</strong> where structural<br />

control is most evident, def<strong>in</strong>e a zon<strong>in</strong>g where <strong>the</strong><br />

distal zones are characterised with NaFCl, K-rich<br />

assemblages, proximal zones with Ca–FeFCO 2 -<br />

rich assemblages, and <strong>the</strong> core that is probably related<br />

to a structural base, is iron rich. In <strong>deposits</strong><br />

where <strong>the</strong> zon<strong>in</strong>g is well developed, <strong>the</strong> Cu–Au<br />

m<strong>in</strong>eralisation is located <strong>in</strong> <strong>the</strong> ironstone core whereas<br />

<strong>in</strong> <strong>deposits</strong> with no clear zon<strong>in</strong>g Cu–Au m<strong>in</strong>eralisation<br />

is <strong>in</strong> <strong>the</strong> magnetite-dissem<strong>in</strong>ated albitite.<br />

3. Highly sal<strong>in</strong>e aqueous-carbonic fluids that circulated<br />

<strong>in</strong> <strong>the</strong> system dur<strong>in</strong>g <strong>the</strong> ma<strong>in</strong> alteration and m<strong>in</strong>eralisation<br />

event and, subsequently, dur<strong>in</strong>g a later<br />

brittle stage.<br />

4. The <strong>deposits</strong> are controlled by <strong>the</strong> structures of a<br />

major shear zone system ra<strong>the</strong>r than be<strong>in</strong>g bound to<br />

certa<strong>in</strong> stratigraphical position.<br />

Based on <strong>the</strong> data presented <strong>in</strong> this work, we suggest<br />

that an IOCG m<strong>in</strong>eralisation model would best expla<strong>in</strong><br />

<strong>the</strong> features of <strong>the</strong> <strong>Kolari</strong> Fe–Cu–Au <strong>deposits</strong>.


104<br />

T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105<br />

Acknowledgements<br />

We are grateful to Dr. Pasi Eilu, Dr. Juhani Ojala, Dr.<br />

Raimo Laht<strong>in</strong>en for <strong>the</strong> general assistance and constructive<br />

comments on <strong>the</strong> manuscript. We also wish to<br />

express our gratitude to <strong>the</strong> staff of <strong>the</strong> GSF isotope<br />

laboratory. Reviews by Dr. Hamid Mum<strong>in</strong> and Dr. P<strong>in</strong>g<br />

Dong are highly appreciated. Senior author wishes to<br />

thank <strong>the</strong> management of <strong>the</strong> Geological Survey of<br />

F<strong>in</strong>land <strong>in</strong> Espoo and Rovaniemi offices for responsiveness<br />

that allowed <strong>the</strong> f<strong>in</strong>ish<strong>in</strong>g of this work. Outokumpu<br />

Oyj Foundation and F<strong>in</strong>nish Graduate School <strong>in</strong> Geology<br />

provided <strong>the</strong> fund<strong>in</strong>g for <strong>the</strong> senior author’s work.<br />

References<br />

Barton, M.D., Johnson, D.A., 1996. Evaporitic-source model for<br />

igneous-related Fe <strong>oxide</strong>–(REE–Cu–Au–U) m<strong>in</strong>eralization. Geology<br />

24, 259–262.<br />

Barton, M.D., Johnson, D.A., 2000. Alternative br<strong>in</strong>e sources for Fe<strong>oxide</strong>(–Cu–Au)<br />

systems: implications for hydro<strong>the</strong>rmal alteration<br />

and metals. In: Porter, T.M. (Ed.), Hydro<strong>the</strong>rmal <strong>Iron</strong> Oxide<br />

Copper–Gold and Related Deposits: A Global Perspective. Australian<br />

M<strong>in</strong>eral Foundation, Adelaide, pp. 43–60.<br />

Bergman, S., Skiöld, T., 1998. Implications of ca 1.8 Ga metamorphic<br />

ages <strong>in</strong> <strong>the</strong> Pajala area, nor<strong>the</strong>rnmost Sweden. Proceed<strong>in</strong>gs, 23rd<br />

