Sei sulla pagina 1di 28

2008 Society of Economic Geologists, Inc. Economic Geology, v. 103, pp.

241267

The Cu Stockwork and Massive Sulfide Ore of the Feitais Volcanic-Hosted Massive Sulfide Deposit, Aljustrel, Iberian Pyrite Belt, Portugal: A Mineralogical, Fluid Inclusion, and Isotopic Investigation
CARLOS M. C. INVERNO,
Servios Geolgicos, Instituto Nacional de Engenharia, Tecnologia e Inovao, Apartado 7586, 2720 Alfragide (Lisboa), Portugal, and Centro de Recursos Minerais, Mineralogia e Cristalografia (CREMINER), Faculdade de Cincias, Universidade de Lisboa, Edifcio C6, 1749-016 Lisboa, Portugal

MICHAEL SOLOMON,
Centre for Ore Deposit Research (CODES), University of Tasmania, GPO Private Bag 79, Hobart, Tasmania 7001, Australia

MARK D. BARTON,
Department of Geosciences, Center for Mineral Resources, University of Arizona, Tucson, Arizona 85721
AND JOHN

FODEN

Department of Geology, University of Adelaide, Adelaide, South Australia 5005, Australia

Abstract
The Variscan Feitais volcanic-hosted massive sulfide deposit in the Aljustrel district of the Iberian Pyrite Belt consists of 55 million metric tons of Zn-Pb-Cu massive sulfide overlying a Cu-rich stockwork. The massive ore is overlain by up to 30 m of feldspar-phyric, rhyolitic volcaniclastic rock and locally by a jasper and/or chert layer up to 15 m thick. The massive sulfide orebody consists dominantly of pyrite, sphalerite, galena, chalcopyrite, tetrahedrite-tennantite, arsenopyrite, and bournonite, together with minor quartz, chlorite, sericite, carbonate, and barite. The orebody is up to 100 m thick and is underlain by a tabular alteration zone of chloritedominated, locally silicified, felsic volcanic rock, the upper 30 to 60 m of which contains chalcopyrite-quartzchlorite-sericite-carbonatebearing stockwork vein(let)s that prior to deformation were at a shallow angle to the base of the massive orebody. Chloritized footwall rocks extend up to 20 m below the Cu stockwork zone and are underlain by up to 50 m of quartz-sericite-pyritealtered rhyolitic rock. The stockwork veins also contain pyrite, tetrahedrite-tennantite, sphalerite, and arsenopyrite. Pyrite, both in stockwork and massive ore, locally displays partly recrystallized framboidal, reniform, and cellular textures. Two generations of quartz, Q1 and Q2, and carbonate in the stockwork veins contain primary (in growth zones) and pseudosecondary fluid inclusions, with homogenization temperatures of 270 to 315C and salinities of 2.2 to 8.1 wt percent NaCl equiv. The 34S(CDT) values of massive and stockwork ores range from 15.4 to +4.7 (mean, 2.8) and 11.2 to +11.9 (mean, 0.4) per mil, respectively, the lowest values from colloformtextured pyrite. With no evidence of oxidation of sulfide sulfur during mineralization, the most negative values indicate an origin by biogenic reduction of seawater sulfate. The 13C(PDB) values for carbonates, 7.5 to 13.7 and +9.3 to 14.3 per mil in massive and stockwork ore, respectively, indicate an origin mostly by oxidation of methane derived from organic matter in underlying sedimentary rocks and possibly a contribution of magmatic carbon. There are no significant lateral or vertical variations in S isotope values in sulfides or C-O isotope values in carbonates, either in massive or stockwork ore. The 18O(SMOW) values for quartz in stockwork and massive sulfide are 11.6 to 13.9 and 16.7 to 17.9 per mil, respectively. Coexisting, and texturally contemporaneous, carbonate and quartz in stockwork veins are not in isotopic equilibrium, indicating that the C-O isotope values may have been reset. The 18O values of fluid calculated to be in equilibrium with quartz at fluid inclusion homogenization temperatures are 4.2 to 5.2 per mil. Barite from the hanging wall and massive ore yields 34S values (21.927.9) equal to or slightly higher than those of coeval seawater; 87Sr/86Sr ratios (0.708438 0.709063) are slightly more radiogenic than those of coeval seawater (0.70800.7085), and much more radiogenic than those of coeval volcanic rocks (0.7033040.706642), probably representing mixtures between seawater Sr and radiogenic Sr in fluids sourced in the crustal pile. Deposition of the massive sulfide on the sea floor is suggested by its stratiform nature, the stronger alteration of footwall relative to hanging-wall rocks, the stockwork system terminating sharply at the base of the massive sulfide, the presence of sedimentary-like textures in the massive sulfide, the absence of replacement fronts, and the presence of framboidal and other sea-floor depositional textures indicative of fluid quenching. The sheetlike form, lack of rubble mounds and chimneys, scarcity of barite, reduced mineral assemblage, and metal zoning distinguish Feitais from Kuroko-type deposits. It shares most of the characteristics of those Iberian Pyrite Belt deposits for which a brine-pool origin has been proposed based on fluid inclusion data,
Corresponding

author: e-mail, carlos.inverno@ineti.pt

0361-0128/08/3725/241-27

241

242

INVERNO ET AL.

suggesting a similar depositional origin, although the evidence from fluid inclusions in this study is equivocal. The sulfate that underwent biogenic reduction may have been derived from mixing with seawater during early filling of the brine pool; diffusion across the brine-seawater interface; and sulfate reduction in the footwall volcaniclastic rocks. Stable and radiogenic isotope compositions of sulfates, sulfides, and carbonates suggest involvement of modified seawater and crustal fluids convecting due to magmatic heating, but the calculated high fluid pressures in the stockwork may indicate the additional involvement of magmatic fluids.

Introduction THE ALJUSTREL DISTRICT, in the southern Portuguese part of the Iberian Pyrite Belt (Fig. 1), contains several large volcanic-hosted massive sulfide (VHMS) deposits, with known resources of massive ore totalling 190 million metric tons (Mt) grading 45 to 47 wt percent S, 38 to 41 wt percent Fe, 3.4 wt percent Zn, 1.2 wt percent Pb, 0.8 wt percent Cu, 36 g/t Ag, and 1 g/t Au. An estimated 50 Mt has been mined or left in mining pillars (Carvalho et al., 1976; Barriga, 1983; Leito, 1997). In addition, stockwork ores also occur in three of the district orebodies (Feitais, Moinho, and Algares, most significantly in the first two). A recent drilling program, conducted by Eurozinc Mining Corp. (now Lundin Mining Corp.), outlined within the 54.5-Mt resources of one of its orebodies, Feitais, 18.4 Mt of massive zinc ore grading 6.0 wt percent Zn, 1.8 wt percent Pb, 0.2 wt percent Cu, 67 g/t Ag, and 0. 73 g/t Au (at 4.5 wt % Zn cutoff; Barrett et al., 2008), 4.1 Mt of massive copper ore grading 2.1 wt percent Cu, 0.7

wt percent Zn, 0.2 wt percent Pb, 12 g/t Ag, and 0.17 g/t Au (at 1.5 wt % Cu cutoff), and 2.6 Mt of stockwork copper ore grading 2.1 wt percent Cu, 1.0 wt percent Zn, 0.3 wt percent Pb, 14 g/t Ag, and 0.37 g/t Au (Dawson et al., 2001a, b; Barrett et al., 2008). In this paper we present the results of a metallogenic study, involving geology, mineralogy, fluid inclusion, and stable and radiogenic isotope data, centered on the Feitais copper stockwork and partially extending to the massive ores. Field work included logging of drill cores (Feitais drilling sections 160, 240, 320, 520, 560, 600, 640, and 800 ft) and also mapping, at scales of 1:50 and 1:100, of the scarce underground exposures of the footwall on levels 140, 190, and 200 m. Previous Work The geology of the Aljustrel area and its VHMS deposits was described by Schermerhorn and Stanton (1969), Andrade and Schermerhorn (1971), Carvalho et al. (1976) and Schermerhorn et al. (1987). Several comprehensive ore microscopy,

Lagoa Salgada

8 OO'

7 OO'
OC

Gra ndola

SA

SP AI

BEJA

ZONE

PORTUGAL

RE

AT LA NT IC

MO

EA

OS

FRAN

CE

NA

MADRID

IN SPA

Sines

Lousal

LISBOA

ALJUSTREL
Gavio Aguas Teidas

38 OO'

Cercal
Salgadinho

San Telmo

Concepcin Rio Tinto

Castro Verde
S. Domingos

Romanera La Zarza

Ourique

Mrtola
Tharsis Neves-Corvo Sotiel

Migollas Los Frailes

Pomaro Almodo var


Valverde Aznalcollar Las Cruces

SEVILLA

Monchique
HUELVA Castro Marim

VMS DEPOSITS
FARO

NEW VMS DISCOVERIES (1985-1995) MONCHIQUE MASSIF (76 Ma)

MESOZOIC AND CENOZOIC NAMURIAN - L. WESTPHALIAN NAMURIAN VISEAN U. DEVONIAN - L. VISEAN U. DEVONIAN AND OLDER

SEDIMENTARY ROCKS BREJEIRA FM. MIRA FM. MERTOLA FM. FLYSCH GROUP

37 OO'
GRANITIC ROCKS HERCYNIAN BEJA-ACEBUCHES BASIC-ULTRABASIC COMPLEX CARRAPATEIRA GROUP (U. DEVONIAN - L. NAMURIAN) VOLCANIC SILICEOUS COMPLEX FERREIRA-FICALHO GROUP (Mainly U. DEVONIAN) PHYLLITE-QUARTZITE GROUP; PULO DO LOBO GROUP ( )

20 Km

FIG. 1. Regional geology map of the Iberian Pyrite Belt, showing the location of Aljustrel (modified after Carvalho et al., 1999). 0361-0128/98/000/000-00 $6.00

242

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

243

textural, and beneficiation studies have been conducted on the Aljustrel massive ores, including those of Feitais (e.g., Gaspar and Conde, 1978; Gaspar, 1984, 1996, 1997). Barrigas (1983) study of Feitais involving transmitted light petrography, whole-rock geochemistry, and O isotope analysis of the hydrothermally altered rocks and ores, including stockwork and massive ores and electron microprobe mineral analysis, concluded that the ore-forming fluid was seawater that had undergone deep convective circulation in older rocks. The same isotope data were used by Barriga and Kerrich (1984) and Munh et al. (1986), the latter authors also including new D/H data on stockwork chlorite, to reach the same conclusion. These authors calculated from oxygen isotope equilibrium fractionations, temperatures of 250 to 270 and 200 to 215C for stockwork ore and 200 to 220 and 160 to 175C for massive ore, respectively. Barriga and Fyfe (1988, 1991) and Carvalho et al. (1999), following Barriga (1983), proposed that the massive sulfide orebody formed in open space under an older, overlying, jasper and/or chert layer, which acted as a seal during the cooling of the ore-forming fluids. Hutchinson (1991) disputed this conclusion based on the presence of fragmental ore within the massive sulfide orebody. Barriga and Fyfe (1998) studied the metasomatism of Aljustrel volcaniclastic rocks not affected by ore zone hydrothermal alteration, concluding that they were altered by seawater that, with increasing depth, changed from a cold, oxidizing, Mg-bearing fluid to a hot, reducing, Mg-free, and Si-Fe-Cu-Znrich fluid, with water/rock ratios (w/r) in the whole area progressively decreasing from very high to low values. The probable amount of Cu and Zn leached from volcaniclastic rocks was shown to be compatible with the Cu and Zn metal contents in the massive sulfide body. Barriga (1983) and Carvalho and Barriga (2000) recognized a sericitic alteration zone underlying the footwall chloritic alteration zone. Dawson et al. (2001a, b) presented the main results of a Eurozinc drilling program, with new resources for massive Zn ore, massive Cu ore and stockwork Cu ore, and proposed that a volumetrically important percentage of the rhyolites in the area are intrusive and not of volcaniclastic nature, as suggested by previous authors. More recent information is contained in Barrett et al. (2008) that also presents a lithogeochemical study of the Aljustrel volcanic rocks and proposes that the massive sulfides and the overlying volcanic rocks and chert formed in a growth fault-controlled basin. Regional Geology The Iberian Pyrite Belt, located in the southwest of the Iberian Peninsula (Fig. 1), with an estimated pre-erosional 1.7 billion tons of sulfides, is the largest (in tonnage) VHMS province in the world (Carvalho et al., 1999). It is 250 km long and 30 to 60 km wide. The stratigraphic sequence in the belt has been divided into (1) the Late Fammenian (Late Devonian) Phyllite-Quartzite Group at the base, (2) the Late Fammenian to earliest Late Visean intermediate Volcanic Siliceous Complex, consisting of magmatic and sedimentary rocks that host the massive sulfide deposits, and (3) at the top, the Late Visean to Early Westphalian (Carboniferous) (Baixo Alentejo) Flysch Group (Carvalho et al., 1971, 1976, 1999; Oliveira, 1990; Oliveira et al., 2004).
0361-0128/98/000/000-00 $6.00

During the 1970s most of the models for the geotectonic setting of the Iberian Pyrite Belt involved various plate tectonic scenarios (reviewed in Silva et al., 1990, and Carvalho et al., 1999), and in the 1980s the role of oblique, Variscan continent-continent collision and related sinistral transcurrent movements became important (Badham, 1982; Brun and Burg, 1982). From about 1990 all models envisaged the Iberian Pyrite Belt as part of the South Portuguese zone subducting northward beneath the Iberian massif, part of the Gondwana continent (Silva et al., 1990; Fonseca and Ribeiro, 1993; Quesada et al., 1994; Dias and Ribeiro, 1995). Fonseca and Ribeiro (1993) and Quesada et al. (1994) found evidence of Early to Middle Devonian northward obduction of remnant oceanic material (the Beja-Acebuches Ophiolite Complex) during early phases of oblique continent-continent collision, followed by Devonian-Carboniferous sinistral faulting and southward thrusting and folding, and associated flysch sedimentation (e.g., Quesada, 1998; Onzime et al., 2003). Transpression during oblique continental collision resulted in the formation within the South Portuguese zone of extensional, pull-apart basins in which magmatism, massive sulfide mineralization, and marine sedimentation occurred (Munh, 1983; Dias and Ribeiro, 1995; Mitjavila et al., 1997; Oliveira and Quesada, 1998; Fonseca et al., 1999; Rosa et al., 2004). Solomon and Quesada (2003) suggested that steep transcurrent faults were conduits for crust- and mantle-derived magmas and that their exsolved fluids were important components of the fluids responsible for mineralization. In the Aljustrel area, Phyllite-Quartzite Group rocks are absent but crop out in Corte Vicente Anes, 3 km northeast of Aljustrel. The lower part of the Volcanic Siliceous Complex sequence (Fig. 2), including that at Feitais, consists of submarine crystal and lapilli rhyolitic volcaniclastic rocks, and a few massive or autobrecciated lavas, with a total thickness of over 300 m in the central parts and less than 80 m laterally (Andrade and Schermerhorn, 1971; Carvalho et al., 1976; Barriga, 1983; Dawson et al., 2001b). Six orebodies (Algares, S. Joo, Gavio, Moinho, Feitais, and Estao; Fig. 2) lie near the top of this sequence. A U-Pb zircon age of 352.9 1.9 Ma was obtained for the hydrothermally altered footwall rhyolite of the Algares orebody by Barrie et al. (2002). This age is very close to that of the hanging-wall rock at Las Cruces, which is on the same horizon as Aznalcllar, dated by palinology of the host shales as Late Strunian (Pereira et al., 1996; Conde et al., 2003); a Late Strunian age has been assumed for Feitais in discyssing isotope studies. Sedimentary rocks, mainly black or siliceous shales, are interbedded with the volcaniclastic rocks in places (Andrade and Schermerhorn, 1971; Leito, 1997). T.J. Barrett and W.H. MacLean (unpub. report for Eurozinc Mining Corp., 1999, 65 p.), Dawson et al. (2001b), and Barrett et al. (2008) indicated the existence of intrusive rhyolites in this lower part of the Aljustrel Volcanic Siliceous Complex sequence, but our interpretation is that the field evidence remains inconclusive. The ore horizon at Feitais is overlain by up to 30 m of feldspar-phyric, rhyolitic volcaniclastic rocks (Barrett et al., 2008) that are themselves overlain by a 0- to 15-m-thick, hematitic to manganiferous jasper and/or chert layer that marks the base of the Paraso Formation (Fig. 3; Barriga and Kerrich, 1984). This layer directly caps the massive sulfide in

243

244

INVERNO ET AL.

8o 12'

8o 10'
S S

8o 08'

S S

37 o 53'

Estao S. Joo
Roldana

Moinho

Gavio

Sto. Anto Feitais

ALJUSTREL
0
VMS deposits (not in production)
PLEISTOCENE

1000 m

Algares
37o 52'

Clastic Sedimentary Cover Rocks


MIO-PLIOCENE MID-JURASSIC UPPER VISEAN

Dolerite Dike Flysch Group - Mrtola Fm Siliceous Phyllite, Volcanic Sandstone and Shale

TOURNAISIAN-LOWER VISEAN
S S

Jasper Rhyolitic Volcaniclastic Rock (hosting Ore on top); Lateral Sedimentary Rock ( )

VOLCANIC SILICEOUS COMPLEX Road Fault Thrust fault

FIG. 2. Geology map of Aljustrel (modified after Andrade and Schermerhorn, 1971; Sto. Anto area after Leito, 1992).

