The Jabali nonsulfide Zn–Pb–Ag deposit, western Yemen Ore
Transcription
The Jabali nonsulfide Zn–Pb–Ag deposit, western Yemen Ore
Ore Geology Reviews 61 (2014) 248–267 Contents lists available at ScienceDirect Ore Geology Reviews journal homepage: www.elsevier.com/locate/oregeorev The Jabali nonsulfide Zn–Pb–Ag deposit, western Yemen N. Mondillo a,⁎, M. Boni a, G. Balassone a, M. Joachimski b, A. Mormone c a b c Dipartimento di Scienze della Terra, dell'Ambiente e delle Risorse, Università degli Studi di Napoli Federico II, Via Mezzocannone 8, 80134 Napoli, Italy GeoZentrum Nordbayern, University of Erlangen-Nuremberg, Schlossgarten 5, 91054 Erlangen, Germany INGV Osservatorio Vesuviano, Via Diocleziano 328, 80124 Napoli, Italy a r t i c l e i n f o Article history: Received 19 August 2013 Received in revised form 30 January 2014 Accepted 5 February 2014 Available online 14 February 2014 Keywords: Yemen Nonsulfide deposit Pb–Zn–Ag Mineralogy Stable isotopes a b s t r a c t The Jabali Zn–Pb–Ag deposit is located about 110 km east of Sana'a, the capital of Yemen, along the western border of the Marib-Al-Jawf/Sab'atayn basin. The economic mineralization at Jabali is a nonsulfide deposit, consisting of 8.7 million tons at an average grade of 9.2% zinc, derived from the oxidation of primary sulfides. The rock hosting both primary and secondary ores is a strongly dolomitized carbonate platform limestone of the Jurassic Shuqra Formation (Amran Group). The primary sulfides consist of sphalerite, galena and pyrite/ marcasite. Smithsonite is the most abundant economic mineral in the secondary deposit, and is associated with minor hydrozincite, hemimorphite, acanthite and greenockite. Smithsonite occurs as two main generations: smithsonite 1, which replaces both host dolomite and sphalerite, and smithsonite 2, occurring as concretions and vein fillings in the host rock. At the boundary between smithsonite 1 and host dolomite, the latter is widely replaced by broad, irregular bands of Zn-bearing dolomite, where Zn has substituted for Mg. The secondary mineralization evolved through different stages: 1) alteration of original sulfides (sphalerite, pyrite and galena), and release of metals in acid solutions; 2) alteration of dolomite host rock and formation of Zn-bearing dolomite; 3) partial dissolution of dolomite by metal-carrying acid fluids and replacement of dolomite and Zn-bearing dolomite by a first smithsonite phase (smithsonite 1). To this stage also belong the direct replacement of sphalerite and galena by secondary minerals (smithsonite and cerussite); 4) precipitation of a later smithsonite phase (smithsonite 2) in veins and cavities, together with Ag- and Cd-sulfides. The δ18O composition of Jabali smithsonite is generally lower than in other known supergene smithsonites, whereas the carbon isotope composition is in the same range of the negative δ13C values recorded in most supergene nonsulfide ores. Considering that the groundwaters and paleo-groundwaters in this area of Yemen have negative δ18O values, it can be assumed that the Jabali smithsonite precipitated in different stages from a combination of fluids, possibly consisting of local groundwaters variably mixed with low-temperature hydrothermal waters. The carbon isotope composition is interpreted as a result of mixing between carbon from host rock carbonates and soil/atmospheric CO2. The most favorable setting for the development of the Jabali secondary deposit could be placed in the early Miocene (~17 Ma), when supergene weathering was favored by major uplift and exhumation resulting from the main phase of Red Sea extension. Low-temperature hydrothermal fluids may have also circulated at the same time, through the magmatically-induced geothermal activity in the area. © 2014 Elsevier B.V. All rights reserved. 1. Introduction The Jabali zinc deposit is located at an altitude of 1800 m above sea level in a mountainous desert terrain along the western border of the Marib-Al-Jawf/Sab'atayn basin, about 110 km east of Sana'a, the capital of Yemen. It represents the most significant base metal deposit of Yemen (Grist, 2006; Watts, Griffis, and McOuat, 1993; Yemen Geological Survey and Mineral Resources Board, 1994, 2009). The artisanal mine workings in the area are thought to be over 2500 years old. The old workings for silver and lead were extended over an area of about ⁎ Corresponding author. Tel.: +39 081 2535063. E-mail address: nicola.mondillo@unina.it (N. Mondillo). http://dx.doi.org/10.1016/j.oregeorev.2014.02.003 0169-1368/© 2014 Elsevier B.V. All rights reserved. 