Nordic Geological W<strong>in</strong>ter Meet<strong>in</strong>g. Department of Earth Sciences,<br />

University of Aarhus, Denmark, p. 32.<br />

Ber<strong>the</strong>lsen, A., Marker, M., 1986. 1.9–1.8 Ga old strike-slip megashears<br />

<strong>in</strong> <strong>the</strong> Baltic shield, and <strong>the</strong>ir plate tectonic implications.<br />

Tectonophysics 128, 163–181.<br />

Bodnar, R.J., 1994. Syn<strong>the</strong>tic fluid <strong>in</strong>clusions: XII. The system H 2 O–<br />

NaCl. Experimental determ<strong>in</strong>ation of <strong>the</strong> halite liquidus and isochors<br />

for a 40 wt% NaCl solution. Geochimica et Cosmochimica<br />

Acta 58, 1053–1063.<br />

Brown, P.E., 1989. A microcomputer program for <strong>the</strong> reduction and<br />

<strong>in</strong>vestigation of fluid <strong>in</strong>clusion data. American M<strong>in</strong>eralogist 74,<br />

1390–1393.<br />

Cl<strong>in</strong>e, J.S., Bodnar, R.J., 1994. Experimental Determ<strong>in</strong>ation of <strong>the</strong><br />

PVTX Properties of 30 wt.% NaCl–H 2 O Us<strong>in</strong>g Syn<strong>the</strong>tic Fluid<br />

Inclusions. PACROFI V, Pan-American Conference on Research<br />

on Fluid Inclusions, Instituto de Investigaciones Eléctricas. Departamento<br />

de Geotermia, Cuernavaca, Morelos, Mexico, p. 12.<br />

Crawford, M.L., 1981. Phase equilibria <strong>in</strong> aqueous fluid <strong>in</strong>clusions.<br />

Hollister, L.S., Crawford, M.L. Short Course <strong>in</strong> Fluid Inclusions:<br />

Applications to Petrology: M<strong>in</strong>eralogical Association of Canada<br />

Short Course Handbook, vol. 6, pp. 75–100.<br />

E<strong>in</strong>audi, M.T., Me<strong>in</strong>ert, L.D., Newberry, R.J., 1981. Skarn <strong>deposits</strong>.<br />

Economic Geology, 75th Anniversary Issue, pp. 317–391.<br />

Frietsch, R., Billström, K., Perhdahl, J.A., 1995. Sulphur isotopes <strong>in</strong><br />

lower Proterozoic iron and sulphide ores <strong>in</strong> nor<strong>the</strong>rn Sweden.<br />

M<strong>in</strong>eralium Deposita 30, 275–284.<br />

Grant, J.A., 1986. The isocon diagram—a simple solution to Gresens’<br />

equation for metasomatic alteration. Economic Geology 81,<br />

1976–1982.<br />

Hanski, E., Perttunen, V., Sorjonen-Ward, P., 1997. Overview of <strong>the</strong><br />

geology of nor<strong>the</strong>rn F<strong>in</strong>land: geological survey of F<strong>in</strong>land. Guide<br />

43, 7–9.<br />

Hanski, E., Huhma, H., Vaasjoki, M., 2001. Geochronology of nor<strong>the</strong>rn<br />

F<strong>in</strong>land: a summary and discussion. In: Vaasjoki, M. (Ed.),<br />

Radiometric Age Determ<strong>in</strong>ations from F<strong>in</strong>nish Lapland and <strong>the</strong>ir<br />

Bear<strong>in</strong>g on <strong>the</strong> Tim<strong>in</strong>g of Precambrian Volcano-Sedimentary<br />