SW 0 level

NE

A'

FLYSCH GROUP
Late Visean
Mrtola Fm: graywacke and shale

LC13 (124 m) 100 level SEA LEVEL

F063 (50 m)

L064 (132 m)

VOLCANIC SILICEOUS COMPLEX


Tournaisian to early Visean
Paraso Fm: siliceous shale, phyllite, volcanic sandstone and shale Jasper / chert Feldspar-phyric rhyolitic volcaniclastic rock ( hanging wall ) Massive sulfide orebody Rhyolitic volcaniclastic rock (footwall) Stockwork Fault Drill-hole trace
F003 (52 m)

200 level F003 (52 m) F014 (59 m) FS9803 (365 m) F015 (99 m) 300 level F016 (126 m) F065A (177 m)

FS9944 (389 m) 400 level FS9942 (422 m) FS11 (506 m) FS9802 (492 m) 500 level FS21 (574 m) FS13 (520 m)

600 level

198,400E 100,600N

100 m

FS22 (642 m)

FIG. 3. Feitais drilling section 560 ft Northwest. Position of profile A-A' is shown in Figure 4 (modified after Sousa, 1999; Eurozinc 1:1,000 scale section, 2000). 0361-0128/98/000/000-00 $6.00

244

198,700E 100,900N

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

245

some areas, but for the most part does not, being separated from it by up to 30 m of hydrothermally altered (chlorite, sericite, pyrite) rhyolitic, volcaniclastic rock (Fig. 3). In the vicinity of, and above, the orebody, red hematitic jasper grades into a gray, black, or dark brownish manganiferous chert, with disseminated carbonates, magnetite, and pyrite; local Mn oxide-rhodocrosite-rhodonite accumulations have been exploited in small open pits (Barriga and Fyfe, 1988). Chlorite is at least locally abundant in the layer, parts of which also contain intercalated altered feldspar-phyric, rhyolitic volcaniclastic rock; the layer is crosscut by abundant veins filled with quartz and later carbonate (-barite). Leistel et al. (1998) thought that the Iberian cherts in general were emplaced below the sea floor through chemical precipitation or replacement of a preexisting rock, but Barriga and Fyfe (1988) considered the jaspers and associated rocks in the Aljustrel area to be hydrothermal precipitates deposited on the sea floor. The jasper and/or chert layer at Feitais is locally overlain by basaltic lavas and volcaniclastic rock, and volcanic sandstones and shales (Leito, 1997). The remainder of the Paraiso Formation, the top of the Volcanic Siliceous Complex sequence, consists of 50 m of siliceous shales, volcanic sandstones and shales, and minor, thin, rhyolitic volcaniclastic rock layers in the lowermost part (Carvalho et al., 1976; Barrett et al., 2008). Members of the Paraso Formation above the jasper and/or chert layer have been hydrothermally altered, with significant development of chlorite, sericite, carbonate, and sulfides (Schermerhorn, 1978; Barriga, 1983). Carbonate-quartz (barite) veins occur in the feldspar-phyric, rhyolitic volcaniclastic rocks, the jasper and/or chert layer and at least the lowermost 20 m of the Paraso Formation rocks. Barriga (1983) noted the presence of cymrite, a hydrated barium silicate, in a barite-free vein cutting the jasper and/or chert layer. There are barite-rich layers, up to over a decimeter thick, in the feldspar-phyric, rhyolitic volcaniclastic hanging-wall rocks. The Tournaisian to early Visean Volcanic Siliceous Complex sequence (Schermerhorn et al., 1987; Barrie et al., 2002) is conformably overlain by and in gradational contact with the Mrtola Formation of the Flysch Group, consisting of several hundred meters of a turbiditic sequence of graywackes and shales, of late Visean age, but the Volcanic Siliceous Complex rocks have been thrust over the flysch along much of the contact (Andrade and Schermerhorn, 1971; Leito, 1997). The whole sequence was deformed during the Hercynian orogeny, with a pre-late Westphalian main deformation phase (Carvalho et al., 1999), involving development of northwestsoutheast folds with steep regional slaty cleavage dipping to the northeast, as well as overthrusts. The Aljustrel orebodies lie in the steep limbs of four asymmetrical anticlines that distribute the orebodies into three lineaments (Fig. 2), from northeast to southwest, Feitais-Estao, Algares-Moinho-S. Joo-Gavio (Schermerhorn et al., 1987), and Gavio SW (Conde and Leito, 1984; Conde et al., 1986). S. Joo and Gavio, and Feitais and Estao, are displaced relative to each other by an offset of 3 and 0.5 km, respectively, along northeast-southweststriking (Messejana) and north-northeast-south-southweststriking (Represa) faults, respectively. Each pair of sulfide lenses was initially a single orebody, reaching 1.7 km along strike in the case of Feitais-Estao (Andrade and Schermerhorn, 1971; Dawson et al., 2001b).
0361-0128/98/000/000-00 $6.00

Dawson et al. (2001a, b) and Barrett et al. (2008) suggested that the Feitais orebody formed in a restricted basin within a major basin, adjacent to a growth fault that focused the fluids responsible for alteration and veining in the footwall. Feitais Orebody and Footwall The Feitais massive sulfide orebody, in the northeast flank of the Feitais anticline, is 750 to 1,000 m long (remains open), 500 m wide, and 60 m thick (max 100 m; Figs. 3, 4). It has a Cu-(Bi-Sb)enriched base of medium- to coarse-grained homogeneous to banded massive ore, an intermediate, lowgrade, pyritic zone (fine-grained massive ore), and an upper Zn-Pbrich fine-grained banded ore zone (Conde and Leito, 1984; Gaspar, 1996, 1997; Dawson et al., 2001a; Barrett et al., 2008). The massive ore consists predominantly of pyrite, sphalerite, galena, chalcopyrite, tetrahedrite-tennantite, arsenopyrite, and bournonite, with minor pyrrhotite, jamesonite, cassiterite, stannite, meneghinite, marcasite, cobaltite, native bismuth, gudmundite, kobellite, magnetite, and hematite as metallic minerals (Gaspar and Conde, 1978; Gaspar, 1996). Where the stratigraphic base of the massive ore is underlain by a stockwork zone, sphalerite contains abundant chalcopyrite inclusions (chalcopyrite disease). The highest Bi contents, both in Feitais and Moinho, are in the base of the massive sulfide orebody, where native bismuth occurs locally in the latter (Conde and Leito, 1984; Gaspar, 1997). The massive ore gangue (a few percent in the Moinho orebody: Gaspar, 1997) comprises quartz, chlorite, sericite, carbonate, and minor barite (Barriga, 1983; Gaspar, 1996). Barite is more common near the hanging wall and in the periphery of the orebody (Barriga, 1983; Mitsuno et al., 1988) but overall is scarce within massive ore. It also occurs in carbonate-quartz-barite veins that cut the orebody. The massive ore has been recrystallized, consisting mostly of an aggregate of idiomorphic, fine-grained (rarely mediumgrained) pyrite (Barriga, 1983). The massive ore in places displays mineral layering consisting of sphalerite-rich layers in pyrite and also bands of different grain size or texture in pyrite; in places, the bands are parallel to slaty cleavage and may be due to deformation, but Barriga (1983) and Gaspar (1996) believed the layering to be in most cases primary. Structures such as graded bedding, convolute lamination, and small slumps are not abundant in massive ore according to Barriga (1983) but are common in its stratigraphic lower portion according to Gaspar (1996, 1997). Framboidal, as well as cellular and zoned botryoidal (colloform) pyrite textures observed in the present study are shown in Figure 5AD. The massive ore, in places, also grades locally into fragmental, microconglomeratic ore, commonly with pyrite (or rarely quartz) fragments up to 1 cm wide in a pyrite-rich, or locally quartz- or carbonate-rich matrix. Domains of altered rhyolitic volcaniclastic rock up to 2 m long are found locally within massive sulfide ore immediately adjacent to the stratigraphic footwall, e.g., in the 190-m level (depth relative to a reference level 0 m that is 150 m above sea level; surface is 065 m above level 0 m). The position and attitude of the stope faces containing those domains suggests that such faces may represent the intersection with irregular terminations of the massive sulfide orebody, and are not evidence for replacement, as claimed by Barriga (1983), although, in a few

245

246

INVERNO ET AL.

10

00

Ft

0 96

N
0 92 88 0 84 0 80 0 76 0 72 0 0 68 0 64 60

FS9801 FS17
101,000

FS22 FS9975 FS13 FS9951 / FS9956 / FS9949 / FS9953 FS9947 / FS9950 FS11

0 56 0 0 52 48 0 44 0 40 0 36 0 0 32 28 0 0 24 20 0 0 16 12 0 80

FS21

A'

FS9802 FS9803 FS9942 / FS99444 FS9963 FS9932 / FS9933

Repres

a fault

LC3 FS9934

40 0

Ft

100,500

FS9939

FS9822 Surface projection of massive sulfide orebody at 200 level Fault Fault (approx. position) Surface drill holes mentioned in text
198,000 198,500

200 m

100,000

FIG. 4. Plan of Feitais deposit, showing the drill sections and surface projection of the massive sulfide body outlined on the 200-m level. Line A-A' denotes position of profile shown in Figure 3 (after Eurozinc maps, 1999).

other cases reported in that study, there is evidence of minor, local replacement. Most of the massive sulfide orebody overlies a broadly tabular zone of altered volcaniclastic rock, the upper part of which, 30 m to locally 60 m thick (Fig. 3), contains chalcopyrite-bearing stockwork quartz veins and veinlets. In places this stockwork extends for some tens of meters beyond the limits of the massive sulfide orebody, but particularly at its southwest margin and in the northwest and southeast ends (Fig. 4), the orebody is not underlain by a stockwork zone; in these portions, up to 220 m in extent, the orebody commonly has neither a Cu-rich stratigraphic base nor chalcopyrite disease in sphalerite. The stockwork veins lie mostly within strongly chloritized rhyolitic volcaniclastic rock, and this altered rock extends downward up to an additional 20 m without veins, then grades into a dominantly sericitized, pyrite-bearing, rhyolitic volcaniclastic rock, which continues for at least 50 m below the chlorite-altered zone. Carvalho and Barriga (2000) recognized both the sericitic- and chloritic-altered zones in the Feitais orebody footwall. A paragonite-quartz-sulfide assemblage (Relvas et al., 1990; Massano et al., 1991) in the outermost
0361-0128/98/000/000-00 $6.00

stockwork rocks of the Gavio orebody footwall was not recognized in Feitais by those authors or in our study. In places the chloritized volcaniclastic rock within the Feitais stockwork grades into a medium- to pale-gray, chloritized, partly silicified, or even completely silicified, whitish rhyolitic rock. In the uppermost stockwork, local silicification also produced a white to buff silicified and sericitized rhyolitic volcaniclastic rock, with millimeter- to centimeter-scale bands and intervening anastomosing stringers of pyrite (chalcopyrite; Fig. 5E, F). Patches of this rock are also locally present within the underlying chloritized stockwork. Dominant sericitic alteration, and minor chloritic alteration (more intense in association with the jasper and/or chert layer) occurred in the feldspar-phyric, rhyolitic volcaniclastic hanging-wall rocks up to the lowermost 10 to 20 m of the Paraso Formation shales, with enrichment in Fe, Mn, and Ba (Barriga, 1983) and development of quartz-carbonate (calcite, Fe dolomite) veins. These hydrothermal alteration assemblages overprinted a lower temperature regional alteration event that converted basalts into spilites and rhyolites into quartz keratophyres. This alteration is thought to result from the interaction between seawater and the volcanic rocks (Munh and Kerrich,

246

199,000

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

247

1980) and was followed by prehnite-pumpellyite (to lower greenschist) facies regional metamorphism during the Hercynian orogeny (Munh et al., 1986; Munh, 1990). Chlorite and white mica compositions Hydrothermal choritization altered the pale gray porphyritic (albite phenocrysts) rhyolitic volcaniclastic rock (Andrade and Schermerhorn, 1971) to a medium to dark graygreen fine-grained (<100 m) rock, consisting of quartz and interstitial chlorite (i.e., a chloritite). This chlorite is commonly green, with shades of either gray or blue in hand specimen and pale to medium green in plane-polarized light under the optical microscope, and with gray or bluish interference colors (referred to hereafter as chlorite I). Not uncommonly, ghosts of former millimeter-sized phenocrysts, fully replaced by both minerals, are recognizable (Fig. 5G). However, the most distinctive footwall chlorite, generally accompanying chalcopyrite in stockwork veins, is massive, medium green (to rare dark blue), has a grain diameter of 10 to 300 m and has almost no intergranular quartz (Fig. 5H); it is pale green under the microscope microscope in plane-polarized light, with anomalous blue birefringence colors. This is hereafter referred to as chlorite II. Type II chlorite also occurs in the jasper and/or chert layer and, locally, in the chloritized host-rock selvage to the stockwork veins, whereas both types occur as minor phases in host rock to the silicified and sericitized stockwork and in massive ore gangue. Within the stockwork, chlorite II in the veins and chlorite I in the chloritized volcanic rock are chemically indistinguishable (Fig. 6; Table 1), with 20 to 35 wt percent FeO and 6 to 16 wt percent MgO. Chlorites I and II of the silicified and sericitized rock of the uppermost stockwork have narrower ranges (3035 wt % FeO and 59 wt % MgO). Chlorite in this rock is also enriched in MnO (up to 1 wt %) relative to chlorite in the lower stockwork. There is a general increase in MgO in chlorites II and I downward within the stockwork, either in veins or altered host rock, indicating a more Mg-rich chlorite in the core, as Urabe et al. (1983) described for alteration pipes in several VHMS deposits. Chlorite compositions in the massive sulfide ore are more variable than in the stockwork, with chlorite II having 22 to 32 wt percent FeO and 7 to 10 wt percent MgO, and chlorite I having 20 to 25 wt percent FeO and 12 to 15 wt percent MgO (Fig. 6; Table 1). With the exception of data for sample FXI, adjacent to the massive ore footwall, there is a decrease in MgO in both chlorite I and II in massive ore toward the center of the massive sulfide body. In general, there is an upward increase in MnO in any chlorite type, from the lower stockwork to the massive sulfide body, as noted by Barriga (1983). Chlorite from altered rhyolitic volcaniclastic rock away from the orebody, at S. Joo quarry and Esteval da SerraMalhadinha, 2 km southeast of Feitais, contains 19 to 27 wt percent FeO and 13 to 19 wt percent MgO (Fig. 6), virtually indistinguishable from the above-mentioned chlorites. All these chlorites have relatively constant and similar Si and Al contents (Table 1) and plot in the triangular diagram for trioctahedral chlorites (Zane and Weiss, 1998) as dominant Fe chlorites and minor Mg chlorites, and away from Al chlorites (i.e., mostly magnesium-aluminian chamosite to rare magnesian chamosite and also minor iron-aluminian clinochlore to
0361-0128/98/000/000-00 $6.00

rare iron clinochlore, according to Bayliss, 1975, and Bailey, 1980). Using Cathelineaus (1988) geothermometer, the composition of chlorite II from the stockwork veins indicates temperatures of 340 to 404C, whereas Kranidiotis and MacLeans (1987) geothermometer, which also takes into account Fe and Mg, gives 320 to 380C. The same geothermometers applied to chlorite I from the host rock of the chloritized stockwork yield 296 to 417 and 304 to 385C, respectively. Footwall chlorites in other VHMS deposits of the Iberian Pyrite Belt, unlike those at Feitais, differ from regional chlorites in having variable Fe/(Fe + Mg) ratios and in some cases also variable Al/Si ratios. Deposits such as Concepcin, San Miguel, Aguas Teidas Este, and Cueva de la Mora have footwall chlorites with Fe/(Fe + Mg) ratios of 0.3 to 0.7, lower than those of regional chlorite (0.40.8, mostly in the higher range), and more or less constant Al/Si ratios throughout (Snchez-Espaa et al., 2000). In the Aznalcllar deposits, the footwall chlorite has a higher Fe/(Fe + Mg) ratio of 0.7, decreasing toward regional chlorites (0.350.20), and the footwall Al/Si ratios are also higher than regional values (Almodvar et al., 1998). The application of Cathelineaus (1988) geothermometer to the above-mentioned northern Iberian Pyrite Belt deposits yields formation temperatures of 220 to 380C for central stockwork chlorites (Snchez-Espaa et al., 2000), and Walshes (1986) geothermometer yields formation temperatures of 240 to 272C for Aznalcllar footwall chlorite (Almodvar et al., 1998). Electron microprobe analyses from this study show that the composition of white mica in massive ore and in the footwall rock at Feitais corresponds to that of muscovite (Deer et al., 1992), and X-ray diffraction of sericite from the Feitais ore zone shows this to be 2M1 muscovite (Carvalho and Barriga, 2000). Our analyses of white mica in Feitais massive ore and footwall rock yield 3.6 and 2.6 wt percent BaO, respectively, but Barriga (1983) found values up to 9.2 wt percent BaO in the massive ore white mica. Mineralized footwall veins Deformation of the vein network has oriented the veins predominantly parallel to the Hercynian cleavage, but in zones of less intense deformation the randomly oriented vein(lets) cut one another; prior to deformation, these veins were at a shallow angle to the base of the massive sulfide orebody. Most of the stockwork veins and veinlets range in width from a few millimeters up to 20 cm and are comprised of quartz carbonate chlorite, together with chalcopyrite, pyrite, tetrahedrite-tennantite, and sphalerite (Fig. 7). The same metallic minerals occur locally in host rocks up to a few centimeters away from the veins. Chalcopyrite grain size ranges from <10 m in disseminated occurrences up to 300 m, with anhedral grains locally constituting clusters up to 5 cm wide. Pyrite occurs as fine-grained (<50-m diam), subhedral grains forming aggregates up to 3.5 mm wide (pyrite I), followed by a later, more coarse-grained (up to 300-m diam), anhedral to euhedral pyrite II (Fig. 5F, H, I). Later stringers, or recrystallized, coarse, euhedral pyrite is described as pyrite III. Some pyrite II is finely intergrown within chalcopyrite aggregates locally reaching several centimeters in length. The vein chalcopyrite is generally associated with chlorite II. A

247

248

INVERNO ET AL.