10 h, tracing cavities filled by relatively soft oxidized ore, locally rich in silver. The ore was processed on site. Still existing waste dumps contain about 120,000 t, at average grades of 24% Zn, 3.5% Pb, and 160 ppm Ag. The artisanal metallurgical process was not effective, since slags still contain 23% Zn, 6.5% Pb, and 40 ppm Ag (SRK Consulting, 2005). The Jabali site was re-discovered by the Bureau de Recherches Géologiques et Minières (BRGM) and the Yemen Geological Survey and Mineral Resources Board (YGSMRB) in 1980. Between 1981 and 1986, an exploration and evaluation program, based on 57 drill holes, reported an accessible-by-open-pit resource of 3.0 Mt, at 15.2% Zn, and an amenable-by-underground mining tonnage of 1.24 Mt at 13% Zn. The exploration was mainly aimed to nonsulfide ores. During and Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 after the period of exploration at Jabali, BRGM and YGSMRB produced several reports, a few scientific papers on the characterization and genesis of the deposit (Al Ganad et al., 1994; Christmann et al., 1989) and a PhD thesis (Al Ganad, 1991). No further scientific research papers were published in more recent years, with the exception of a preliminary mineralogical study by Mondillo et al. (2011). In June 1993, Watts, Griffis and McOuat (WGM) issued a prefeasibility study indicating reserves of 3.6 Mt at 16.4% Zn. In 1996, Minorco and Ansan Wikfs were granted a license over the deposit. In 1998, ZincOx Resources – whose main interest was in nonsulfides – entered into a joint venture with Minorco and Ansan Wikfs, thus becoming the manager and operator for the development of the deposit. In 2004–2005, the exploitation and development rights of the Jabali zinc deposit were owned by Jabal Salab Company (Yemen) Ltd., a company in which ZincOx previously had a 52% interest, with the balance (48%) held by Ansan Wikfs Investments Ltd. In that period, ZincOx, together with SRK Consulting, concluded the feasibility study of the deposit, and reported resources consisting of 12.6 Mt at 8.9% Zn, 1.2% Pb, and 68 ppm Ag. In March 2013, due to the realignment of the ZincOx strategy towards recycling instead of exploitation of primary natural resources (ZincOx press release, 2013), ZincOx sold its interest in the Jabali Project to Ansan Wikfs, which actually owns the deposit. The feasibility study on the Jabali deposit considered a number of processing routes, owing to its rather complex mineralogical association. The metallurgical problems encountered over the years are essentially a consequence of the incomplete knowledge of the mineralogy and genesis of the Jabali nonsulfide orebody. In the present study, after a brief review on the primary sulfide mineralization, we focus on the mineralogical, geochemical and isotopic characterization of the nonsulfides and their relationships to the host rock, with the aim of identifying the supergene oxidation processes. 2. Regional geology The geology of Yemen is characterized by (i) Precambrian basement transected by a failed Jurassic rift system forming during the break-up of the Gondwana supercontinent; (ii) Jurassic pre-, syn-, and post-rift carbonate and siliciclastic sedimentary sequences; and (iii), Tertiary to recent sedimentary and magmatic rocks associated with the opening of the Gulf of Aden-Red Sea rift (Fig. 1) (Menzies et al., 1994). 249 The basement rocks in Yemen are considered as part of the Proterozoic Arabian–Nubian Shield, which covers northeast Africa and the Arabian Peninsula (Whitehouse et al., 2001). They consist of metavolcanic and metasedimentary suites, deposited in arc environments, deformed and intruded by post-tectonic granites and granodiorites during the Pan-African orogeny (Whitehouse et al., 1998). The southern Arabian Peninsula underwent post-orogenic extension accompanied by magmatism. This was associated with uplift and erosion of the basement rocks, followed by the formation of several basins related to major wrench-fault systems (e.g. Najd fault-system; Ellis et al., 1996). The basins are filled with Paleozoic to early Mesozoic siliciclastic sequences, deposited in marine epicontinental to deltaic environments, locally containing evaporites (Beydoun, 1997). From Triassic to middle Jurassic, Yemen was part of the Afro-Arabian plate of western Gondwanaland. The break-up of Gondwana initiated the extensional tectonics in Yemen and the formation of normal faults oriented along tectonic trends inherited from the Precambrian wrench-fault systems (Ahlbrandt, 2002; Bosence, 1997 and references therein). The Mesozoic extensional tectonic regime resulted in the formation of the Siham-Ad-Dali', Sab'atayn, Say'un–Masilah, Balhaf, and Jiza'–Qamar basins (Fig. 1) (As-Saruri et al., 2010). The Sab'atayn and the Say'un–Masilah basins are the only hydrocarbons producing basins in Yemen; the Jabali mineral concentrations are located at the western border of the Sab'atayn basin. Pre-rift sedimentation within the basins (Toarcian–Bathonian) is initially represented