Sequences, Special Paper-Geological Survey of F<strong>in</strong>land, vol. 33,<br />

pp. 255–279.<br />

Hedenquist, J.W., Henley, R.W., 1985. The importance of CO 2 on<br />

freez<strong>in</strong>g po<strong>in</strong>t measurements of fluid <strong>in</strong>clusions; evidence from<br />

active geo<strong>the</strong>rmal systems and implications for epi<strong>the</strong>rmal ore<br />

deposition. Economic Geology 80, 1379–1406.<br />

Hiltunen, A., 1982. The Precambrian geology and skarn iron ores of<br />

<strong>the</strong> Rautuvaara area, nor<strong>the</strong>rn F<strong>in</strong>land. Bullet<strong>in</strong>-Geological Survey<br />

of F<strong>in</strong>land 318 (133 pp.).<br />

Hitzman, M.W., 2000. <strong>Iron</strong> <strong>oxide</strong>–Cu–Au <strong>deposits</strong>: what, where,<br />

when and why. In: Porter, T.M. (Ed.), Hydro<strong>the</strong>rmal <strong>Iron</strong> Oxide<br />

Copper–Gold and Related Deposits: A Global Perspective. Australian<br />

M<strong>in</strong>eral Foundation, Adelaide, pp. 9 – 25.<br />

Hitzman, M.W., Oreskes, N., E<strong>in</strong>audi, M.T., 1992. Geological characteristics<br />

and tectonic sett<strong>in</strong>g of Proterozoic iron <strong>oxide</strong> (Cu–U–<br />

Au–REE) <strong>deposits</strong>. Precambrian Research 58, 241–287.<br />

Hölttä, P., Väisänen, M., Väänänen, J., Mann<strong>in</strong>en, T., <strong>in</strong> press. Paleoproterozoic<br />

metamorphism and deformation <strong>in</strong> Central F<strong>in</strong>nish<br />

Lapland. Special Paper-Geological Survey of F<strong>in</strong>land.<br />

Ke<strong>in</strong>änen, V., 1995. Tutkimustyöselostus Kolar<strong>in</strong> kunnassa valtausalueella<br />

Kuervitikko 1 (Kaivosrekister<strong>in</strong>ro 5082/1) suoritetuista<br />

kulta-kuparimalmitutkimuksista vuos<strong>in</strong>a 1992–93. Geological<br />

Survey of F<strong>in</strong>land report M06/2714/-95/1/10. 2 pp. (<strong>in</strong> F<strong>in</strong>nish).<br />

K<strong>in</strong>nunen, A., 1980. Geolog<strong>in</strong>en kairausseloste, Laur<strong>in</strong>oja, Reikä no.<br />

170. Rautaruukki Oy <strong>in</strong>ternal report, 5 pp. (<strong>in</strong> F<strong>in</strong>nish).<br />

Korsman, K., Koist<strong>in</strong>en, T., Kohonen, J., Wennerström, M., Ekdahl,<br />

E., Honkamo, M., Idman, H., Pekkala, Y. (Eds.), 1997. Bedrock<br />

Map of F<strong>in</strong>land 1:1,000,000. Geological Survey of F<strong>in</strong>land,<br />

Espoo.<br />

Krogh, T.E., 1973. A low-contam<strong>in</strong>ation method for hydro<strong>the</strong>rmal<br />

decomposition of U and Pb for isotopic age determ<strong>in</strong>ations.<br />

Geochimica et Cosmochimica Acta 37, 485–494.<br />

Krogh, T.E., 1982. Improved accuracy of U–Pb zircon ages by <strong>the</strong><br />

creation of more concordant systems us<strong>in</strong>g an air abrasion technique.<br />

Geochimica et Cosmochimica Acta 46, 637–649.<br />

Laht<strong>in</strong>en, R., Nironen, M., Korja, A., 2003. Palaeoproterozoic orogenic<br />

evolution of <strong>the</strong> Fennoscandian Shield at 1.92–1.77 Ga<br />

with notes on <strong>the</strong> metallogeny of FeOx–Cu–Au, VMS, and<br />

orogenic <strong>gold</strong> <strong>deposits</strong>. In: Eliopoulos, D.G., et al., (Eds.), M<strong>in</strong>eral<br />