D C

E
cpy

py

FIG. 5. Photomicrographs of polished thin sections from Feitais massive sulfide and stockwork ore samples. Sample locations are indicated in tables. A. Microconglomeratic massive sulfide ore, with fragment of pyrite with colloform texture in quartz (black) matrix. Sample F57, reflected light (RL). B. Microconglomeratic massive sulfide ore, with fragment of botryoidal pyrite, with an outer more recrystallized pyrite shell. Tetrahedrite-tennantite (tt) and chalcopyrite (cpy) are also present in quartz (black) matrix. Sample F57 (RL). C. Massive sulfide ore with zoned, colloform texture; shells of tetrahedrite-tennantite (tt), chalcopyrite (cpy), chlorite (chl), and quartz (qtz), in pyrite. Sample FXIA (RL). D. Silicified and sericitized peripheral stockwork rock, 1 m below MS body footwall, 140-m level. Partly recrystallized pyrite (py) framboids occur within quartz-sericite (black) and also rimming sphalerite I (sp). Sample FXXII (RL). E. Silicified and sericitized stockwork containing anastomozed pyrite (py)-minor chalcopyrite (cpy) stringers in quartz-sericite rock (dark gray); the pyrite grains have abundant chalcopyrite inclusions and microveinlets. Sample F32a, transmitted light (TL + RL). F. Silicified and sericitized peripheral stockwork, 190-m level, with pyrite II (py)-chalcopyrite (cpy)-quartz (qtz)-sericite microveinlet. Quartz-sericite host rock in the upper left corner. Sample FVIB (TL + RL), crossed polars. G. Quartz (white)-chlorite host rock to stockwork veins, with ghosts of phenocrysts, now fully replaced by quartz and chlorite. Sample F33 (TL). H. Q2 (qtz)chlorite II (dchl)-chalcopyrite (cpy)-pyrite (py) stockwork vein. Sample F60 (TL + RL). I. Stockwork vein with chalcopyrite (cpy) replacing pyrite II (py). The arrowed sphalerite I (sp) microveinlet is later than the narrow chalcopyrite microveinlet extending from the upper right to the lower middle of the photomicrograph. Sample F47 (RL). J. Colloform texture in stockwork, with stringers of chlorite II-pyrite-chalcopyrite (all black) within very fine grained chlorite (gray). Sample F38 (TL). K. Stockwork with chlorite II (black) with arsenopyrite grains (apy), mostly coarser than pyrite (py), both corroded by chalcopyrite (cpy). Sample F51 (RL). (L) Stockwork quartz vein containing colloform pyrite-chalcopyrite (py-cpy) band, partly recrystallized. Sample F58 (RL). M. Detail of left end of band in L. Zones, within quartz (black), of very fine grained intergrown pyrite-chalcopyrite and intercalated more recrystallized zones (whiter) of both minerals. Sample F58 (RL). N. Stockwork containing pyrite with cellular, colloform texture, with inner, finer grained zone and outer, more (solid) inclusion-free recrystallized zone. Surrounded by finer pyrite, chlorite II (black), and a quartz (qtz) grain. Sample F22 (RL). 0361-0128/98/000/000-00 $6.00

248

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

249

H
py

py

p y- c p y

FIG. 5. (Cont.) 0361-0128/98/000/000-00 $6.00

249

250

INVERNO ET AL.
VI

Al +
Chl . II in stockwork veins Chl . I in stockwork host rocks Chl . II in stockwork host rocks Chl . II in upper sericitized/silicified s tockwork rocks Chl . I in upper sericitized/silicified stockwork rocks Chl . in stockwork rocks (Barriga, 1983) Chl . I in massive ore Chl . II in massive ore Chl . in massive ore (Barriga, 1983) Chl . II in chert Chl . in chert (Barriga, 1983) Chl . in other hanging wall rocks (Barriga, 1983) Chl . I in regional rocks Chl . in regional rocks (Barriga, 1983)

Mg

Fe

FIG. 6. Triangular plot of tri-octahedral chlorite compositions from Feitais (from Zane and Weiss, 1998), showing dominant Fe chlorites and minor Mg chlorites. Number of vacancies in octahedral sites based on 28 anhydrous oxygen [12 (R2+ + R3+ + R4+)].

chalcopyrite-chlorite II association is also present in localized, rosary-like stringers of fine- to medium-grained chlorite II, pyrite and, locally, abundant chalcopyrite, within massive, very fine grained chlorite II, in a partially recrystallized early texture (Fig. 5J). Two varieties of sphalerite are recognized, both commonly fine grained. Sphalerite I is medium brown (medium green in plane-polarized light under the microscope) with abundant inclusions of chalcopyrite (chalcopyrite disease) and yellow internal reflections. Sphalerite I typically precedes, but in places postdates, chalcopyrite (Figs. 5I, 7). Sphalerite II is darker in
Quartz H Sericite I Chlorite I Quartz 1 Quartz 2 Chlorite II Sericite II/ Muscovite Arsenopyrite Pyrite I Pyrite II ? Carbonate Tetrahedrite - tennantite Sphalerite I Chalcopyrite Sphalerite II Boulangerite

TIME

EARLY

Bournonite ?

LATE

Galena Quartz 3 ? ? Pyrite III ? ? Goethite/Lepidocrocite

FIG. 7. Paragenetic sequence of Feitais stockwork vein minerals. 0361-0128/98/000/000-00 $6.00

color (red in plane-polarized light under the microscope), with few or no chalcopyrite inclusions and enriched in Fe, as inferred from its red internal reflections. It is generally paragenetically later and less common than sphalerite I (Fig. 7). Barriga (1983) reported 1.4 to 2.5 wt percent Fe in sphalerite from the stockwork, and this likely corresponds to our sphalerite I. Tetrahedrite-tennantite is a distinctive and relatively abundant mineral in the Feitais stockwork. It is commonly anhedral, forming aggregates up to 2 mm long, and, like sphalerite, is nearly synchronous with chalcopyrite (Fig. 7). Fineto medium-grained euhedral to anhedral arsenopyrite is abundant in the pyrite-chalcopyrite stockwork bands and intervening stringers mentioned above in silicified and/or sericitized volcaniclastic rock, as well as, locally, with chlorite II in stockwork quartz veins and predominantly in their chloritized walls (Fig. 5K). Galena, bournonite, and boulangerite are minor minerals. Pyrrhotite was identified in three chloritized rhyolitic volcaniclastic rock samples, where it occurs as minute grains within chalcopyrite and in one case is partly converted to marcasite and pyrite. Native bismuth is present locally in stockwork veins (Mitsuno et al., 1988). Several generations of vein quartz are present (Fig. 7), postdating the anhedral, fine-grained quartz of the hydrothermally altered host rock (Quartz H). Quartz 1 (Q1), commonly anhedral but locally displaying a comb structure, is up to 6 mm long and contains abundant fluid inclusions that impart a milky white appearance. In these same veins and veinlets, in proximity to chalcopyrite, Q1 is cut by a more transparent, anhedral to euhedral quartz 2 (Q2). Q2 is typically more fine grained (<1 mm long) and contains far fewer fluid inclusions. A third generation, quartz 3 (Q3), is more localized and related to deformation. It is either the product of recrystallization along Q1 grain boundaries and fractures, developing new quartz grains with 120 boundaries due to annealing or

250

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

251

formed as quartz fibers aligned in the pressure shadows of coarse pyrite grains (e.g., cubes). Q2 is also accompanied by up to 2-mm carbonate crystals in proximity to chalcopyrite (Fig. 7); both the quartz and carbonate locally form growthzoned, idiomorphic crystals with terminations pointing into the vein center, characteristic of open-space growth. At the periphery of the stockwork, centimeter-long bands in chloritized host rock to the stockwork (Fig. 5L) consist of very fine grained intergrown pyrite and chalcopyrite (both <2 mm wide), in a zoned, colloform texture. Zones of more coarsely grained, recrystallized pyrite and chalcopyrite are intercalated within the bands (Fig. 5M). Other colloform, cellular textures occur in pyrite from stockwork ore (Fig. 5N), all typical of early, rapid crystallization. The presence of Pb (-Cu) sulfosalts, boulangerite, and bournonite in the stockwork veins, and Bi enrichment in both stockwork veins and base of the massive sulfide, in both cases accompanying chalcopyrite, and the chalcopyrite disease at the base of the massive sulfide match the observations of Marcoux et al. (1996) from other VHMS deposits in the Iberian Pyrite Belt. These authors concluded that these minerals textures were late and due to reworking (zone refining) of earlier sulfide minerals by high-temperature, Cu-bearing fluids. Fluid Inclusion and Stable and Radiogenic Isotope Studies Sampling and analytical methods Samples were collected away from faults, from diamond drill hole cores in sections 160, 520, 560, 600, 640, 680, 800 and 880 ft (Fig. 4), from faces on underground levels 140, 190, and 200 m, and from outcrops. About 100 polished thin sections were scanned to evaluate the nature and distribution of fluid inclusions in quartz (Q1 and Q2), carbonate, and sphalerite (Fig. 7) from stockwork ore. Determinations of final ice-melting (Tm) and homogenization temperatures (Th) of fluid inclusions were obtained with a Fluid Inc.-adapted U.S.G.S. heating-freezing stage. The accuracy of the thermocouple was estimated at 0.1C from the recommended values for standards, and the maximum variation of replicate determinations was 0.1C. Sixty 34S determinations by laser ablation-mass spectrometer were conducted at the Central Science Laboratory at the University of Tasmania on massive ore and stockwork sulfides (mainly pyrite, but also chalcopyrite, sphalerite, galena, and arsenopyrite). Analytical procedures were those described by Huston et al. (1995). A 200-m-thick polished slice of each sample was used for analysis. CO2 separation was conducted on gas derived from samples containing carbonates prior to sulfur isotope analysis. Ablation of sulfides by a Nd-YAG laser sampled a 150- to 500-(commonly 300-)m-wide and about 200m-deep cone. A VG SIRA series 2 mass spectrometer was used for isotopic analysis of SO2. Correction factors were used for the different sulfides (Huston et al., 1995), and where the ablation cone included more than one sulfide because of the fine-grain size, a visual volume estimate of each mineral was weighted by the respective mineral correction factor. The mass spectrometer precision (1) was 0.1 per mil, and, ignoring errors in estimating the proportions of minerals ablated, the overall precision of the ablation results was probably 0.4 to 0.5 per
0361-0128/98/000/000-00 $6.00

mil (Huston et al., 1995). The sulfur isotope composition of barite was determined by reaction with silica and cuprous oxide in vacuo at 1,120C, and passing the gas over copper at 600C to produce SO2 for analysis in the mass spectromer; the precision (1) was estimated to be 0.2 per mil. The 18O compositions of six stockwork quartz (Q1 and Q2) samples and two of quartz in massive ore were determined at the University of Arizona. Analytical procedures were those of Clayton and Mayeda (1963). Repeated analyses of the standards yield a 1 error of 0.15 percent, and overall precision for mineral separates is estimated to be 0.2 per mil (1). Carbon and oxygen isotope analyses were conducted on 21 carbonate samples from quartz-carbonate stockwork veinlets and on five samples from carbonate matrix and veinlets in massive ore, also at the Central Science Laboratory at the University of Tasmania. CO2 released by reaction with H3PO4 (McCrea, 1950) was analyzed with the same mass spectrometer used for S isotope analysis. Precision (1) was estimated to be 0.03 and 0.04 per mil for C and O, respectively. The 87Sr/86Sr ratios were determined on six barite samples at the University of Adelaide, South Australia, by methods described in Foden et al. (1995, 2001), using a Finnigan MAT 262 multicollector mass spectrometer operated in combined static-dynamic mode. Precision (2) was estimated to be 0.000047. The 34S determinations were also made on these barite samples. Fluid inclusions Sphalerite grains in the stockwork veins and veinlets are almost devoid of visible fluid inclusions. Most Q1 quartz grains exhibit a dense cluster of fissures filled with secondary fluid inclusions. However, in a few of them there are well-developed growth zones with primary two-phase liquid (L) + vapor (V) inclusions (Fig. 8A). Growth zones defined by primary fluid inclusions are even more abundant in Q2 quartz and carbonate grains (Fig. 8B, E), and these minerals also have inclusion-free overgrowths with L + V inclusions in the inner portion of the grain that are interpreted to be pseudosecondary in origin (Fig. 8C, D, F). All these primary and pseudosecondary aqueous fluid inclusions are similar, mostly 1 to 4 m, and in some cases in Q1 quartz up to 6 to 8 m in diameter. They are mostly rounded to elongated or negative-crystal shaped and predominantly rhombohedral in carbonate. Only portions along growth zones or other inclusion assemblages, having inclusions with consistent vapor (V) to total inclusion (T) volumetric ratios (V/T 5%), were used for microthermometry (see Bodnar et al., 1985; Goldstein and Reynolds, 1994). These L + V inclusions coexist in the same inclusion assemblage (e.g., growth zone) with larger, irregularly shaped, vapor-only inclusions, up to 10 m wide, indicating that necking occurred locally (Fig. 8E close-up, G, H; Sheperd et al., 1985). The presence in the same growth zones of additional liquid-only inclusions, formed by necking, and the absence throughout the stockwork of healed microfractures (in quartz) containing only vapor-rich inclusions suggests that boiling did not occur (Bodnar et al., 1985). Due to the small size of the two-phase L + V inclusions, only 122 (46 in Q1, 33 in Q2, and 43 in carbonate grains) usable determinations were completed from 22 doubly polished

251

TABLE 1. Electron Microprobe Analyses of Chlorite from Feitais

252

Rock type F9a F14 F21 F33 F42(R)

Stockwork ore

Sample no.

F31

Drill hole section, (m) or level/exposure FS9803-560, 357.2 Chlorite II Stw veins 3 3 3 3 3 2 5 1 4 Chlorite I Stw host rock Chlorite II Stw veins Chlorite I Stw host rock Chlorite II Stw veins Chlorite II Stw host rock Chlorite II Stw veins Chlorite I Stw host rock Chlorite II Stw veins FS9944-560, 365.1 FS9947-600, 412.3 FS9953-640, 330.2

FS9963-540, 452.6

FS9949-640, 387.7 Chlorite II Stw host rock 3

0361-0128/98/000/000-00 $6.00 3 24.85 24.66 0.03 27.83 10.68 0.23 0.08 0.003 0.02 0.02 88.40 Cations normalized to 28 oxygens 5.238 2.762 8.000 3.364 0.005 4.904 3.356 0.041 0.018 0.0003 0.008 0.005 11.701 0.594 11.802 0.666 11.823 0.679 11.705 0.694 11.827 0.697 11.859 0.747 5.204 2.796 8.000 3.199 0.006 5.658 2.842 0.076 0.003 0.001 0.013 0.003 5.175 2.825 8.000 3.194 0.012 5.773 2.735 0.063 0.006 0.001 0.034 0.005 5.112 2.888 8.000 3.511 0.006 5.613 2.472 0.043 0.014 0.0003 0.008 0.038 5.026 2.974 8.000 3.317 0.005 5.928 2.510 0.054 0.005 0.0003 0.005 0.003 5.425 2.575 8.000 2.855 0.003 6.650 2.253 0.086 0.003 0.001 0.005 0.003 5.389 2.611 8.000 2.907 0.000 6.685 2.183 0.067 0.006 0.003 0.000 0.003 11.854 0.754 89.25 89.22 90.74 90.07 86.07 84.95 24.35 23.80 0.04 31.66 8.92 0.42 0.01 0.01 0.03 0.01 24.12 23.81 0.07 32.18 8.55 0.35 0.03 0.01 0.08 0.02 24.39 25.91 0.04 32.03 7.91 0.24 0.06 0.003 0.02 0.14 23.56 25.02 0.03 33.23 7.89 0.30 0.02 0.003 0.01 0.01 23.83 20.21 0.03 34.88 6.63 0.45 0.01 0.01 0.01 0.01 23.33 20.27 <0.01 34.61 6.34 0.34 0.02 0.03 <0.01 0.01 23.98 22.31 0.01 31.07 8.48 0.32 0.03 0.03 0.04 0.08 0.03 86.38 26.78 25.36 0.07 24.51 6.60 0.36 0.01 0.51 0.16 0.38 0.00 84.75 25.96 23.08 0.01 23.27 13.02 0.15 0.04 0.08 0.06 0.04 0.01 85.89 26.92 23.44 0.03 22.57 12.95 0.17 0.08 0.02 0.07 0.48 0.02 86.75 INVERNO ET AL. 5.358 2.642 8.000 3.179 0.001 5.751 2.798 0.060 0.007 0.003 0.016 0.021 0.005 11.841 0.673 5.779 2.221 8.000 4.229 0.012 4.423 2.124 0.066 0.003 0.043 0.067 0.104 0.000 11.071 0.676 5.504 2.496 8.000 3.272 0.001 4.126 4.116 0.027 0.019 0.001 0.025 0.056 0.001 11.644 0.501 5.620 2.380 8.000 3.386 0.005 3.939 4.029 0.030 0.018 0.001 0.028 0.128 0.003 11.567 0.494

Chlorite type and host

Chlorite II Stw veins

Chlorite I Stw host rock

No. of analyses

Wt %

SiO2 Al2O3 TiO2 FeO MgO MnO CaO BaO Na2O K2O P2O5 Total

24.65 25.64 0.01 28.75 10.88 0.33 0.02 0.01 0.003 0.003

90.30

252

At. %

Si Al(IV) Sum Al(VI) Ti Fe Mg Mn Ca Ba Na K P Sum Fe/(Fe+Mg)

5.104 2.896 8.000 3.362 0.001 4.978 3.359 0.059 0.005 0.001 0.001 0.001

11.767 0.597

TABLE 1. (Cont.) Sil/ser upper stw ore AF123b 200 level Chlorite II Stw veins 4 3 1 4 4 4 3 3 Chlorite I Stw veins Chlorite I Chlorite II Chlorite II Chlorite I Chlorite II Chert Chlorite I FS9949640, 373.9 FS9822160, 250.1 140 level 190 level 200 level FS9801800, 469.9 FS9950600, 214.1 S. Joo quarry F40 F54a FXIA FVB FXXXIIIA F57 F67 F(SJ)70 F(EDS)65 EDS0001, 574.9 Chlorite I (Est. Serra) 3 Massive Cu ore Massive pyritic ore Hanging Wall rock Regional rhyol. rock

Rock type

Stockwork ore

Sample no.