by the continental fluviatile red bed sediments of the Kuhlan Formation. The red beds grade upward into shallowmarine sediments, representing the early transgressive phase of the Middle Jurassic sea (Bathonian–Callovian) (Beydoun et al., 1998). The most widespread Jurassic marine deposits correspond to the Amran Group, which consists of pre-rift sediments, as well as of syn- and post-rift sequences (Csato et al., 2001). Beydoun et al. (1998), in a review on the stratigraphy of Yemen, established the “official” subdivision of the Amran Group. They distinguished: (i) the Shuqra Formation (Callovian–Oxfordian); (ii) the Madbi Formation (Kimmeridgian– Tithonian); (iii) the Sab'atayn Formation; and (iv) the Naifa Formation (upper Tithonian–Berriasian). These formations have been further subdivided into several members, based on the heteropic facies of the sediments (Beydoun et al., 1998). After Csato et al. (2001) the Shuqra Formation was deposited in a pre-rift regime, whereas the Madbi and Fig. 1. Geological map of Yemen, showing the location of Jabali (Yemen Geological Survey and Mineral Resources Board, 2009, modified). Hydrocarbon producing Mesozoic basins: 1 =Sab'atayn basin; 2 = Say'un–Masilah basin. Author's personal copy 250 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Sab'atayn Formations have the characteristics of syn-rift successions. The Naifa Formation is interpreted as a post-rift sedimentary sequence. In the following descriptions of the sedimentary sequences, we will keep the original terminology of the authors quoted in the literature (Al Ganad et al., 1994; As-Saruri et al., 2010; Beydoun et al., 1998; Youssef, 1998). The Shuqra Formation, which is the main host of the Jabali ores, is characterized by the following lithological units: 1) basal intertidal limestones; 2) marls and thinly bedded carbonaceous biomicrites; 3) foraminiferal biomicrite with chert nodules; 4) dolomitic marls interbedded with dolostones, and oolitic–oncolitic limestones interbedded with fossiliferous limestones and bioclastic sandstones; 5) coral–algal stromatolitic limestones and black foraminiferal biomicrites (Youssef, 1998). The Madbi Formation consists of bituminous marine shales, deltaic sandstones, debris flow breccias, well-bedded limestones, and turbidites (As-Saruri et al., 2010). The Sab'atayn Formation consists mainly of evaporites (Beydoun et al., 1998). In the Naifa Formation dolostones, dolomitic shales, and fine-grained limestones occur, locally yielding a rich ammonite fauna (Beydoun et al., 1998; Menzies et al., 1994). In the Sab'atayn and Say'un–Masilah basins, the siliciclastic and carbonate members of the Madbi Formation represent the main onshore petroleum reservoirs of Yemen (Ahlbrandt, 2002; As-Saruri et al., 2010). The oil generation may have started in the Late Cretaceous (Csato et al., 2001). Cretaceous sediments, ranging in age from Berriasian to Maastrichtian (Tawilah Group), completed the infilling of the Jurassic graben (Beydoun et al., 1998). In western Yemen, the main Cretaceous lithofacies consist of medium- to very coarse-grained, cross-bedded sandstones, interrupted by numerous paleosols. The total thickness of the Tawilah Group can vary from a few hundred meters in the Sana'a area to a few thousand meters in the Sab'atayn basin (Beydoun et al., 1998). A second important rifting phase was associated with the opening of the Red Sea and the Gulf of Aden during Oligocene and Miocene. At that time, widespread volcanism followed by uplift and denudation affected the whole of western Yemen, whereas the eastern part of the country was characterized by continuous sedimentation and complete absence of magmatic activity (Menzies et al., 1994). The volcanism was associated with the upwelling of the Afar plume beneath the Africa-Arabian plate (Bosworth et al., 2005). The Oligocene igneous rocks, known as the Yemen volcanic group or Yemen trap series (Moseley, 1969), consisting of more than 3 km-thick flood basalts to massive ignimbrites, erupted between ~31 Ma and ~29 Ma, and between ~29 Ma and 26 Ma (Baker et al., 1996; Coulié et al., 2003). A stratigraphic gap (~ 26 to ~ 19 Ma), marked by an unconformity at the top of the deformed Yemen volcanic group, indicates that the main rifting phase started in late Oligocene or early Miocene (Ukstins et al., 2002). Early Miocene magmatic rocks consist of small plutonic bodies (age ~22–21 Ma), and abundant mafic and felsic dykes (age ~ 25–16 Ma), both emplaced along the Red Sea margin (Zumbo et al., 1995). Pliocene to Quaternary volcanism is also fairly widespread in the areas NW of Sana'a–Amran, Dhamar–Rada, Marib–Sirwah, Balhalf–Bir Ali, and Shuqra (Menzies et al., 1994). Fission track ages and length data for apatite specimens from Pan-African basement rocks of Yemen indicate a