Exploration and Susta<strong>in</strong>able Development. Millpress, Rotterdam,<br />

pp. 1057–1060.<br />

Lehtonen, M.I., Airo, M.-L., Eilu, P., Hanski, E., Kortela<strong>in</strong>en, V.,<br />

Lanne, E., Mann<strong>in</strong>en, T., Rastas, P., Räsänen, J., Virransalo, P.,<br />

1998. The stratigraphy, petrology and geochemistry of <strong>the</strong> Kittilä<br />

greenstone area, nor<strong>the</strong>rn F<strong>in</strong>land. A Report of <strong>the</strong> Lapland Volcanite<br />

Project, Geological Survey of F<strong>in</strong>land, Report of Investigation,<br />

vol. 140, 144 pp. (<strong>in</strong> F<strong>in</strong>nish with summary <strong>in</strong> English).<br />

Ludwig, K.R., 1991. PbDat 1.21 for MS-DOS: a computer program<br />

for IBM-PC compatibles for process<strong>in</strong>g raw Pb–U–Th isotope<br />

data. Version 1.07. Open-File Report (United States Geological<br />

Survey) 88–542 (35 pp.).<br />

Ludwig, K.R., 2003. Isoplot/Ex 3. A geochronological toolkit for<br />

Microsoft Excel. Berkeley Geochronology Center Special Publication<br />

No. 4.<br />

Mäkelä, M., Tammenmaa, J., 1978. Sulfur isotope studies <strong>in</strong> F<strong>in</strong>nish<br />

Lapland, 1974–1976. Geological Survey of F<strong>in</strong>land Report of<br />

Investigation, vol. 24, 64 pp. (<strong>in</strong> F<strong>in</strong>nish with summary <strong>in</strong><br />

English).


T. Niiranen et al. / Ore Geology Reviews 30 (2007) 75–105 105<br />

Mark, G., Oliver, N.H.S., Williams, P.J., Valenta, R.K., Crookes,<br />

R.A., 2000. The evolution of <strong>the</strong> Ernest Henry Fe-Oxide–(Cu–<br />

Au) hydro<strong>the</strong>rmal system. In: Porter, T.M. (Ed.), Hydro<strong>the</strong>rmal<br />

<strong>Iron</strong> Oxide Copper–Gold and Related Deposits: A Global Perspective.<br />

Australian M<strong>in</strong>eral Foundation, Adelaide, pp. 123–136.<br />

McDonough, W.F., Sun, S.-S., 1995. The composition of <strong>the</strong> earth.<br />

Chemical Geology 120, 223–253.<br />

Mellquist, C., Öhlander, B., Weihed, P., Schöberg, H., 2003. Some<br />

aspects on <strong>the</strong> subdivision of <strong>the</strong> Haparanda and Jörn <strong>in</strong>trusive<br />

suites <strong>in</strong> nor<strong>the</strong>rn Sweden. GFF 125, 77–85.<br />

Mutanen, T., Huhma, H., 2001. U–Pb geochronology of <strong>the</strong> Koitela<strong>in</strong>en,<br />

Akanvaara and Keivitsa mafic layered <strong>in</strong>trusions and<br />

related rocks. In: Vaasjoki, M. (Ed.), Radiometric Age Determ<strong>in</strong>ations<br />

from F<strong>in</strong>nish Lapland and <strong>the</strong>ir Bear<strong>in</strong>g on <strong>the</strong> Tim<strong>in</strong>g of<br />

Precambrian Volcano-Sedimentary Sequences, Special Paper-<br />

Geological Survey of F<strong>in</strong>land, vol. 33, pp. 229–246.<br />

Oliver, N.H.S., Cleverley, J.S., Mark, G., Pollard, P.J., Fu, B., Marshall,<br />