F60

Drill hole section (m) or level/ exposure

FS9801-800, 499.2

0361-0128/98/000/000-00 $6.00 3 3 26.31 22.00 0.03 25.35 14.27 0.25 0.04 0.04 0.06 0.02 88.37 Cations normalized to 28 Oxygens 5.475 2.525 8.000 2.872 0.006 4.411 4.459 0.044 0.009 0.008 0.025 0.005 11.839 0.497 11.748 0.723 11.762 0.739 5.022 2.978 8.000 3.480 0.004 5.834 2.240 0.132 0.051 0.001 0.005 0.001 5.090 2.910 8.000 3.378 0.008 6.056 2.141 0.161 0.009 0.001 0.005 0.003 5.426 2.574 8.000 4.003 0.001 4.293 2.733 0.193 0.023 0.004 0.032 0.037 11.319 0.611 5.334 2.666 8.000 3.346 0.006 4.007 4.171 0.102 0.009 0.004 0.012 0.007 11.664 0.490 5.616 2.384 8.000 2.794 0.041 7.081 1.747 0.049 0.030 0.001 0.011 0.011 11.765 0.802 5.499 2.501 8.000 2.918 0.001 4.702 4.085 0.074 0.006 0.001 0.007 0.002 11.796 0.535 88.40 89.34 88.03 89.28 23.13 25.24 0.02 32.13 6.92 0.72 0.22 0.01 0.01 0.003 23.54 24.67 0.05 33.49 6.64 0.88 0.04 0.01 0.01 0.01 26.13 26.88 0.01 24.72 8.83 1.10 0.10 0.04 0.08 0.14 26.19 25.05 0.04 23.53 13.74 0.59 0.04 0.04 0.03 0.03 25.49 19.94 0.25 38.44 5.32 0.26 0.13 0.02 0.03 0.04 89.92 26.49 22.15 0.003 27.09 13.20 0.42 0.03 0.02 0.02 0.01 89.43 24.57 22.70 0.03 28.33 11.51 0.34 0.04 0.03 0.01 0.02 0.01 87.59 25.10 21.03 <0.01 25.21 13.68 0.08 <0.01 <0.01 0.04 <0.01 0.02 85.17 23.75 21.84 0.01 31.08 8.63 0.43 0.02 <0.01 0.05 0.01 0.02 85.84 26.34 22.48 0.03 18.95 18.36 0.24 0.05 <0.01 0.05 0.02 0.01 86.53 5.270 2.730 8.000 3.006 0.005 5.081 3.680 0.062 0.009 0.003 0.005 0.005 0.002 11.858 0.580 5.448 2.552 8.000 2.829 0.000 4.576 4.427 0.014 0.000 0.000 0.016 0.000 0.003 11.865 0.508 5.167 2.833 8.000 2.767 0.001 5.655 2.799 0.080 0.005 0.000 0.021 0.003 0.003 11.334 0.669 5.422 2.578 8.000 2.875 0.004 3.261 5.634 0.042 0.011 0.000 0.020 0.005 0.003 11.855 0.367 FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

Chlorite type and host

Chlorite II Stw veins

Chlorite II Chlorite II Stw host rock Stw host rock

No. of analyses

Wt %

SiO2 Al2O3 TiO2 FeO MgO MnO CaO BaO Na2O K2O P2O5 Total

26.60 21.93 0.03 25.68 15.11 0.25 0.02 0.05 0.02 0.01

89.70

253

At. %

Si Al(IV) Sum Al(VI) Ti Fe Mg Mn Ca Ba Na K P Sum Fe/(Fe+Mg)

5.464 2.536 8.000 2.775 0.006 4.412 4.628 0.043 0.005 0.009 0.007 0.002

11.887 0.488

Abbreviations: regional rhyol. rock = regional rhyolitic volcaniclastic rock, sil/ser = silicified/sericitized, Stw = stockwork

253

254

INVERNO ET AL.

qt z

0361-0128/98/000/000-00 $6.00

254

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

255

thin sections prepared from 14 selected samples. For 57 of these determinations, only either Tm or Th was determined for each inclusion; in a temperature-salinity plot, the other missing parameter was assumed to be equal to that of an adjacent primary inclusion (or the average of values of adjacent primary inclusions), for example, along the same growth zone. However, for 65 determinations (24 in Q1, 19 in Q2, and 22 in carbonate grains) both Tm and Th were obtained for each fluid inclusion. Tm and Th measurements in these very small inclusions were only possible through the use of a cycling technique in both freezing and heating runs (Goldstein and Reynolds, 1994). This cycling technique works in sequential steps, by checking in each step the inclusion phases present at a certain temperature, T1, and then checking again at a slightly higher temperature, T2. If it is concluded that last icecrystal melting or homogenization occurred between T1 and T2, a temperature range, T1 to T2, or a midpoint temperature, is reported for Tm and Th (see Table 2 where cases of midpoint temperature are specified), respectively, because the accurate recording of the moment of disappearance of the last ice crystal or vapor bubble is extremely hard to detect in these very small inclusions. Only in a few cases of either the largest inclusions or under exceptional optical conditions is a unique temperature reported for Tm in Table 2. Neither CO2 nor clathrates were seen in these primary and pseudosecondary inclusions during freezing and/or heating runs, and daughter minerals are absent. Although the temperature of first melting of ice (correlative of eutectic melting) was virtually impossible to discern in these very small inclusions, some liquid phase was detected in a few of them at a temperature of 20C, and considering the Tm range obtained (see below), the Ca or Mg chloride species are excluded and the only possible common solutions within the inclusions are NaCl-H2O and NaCl-KCl-H2O (Crawford, 1981; Goldstein and Reynolds, 1994), with the former predominating in VHMS systems (de Ronde, 1995; Scott, 1997). Therefore, we express the bulk salinity in wt percent NaCl equiv, calculated from Tm using the equation of Bodnar (1993). The primary and pseudosecondary fluid inclusions yield salinities and Th of 3.0 to 7.9 wt percent NaCl equiv and 270 to 315C for Q1, 2.2 to 7.9 wt percent NaCl equiv and 270 to 315C for Q2, and 2.2 to 8.1 wt percent NaCl equiv and 270 to 315C for carbonate, respectively (Table 2). A Thsalinity plot is shown in Figure 9. Because Tm and Th data

from primary fluid inclusions in carbonates are commonly suspect, due to the possible coincidence of growth zones with cleavage planes (Roedder, 1984), all such cases are marked in Table 2. Other than this, the results for carbonates are consistent with those from quartz (Q1 and Q2) and are taken as representative, given the textural relationships between Q1, Q2, and carbonate (Fig.7). No significant variation in salinity or homogenization temperature exists laterally and vertically in the stockwork zone. A comparison was made, in core samples from hole FS9953640, between salinity values for quartz and carbonate from the uppermost silicified and sericitized stockwork veins (sample F32a) and from chloritized stockwork veins (samples F34, F36, and F37; Table 2) some 11 to 14 m below. The latter are slightly more saline (differences clearly above the error and precision of results), but homogenization temperatures are comparable. The Th range does not extend as high as that estimated from the stockwork chlorite (i.e., 285385C: Table 1, Fig. 7) but is higher than that calculated from the oxygen isotope equilibrium temperatures (250270C) obtained by Barriga and Kerrich (1984) for quartz-chlorite pairs from three Feitais stockwork samples, using the fractionation equations of Clayton et al. (1972) and Wenner and Taylor (1971). Munh et al. (1986) obtained a temperature range of 200 to 215C from the same quartz-chlorite data, using additional fractionation equations (Matsuhisa et al., 1979; Matthews and Beckinsale, 1979). Larger, apparently isolated inclusions in the cores of crystals were not used because a detailed inspection showed that they belong to fluid inclusion assemblages in tiny microfissures transecting the whole crystal, together with much smaller secondary fluid inclusions, and thus are not primary. Both the larger and smaller inclusions yield similar Tm and Th to those obtained from other secondary inclusions along better developed microfractures in the crystal. Several generations of microfractures with dilute secondary fluid inclusions transect the quartz and carbonate crystals, yielding Th below those reported above. CO2-bearing secondary inclusions in other fissures in the same grains probably relate to regional metamorphism (Inverno and Solomon, 2001). Sulfur isotope compositions of sulfides and barite The 32 34S determinations from massive ore sulfides range from 15.4 to +4.7 per mil (mean, 2.8), with 10 values

FIG. 8. Transmitted-light photomicrographs of Feitais stockwork quartz and carbonate crystals and their fluid inclusions. A. Prominent growth zones, with primary fluid inclusions, in central to lower right portion of Q1 crystal, that are cut by a multitude of secondary inclusion planes. Sample MP61, crossed polars. B. Q2 crystal exhibiting growth zones with primary fluid inclusions and NE-SW planes of secondary inclusions. Sample F10. C. Quartz (qtz) crystal in center, with a clear, inclusionfree overgrowth and an inner, dark portion containing abundant pseudosecondary fluid inclusions. Adjacent carbonate (car) and pyrite (py) grains are surrounded by finer grained quartz-carbonate matrix, in peripheral stockwork host rock. Sample FI. D. Close-up of rectangular inset in (D), showing the inclusion-free overgrowth in the quartz crystal, contrasting the inner portion of the crystal packed with pseudosecondary fluid inclusions. Sample FI. E. Growth zones in carbonate crystals along a microveinlet crosscutting Q1 (qtz) grains. Close-up of rectangular inset in lower left of crystal, with detail of carbonate crystal containing minute aqueous L + V inclusions along growth zone, together with larger, dark vapor-only inclusions. Sample F58. F. Inclusion-free overgrowth in carbonate crystal in center, with its inner, darker portion containing very abundant pseudosecondary fluid inclusions. Q1 on the bottom. Sample F37. G. Detail of central part of Q1 crystal in (A). NE-SW growth zones with aqueous L + V inclusions and a few dark gray, larger vapor-only inclusions. Sample F61. (H) Detail of carbonate crystal in center of F. Inclusion-free overgrowth in lower right corner. Rounded or elongated (e.g., arrowed) aqueous L + V and larger, irregularly shaped, dark vapor-only pseudosecondary inclusions in inner part of crystal. Sample F37. 0361-0128/98/000/000-00 $6.00

255

256

INVERNO ET AL.

TABLE 2. Microthermometric Data from Single Determinations from Primary and Pseudosecondary Inclusions from Feitais Stockwork Quartz and Carbonate Sample no. F34 Drill hole section (m) or UG level FI Host genetic mineral type P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P P PS P PS P P P P P P P P P P P P P P Salinity (wt % NaCl equiv) 6.1 6 6.1 6 4.6 6 4.6 6 6.1 6 4.6 6 5.7 6.85 7 6.1 6 6.85 7 5.35 7 6.85 7 5.35 7 4.6 6 4.4 8 3.8 6 3.0 6 6.1 6 6.85 7 6.85 7 4.6 6 5.35 7 7.55 7 6.1 6 6.85 7 7.55 7 7.2 7.55 7 6.1 6 6.1 6 3.8 6 7.9 6.1 6 3.0 6 4.7 7.55 7 4.6 6 4.6 6 3.0 6 6.85 7 5.35 7 7.55 7 5.0 3.0 6 6.1 6 3.8 6 2.15 8 3.0 6 6.1 6 3.8 6 6.1 6 5.35 7 5.7 3.8 6 7.55 7 5.35 7 6.1 6 5.35 7 3.8 6 3.8 6 7.55 7 Sample no. F60 Drill hole section (m) or UG level FI Host genetic mineral type P P P P P P P P P P P P P P P PS PS PS P P P P P P PS P P P P P P P P P P P? P? P? P P P P P P P P P P P P P PS PS P? PS PS PS PS Salinity (wt % NaCl equiv) 6.65 9 6.7 3.0 6 3.8 6 4.6 6 5.35 7 3.8 6 6.85 7 6.85 7 7.55 7 7.55 7 6.85 7 6.1 6 7.55 7 2.6 6.1 6 6.85 7 7.55 7 6.1 6 4.6 6 6.1 6 8.1 3.8 6 4.6 6 6.5 7.2 7.55 7 2.6 6.85 7 6.85 7 6.1 6 5.35 7 2.2 6.1 6 5.35 7 6.1 6 5.35 7 4.6 6 6.1 6 5.0 5.35 7 5.35 7 7.55 7 3.8 6 3.0 6 6.1 6 5.35 7 6.1 6 6.1 6 5.9 5.35 7 7.55 7 5.35 7 6.85 7

Th

(oC)
4

Tm

(oC)1

Th

(oC)
5

Tm

(oC)1

FS9953-640 Qtz 1 (333.5)

287.5

F36

FS9953-640 Qtz 1 (336.2)

295 5 305 5 305 5 307.5 4 297.5 4 285 5 305 5 295 5

307.5

3.75 3.75 2.75 2.75 3.75 2.75 3.5 3 4.25 3.75 4.25 3.25 4.25 3.25 2.75 2.65 2 2.25 1.75 3.75 4.25 4.25 2.75 3.25 4.75 3.75 4.25 4.75 4.5 3 4.75 3.75 3.75 2.25 5.0 3 3.75 1.75 2.8 3 4.75 2.75 2.75 1.75 4.25 3.25 4.75 3.0 3 1.75 3.75 2.25 1.25 1.75 3.75 2.25 3.75 3.25 3.5 + 2.25 4.75 3.25 3.75 3.25 2.25 2.25 4.75

FS9801-800 Qtz 2 (499.2)

275 295 5 285 5

F61

FS9801-800 Qtz 2 (500.6)

305 5 295 5 305 5 312.5 4 312.5 4 305 5 305 5 295 5 305 5 275 7 312.5 4 302.5 4 305 5 275 5 297.5 4 305 5 277.5 4 275 5 305 5

4.15 2 4.2 3 1.75 2.25 2.75 3.25 2.25 4.25 4.25 4.75 4.75 4.25 3.75 4.75 1.5 3 3.75 4.25 4.75 3.75 2.75 3.75 5.2 3 2.25 2.75 4.0 3 4.5 3 4.75 1.5 3 4.25 4.25 3.75 3.25 1.3 3 3.75 3.25 3.75 3.25 2.75 3.75 3.0 3 3.25 3.25 4.75 2.25 1.75 3.75 3.25 3.75 3.75 3.6 3 3.25 4.75 3.25 4.25

F3 F19

FS9942-560 Fe-dol. (382.2) FS9950-600 Calc. (363.2)

292.5 4 302.5 4 295 5 295 5 275 5 295 5 312.5 4 307.5 4 302.5 4 312.5 4 295 5 295 5 305 5 285 5 312.5 4 312.5 4 277.5 4

F61

FS9801-800 Qtz 1 (500.6)

F24a

FS9933-520 Calc. (305.9)

F37 F43a F58

F62

FS9801-800 Qtz 1 (505.6)

FS9953-640 Calc. (336.4) FS9949-640 Dol. (393.0) FS9801-800 Dol. (473.9)

285 5 275 5 275 5 305 5 312.5 4 312.5 4 312.5 4 305 5

F10

FS9800-600 Qtz 2 (452.6) FS9953-640 Qtz 2 (322.1)

287.5 4 287.5 4 285 5 295 5 297.5 4 285 5 285 5 275 5 285 5 295 5 302.5 4

F32a

FI

level 190

Dol.

307.5 4 312.5 4 312.5 4 312.5 4

Abbreviations: FI = fluid inclusion, P = primary (in growth zone) fluid inclusion, PS = pseudosecondary fluid inclusion, P? = fluid inclusion along growth zone coincident with cleavage plane, Qtz. 1 = quartz 1, Qtz. 2 = quartz 2, Calc. = calcite, Dol. = dolomite, Fe-dol. = ferroan dolomite; UG = underground 1 Midpoint temperature (T ) of a 0.5C interval, unless stated otherwise m 2 Midpoint temperature (T ) of a 0.1C interval m 3 One-number temperature (T ) m 4 Midpoint temperature (T ) of a 5C interval h 5 Midpoint temperature (T ) of a 10C interval h 6 Midpoint salinity of a 0.8 wt percent NaCl equiv interval 7 Midpoint salinity of a 0.7 wt percent NaCl equiv interval 8 Midpoint salinity of a 0.2 wt percent NaCl equiv interval 9 Midpoint salinity of a 0.1 wt percent NaCl equiv interval

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES
9

257

Salinity (wt % NaCl equiv.)