significant exhumation phase at ~ 17–16 Ma, associated with the main continental rifting (Menzies et al., 1992). The fossil beaches along the Gulf of Aden record a continuous tectonic uplift during most of the Plio-Quaternary (Brannan et al., 1997). About twenty zinc–lead occurrences have been reported in Yemen, most of them located along the margin of the rifts, or in rift-affected blocks. Lead and zinc deposits are hosted in Jurassic to Paleocene carbonate rocks within the Sab'atayn basin, and form two main clusters in the Jabali and Tabaq areas (Fig. 1) (Yemen Geological Survey and Mineral Resources Board, 1994, 2009). The Jabali area includes the Jabali deposit and several other mineral occurrences, which are characterized by maximum grades of 16.5% Zn and 6% Pb (Yemen Geological Survey and Mineral Resources Board, 1994, 2009). The Tabaq area, where nine small Pb–Zn occurrences (with some barite) have been identified, is located in southern Yemen, approximately 360 km east-northeast of Aden and 500 km east of the Jabali mine site, in the same rift system hosting the Jabali deposit (Fig. 1). The ore deposit is characterized here by maximum values of 12% Zn and 3.8% Pb. Another base-metal mineralized district is located in the Mukalla area, in the southwestern part of the Say'un–Masilah basin (Fig. 1). The ore concentration is fault-controlled, and consists of barite and galena in veins. A secondary mineralization containing willemite, smithsonite, cerussite, descloizite, calcite, pyrolusite and celestine, with anomalous grades of Ag, Cd, Ga, Ge, and Mo was also identified in this area (Mattash, 2008; Mattash et al., 2005). 3. Jabali geology and ores — previous studies 3.1. Local geology The Jabali deposit (15°37′ N latitude, 44°46′ W longitude) covers an area of about 2 km2, at an altitude between 1850 and 1950 m.s.l. (Fig. 2). The deposit is located on a small plateau on the eastern flank of a NW–SE-elongated mountainous area that is a segment of the western boundary of the Sab'atayn basin. In the southeast the plateau is Fig. 2. Geological map of the Jabali mining site with the location of analyzed drill holes, and the future open pit area (from SRK Consulting, 2005, modified). Description of the units in text and Fig. 3. Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 dissected by the Wadi Khaynar valley, which cuts the mineralized successions exposed along its flanks. The deposit area is dominated by the sedimentary rocks of the Amran Group, hosts of the deposit (Al Ganad, 1991; Al Ganad et al., 1994; Christmann et al., 1983, 1989). The maximum thickness of the Amran Group at the mine site is 300 m. The Shuqra Formation (Callovian–late Kimmeridgian) is locally subdivided in seven units (members), which from the base of the succession are (Al Ganad et al., 1994) (Figs. 2, 3): Unit 1: sandstones and conglomerates, transgressive on the Proterozoic basement (10 m); Unit 2: gypsiferous mudstones and dolomitized calcarenites, interbedded with marls and nodular limestones (25 m); Unit 3: biomicritic limestones (Callovian) with nodular concretions and chert layers (50 m); Unit 4: micritic limestones and laminated dolomites (15 m); Unit 5: partly dolomitized bryozoan calcarenites (late Oxfordian– early Kimmeridgian), overlain by coral-bearing oolitic and oncolitic limestones. A local disconformity occurs at the top of the unit (40 m); Unit 6: gypsiferous mudstones, followed by ammonite-bearing limestones interbedded with marls and calcareous sandstones (Kimmeridgian) (80 m); Unit 7: partly dolomitized massive bioclastic and biomicritic limestones, locally oolitic limestone with coral bioherms (Kimmeridgian). This unit is exposed at the top of the Jabali plateau, and is strongly affected by karstic erosion (80 m). The Madbi Formation (maximum thickness of 30 m) is locally called Unit 8 and consists of black mudstones and argillites with gypsum Fig. 3. Stratigraphy of the Jabali area with the units established by Al Ganad et al. (1994) (modified), correlated with the corresponding formations of the “standard” Amran Group of Beydoun et al. (1998). Shuqra Formation — Unit 1: sandstone and conglomerate, transgressive on the Late Proterozoic basement; Unit 2: gypsiferous mudstone overlain by dolomitized calcarenite, marl and nodular limestone; Unit 3: micritic-biomicritic limestone (Callovian), with nodular concretions and chert layers; Unit 4: micritic limestone and lagoonal/lacustrine dolomite; Unit 5: partly dolomitized bryozoan calcarenite (Late Oxfordian–Early Kimmeridgian), overlain by coral-bearing oolitic limestone; Unit 6: gypsiferous mudstone grading into micritic limestone (Kimmeridgian) and marl; Unit 7: massive bioclastic–biomicritic limestone, locally oolitic with coral bioherms (Kimmeridgian). Madbi Formation — Unit 8: black mudstone and argillite with gypsum and dolomite intercalations, grading laterally into micritic ammonite-bearing limestone (Late Kimmeridgian–Tithonian). Sab'atayn Formation — Unit 9: biomicrite with oncolites and bio-oocalcarenite (Late Jurassic). 