L.J., Rubenach, M.J., Williams, P.J., Baker, T., 2004. Model<strong>in</strong>g<br />

<strong>the</strong> role of sodic alteration <strong>in</strong> <strong>the</strong> genesis of iron <strong>oxide</strong>–<br />

<strong>copper</strong>–<strong>gold</strong> <strong>deposits</strong>, Eastern Mount Isa Block, Australia. Economic<br />

Geology 99, 1145–1176.<br />

Perr<strong>in</strong>g, C.S., Pollard, P.J., Dong, G., Nunn, A.J., Blake, K.L., 2000.<br />

The Lightn<strong>in</strong>g Creek Sill Complex, Cloncurry District, Northwestern<br />

Queensland: a source of fluids for <strong>the</strong> Fe <strong>oxide</strong> Cu–Au<br />

m<strong>in</strong>eralization and sodic–calcic alteration. Economic Geology 95,<br />

1067–1089.<br />

Polland, P.J., 2000. Evidence of a magmatic fluid and metal source for<br />

Fe-<strong>oxide</strong> Cu-Au m<strong>in</strong>eralization. In: Porter, T.M. (Ed.), Hydro<strong>the</strong>rmal<br />

<strong>Iron</strong> Oxide Copper-Gold & Related Deposits: A Global<br />

Perspective, Australian M<strong>in</strong>eral Foundation, Adelaide, pp. 27–41.<br />

Pollard, P.J., 2001. Sodic (–calcic) alteration <strong>in</strong> Fe-<strong>oxide</strong>–Cu–Au<br />

districts: an orig<strong>in</strong> via unmix<strong>in</strong>g of magmatic H 2 O–CO 2 –<br />

NaClFCaCl–KCl fluids. M<strong>in</strong>eralium Deposita 36, 93–100.<br />

Puust<strong>in</strong>en, K., 2003. Suomen kaivosteollisuus ja m<strong>in</strong>eraalisten raakaa<strong>in</strong>eiden<br />

tuotanto vuos<strong>in</strong>a 1530–2001, historiall<strong>in</strong>en katsaus erityisesti<br />

tuotantolukujen valossa. Geological Survey of F<strong>in</strong>land<br />

Report M 10.1/2003/3, 578 pp. (<strong>in</strong> F<strong>in</strong>nish).<br />

Rastas, P., Huhma, H., Hanski, E., Lehtonen, M.I., Paakkola, J.,<br />

Mänttäri, I., Härkönen, I., 2001. U–Pb isotopic studies on <strong>the</strong><br />

Kittilä greenstone area, central Lapland, F<strong>in</strong>land. In: Vaasjoki, M.<br />

(Ed.), Radiometric Age Determ<strong>in</strong>ations from F<strong>in</strong>nish Lapland and<br />

<strong>the</strong>ir Bear<strong>in</strong>g on <strong>the</strong> Tim<strong>in</strong>g of Precambrian Volcano-Sedimentary<br />

Sequences, Special Paper-Geological Survey of F<strong>in</strong>land, vol. 33,<br />

pp. 95–141.<br />

Schmidt, C., Bodnar, R.J., 1994. Two-phase <strong>region</strong> and l<strong>in</strong>es of<br />

constant homogenization temperature for H 2 O–40 wt.% NaCl–5<br />

mol% CO 2 . PACROFI V, Pan-American Conference on Research<br />

on Fluid Inclusions. Instituto de Investigaciones Eléctricas, Departamento<br />

de Geotermia, Cuernavaca, Morelos, Mexico, p. 92.<br />

Sorjonen-Ward, P., Nurmi, P., Härkönen, I., Pankka, H., 1992. Epigenetic<br />

<strong>gold</strong> m<strong>in</strong>eralization and tectonic evolution of a lower<br />

Proterozoic greenstone terra<strong>in</strong> <strong>in</strong> <strong>the</strong> nor<strong>the</strong>rn Fennoscandian<br />