Quartz 1 Quartz 2 Carbonate

2 270

280

290

300

310

320

Homogenization temperature ( oC)

FIG. 9. Bivariate plot of salinity vs. homogenization temperature from primary and pseudosecondary inclusions from Feitais stockwork quartz and carbonate for all 122 determinations; the data of the 65 inclusions with dual, Tm and Th determinations are shown in solid patterns. Line separates reversing buoyancy fluids and permanently buoyant fluids on mixing with seawater (after Turner and Campbell, 1987).

below 5.0 per mil (Table 3, Fig. 10; Inverno et al., 2002), extending the total range of Mitsuno et al. (1988) and Yamamoto et al. (1993). The 28 stockwork ore sulfide values range from 5.5 to +4.7 per mil, but with two higher 34S

values of +6.3 and +11.9 and one low value at 11.2 per mil (mean, 0.4; Table 3). The total range of 34S values is typical of VHMS ores, except for the low stockwork values and the similarity of stockwork and massive sulfide values, with stockwork values commonly being higher (Huston, 1999). There are no significant lateral or vertical variations in the sulfur isotope compositions in the massive sulfide body or the stockwork zone. There is up to 10 per mil variation within individual samples for grains of even the same sulfide mineral, and no observable regular trend. A variation of 3 per mil was found for three spots along a 4.5-mm-long profile within a pyrite clast in microconglomeratic massive ore. The most negative 34S values in Feitais massive ore (Fig. 10) are generally associated with very fine grained, reniform, cellular pyrite and associated minerals, as Velasco et al. (1998) noted for framboidal and other early pyrites at Aznalcllar, Tharsis, and Concepcin in the eastern Iberian Pyrite Belt. At Feitais, recrystallized pyrite grains adjacent to the early, fine-grained pyrites in massive ore have up to 10 per mil higher 34S values. Similarly, in stockwork ore at Feitais, pyrite grains with identical textures have negative 34S values down to 11.2 per mil, and adjacent recrystallized pyrite grains have values up to 8 per mil higher. Barite from three thin barite-bearing layers and a meterthick lens within rhyolitic volcaniclastic hanging-wall rock yield 34S values of 22.2 to 23.0 per mil (Table 3), similar to a single value of 21.9 per mil recorded previously (Mitsuno et al., 1988; Yamamoto et al., 1993). All of these values for barite are close to that of early Carboniferous seawater sulfate (22; Strauss, 1997; Huston, 1999). Barite from baritecarbonate quartz veins in the massive sulfide orebody, however, yield higher 34S values of 23.9 and 27.9 per mil.

TABLE 3. Sulfur Isotope Data from Feitais Sulfides and Barite Sample/spot F53/2A F53/3 F29/1 F1/1A F2/1 F2/2 F2/3 F2/5 F2/6 F2/7 F15/3 F15/5 F15/6 F15/7A F15/8 F44/1 F48/1A F48/1B F48/1C F57/3 F57/4 0361-0128/98/000/000-00 $6.00 FS9801-800, 469.9 Microconglomeratic py MS Drill hole section (m) or underground level FS9822-160, 247.5 FS9963-540, 426.9 FS9942-560, 287.2 FS9942-560, 308.3 Ore or rock type Microconglomeratic py MS Cu-MS Banded py MS Banded Zn-MS Mineral Py + cpy (1%) Apy Py + cpy (5%) + te (0.5%) Py Gn Py + gn (0.1%) Gn + py (20%) Sp + py (1%) Sp + py (5%) Py + sp (1%) Cpy + py (0.5%) Cpy Cpy + py (5%) Py Py + cpy (5%) Py + cpy (1%) + sp (1%) Py Py Py Py Py Texture Recrystallized py Aggregate of very fine to fine idiomorphic grains Subidiomorphic coarse-grained py crystal Recrystallized py band Fine-grained gn veinlet Fine-grained Gn veinlet with py grains Fine-grained Fine-grained Fine-grained Cpy veinlet Cpy veinlet Cpy veinlet Recrystallized idiomorphic py Recrystallized py Spongy py, 100% recrystallized Py fragment, 35% finely recrystallized Same py fragment (at 2 mm from 1A), 25% finely recrystallized Same py fragment (at 2.7mm from 1B), 25% finely recrystallized Spheroidal, zoned, colloform py, 10% recrystallized Subidiomorphic py crystal, 100% recrystallized 34S () 0.6 2.0 7.0 5.2 7.0 9.6 4.2 6.8 2.9 10.0 3.5 4.7 0.3 3.4 0.2 0.1 3.4 0.9 0.4 1.1 1.3

FS9950-600, 315.3

Cu-MS

FS9956-640, 424.9 FS9951-640, 427.9

Banded Cu-MS Microconglomeratic py MS

257

258

INVERNO ET AL. TABLE 3. (Cont.)

Sample/spot F57/5 F57/6 F57/7 FXVIB/1 FXVIB/2 FXVIB/3 FXVIB/4 FXXXVIIA/1 FVB/1 FVIII/1 FXXX/1 F55/5 F25/1 F25/2 F25/4 F25/5 F25/6 F25/7A F27/1 F8a/1 F8a/2 F8a/3 F8a/4 F13/4 F13/5 F13/7 F13/8 F13/9 F17/1A F32b/1 F50/1 F51/1 F59/1 F59/2 F59/3 F59/4A F59/5 F59/6 FVID/1 F71B F72 F73 F74 F75 F76

Drill hole section (m) or underground level

Ore or rock type

Mineral Cpy + py (2%) + te (0.5%) Py Py + cpy (1%) Py Py Py Py Py Py + cpy (5%) Py Py + cpy (0.25%) + sp (0.25%) Py + cpy (5%) Cpy Cpy Py + cpy (1%) Py + cpy (3%) Cpy Cpy Py+ cpy (5%) + sp (1%) + te (0.2%) Py + cpy (0.5%) + sp (0.5%) Py + cpy (25%) Cpy (95%) + py + qtz Cpy + py (8%) Cpy + py (1%) Py Cpy Cpy Py + cpy (1%) Py + cpy (5%) Py + cpy (4%) Py + cpy (2%) Py + cpy (2%) Py + cpy (0.1%) Cpy Py Cpy + py (1%) Sp + cpy (5%) Sp + cpy (3%)

Texture Aggregate of fine cpy grains Banded, colloform py, 1% recrystallized Recrystallized py Banded, colloform py, mildly-finely recrystallized Py with traces of reniform banding Mildly recrystallized py Mildly recrystallized py (at 1 mm from spot 3) Cellular py, 25% recrystallized Aggregate of very fine to fine subidiomorphic to allotriomorphic py grains Aggregate of very fine to fine subidiomorphic py grains Very fine to fine-grained Fine-grained py, fully recrystallized Aggregate of very fine to medium-grained cpy Aggregate of very fine to medium-grained cpy Idiomorphic medium-grained py crystal Subidiomorphic medium-grained py crystal Aggregate of very fine to medium cpy grains Aggregate of very fine to medium cpy grains Idiomorphic medium-grained py crystal with other sulfide inclusions Subidiomorphic medium-grained py crystal Aggregate Aggregate Aggregate Cpy veinlet Subidiomorphic medium-grained py crystal Cpy veinlet Aggregate Fine-grained py within chlorite II Subidiomorphic medium-grained py crystal Aggregate of idiomorphic medium-grained py (90%)+ very fine grained py, 30% recrystallized Allotriomorphic medium-grained py I grain Allotriomorphic medium-grained py grain within chlorite II Subidiomorphic medium-grained py crystal Cpy pod in qtz vein Fine-grained mass in qtz vein Cpy pod in qtz vein Band of allotriomorphic fine sp(-cpy) grains in quartz vein Band of allotriomorphic fine sp(-cpy) grains in quartz vein Aggregate of fine-grained idiomorphic py grains Subidiomorphic medium-coarse barite grains in qtz-dominated band Vein of allotriomorphic fine-medium grained barite; minor qtz-carb. in vein Subidiomorphic fine-coarse barite grains in qtz-dominated band Subidiomorhpic fine-medium barite grains in qtz-dominated band Coarse and fine barite grains in barite (minor qtz, carb.) band Allotriomorphic fine-medium barite grains in carb. (-qtz) vein

34S () 1.0 4.0 0.1 15.4 12.9 4.7 5.0 6.8 2.0 4.4 3.6 0.9 3.0 1.3 1.4 1.7 1.5 1.1 11.9 11.2 2.3 2.4 1.4 0.8 1.1 3.3 1.6 0.8 2.6 4.7 4.9 5.5 6.3 1.7 1.8 2.3 0.8 0.7 4.4 22.2 27.9 21.2 23.0 22.2 23.9

140 level

Py MS

140 level 190 level 190 level 200 level FS9822-160, 256.5 FS9932-520, 323.3

Banded pyr. MS Banded Cu-MS Py MS Py MS Sil/ser upper Stockwork Sil/ser upper Stockwork

FS9932-520, 350.4 FS9803-560, 352.4

Stockwork Stockwork

FS9944-560, 364.0

Stockwork

FS9950-600, 348.5 FS9953-640, 322.2 FS9951-640, 442.2 FS9951-640, 477.1 FS9801-800, 499.0

Stockwork Sil/ser upper Stockwork Sil/ser upper Stockwork Stockwork Stockwork

190 level FS9975-880, 310.9 FS9975-880, 355.7 FS9954-680, 310.1 FS9954-680, 309.7 FS17-800, 417.6 FS9975-880, 374.4

Sil/ser upper Stockwork Thin stratified band in HW rock Zn-MS Thin stratified band in HW rock Thin stratified band in HW rock Thick stratified band in HW rk Py MS

Py + sp (2%) + cpy (1%) Barite Barite Barite Barite Barite Barite

Notes: Granularity: very fine = <10 m, fine = 10100 m, medium = 1001,000 m, coarse = >1 mm; Py MS = pyritic massive sulfide, Carb. = carbonate, Cu-MS = Cu massive sulfide ore, Zn-MS = Zn massive sulfide ore; HW = hanging wall; Sil/ser = silicified/sericitized; Apy = arsenopyrite, Cpy = chalcopyrite, Gn = galena, Py = pyrite, Qtz = quartz, Sp = sphalerite, Te = tetrahedrite-tennantite 0361-0128/98/000/000-00 $6.00

258

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

259

10 Massive sulfide ore Stockwork ore 8

rocks unaffected by the mineralizing fluids. The outward increase in whole-rock 18O values was interpreted as reflecting a progressive decrease of water/rock ratios and declining temperature (Barriga and Kerrich, 1984; Barriga and Fyfe,1998; see also Huston, 1999). Carbon and oxygen isotope compositions of carbonates Calcite is the prevailing carbonate mineral, with minor Mn or Fe (Barriga, 1983), but in six stockwork samples the carbonate is dolomite or Fe dolomite (Table 5). The 13C values for carbonate in stockwork and massive ore are similar, as are those for 18O. Carbonate in the stockwork and massive ores has 13C of 9.3 to 14.3 and 7.5 to 13.7 per mil, respectively, and 18O of 10.4 to 25.9 and 14.1 to 23.9 per mil, respectively (Table 5, Fig. 12). Six 18O values above 20 per mil are from samples (among others with lower values) in the central part of the stockwork and in one case at the base of the massive sulfide orebody. No other lateral or vertical general variation was detected in either the massive orebody or the stockwork zone.

Number of analyses

0 -16 -12 -8 -4
34

12

S ()
FIG. 10. Frequency histogram of 34S compositions of sulfides from Feitais massive sulfide and stockwork ores, showing most stockwork and massive sulfide values in the range from 5.5 to +4.7 per mil, typical of VHMS ores but with additional very negative values associated with colloform pyrite and biogenically derived sulfur.

Strontium isotope compositions in barite Barite from the three thin barite-bearing bands and the thick barite lens in feldspar-phyric, rhyolitic volcaniclastic hanging-wall rock yields 87Sr/86Sr ratios of 0.708438 to 0.709063, and barite from barite-carbonate quartz veins in massive ore has values of 0.708836 and 0.708877 (Table 6). Discussion

Oxygen isotope compositions of quartz Stockwork quartz (Q1) yields 18O values of 12.0 to 12.3 per mil, with the exception of 13.9 per mil for a single sample beneath a 0.7-m-thick massive ore layer at the northwest end of the Feitais body. Values for stockwork quartz (Q2) are 11.6 and 11.7 per mil, and for quartz from the massive sulfide ore are 16.7 and 17.9 per mil (Table 4, Fig. 11; Inverno et al., 2002). These results are not significantly different from those of Barriga (1983), Barriga and Kerrich (1984), and Munh et al. (1986) for similar Feitais stockwork quartz (14.915.4) and quartz from massive ore (17.118.3), and for Estao stockwork quartz (13.313.9). These authors reported whole-rock 18O values of 11.7 to 12.4 per mil for Feitais stockwork rocks, compared to higher values (13.914.4) for peripheral stockwork rocks at the margins of the stockwork, and even higher values (15.618.1) for volcaniclastic

Significance of the fluid inclusion data The uncorrected fluid inclusion data suggest that stockwork fluids were at temperatures up to about 315C, somewhat lower than the maxima recorded at some modern sea-floor vents and in the deposits of the Hokuroku basin (PisuthaArnond and Ohmoto, 1983; Scott, 1997) and also lower than the maxima found in other massive sulfide deposits of the Iberian Pyrite Belt (Toscano et al., 1997; Almodvar et al., 1998; Nehlig et al., 1998; Knight, 2000; Snchez-Espaa et al., 2000, 2003). The Feitais chlorites yielded higher formation temperatures (see above). Feitais trapping temperatures would be slightly higher than Th if pressure corrections for the overlying water column could be applied. Ocean depth is not known anywhere within the Iberian Pyrite Belt, but fluids having 7 to 8 wt percent NaCl equiv at a temperature of about 315C would, under hydrostatic conditions, boil (see below)

TABLE 4. Oxygen Isotope Data from Quartz from Feitais Stockwork and Massive Sulfide Ores Sample no. F34 F36 F58 F61 F10 F60 F7 F53 Drill hole section, m FS9953-640, 334.6 FS9953-640, 336.2 FS9801-800, 473.8 FS9801-800, 500.6 FS9802-600, 453.0 FS9801-800, 499.2 FS9803-560, 300.3 FS9822-160, 247.5 Ore type Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Massive pyrite Massive pyrite Quartz type Quartz 1 Quartz 1 Quartz 1 Quartz 1 Quartz 2 Quartz 2 Quartz matrix Quartz matrix 18O () 12.3 12.0 13.9 12.3 11.7 11.6 16.7 17.9 Th (oC) 287.5 1 280-310 (270-315) 270-315 287.5 1 270-300 18Ofluid () 5.0 5.0 6.7 5.2 4.4 4.2

Note: The Th value in parentheses represents the whole Th interval for that mineral in that ore type 1 Midpoint temperature (T ) of a 5C interval h 0361-0128/98/000/000-00 $6.00

259

260
6

INVERNO ET AL.

Number of analyses

5 4 3 2 1 0 11 12 13 14 15 16 17 18 19

Quartz from massive sulfide ore Quartz from massive sulfide ore (Barriga, 1983) Stockwork quartz 1 Stockwork quartz 2 Stockwork quartz (Barriga, 1983)

O ()
FIG. 11. Frequency histogram of 18O values from quartz from Feitais stockwork and massive sulfide ores, showing lower values for the latter, probably reflecting the lower fluid temperatures of silica precipitation on the sea floor.

18

unless covered by about 1 km of seawater (Bischoff and Pitzer, 1989). This would raise the lowest Th values (for the lowest salinities) by no more than 5C but not the highest Th values (Potter, 1977; Potter and Brown, 1977; Roedder and Bodnar, 1980). The salinities of the fluid inclusions range between that of modern seawater (3.2 wt % NaCl equiv; Bischoff and Rosenbauer, 1985) and 2.5 times that of seawater, higher than recorded from most modern sea-floor vents and higher than found in and below the Hokuroku ores (3.55.5 wt % NaCl equiv except for one sample which yielded 7.9 wt %: PisuthaArnond and Ohmoto, 1983; Scott, 1997). The Th-salinity plots

(Fig. 9A, B) do not suggest mixing between low- and highsalinity end members. The maximum salinities are also similar to those for Lagoa Salgada (Fig. 1; Jaques and Noronha, 2002), but considerably lower than those for several other massive sulfide deposits in the Iberian Pyrite Belt, which locally have fluid inclusion salinities up to about 24 wt percent NaCl equiv but mostly up to 12 to 14 wt percent NaCl equiv (summary in Solomon et al., 2002). Most of these fluids are unlike the lowsalinity fluids thought to be of metamorphic origin (5 wt % NaCl equiv, 314 mol % CO2: Moura et al., 1997; Moura, 2002; Snchez-Espaa et al., 2000, 2003).