251 crystals and dolomite intercalations. It grades laterally into ammonitebearing limestones. The Sab'atayn Formation (locally known as Unit 9) consists of biomicrites with oncolites and bio-oocalcarenites; intercalations of gypsum lenses and arkosic sandstones also occur. In the Jabali area, early Miocene trachytic sills and dykes (dated at 22 Ma) cut the sedimentary rocks (Al Ganad et al., 1994). Several Tertiary alkaline granite bodies occur at Jabal as Saad, 15 km to the west of the mine site. Less than a kilometer south of Jabal as Saad, a thick Pliocene to Holocene travertine deposit indicates that thermal springs were active in this area, in analogy to travertine deposits in the Sirwah area, associated to the Marib volcanic field (Weiss et al., 2009). The structural setting at Jabali is dominated by extensional rift tectonics and the rocks were affected mostly by brittle deformation. The sedimentary succession is almost horizontal at the periphery of the basin, and only small areas are characterized by west-slightly dipping strata. The most prominent normal faults strike 120 to 140° (Fig. 2). This fault set includes the main Jabal Salab fault that borders the shoulder of the plateau below the Jabal Salab peak. Another set of normal faults has 65° to 80° trends. A further major system is developed at 0° to 5°, while a more subtle fracture trend ranges between 25° and 40° (SRK Consulting, 2005). All fault planes have dip angles between 60° and 80°. 3.2. Mineralization The Jabali ores (Fig. 4A) are hosted in the higher part of the Shuqra Formation. The ore is almost completely oxidized (Fig. 4B, C), even though primary sulfides have been locally preserved in the interval underlying the Madbi Formation (Unit 8) (Fig. 4D). The nonsulfide ore at Jabali is massive, semi-massive and disseminated, and consists of vuggy to highly porous, brown-orange to white nonsulfide mineral aggregates (Fig. 4E, F). A porous cellular boxwork structure, accompanied by numerous cavities coated with secondary zinc minerals is also widespread. The ore bodies are only partly exposed, since at least half of the mineralized lithologies occur in the subsurface below Jabal Barrik (SRK Consulting, 2005). The mineralization is structurally and lithologically controlled. This is reflected in the stratiform-to-tabular architecture of the ore bodies, and in the mineral enrichments occurring along vertical fractures, faults and at the intersection of the above structures. Specifically, the mineralization is developed along the three main structural trends recognized in the field. At the intersection of faults, big vertical ore bodies occur, called “chimneys” by local exploration geologists. The stratiform bodies occur in three different zones: a laterally extensive upper zone, and more sporadic lower and middle zones. These bodies are generally flat and manto-like but, at the base of the Jabal Salab massif, along the NW–SE fault they dip towards NE with angles greater than 30° (SRK Consulting, 2005). The more extensive studies on the Jabali primary mineralization were performed by Al Ganad et al. (1994). According to these authors the primary sulfide association consists of sphalerite (predominant), galena, and pyrite/marcasite. Sphalerite occurs as two distinct generations: a first dark-colored, and a second represented by zoned euhedral to subhedral honey-colored or brownish-red crystals. The second generation is the most abundant; it is not pure, but contains iron, silver, cadmium, copper, germanium and mercury. The ore deposit at Jabali is hosted in the topmost horizons of the Shuqra Formation (Unit 7), specifically when this interval is dolomitized (Al Ganad et al., 1994). Two dolomitization stages have been distinguished: an early district-scale diagenetic (dolomite phases D1, D2) and a second one, more local, associated with sulfide deposition (dolomite phases D3, D4). The dolomite phases are strongly fabric destructive and have obliterated both depositional and diagenetic features of the precursor limestone. In detail, D3 and D4 dolomites are characterized by a saddle structure (Radke and Mathis, 1980), with crystal sizes Author's personal copy 252 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Fig. 4. A) View of Wadi Jabali, looking E: old mine workings. B) JS MON 3. Smithsonite in outcrop, with a vuggy-highly porous texture. C) JS MON 28. Hydrozincite coating smithsonite and host rock. D) J125-2. Partly oxidized ore, with remnants of sphalerite and galena. E) J125-15. Massive smithsonite, replacing Zn-dolomite, and smithsonite crusts in cavities. F) J125-3. Gypsum veins, cutting both dolomite and smithsonite. Dol = dolomite; Gn = galena; Gp = gypsum; Hz = hydrozincite; Sm = smithsonite; Sp = sphalerite; Zn-Dol = Zn-bearing dolomite. comprised from a few hundred microns to few millimeters. The saddle crystals occur within cavities and fill fractures crosscutting the previous dolomite generations. A calcite phase is associated with saddle dolomite, and mainly occurs as late cement completely filling the porosity. Fluid inclusions in sphalerite (Al Ganad et al., 1994) show a bimodal grouping (60–85 °C and 85–110 °C) of homogenization temperatures to the liquid, mirrored by salinities of 10–14 eq. wt.