(Baltic) Shield. In: Sarkar, S.C. (Ed.), Metallogeny Related to<br />

Tectonics of <strong>the</strong> Proterozoic Mobile Belts. Oxford and IBH<br />

Publish<strong>in</strong>g, New Delhi, pp. 37–52.<br />

Sorjonen-Ward, P., Ojala, V.J., Airo, M.-L., 2003. Structural modell<strong>in</strong>g<br />

and magmatic expression of hydro<strong>the</strong>rmal alteration <strong>in</strong><br />

<strong>the</strong> Paleoproterozoic Lapland greenstone belt, nor<strong>the</strong>rn Fennoscandian<br />

Shield. In: Eliopoulos, D.G., et al., (Eds.), M<strong>in</strong>eral<br />

Exploration and Susta<strong>in</strong>able Development. Millpress, Rotterdam,<br />

pp. 1107–1110.<br />

Stacey, J.S., Kramers, J.D., 1975. Approximation of terrestrial lead<br />

isotope evolution by a two-stage model. Earth and Planetary<br />

Science Letters 26, 207–221.<br />

Väänänen, J., Lehtonen, M., 2001. Isotopic age determ<strong>in</strong>ations from<br />

<strong>the</strong> <strong>Kolari</strong>-Muonio area, western F<strong>in</strong>nish Lapland. In: Vaasjoki,<br />

M. (Ed.), Radiometric Age Determ<strong>in</strong>ations from F<strong>in</strong>nish Lapland<br />

and <strong>the</strong>ir Bear<strong>in</strong>g on <strong>the</strong> Tim<strong>in</strong>g of Precambrian Volcano-Sedimentary<br />

Sequences, Special Paper-Geological Survey of F<strong>in</strong>land,<br />

vol. 33, pp. 85–93.<br />

Väisänen, M., 2002. Structural features <strong>in</strong> <strong>the</strong> Central Lapland greenstone<br />

belt, nor<strong>the</strong>rn F<strong>in</strong>land. Geological Survey of F<strong>in</strong>land Report<br />

M21.42/2002/3, 20 pp.<br />

Wang, S., Williams, P.J., 2001. Geochemistry and orig<strong>in</strong> of Proterozoic<br />

skarns at <strong>the</strong> Mount Elliott Cu–Au(–Co–Ni) deposit, Cloncurry<br />

district, NW Queensland, Australia. M<strong>in</strong>eralium Deposita<br />

36, 109–124.<br />

Ward, P., Härkönen, I., Pankka, H.S., 1989. Structural studies <strong>in</strong> <strong>the</strong><br />

Lapland greenstone belt, nor<strong>the</strong>rn F<strong>in</strong>land and <strong>the</strong>ir application to<br />

<strong>gold</strong> m<strong>in</strong>eralization. Special Paper-Geological Survey of F<strong>in</strong>land<br />

10, 71–78.<br />

Williams, P.J., 1994. <strong>Iron</strong> mobility dur<strong>in</strong>g synmetamorphic alteration<br />

<strong>in</strong> <strong>the</strong> Selwyn Range area, NW Queensland: implications for <strong>the</strong><br />

orig<strong>in</strong> of ironstone-hosted Au–Cu <strong>deposits</strong>. M<strong>in</strong>eralium Deposita<br />

29, 250–260.<br />

Williams, P.J., Pollard, P.J., 2003. Australian Proterozoic iron <strong>oxide</strong>–<br />

Cu–Au <strong>deposits</strong>: an overview with new metallogenic and exploration<br />

data from <strong>the</strong> Cloncurry district, northwest Queensland.<br />

Exploration and M<strong>in</strong><strong>in</strong>g Geology 10, 91–213.<br />

W<strong>in</strong>chester, J.A., Floyd, P.A., 1977. Geochemical discrim<strong>in</strong>ation of<br />

different magma series and <strong>the</strong>ir differentiation products us<strong>in</strong>g<br />

immobile elements. Chemical Geology 20, 325–343.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!