TABLE 5. Carbon and Oxygen Isotope Data from Carbonates from Feitais Ores Sample no. F3 F4 F8a F11b F13 F14 F16 F17 F19 F22 F25 F32b F35 F37 F38 F43a F51 F58 F61 FI FXXXI F29 F54b FVA FXVIA FXXXIIIB Drill hole section (m) or underground level FS9942-560, 382.3 FS9942-560, 384.1 FS9803-560, 352.4 FS9802-600, 472.4 FS9944-560, 364.0 FS9944-560, 365.1 FS9950-600, 336.5 FS9950-600, 348.5 FS9950-600, 363.2 FS9979-600, 530.1 FS9932-520, 323.3 FS9953-640, 322.2 FS9953-640, 335.5 FS9953-640, 336.4 FS9953-640, 348.1 FS9949-640, 392.0 FS9951-640, 477.1 FS9801-800, 473.9 FS9801-800, 500.6 190 level 200 level FS9963-540, 426.9 FS9822-160, 251.3 190 level 140 level 200 level Rock type Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Stockwork Peripheral stw FW rhyolite with carb. veinlets Massive ore Massive ore Massive ore Massive ore Massive ore Carbonate Fe-dolomite Fe-dolomite Calcite Calcite Calcite Calcite Calcite Calcite Calcite Calcite Fe-dol. + dol. Calcite Calcite Calcite Calcite Dolomite Calcite Dolomite Calcite Dolomite Calcite Calcite Calcite Calcite Calcite Calcite 13C () 11.8 11.4 14.2 12.5 12.9 12.4 12.0 12.7 11.7 9.8 11.9 12.5 13.4 11.7 11.1 10.4 12.1 11.2 11.4 9.3 13.7 10.2 7.5 13.7 9.7 13.3 Th (oC) 290-310 (270-315) (270-315) (270-315) (270-315) (270-315) (270-315) (270-315) 270-315 (270-315) (270-315) (270-315) (270-315) (270-315) (270-315) 285 1 (270-315) 270-315 (270-315) 305-315 (270-315) 13Cfluid () 9.2 8.9 11.8 10.0 10.4 10.0 9.5 10.1 9.2 7.3 9.4 10.0 10.9 9.2 8.6 8.0 9.6 8.7 8.9 6.6 11.2 18O () 25.9 25.0 12.4 13.0 10.4 11.5 14.1 10.5 21.6 15.2 17.3 15.1 15.4 20.2 14.2 14.5 11.1 18.2 18.6 22.5 16.8 14.2 19.3 16.9 23.9 15.0 18Ofluid () 17.2 16.1 6.6 7.2 4.6 5.7 8.3 4.4 15.8 9.5 8.6 9.3 9.6 14.4 8.4 5.6 5.3 9.6 12.9 14.5 11.0

Note: The Th values in parentheses represent the whole Th interval for that mineral in that ore type Abbreviations: dol. = dolomite, FW = footwall, stw = stockwork 1 Midpoint temperature (T ) of a 10C interval h 0361-0128/98/000/000-00 $6.00

260

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

261

18O ()
5 -5 10 15 20 25 30

Massive sulfide ore Stockwork ore


C ()

-10

-15
FIG. 12. Bivariate plot of 13C vs. 18O for carbonates from Feitais massive sulfide and stockwork ores, showing that the 18O values are typical of but with a wider range than in most VHMS deposits, due to possible involvement of different fluids; and that the 13C values are clearly lower than in those deposits, probably due to mixing of seawater carbonate with organically derived or igneous carbonate.

The salinities could have been increased above that of seawater through (1) phase separation at depths of a few kilometers, (2) leaching of evaporites in the pre-Phyllite-Quartzite Group sedimentary rocks, (3) subsurface hydration reactions, and/or (4) mixing of ground water with fluid exsolved from underlying magmas (Scott, 1997). Evaporites have not been identified in the Phyllite-Quartzite Group, but only a small part of the continental margin succession is exposed, and they may be present at greater depths. Increases in salinity through hydration reactions are unlikely to be sufficient to account for the data (e.g., Peter and Scott, 1993) but could have occurred through boiling. The lack of evidence in the fluid inclusions for boiling is not a strong indicator that this process did not take place (Roedder, 1984, p. 1104); however, near-surface boiling of seawater to more than double the salinity would probably have led to extensive sulfide precipitation and oxidation (Drummond and Ohmoto, 1984; Reed and Spycher, 1985) and most probably would have caused significant brecciation. This leaves evaporitic and/or magmatic origins to explain the relatively high salinities. Significance of the isotopic data All but one of the stockwork quartz samples used in oxygen isotope analysis yield Th values for primary and pseudosecondary fluid inclusions, so the 18O value of the fluid can be

calculated using the quartz-water equations of Matsuhisa et al. (1979). As the Th range for each sample is narrow (up to 45C), the mean temperature was used for calculations. For sample F58 (Tables 2, 4), which lacks Th data for quartz, the mean of the overall Th interval for Q2 (270315C) was used. The Th values of quartz Q1 and Q2 yield 18Ofluid values of 5.0 to 5.2 and 4.2 to 4.4 per mil, respectively (Table 4). These values are probably more reliable than those calculated from temperatures determined from quartz-chlorite oxygen isotope fractionation by Barriga and Kerrich (1984) and Munh et al. (1986). Correcting for fluid salinity could raise or lower the higher values by less than 1.0 per mil (Rye and Sawkins, 1974; Truesdell, 1974). The 18Ofluid range of 4.2 to 5.2 per mil, determined for the Feitais deposit, is common for massive sulfide deposits (e.g., Huston, 1999), but considering the complexities of isotopic variation due to rock-water interaction at depth, and that Late Strunian seawater could be depleted by as much as 8 per mil (Taylor, 1997; Veizer et al., 1999), the primary source(s) of the oreforming fluids is (are) obscure. The D values of chlorite from the footwall (1 to +18; Munh et al., 1986) do not elucidate the source. The higher 18O values for quartz in the massive sulfide compared to stockwork veins could reflect the lower fluid temperatures of silica precipitation on the sea floor. Using the 18Ofluid values determined from stockwork quartz and, including data from Barriga and Kerrich (1984), an average 18O of 17.5 per mil for quartz in the massive sulfide, a temperature range for silica precipitation of 175 to 189C is estimated. This is probably the temperature of quartz precipitation during cooling of the mineralizing fluids responsible for the massive sulfide. The sulfur isotope analyses were obtained from small samples, but the large number of these is thought to give a reasonable indication of the isotopic composition of sulfur in the whole deposit. The 34S values between 5 and +5 per mil may reflect derivation from magmatic sulfur or that leached from subsurface magmatic rocks (Ohmoto and Goldhaber, 1997). A contribution from seawater sulfate reduced during deep circulation of seawater seems unlikely, as more positive values might be expected, given that Late Devonian seawater sulfate has values near 22 per mil (Strauss, 1997) and sulfate in the system, if originally present, would have been completely reduced. The abundance of negative 34S values in sulfides might have resulted from leaching of biogenically derived sulfides in older sedimentary rocks, but a major contribution from biogenically mediated seawater sulfate reduction, as in other VHMS deposits in the Iberian Pyrite Belt (Velasco et al., 1998), presumably at or near the sea floor,

13

TABLE 6. Strontium Isotope Ratios of Barite from Feitais Massive Sulfide Ore and Hanging-Wall Rocks Sample no. F71B F72 F73 F74 F75 F76 Drill hole section, m FS9975-880, 310.9 FS9975-880, 355.7 FS9954-680, 310.1 FS9954-680, 309.7 FS17-800, 417.6 FS9975-880, 374.4 Rock type Thin stratiform barite-bearing band (hanging-wall rock) Barite+carb+qtz veins in massive Zn ore Thin stratiform barite-bearing band (hanging-wall rock) Thin stratiform barite-bearing band (hanging-wall rock) Thick stratiform barite lens (hanging-wall rock) Barite+carbonate+qtz veins in massive pyritic ore
87Sr/86Sr

2 0.000013 0.000013 0.000014 0.000047 0.000013 0.000013

0.709063 0.708836 0.708851 0.709150 0.708438 0.708877

0361-0128/98/000/000-00 $6.00

261

262

INVERNO ET AL.

seems a more likely process, as proposed by Yamamoto et al. (1993). The proportion of negative 34S values in the massive sulfide body is similar to that in the Tharsis lenses (Tornos et al., 1998), but Feitais is unique among the worlds massive sulfide ores in having a high proportion of negative values in the stockwork veins (Table 3). The 34S values of stratiform barite in the hanging-wall volcaniclastic rocks testify to late circulation of seawater that had undergone little or no sulfate reduction. The higher 34S values in veins in the orebody, however, suggest that some sulfate reduction occurred by interaction with volcanic rocks during deeper seawater circulation, and/or by reaction with Fe2+ in the ore-forming fluid. The Sr isotope values of barite (0.7084380.709063) are more radiogenic than the range for Late Strunian seawater (about 0.70800.7085: Veizer et al., 1999) and considerably more radiogenic than those of stratigraphically equivalent felsic volcanic rocks in the Iberian Pyrite Belt in Spain (0.7033040.706642: Mitjavila et al., 1997). Sr isotope data from various ore types at Neves Corvo (0.7108810.73556) are more radiogenic than any Feitais values and suggest that the ore-forming fluids were highly radiogenic and probably of crustal origin (Relvas et al., 2001). The Feitais Sr isotope values could have resulted from mixing of such fluid with seawater, provided that the former contained negligible sulfate at that stage. Two of the Feitais quartz-carbonatebearing stockwork samples, F58 and F61 (Table 5), in the northwest end of the body, yielded 18O values of 18.2 and 13.9 per mil in dolomite and quartz, and 18.6 and 12.3 per mil in calcite and quartz, respectively, showing that carbonate and quartz are not in O isotope equilibrium (Taylor, 1997). As textural evidence shows that quartz (Q1 and Q2), carbonate, and chalcopyrite formed at much the same time (Fig. 7; Inverno and Solomon, 2001), there is a possibility that the carbonate O isotope compositions have been reset after deposition. The 18O values of the Feitais carbonates are typical of those in massive sulfide deposits but show a much wider range than in other VHMS deposits (Huston, 1999), suggesting the involvement of fluids with variable 18O. The spread cannot be explained by variation of temperatures because of the constraints provided by the Th values. Using the six analyzed stockwork carbonate samples with Th results (Tables 2, 5), and the mean temperature of the overall interval, 270 to 315C, for all the other stockwork samples, the 18O value of the fluid responsible for depositing carbonate is calculated as 4.6 to 17.2 per mil, using the calcite-H2O, dolomite-H2O, and Fe-dolomite-H2O oxygen isotope fractionation equations given in Longstaffe (1989). The highest 18Ofluid values (11.017.2), corresponding to carbonate samples with high 18O compositions (16.825.9: Table 5), do not match the highest 18Ofluid and corresponding 18O values for stockwork quartz (Table 4), indicating that O isotope compositions of the carbonates were partly reset after deposition. The 13C values of carbonates (14.2 to 7.5) are clearly lower than those of carbonates in most other VHMS deposits (Huston, 1999). For the same six analyzed stockwork carbonate samples with Th data used above, and assuming acidic (sericite stable) conditions and using the C isotope calciteCO2 fractionation equation (Chacko et al.,1991), the 13Cfluid is calculated as 9.2 to 6.6 per mil. For those carbonate
0361-0128/98/000/000-00 $6.00

samples without Th data, estimates using the mean temperature of the overall interval yield 13Cfluid values as low as 11.8 per mil, giving an overall range of 11.8 to 6.6 per mil ( Table 5). Carbonic acid in Late Strunian seawater had a 13C range of about 0 to 6 per mil (Veizer et al., 1999). Thus, the 13C values of carbonate could have been derived by mixing seawater carbonate with that derived from oxidation of methane resulting from degradation of organic carbon in underlying sedimentary rocks (organically derived carbon typically has 13C values of 20 to 30: Berger and Vincent, 1986), as suggested for other VHMS deposits (Huston, 1999), or by mixing of seawater carbonate with that of igneous derivation (13CCO2 values of 5 to 10: Ohmoto and Goldhaber, 1997). Environment and manner of sulfide deposition For the Feitais body, sea-floor deposition is indicated by its stratiform nature, the greater degree of hydrothermal alteration in the footwall rocks compared to that in the hangingwall rocks, the abrupt termination of the main stockwork system at the base of the massive sulfide, the absence of evidence for replacement, such as replacement fronts (Barriga, 1983), the presence of sedimentary features, such as stratigraphic layering and small-scale slumps and breccias, and colloform and other pyrite textures common in sea-floor deposits (e.g., Hokuroku basin ores: Eldridge et al., 1983; Hellyer: Solomon and Gaspar, 2001). Further, hydrothermal activity continued during deposition of the hanging-wall rocks, including the jasper layer (Schermerhorn, 1978; Barriga and Kerrich, 1984), because secondary sericite, chlorite, and carbonate, and enrichments in Fe, Mn, and Ba, extend into the lowermost 20 m of the Paraso Formation. Barriga (1983), Barriga and Fyfe (1988, 1991) and Carvalho et al. (1999) all proposed that the Feitais massive sulfide body formed beneath the chert and/or jasper layer, not by replacement of the underlying rocks but by deposition from oreforming fluid trapped below the layer. The justification for this theory was that the chert above the massive sulfide is converted to a silica-pyrite assemblage, there is no sedimentary dilution of the ore, there is no evidence of oxidation of the massive sulfide, and there is little or no evidence that sulfides replaced the volcanic rocks (Barriga, 1983; Barriga and Fyfe, 1991). However, the rock capping the massive sulfide is only locally jasper and/or chert, with the majority an altered feldspar-phyric, rhyolitic volcaniclastic rock, with jasper and/or chert lying up to 30 m above the top of the massive sulfide (Fig. 4). If the presence of this altered hanging-wall rock was due to synsedimentary or tectonic movements, as Barriga and Fyfe (1988) maintained, it would be unlikely that only that very same rock occurred between the jasper and/or chert layer and the massive sulfide orebody along and beyond its length, and that the faults responsible for this would not be evident. Sulfide deposition would presumably have taken place by conductive cooling, a slow process incompatible with the quench textures observed in the orebody. If the footwall stockwork veins filled hydraulic fractures (see below), then any immediately overlying rocks would also be fractured, and the ore-forming fluids released. A hydraulic origin is indicated because the veins appear to have opened episodically at right angles to their walls, and there is otherwise no displacement of rock or brecciation.

262

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES

263

Despite the presence of fragmental ore due to reworking of massive sulfide on the sea floor (Hutchinson, 1991), the absence of rubble mounds and lack of evidence of chimney formation and collapse, and the sheetlike form of the Feitais-Estao orebody suggest that processes observed in modern sea-floor massive sulfide deposits, and inferred for the Tertiary Hokuroku ores, were not significant in Feitais ore formation. The low barite content of the massive orebody, despite enrichment of Ba in some altered footwall and hangingwall rocks (Barriga, 1983; Barrett et al., 2008), suggests that seawater was involved to only a limited extent during ore formation, contrasting with the sulfate-rich kuroko of the Hokuroku basin that was formed by the mixing of ore-forming fluids with ambient seawater (Ohmoto, 1996). That sulfate was present in ambient seawater is indicated by the 34S values of barite near the top of and above the massive sulfides, and also by the negative 34S values of some sulfides. The Feitais orebody also differs from the mound deposits of the Hokuroku district, Japan, in its large size (3.2 Mt combined Zn + Pb + Cu at Moinho plus Feitais), high Fe/Cu ratio of the massive sulfides, reduced mineral assemblage (pyritearsenopyrite), vertical metal zoning, and degree of zone refining. These properties, together with the sheetlike form of the massive sulfides, lack of rubble mounds and chimney structures, and scarcity of barite, are shared by most other Iberian Pyrite Belt orebodies and particularly those for which a brine-pool origin has been postulated, based on the fluid inclusion data summarized for stockwork veins by Solomon et al. (2002, 2004), i.e., San Miguel, San Telmo, and Aguas Teidas Este within the northern part of the Iberian Pyrite Belt (Snchez-Espaa et al., 2000, 2003); Rio Tinto farther south (Nehlig et al., 1998; Snchez-Espaa et al., 2003); Masa Valverde (Toscano et al., 1997); and Aznalcllar and Los Frailes (Almodvar et al., 1998). The same characteristic features are also found in primary ore at Las Cruces, where the majority of fluid inclusions found in stockwork veins also contain fluids with reversing buoyancy (Knight, 2000), and also at Tharsis for which sulfide deposition in a basin has been proposed by Tornos et al. (2008). Dawson et al. (2001a, b) and Barrett et al. (2008) have also suggested that the Feitais body was deposited in a basin. Thirty-eight percent (n = 46) of the quartz, or carbonate, Th-salinity data lie at least slightly above the line determined for NaCl solutions by Turner and Campbell (1987) that separates reversing fluids and permanently buoyant fluids (Fig. 9, Table 2), assuming a seawater salinity of 3.2 wt percent NaCl equiv at 0C. If at some stage during mineralization the exhaling fluids had temperatures and salinities largely above the line then, assuming a basin existed, as proposed by Barrett et al. (2008), a brine pool would have developed and may well have survived significant subsequent lowering of input salinity, as discussed in the case of Hellyer by Solomon and Khin Zaw (1997). However, the fluid paragenesis and relative volumes of high and low salinities at Feitais are not known, and if data are taken collectively, the average Th and salinity values are 297.3C and 5.96 wt percent NaCl equiv, respectively (Fig. 9), too close to the reversing and buoyancy line to predict reliably their behavior. Sulfide deposition in the proposed brine pool was probably largely the result of rapid cooling or quenching of the exhaling
0361-0128/98/000/000-00 $6.00