% NaCl and 19–23 eq. wt.% NaCl, respectively. These data are compatible with a basinal fluid origin: salinities and temperatures are similar to many other carbonate-hosted zinc–lead Mississippi Valley-type deposits (Roedder, 1976). Lead isotope ratios of galena and cerussite are in the range of: 18.85 and 18.95 206Pb/204Pb, 15.66 and 15.72 207Pb/204Pb, and 39.71 and 39.92 208Pb/204Pb (Al Ganad et al., 1994). After these authors the lead isotope compositions point to an original metal source in the early Proterozoic crustal basement of the Sab'atayn basin. Based on these observations, Al Ganad et al. (1994) stated that the primary mineralization was deposited in karst cavities related to the emersion surface at the top of the Shuqra Formation (Unit 7), from fluids migrating from the Sab'atayn basin during Mesozoic rifting. They also proposed that sulfide ores were emplaced slightly after sedimentation of the Madbi Formation (Unit 8) (Late Jurassic– Cretaceous). The black shales of the Madbi Formation acted as an impermeable barrier to fluid migration. Al Ganad et al. (1994) also suggested that oxidation of the primary deposit may have begun during Cretaceous, developed during Paleogene and continued up to the present. C. Allen (unpublished data, 2000) speculated on a different genetic concept, hypothesizing a hybrid Carbonate Replacement Deposit–Mississippi Valley-type character for the primary Jabali mineralization. Reynolds and Large (2010), in their revision on the Tethyan zinc– lead metallogeny in Europe, northern Africa and Asia, have considered Jabali as a Mississippi Valley-type deposit formed in a carbonate platform during the Sab'atayn basin evolution. 4. Materials and methods The mineralogical study was carried out on 40 sections (each one ca. 1 m in length) of drillcores from the Jabali drill holes J109, J125 and J138 Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 (Table 1) (Mondillo et al., 2011). The sampled core intervals are on the drill hole logs shown in Fig. 5. The selected holes were drilled by ZincOx Resources in 2004 in the Jabal Barrik area, near the center of the planned pit (Fig. 2). Several mineralized outcrop samples (49) (Table 2; Fig. 2) were analyzed as well. Core and outcrop samples have been studied using petrographic and cathodoluminescence (CL) microscopy. Cathodoluminescence measurements were done on carbon-coated, polished thin sections using a hot cathode cathodoluminescence microscope HC1-LM (cf., Neuser et al., 1995), at TU Bergakademie Freiberg, Germany. The system was operated at 14 kV accelerating voltage and a current of 0.2 mA (current density of about 10 μA/mm2). Luminescence images were captured on-line during cathodoluminescence operations using a peltier cooled digital video camera (OLYMPUS DP72). Cathodoluminescence spectra in the wavelength range 380 to 1000 nm were recorded with an Acton Research SP-2356 digital triple-grating spectrograph with a Princeton Spec-10 CCD detector that was attached to the cathodoluminescence microscope by a silica-glass fiber guide. Cathodoluminescence spectra were measured under standardized conditions (wavelength calibration by a Hg-halogen lamp, spot width 30 μm, measuring time 2 s). Scanning electron microscopy (SEM) examination was carried out using a Jeol JSM 5310 instrument at the University of Napoli Federico II (CISAG). Element mapping in backscattered electron mode (BSE) and qualitative and quantitative analyses by energy-dispersive spectrometry (EDS) were obtained with the INCA X-stream pulse processor Table 1 Jabali drillcore samples. Drillcore location from Fig. 2. Drillcore n. From (m) To (m) ZincOx assay n. Sample n. J109 57.3 58.3 59.3 60.3 61.65 50.78 51.78 53.1 54.73 55.73 57.92 59.45 60.97 62 64 65 66 67 68 69 74.5 75.5 76.5 77.5 78.5 79.5 84.5 85.5 86.5 87.5 88.5 89.5 90.5 68 69 70 71 72 73 74 58.3 59.3 60.3 61.65 62.7 51.78 53.1 54.73 55.73 56.73 59.45 60.97 62 64 65 66 67 68 69 70 75.5 76.5 77.5 78.5 79.5 80.5 85.5 86.5 87.5 88.5 89.5 90.5 91.5 69 70 71 72 73 74 75 1 2 3 4 5 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 19 20 21 22 23 24 29 30 31 32 33 34 35 4 5 6 7 8 9 10 J109-1 J109-2 J109-3 J109-4 J109-5 J125-1 J125-2 J125-3 J125-4 J125-5 J125-6 J125-7 J125-8 J125-9 J125-10 J125-11 J125-12 J125-13 J125-14 J125-15 J125-19 J125-20 J125-21 J125-22 J125-23 J125-24 J125-29 J125-30 J125-31 J125-32 J125-33 J125-34 J125-35 J138-4 J138-5 J138-6 J138-7 J138-8 J138-9 J138-10 J125 J138 253 Fig. 5. Stratigraphic logs of selected drill holes. The sampled intervals are indicated. and the 4.08 version Inca software (Oxford Instruments detector), interfaced with the JEOL JSM 5310. The following reference standards were used: albite (Si, Al, Na), orthoclase (K), wollastonite (Ca), diopside (Mg), almandine (Fe), rutile (Ti), barite (Ba), strontianite (Sr), Cr2O3 (Cr), rhodonite (Mn), sulfur (pyrite), sphalerite (Zn), galena (Pb), fluorite (F), apatite (P), sylvite (Cl), Smithsonian phosphates (La, Ce, Nd, Sm, Y), pure vanadium (V) and Corning glass (Th and U). Analytical errors are 1% relative for major elements and 3% relative for minor elements. X-ray diffraction analysis was carried out on all samples. The core sections and the outcrop samples have been crushed to 1 mm, milled and homogenized. X-ray diffraction analyses have been carried out with a Philips PW 3020 automated diffractometer at the University of Heidelberg, with CuKα radiation, 40 kV and 30 mA, 10 s per step and a step scan of 0.02° 2θ. The data were collected from 3 to 110° 2θ. X-ray diffraction quantitative phase analysis of the core samples (Mondillo et al., 2011) was performed on the X-ray diffraction patterns using the Rietveld method (Bish and Howard, 1988; Bish and Post, 1993; Hill, 1991; Rietveld, 1969). X-ray powder diffraction data were analyzed using the GSAS package (General Structure Analysis System, Larson and Von Dreele, 2000) and its graphical interface EXPGUI (Toby, 2001). In this paper we report only the mean values of the data previously published by Mondillo et al. (2011), calculated for each mineralized interval occurring in the drillcores. The mean values were calculated weighting the data relatively to the length of the samples, and to the total length of the mineralized intervals, using the formula: Mx ¼ ð∑I Wxi Li Þ=∑I Li where Mx is the mean amount of a mineral x in a specific mineralized interval I, Wxi is the amount of mineral x in the sample i, and Li is the length of the core sample i; the sums ∑I have been obtained considering all the samples i of the interval I. X-ray diffraction qualitative analyses on outcrop samples were performed with the PANalytical B.V. software HighScore version 3.0e (Osservatorio Vesuviano-INGV) and JCPDS PDF-2 database. For quantitative analysis of X-ray diffraction patterns we used the HighScore Author's personal copy 254 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Table 2 Jabali outcrop samples. Sampling area n. from Fig. 2. Sampling area n. Sample n. 1A JS MON 1 JS MON 2 Sample description Yellowish-brown concretionary smithsonite Weathered red-brown dolomite with sulfide remnants JS MON 3 Smithsonite ore encrusted by hydrozincite JS MON 4 Brownish dolomite, partly dedolomitized by calcite JS MON 5 Porous smithsonite encrusted by hydrozincite JS MAR 1 Massive smithsonite with galena remnants JS MAR 2 Massive smithsonite with macrocrystalline galena remnants 1B JS MON 6 Reddish-brown massive ore JS MON 7 Brown massive ore JS MON 8 Gossan; (hydr)oxide black-red crusts JS MON 9 Reddish-brown massive ore; few smithsonite crusts in voids JS MON 10 Reddish-brown massive ore; few smithsonite crusts in voids JS MON 11 Clay sample from gossan JS MON 12 Gossan sample JS MON 13 Smithsonite–dolomite transition JS MON 36 Dedolomitized dolomite JS MAR 3 Massive smithsonite with hydrozincite coatings 2 JS MON 14 Smithsonite breccia encrusted by hydrozincite and hemimorphite JS MON 15 Smithsonite breccia encrusted by hydrozincite and hemimorphite JS MON 16 Gossan and smithsonite ore JS MON 17 Gossan sample JS MAR 5 Rusty dolomite with galena spots JS MAR 6 Smithsonite breccia with hydrozincite encrustations 3 JS MON 18 Gossan sample JS MON 19 Silicified gossan sample with calcite spots JS MON 20 Smithsonite–hydrozincite massive ore 4 JS MON 21 Smithsonite–hydrozincite massive ore JS MON 22A Smithsonite–hydrozincite massive ore JS MON 22B Smithsonite–hydrozincite massive ore JS MON 23 Massive dolomite JS MON 24 Smithsonite–hydrozincite ore JS MON 25 Gossan sample JS MON 26 Smithsonite–hydrozincite ore JS MON 27 Smithsonite–hydrozincite ore JS MON 28 Smithsonite–hydrozincite massive ore 5 JS MON 29 Smithsonite–hydrozincite massive ore; hemimorphite spots JS MON 30 Gossan sample with hydrozincite JS MON 31 Gossan sample 6 JS MON 32_1 Smithsonite–hydrozincite massive ore; hemimorphite spots JS MON 32_2 Smithsonite–hydrozincite massive ore; hemimorphite spots JS MON 33 Massive dolomite, partly