fluids in the turbulent plume. The framboidal and colloform pyrite probably formed as a result of marked supersaturation of highly soluble Fe monosulfides that evolved to pyrite and framboidal pyrite (Butler and Rickard, 2000); the degree of supersaturation would have been highest near the brine-seawater interface. The lack of evidence for postdepositional oxidation of the massive sulfides might be due to rapid covering by younger rocks but is more likely due to protection by the cool, reduced, spent ore-forming fluids remaining in the basin after exhalation ceased. There is no evidence to suggest that ambient seawater was anoxicthere are no known coeval organic-bearing shales in the vicinity, and in the only study of ore-hosting shales in the Iberian Pyrite Belt, at Tharsis, it was found that prior to sulfide deposition conditions were oxic to dysoxic and only became anoxic during sulfide deposition, a change interpreted as due to establishment of a brine pool (Tornos et al., 2008). The jasperoid chert that is present in much of the Aljustrel area, locally immediately overlying the massive sulfide body (Figs. 2, 3; Barriga, 1983), provides evidence for oxic conditions following massive sulfide deposition at Feitais. Biogenic reduction of aqueous sulfate may have taken place in porous volcanic rocks prior to and during shallow subsurface entrainment of seawater, the resulting sulfur with negative 34S values being incorporated into the veins. Slow mixing with seawater would not have oxidized the rising fluid (Ohmoto et al., 1983). Alternatively, or additionally, reduced sulfur may have been entrained from the overlying brine pool in which sulfate reduction was taking place. Nehlig et al. (1998) showed that because the stockwork veins immediately beneath the San Dionisio massive sulfide at Rio Tinto are near horizontal, the minimum stress, 3, was probably vertical and 1 and 2 horizontal. The original orientation of the veins at Feitais is not clear due to subsequent deformation, but in relatively undeformed rocks they dip at shallow angles to the base of the massive sulfide, showing that 3 was probably near vertical. For new fractures the fluid pressure, Pf, must have exceeded Ph + Pl + T, where Ph is hydrostatic pressure due to the water column, Pl is lithostatic pressure, and T the strength of the rock, the latter probably relatively low due to hydrothermal alteration (Solomon and Groves, 1994; Nehlig et al., 1998). Fractures of this type generally die out at depth due to increasing (Ph + Pl + T), and because shallower rocks have generally lower T values due to their hydrothermally altered character. Following Nehlig et al. (1998), using their (measured) value for T, and assuming a water depth of 1 km to inhibit boiling, Pf at 50 m below the sea floor at Feitais must have been more than 23.6 MPa to have initiated new fractures. This is considerably greater than the likely overpressures in a simple convective circulation system, and as there is no evidence to suggest that boiling was important, fluid pressure created by subcritical fluids exsolved from silicate magmas seems the likely alternative, as suggested for Rio Tinto by Nehlig et al. (1998). Conclusions We conclude that the massive sulfide body at Feitais formed on the sea floor. The form of the massive sulfide lens, lack of rubble mounds and chimney structures, and also of barite, suggest deposition in a basin filled with ore-forming fluid. Although only 38 percent of the stockwork fluid inclusions are

263

264

INVERNO ET AL. Andrade, R.F., and Schermerhorn, L.J.G., 1971, Aljustrel and Gavio, in Carvalho, D., Goinhas, J.A.C., and Schermerhorn, L.J.G., eds., Main ore deposits in S. Portugal, I Congresso Hispano-Luso-Americano de Geologia Econmica, Madrid-Lisboa, September 1971: Lisboa, Direco-Geral de Minas e Servios Geolgicos, Guidebook to Field Trip 4, p. 3259 (in Portuguese). Badham, J.P.N., 1982, Further data on the formation of ores at Rio Tinto, Spain: Transactions of the Institution of Mining and Metallurgy, v. 91, p. B26B32. Bailey, S.W., 1980, Summary of recommendations of the AIPEA Nomenclature Committee: Canadian Mineralogist, v. 18, p. 143150. Barrett, T.J., Dawson, G.L., and MacLean, W., 2008, Volcanic stratigraphy, alteration, and sea-floor setting of the Paleozoic Feitais massive sulfide deposit, Aljustrel, Portugal: ECONOMIC GEOLOGY, v. 103, p. 215239. Barrie, C.T., Amelin, Y., and Pascual, E., 2002, U-Pb geochronology of VMS mineralization in the Iberian Pyrite Belt: Mineralium Deposita, v. 37, p. 684703. Barriga, F.J.A.S., 1983, Hydrothermal metamorphism and ore genesis at Aljustrel, Portugal: Unpublished Ph.D. thesis, London, Canada, University of Western Ontario, 386 p. Barriga, F.J.A.S., and Fyfe, W.S., 1988, Giant pyritic base-metal deposits: The example of Feitais (Aljustrel, Portugal): Chemical Geology, v. 69, p. 331343. 1991, Giant pyritic base-metal deposits: The example of Feitais (Aljustrel, Portugal)Reply: Chemical Geology, v. 90, p. 349352. 1998, Multi-phase water-rhyolite interaction and ore fluid generation at Aljustrel, Portugal: Mineralium Deposita, v. 33, p. 188207. Barriga, F.J.A.S., and Kerrich, R., 1984, Extreme 18O-enriched volcanics and 18O-evolved marine water, Aljustrel, Iberian Pyrite Belt: Transition from high to low Rayleigh number convective regimes: Geochimica et Cosmochimica Acta, v. 48, p. 10211031. Bayliss, P., 1975, Nomenclature of the trioctahedral chlorites: Canadian Mineralogist, v. 13, p. 178180. Berger, W.R., and Vincent, E., 1986, Deep-sea carbonates: Reading the carbon-isotope signal: Geologische Rundschau, v. 75, p. 249269. Bischoff, J.L., and Pitzer, K.S., 1989, Liquid-vapor relations for the NaClH2O system: Summary of the P-T-x surface from 300C to 500C: American Journal of Science, v. 289, p. 217248. Bischoff, J.L., and Rosenbauer, R.J., 1985, An empirical equation of state for hydrothermal seawater (3.2 percent NaCl): American Journal of Science, v. 285, p. 725763. Bodnar, R.J., 1993, Revised equation and table for determining the freezing point depression of H2O-NaCl solutions: Geochimica et Cosmochimica Acta, v. 57, p. 683684. Bodnar, R.J., Reynolds, T.J., and Kuehn, C.A., 1985, Fluid-inclusion systematics in epithermal systems: Reviews in Economic Geology, v. 2, p. 7397. Brun, J.P., and Burg, J.P., 1982, Combined thrusting and wrenching in the Ibero-Armorican arc: A corner effect during continental collision: Earth and Planetary Science Letters, v. 61, p. 319332. Butler, I.B., and Rickard, D., 2000, Framboidal pyrite formation via the oxidation of iron (II) monosulfide by hydrogen sulfide: Geochimica et Cosmochimica Acta, v. 64, p. 26652672. Carvalho, C.M.N., and Barriga, F.J.A.S., 2000, Preliminary report on the detailed mineralogical study at the alteration zones surrounding the Feitais VMS orebody [abs.]: Volcanic Environments and Massive Sulfide Deposits, International Conference and Field Meeting, November 2000, Tasmania, Australia, Program and Abstracts Volume, p. 1921. Carvalho, D., Goinhas, J., Oliveira, V., and Ribeiro, A., 1971, Insights on the geology of S. Portugal and their metallogenic consequences: Estudos Notas Trabalhos, Servio Fomento Mineiro, v. 20, p. 153199 (in Portuguese, with abstract in English). Carvalho, D., Conde, L., Hernandez Enrile, J.H., Oliveira, V., and Schermerhorn, L.J.G.S., 1976, Iberian Pyrite Belt, Guidebook for field trips in III Reunio do Sudoeste Peninsular no Macico Hesprico da Pennsula Ibrica, Huelva-Beja, 1975: Comunicaes Servios Geolgicos Portugal, v. 60, p. 271315 (in Portuguese). Carvalho, D., Barriga, F.J.A.S., and Munh, J., 1999, Bimodal-siliclastic systemsthe case of the Iberian Pyrite Belt: Reviews in Economic Geology, v. 8, p. 375408. Cathelineau, M., 1988, Cation site occupancy in chlorites and illites as a function of temperature: Clay Minerals, v. 23, p. 471485. Chacko, T., Mayeda, T.K., Clayton, R.N., and Goldsmith, J.R., 1991, Oxygen and carbon isotope fractionation between CO2 and calcite: Geochimica et Cosmochimica Acta, v. 55, p. 28672882.

of a type that would reverse buoyancy upon discharge, the similarity of the Feitais ore to other Iberian massive sulfide ores where such an origin has been postulated from the fluid inclusion data, suggests that sulfide deposition probably took place in a brine pool. This interpretation accounts for the lack of sulfide oxidation, the lack of evidence for more than minor involvement of ambient seawater, and the abundant negative 34S sulfide values in both massive sulfide and stockwork. Chloritization, sericitization, and carbonatization of the hanging-wall rocks, including the jasperoid chert layer, suggest that hydrothermal activity continued after massive sulfide deposition had ceased. The 13C and 18O values of carbonate and the 34S of barite in veins in the orebody suggest that coeval seawater was a component of the ore-forming fluids. The heat required to initiate hydrothermal convection was likely to have been derived from a magmatic source. Although the 34S values of sulfides from the orebody are compatible with involvement of modified seawater, they also allow a magmatic component and/or sulfur leached from underlying rocks. The Sr isotope data suggest mixed fluid sources, both seawater and of crustal origin. The higher of the fluid inclusion salinities may be of magmatic origin, as supported by the high Pf values estimated for the stockwork fluids. Acknowledgments C. Inverno acknowledges the financial support of Fundao para a Cincia e Tecnologia (FCT), Portugal, through a scholarship during his postdoctoral studies at the University of Tasmania in 2000 to 2001 under the supervision of the second author. This research was also supported by the Australian Research Councils Research Centres program. Various forms of support by the Centre for Ore Deposit Research (CODES) in Tasmania, the former Instituto Geolgico e Mineiro, and the current Instituto Nacional de Engenharia, Tecnologia e Inovao (INETI), and CREMINER FCT Research Center at Faculdade de Cincias, Universidade de Lisboa (FCUL) in Portugal, are also acknowledged. Eurozinc, namely through Adriano Barros, Joo Carlos Sousa, Garnet Dawson, Paulo Caessa, and Alberto Santos, is thanked for logistic and geologic support during field work in Minas de Aljustrel in 2000 to 2001. Keith Harris and Christine Cook (Central Science Laboratory, University of Tasmania) are acknowledged for sulfur isotope analyses of sulfides and barite and carbon and oxygen isotope analyses of carbonates. Alberto Verde (FCUL) is thanked for preparing a large number of different types of polished thin sections. Pedro Fal and Daniel Oliveira (INETI) helpfully drafted some figures and the latter also assisted in preparing digital photomicrographs. The authors are grateful for discussions with Ross Large, Bruce Gemmell, David Cooke, Garry Davidson, and Tony Crawford, all from CODES, Maria Ondina Figueiredo, from INETI, Antonella Zane, from Universit di Padova (Italy), and Mike Russell, University of Glasgow. Two Economic Geology reviewers, John Slack and Eric Marcoux, guest editor, Jan Peter, and editor Mark Hannington are acknowledged for improving the manuscript.
REFERENCES
Almodvar, G.R, Saez, R., Pons, J.M., Maestre, A., Toscano, M., and Pascual, E., 1998, Geology and genesis of the Aznalcllar massive sulphide deposits, Iberian Pyrite Belt, Spain: Mineralium Deposita, v. 33, p. 111136. 0361-0128/98/000/000-00 $6.00

264

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES Clayton, R.N., and Mayeda, T.K., 1963, The use of bromine pentafluoride in the extraction of oxygen from oxides and silicates for isotopic analysis: Geochimica et Cosmochimica Acta, v. 27, p. 4357. Clayton, R.N., ONeill, J.R., and Mayeda, T.K., 1972, Oxygen isotope exchange between quartz and water: Journal of Geophysical Research, v. 77, p. 30573067. Conde, C., Tornos, F., and Doyle, M., 2003, Encuadre estratigrfico de los sulfuros masivos de la parte suroriental de la faja piritiica: Aznalcllar-Los Frailes, y Las Cruces: Boletin Sociedad Espaola de Mineralogia, v. 26A, p. 161162. Conde, L.N., and Leito, J.R., 1984, Mining development project of Aljustrel. Geological report: Unpublished report, Aljustrel, Pirites Alentejanas, v. 12, 375 p. (in Portuguese). Conde, L.N., Leito, J.R., and Ferreira, A., 1986, Structure of Aljustrel mineralization [abs.]: Maleo, v. 2, no. 13, p. 1516 (in Portuguese). Crawford, M.L., 1981, Phase equilibria in aqueous fluid inclusions: Mineralogical Association of Canada Short Course Handbook, v. 6, p. 75100. Dawson, G.L., Barrett, T.J., Caessa, P., and Alverca, R., 2001a, The Feitais polymetallic massive sulphide deposit, southern Portugal [ext. abs.]: GEODE Workshop Massive Sulfide Deposits in the Iberian Pyrite Belt: New Advances and Comparison with Equivalent Systems, Aracena, Spain, October 2001, Proceedings, p. 15. Dawson, G.L., Caessa, P., Alverca, R., and Sousa, J.C., 2001b, Geology of the Aljustrel mine area, southern Portugal: GEODE Workshop Massive Sulfide Deposits in the Iberian Pyrite Belt: New Advances and Comparison with Equivalent Systems, Aracena, Spain, October 2001, Aljustrel, Eurozinc, Aljustrel Field Trip Guidebook, 28 p. de Ronde, C.E.J., 1995, Fluid chemistry and isotopic characteristics of seafloor hydrothermal systems and associated VMS deposits: Potential for magmatic contribution: Mineralogical Association of Canada Short Course Series, v. 23, p. 479509. Deer, W.A., Howie, R.A., and Zussman, J., 1992, An introduction to the rockforming minerals, 2nd ed.: Essex, Longman, 696 p. Dias, R., and Ribeiro, A., 1995, The Ibero-Armorican arc: A collisional effect against an irregular continent: Tectonophysics, v. 248, p. 113128. Drummond, S.E., and Ohmoto, H., 1984, Chemical evolution and mineral deposition in boiling hydrothermal systems: ECONOMIC GEOLOGY, v. 80, p. 126147. Eldridge, C.S., Barton, P.B., Jr., and Ohmoto, H., 1983, Mineral textures and their bearing on formation of the Kuroko orebodies: ECONOMIC GEOLOGY MONOGRAPH 5, p. 241281. Foden, J., Mawby, J., Kelley, S., Turner, S., and Bruce, D., 1995, Metamorphic events in the eastern Arunta inlier, Part 2. Nd-Sr-Ar isotopic constraints: Precambrian Research, v. 71, p. 207227. Foden, J., Barovich, K., Jane, M., and OHalloran, G., 2001, Sr-isotopic evidence for Late Neoproterozoic rifting in the Adelaide geosyncline at 586 Ma: Implications for a Cu ore-forming flux: Precambrian Research, v. 106, p. 291308. Fonseca, P., and Ribeiro, A., 1993, Tectonics of the Beja-Acebuches ophiolite: A major suture in the Iberian Variscan foldbelt: Geologisches Rundschau, v. 82, p. 440447. Fonseca, P., Munh, J., Pedro, J., Rosas, F., Moita, P., Arajo, A., and Leal, N., 1999, Variscan ophiolites and high-pressure metamorphism in southern Iberia: Ofioliti, v. 24, p. 259268. Gaspar, O.C., 1984, Microscopy applied to beneficiation of Aljustrel ores: Estudos Notas Trabalhos, Servio Fomento Mineiro, v. 26, p. 4961. 1996, Ore microscopy and petrology applied to the genesis, exploitation and beneficiation of Aljustrel and Neves Corvo massive sulphide deposits: Estudos Notas Trabalhos, Instituto Geolgico e Mineiro, v. 38, p. 3195 (in Portuguese, with extended abstract in English). 1997, Ore microscopy applied to the beneficiation of volcanogenic massive sulfides: Exploration and Mining Geology, v. 6, p. 335348. Gaspar, O.C., and Conde, L., 1978, Characterization of Aljustrel sulphides in view of their full use: Porto, Congresso da Ordem dos Engenheiros, Tema 3, Comunicao 12, 23 p. (in Portuguese). Goldstein, R.H., and Reynolds, T.J., 1994, Systematics of fluid inclusions in diagenetic minerals: SEPM (Society for Sedimentary Geology) Short Course 31, 214 p. Huston, D.L., 1999, Stable isotopes and their significance for understanding the genesis of volcanic-hosted massive sulfide deposits: A review: Reviews in Economic Geology, v. 8, p. 157179. Huston, D.L., Power, M., Gemmell, J.B., and Large, R.R., 1995, Design, calibration and geological application of the first operational Australian laser 0361-0128/98/000/000-00 $6.00

265

ablation sulphur isotope microprobe: Australian Journal of Earth Sciences, v. 42, p. 549555. Hutchinson, R.W., 1991, Giant pyritic base-metal deposits: The example of Feitais (Aljustrel, Portugal)comments: Chemical Geology, v. 90, p. 343348. Inverno, C.M.C., and Solomon, M., 2001, New mineralogical and fluid inclusion evidence from the Feitais stockwork, Aljustrel, Iberian Pyrite Belt, Portugal [ext. abs.]: GEODE Workshop Massive Sulfide Deposits in the Iberian Pyrite Belt: New Advances and Comparison with Equivalent Systems, Aracena, Spain, October 2001, Proceedings, p. 2830. Inverno, C.M.C., Solomon, M., and Barton, M., 2002, New sulfur, carbon and oxygen isotope evidence from the Feitais stockwork, Aljustrel, Iberian Pyrite Belt, Portugal [ext. abs.]: GEODE Study Center on Geodynamics and Ore Deposit Evolution, Grenoble, France, October 2002, Proceedings, 3 p. Jaques, L., and Noronha, F., 2002, Mineralizing fluid characteristics at the Lagoa Salgada orebody, Iberian Pyrite Belt: Comunicaes Instituto Geolgico Mineiro, v. 89, p. 209224 (in Portuguese, with abstract in English). Knight, F.C., 2000, The mineralogy, geochemistry and genesis of the secondary mineralisation of Las Cruces, Spain: Unpublished Ph.D. thesis, University of Cardiff (University of Wales), 434 p. Kranidiotis, P., and MacLean, W.H., 1987, Systematics of chlorite alteration at the Phelps Dodge massive sulfide deposit, Matagami, Quebec: ECONOMIC GEOLOGY, v. 82, p. 18981911. Leistel, J.M., Marcoux, E., and Deschamps, Y., 1998, Chert in the Iberian Pyrite Belt: Mineralium Deposita, v. 33, p. 5981. Leito, J.C., 1992, The Aljustrel overthrust problem in view of the new evidence from the Sto. Anto area: Comunicaes Servios Geolgicos Portugal, v. 78, p. 97102. 1997, Geology of the Aljustrel massive sulphide deposits: Society of Economic Geologists Field Trip Guidebook Series, v. 27, p. 8297. Longstaffe, F.J., 1989, Stable isotopes as tracers in clastic diagenesis: Mineralogical Association of Canada Short Course 15, p. 201284. Marcoux, E., Molo, Y., and Leistel, J.M., 1996, Bismuth and cobalt indicators of stringer zones to massive sulphide deposits, Iberian Pyrite Belt: Mineralium Deposita, v. 31, p. 126. Massano, C.M.R., Relvas, J.M.R.S., and Barriga, F.J.A.S., 1991, Na-bearing sericite as a guide to volcanogenic massive sulphide deposits: New data from the Gavio area (South Portugal) [abs.]: Canadian Mineralogist (insert of GAC-MAC Joint Annual Meeting Program with Abstracts [v. 16], Toronto, May 1991), v. 29 (1), p. A81. Matsuhisa, Y., Goldsmith, H.J.R., and Clayton, R.N., 1979, Oxygen isotope fractionation in the system quartz-albite-anorthite-water: Geochimica et Cosmochimica Acta, v. 43, p. 11311140. Matthews, A., and Beckinsale, 1979, Oxygen isotope equilibrium systematics between quartz and water: American Mineralogist, v. 64, p. 232240. McCrea, J.M., 1950, The isotope chemistry of carbonates and a paleotemperature scale: Journal of Chemistry and Physics, v. 18, p. 849857. Mitjavila, J., Marti, J., and Soriano, C., 1997, Magmatic evolution and tectonic setting of Iberian Pyrite Belt volcanism: Journal of Petrology, v. 38, p. 727755. Mitsuno, C., Nakamura, T., Yamamoto, M., Kase, K., Oho, M., Suzuki, S., Thadeu, D., Carvalho, D., and Arribas, A., 1988, Geological studies of the Iberian Pyrite Beltwith special reference to its genetical correlation of the Yanahara ore deposit and others in the inner zone of southwest Japan: Japan, University of Okayama, 300 p. Moura, A.J.N., 2002, Characterization of fluids associated with Neves Corvo mine mineralizations, Portugal: Unpublished Ph.D. thesis, Porto, Portugal, Universidade do Porto, 399 p. (in Portuguese, with abstract in English). Moura, A., Noronha, F., Cathelineau, M., Boiron, and M-C., Ferreira, A., 1997, Evidence of metamorphic fluid migration within the Neves-Corvo ore deposit [abs.]: The fluid inclusion data: Society of Economic Geologists Neves Corvo Field Conference, Abstracts and Program, Lisbon, May 1997, p. 92. Munh, J., 1983, Hercynian magmatism in the Iberian Pyrite Belt: Memrias Servios Geolgicos Portugal 29, p. 3981. 1990, Metamorphic evolution of the South Portuguese/Pulo do Lobo zone, in Dallmeyer, R.D., and Martinez Garcia, E., eds., Pre-Mesozoic geology of Iberia: Berlin, Springer-Verlag, p. 363368. Munh, J., and Kerrich, R., 1980, Sea water-basalt interaction in spilites from the Iberian Pyrite Belt: Contributions to Mineralogy and Petrology, v. 73, p. 191200.

265

266

INVERNO ET AL. Rye, R.O., and Sawkins, F.J., 1974, Fluid inclusion and stable isotope studies on the Casapalca Ag-Pb-Zn-Cu deposit, Central Andes, Peru: ECONOMIC GEOLOGY, v. 69, p. 181205. Snchez-Espaa, J., Velasco, F., and Yusta, I., 2000, Hydrothermal alteration of felsic volcanic rocks associated with massive sulphide deposition in the northern Iberian Pyrite Belt (SW Spain): Applied Geochemistry, v. 15, p. 12651290. Snchez-Espaa, J., Velasco, F., Boyce, A.J., and Fallick, A.E., 2003, Source and evolution of ore-forming hydrothermal fluids in the northern Iberian Pyrite Belt massive sulphide deposits (SW Spain): Evidence from fluid inclusions and stable isotopes: Mineralium Deposita, v. 38, p. 519537. Schermerhorn, L.J.G., 1978, Epigenetic magnesian metasomatism or syngenetic chloritite metamorphism at Falun and Orijrvi: Transactions of the Institution of Mining and Metallurgy, v. 87, p. B162B167. Schermerhorn, L.J.G., and Stanton, W.I., 1969, Folded overthrusts at Aljustrel (South Portugal): Geological Magazine, v. 106, p. 130141. Schermerhorn, L.J.G., Zbyszewski, G., and Veiga Ferreira, O., 1987, Explanatory news of Sheet 42D (Aljustrel), Geological map of Portugal, scale 1:50,000: Lisboa, Servios Geolgicos de Portugal, 55 p. (in Portuguese). Scott, S.D., 1997, Submarine hydrothermal systems and deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York, John Wiley and Sons, p. 797875. Sheperd, T.J., Rankin, A.H., and Alderton, D.H.M., 1985, A practical guide to fluid inclusion studies: Glasgow, Blackie & Son Ltd., 239 p. Silva, J.B., Oliveira, J.T., and Ribeiro, A., 1990, Structural outline, in Dallmeyer, R.D., and Martnez Garca, E., eds., Pre-Mesozoic geology of Iberia: Berlin, Springer-Verlag, p. 348362. Solomon, M., and Gaspar, O.C., 2001, Textures of the Hellyer volcanichosted massive sulfide deposit, Tasmania: The aging of sulfide sediment on the sea floor: ECONOMIC GEOLOGY, v. 96, p. 15131534. Solomon, M., and Groves, D.I., 1994, The geology and origin of Australias mineral deposits: Oxford, Oxford University Press, Monographs on Geology and Geophysics, 24, 951 p. Solomon, M., and Khin Zaw, 1997, Formation of the Hellyer volcanogenic massive sulphide deposit on the sea floor: ECONOMIC GEOLOGY, v. 92, p. 686695. Solomon, M., and Quesada, C., 2003, Zn-Pb-Cu massive sulfide deposits: Brine pool types occur in collisional orogens, black smoker types occur in backarc/arc basins: Geology, v. 31, p. 10291032. Solomon, M., Tornos, F., and Gaspar, O.C., 2002, A possible explanation for many of the unusual features of the massive sulfide deposits of the Iberian Pyrite Belt: Geology, v. 30, p. 8790. Solomon, M., Tornos, F., Large, R.R., Badham, J.N.P., Both, R.A., and Khin Zaw, 2004, Zn-Pb-Cu volcanic-hosted massive sulfide deposits: Criteria for distinguishing brine pool- from black smoker-type sulfide deposition: Ore Geology Reviews, v. 25, p. 259283. Sousa, J.C., 1999, Aljustrel mine complex: The Iberian Pyrite Belt, Joint SGA-IAGOD International Meeting, August-September 1999: Instituto Tecnolgico Geominero de Espaa-Instituto Geolgico e Mineiro, Field Trip B4 Guide, p. 1820. Strauss, H., 1997, The isotopic composition of sedimentary sulfur through time: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 132, p. 97118. Taylor, H.P., Jr., 1997, Oxygen and hydrogen isotope relationships in hydrothermal ore deposits, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York, John Wiley and Sons, p. 229302. Tornos, F., Clavijo, E.G., and Spiro, B.F., 1998, The Filon Norte orebody (Tharsis, Iberian Pyrite Belt): A proximal low-temperature shale-hosted massive sulfide in a thin-skinned tectonic belt: Mineralium Deposita, v. 33, p. 105169. Tornos, F., Solomon, M., Conde, C., and Spiro, B.F., 2008, Formation of the Tharsis massive sulfide deposit, Iberian Pyrite Belt: Geologic, lithogeochemical, and stable isotope evidence for deposition in a brine pool: ECONOMIC GEOLOGY, v. 103, p. 185214. Toscano, M., Sez, R., and Almodvar, G.R., 1997, Multi-scale fluid evolution in the Masa Valverde stockwork (Iberian Pyrite Belt): Evidence from fluid inclusions [abs.]: Society of Economic Geologists Neves Corvo Field Conference, Abstracts and Program, Lisboa, May 1997, p. 101. Truesdell, A.H., 1974, Oxygen isotope activities and concentrations in aqueous solutions at elevated temperatures: Consequences for isotope geochemistry: Earth and Planetary Science Letters, v. 23, p. 387396. Turner, F.J., and Campbell, I.H., 1987, Temperature, density and buoyancy fluxes in black smoker plumes, and the criterion for buoyancy reversal: Earth and Planetary Science Letters, v. 86, p. 8592.

Munh, J., Barriga, F.J.A.S., and Kerrich, R., 1986, High 18O ore-forming fluids in volcanic-hosted base metal massive sulfide deposits: Geologic, 18O/16O, and D/H evidence from the Iberian Pyrite Belt, Crandon, Wisconsin, and Blue Hill, Maine: ECONOMIC GEOLOGY, v. 81, p. 530552. Nehlig, P., Cassard, D., and Marcoux, E., 1998, Geometry and genesis of feeder zones of massive sulphide deposits: Constraints from the Rio Tinto ore deposit (Spain): Mineralium Deposita, v. 33, p. 137149. Ohmoto, H., 1996, Formation of volcanogenic massive sulfide deposits: Ore Geology Reviews, v. 10. p. 135177. Ohmoto, H., and Goldhaber, M.B., 1997, Sulfur and carbon isotopes, in Barnes, H.L., ed., Geochemistry of hydrothermal ore deposits, 3rd ed.: New York, John Wiley and Sons, p. 517611. Ohmoto, H., Drummond, S.E., Eldridge, C.S., Pisutha-Arnond, V., and Lenagh, T.C., 1983, Chemical processes of ore formation: ECONOMIC GEOLOGY MONOGRAPH 5, p. 570604. Oliveira, J.T., 1990, South Portuguese zone, in Dallmeyer, R.D., and Martinez Garcia, E., eds., Pre-Mesozoic geology of Iberia: Berlin, SpringerVerlag, p. 333346. Oliveira, J.T., and Quesada, C., 1998, A comparison of stratigraphy, structure, and paleogeography of the South Portuguese zone and south-west England, European Variscides: Geoscience in South-west England, v. 9, p. 141150. Oliveira, J.T., Pereira, Z., Carvalho, P., Pacheco, N., and Korn, D., 2004, Stratigraphy of the tectonically imbricated lithological succession of the Neves Corvo mine area, Iberian Pyrite Belt, Portugal: Mineralium Deposita, v. 39, p. 422436. Onzime, J., Charvet, J., Faure, M., Bourdier, J.L., and Chauvet, A., 2003, A new geodynamic interpretation of the South Portuguese zone (SW Iberia) and the Iberian Pyrite Belt genesis: Tectonics, v. 22, p. 117. Pereira, Z., Saez, R., Pons, J.M., Oliveira, J.T., and Moreno, C., 1996, Edad devonica (struniense) de las mineralizaciones de Aznalcllar (Faja Piritica Iberica) en base a palinologia: Geogaceta, v. 20, p. 16091612. Peter, J.M., and Scott, S.D., 1993, Fluid inclusion and light stable isotope geochemistry of the Windy Craggy Besshi-type massive sulfide deposit, northwestern British Columbia: Resource Geology Special Issue, v. 17, p. 229248. Pisutha-Arnond, V., and Ohmoto, H., 1983, Thermal history, and chemical and isotopic compositions of the ore-forming fluids responsible for the Kuroko massive sulfide deposits in the Hokuroko district of Japan: ECONOMIC GEOLOGY MONOGRAPH 5, p. 523558. Potter, R.W., II, 1977, Pressure corrections for fluid-inclusion homogenization temperatures based on the volumetric properties of the system NaCl-H2O: U.S. Geological Survey Journal of Research, v. 5, p. 603607. Potter, R.W., II, and Brown, D.L., 1977, Preliminary steam tables for NaCl solutionsthe volumetric properties of aqueous sodium chloride solutions from 0C to 500C at pressures up to 2000 bars based on a regression of available data in the literature: U.S. Geological Survey Bulletin 1421C, 36 p. Quesada, C., 1998, A reappraisal of the structure of the Spanish segment of the Iberian Pyrite Belt: Mineralium Deposita, v. 33, p. 3144. Quesada, C., Fonseca, P.E., Munh, J., Oliveira, J.T., and Ribeiro, A., 1994, The Beja-Acebuches ophiolite (southern Iberia Variscan fold belt): Geological characterization and geodynamic significance: Boletin Geolgico y Minero, v. 105, p. 349. Reed, M.H., and Spycher, N. F., 1985, Boiling, cooling, and oxidation in epithermal systems: A numerical modeling approach: Reviews in Economic Geology, v. 2, p. 249272. Relvas, J.M.R.S., Massano, C.M.R., and Barriga, F.J.A.S., 1990, Ore zone hydrothermal alteration around the Gavio orebodies [ext. abs.]: Implications for exploration in the Iberian Pyrite Belt: Lisboa, Universidade de Lisboa, VIII Semana de Geoquimica, December 1990, 2 p. Relvas, J.M.R.S., Tassinari, C.C. G., Munh, J., and Barriga, F.J.A.S., 2001, Multiple sources for ore-forming fluids in the Neves Corvo VHMS deposit of the Iberian Pyrite Belt (Portugal): Strontium, neodymium and lead isotope evidence: Mineralium Deposita, v. 36, p. 416427. Roedder, E., 1984, Fluid inclusions: Reviews in Mineralogy, v. 20, 646 p. Roedder, E., and Bodnar, R.J., 1980, Geologic pressure determinations from fluid inclusion studies: Annual Review of Earth and Planetary Sciences, v. 8, p. 263301. Rosa, D.R.N., Inverno, C.M.C., Oliveira, V., and Rosa, C.J.P., 2004, Geochemistry of the volcanic rocks of the Albernoa area, Iberian Pyrite Belt, Portugal: International Geology Review, v. 46, p. 366383. 0361-0128/98/000/000-00 $6.00

266

FEITAIS VHMS DEPOSIT, ALJUSTREL, PORTUGAL: MINERALOGY, FLUID INCLUSIONS, AND STABLE ISOTOPES Urabe, T., Scott, S.D., and Hattori, K., 1983, A comparison of footwall-rock interactions and geothermal systems beneath some Japanese and Canadian volcanogenic massive sulfide deposits: ECONOMIC GEOLOGY MONOGRAPH 5, p. 345364. Veizer, J., Ala, D., Azmy, K., Brucksehen, P., Buhl, D., Bruhn, F., Carden, G.A.F., Diener, A., Ebneth, S., Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., and Strauss, H., 1999, 87Sr/86Sr, 13C and 18O evolution of Phanerozoic seawater: Chemical Geology, v. 161, p. 5988. Velasco, F., Snchez-Espaa, J., Boyce, A.J., Fallick, A.E., Sez, R., and Almodvar, G.R., 1998, A new sulphur isotopic study of some Iberian Pyrite Belt deposits: Evidence of a textural control on sulphur isotope composition: Mineralium Deposita, v. 34, p. 418. Walshe, J.L., 1986, A six-component chlorite solid solution model and the conditions of chlorite formation in hydrothermal and geothermal systems: ECONOMIC GEOLOGY, v. 81, p. 681703.

267

Wenner, D.B., and Taylor, H.P., Jr., 1971, Temperatures of serpentinization of ultramafic rocks based on 18O/16O fractionation between co-existing serpentine and magnetite: Contributions to Mineralogy and Petrology, v. 32, p. 165185. Yamamoto, M., Kase, K., Carvalho, D., Nakamura, T., and Mitsuno, C., 1993, Ore mineralogy and sulfur isotopes of the volcanogenic massive sulfide deposits in the Iberian Pyrite Belt: Resource Geology Special Issue, v. 15, p. 6780. Zane, A., and Weiss, Z., 1998, A procedure for classifying chlorites based on microprobe data: Rendiconti Lincei-Accademia dele Szienze Fisiche e Naturali, s. 9, v. 9, p. 5156.

0361-0128/98/000/000-00 $6.00

267

0361-0128/98/000/000-00 $6.00

268

Potrebbero piacerti anche