dedolomitized, with galena veins JS MON 34 Trachyte dyke JS MON 35 Trachyte dyke 7 JS MAR 7 Gossan sample JS MAR 8 Massive smithsonite with hydrozincite spots 8 JS MAR 4 Massive smithsonite with hydrozincite spots Jabal Salab area JS MON 37 Barite vein JS MON 38 Limestone with bioherms of Unit 7 JS MON 39 Gossan sample Plus software with Rietveld structural models based on the American Mineralogical Crystal Structure Database (AMCSD). Whole rock chemical analyses of major and minor elements for the core samples were carried out by OMAC Laboratories Ltd. (Co. Galway, Ireland). Diamond drillcore was cut and the entire half-core samples homogenized and powdered to obtain 30 g aliquots for chemical analysis. After aqua regia digestion, the samples were analyzed by multi-element inductively-coupled plasma mass spectrometry (ICP-MS). Samples with Zn contents N 9% have been also analyzed by atomic absorption spectrometry (AAS), with an excellent agreement between the two data sets (SRK Consulting, 2005). As for the previously mentioned Xray diffraction analyses, we here only report the mean values of chemical assay data from Mondillo et al. (2011), calculated for each mineralized interval recognized in the drillcores. The mean values were calculated weighting the data relatively to the length of the samples and to the total length of the mineralized intervals, using the formula: Ex ¼ ð∑I Axi Li Þ=∑I Li where Ex is the mean amount of element x in a specific mineralized interval I, Axi is the amount of element x in the sample i, and Li is the length of the core sample i; the sums ∑I have been carried out considering all the samples i of the interval I. Whole rock chemical analyses of the outcrop samples were performed by ACME Laboratories (Vancouver) to determine the major and minor elements, using aqua regia digestion and multi-element inductively-coupled plasma energy and mass spectrometry (ICP-ES/ ICP-MS). Loss on ignition was not evaluated. Stable carbon and oxygen isotope analyses were carried out at the University of Erlangen-Nürnberg (Germany). Carbonate powders and picked minerals were reacted with phosphoric acid at 70 °C using a GasBench II connected to a Thermo Finnigan Five Plus mass spectrometer. All values are reported in per mil relative to V-PDB by assigning a δ13C value of +1.95‰ and a δ18O value of −2.20‰ to NBS19. Reproducibility was checked by replicate analysis of laboratory standards and was better than ± 0.07‰ (1σ) for both carbon and oxygen isotope analyses. Oxygen isotope values of dolomite and smithsonite were corrected using the phosphoric acid fractionation factors given by Kim et al. (2007), Rosenbaum and Sheppard (1986) and Gilg et al. (2008). Sulfur isotope analyses of a few sulfide and sulfate samples were carried out at Actlabs (Ancaster, Ontario, Canada). Single crystals and mineral fragments for analysis were collected with a dental drill from core and outcrop samples. Pure BaSO4 and sulfide samples were combusted to SO2 gas under ~ 10−3 tor vacuum. The SO2 was transferred directly from the vacuum line to the ion source of a VG 602 Isotope Ratio Mass Spectrometer (Ueda and Krouse, 1986). Quantitative combustion to SO2 was achieved by mixing 5 mg of sample with 100 mg of a V2O5/SiO2 mixture (1:1). The reaction was carried out at 950 °C for 7 min in a quartz glass tube. Pure copper turnings were used as a catalyst to ensure conversion of SO3 to SO2. Internal Lab Standards (SeaWaterBaSO4 and FisherBaSO4) were run at the beginning and at end of each set of samples and were used to normalize the data as well as to correct for instrumental drift. All results are reported in the permil (‰) notation relative to the international VCDT standard. Precision and reproducibility (1 std.dev.) based on replicate laboratory standard analyses (n = 10) was typically better than ± 0.2‰ (1σ) (www. actalabs.com). 5. Results 5.1. Mineralogy of the nonsulfide orebody The mineralogical results obtained with X-ray diffraction of the core samples set, drilled in different areas of the Jabali minesite, have been published in full by Mondillo et al. (2011). The mean values, calculated from the same cores are reported in Table 3. The interval 57.3–62.7 m from core J109, and the interval 68–75 m from core J138 correspond to the upper zone of the orebody. Dolomite is the predominant lithology of both cores, while smithsonite has been detected in an interval 1–2 m-thick (Fig. 5) with an average grade of 6–7 wt.%. Three mineralized intervals have been sampled from core J125 (Fig. 5): the uppermost interval, immediately below the shales of the Madbi Formation (Unit 8), is comprised between 50.78 and 80 m, the middle interval can be traced between 74.5 and 80.5 m and the deepest Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 255 Table 3 Mineral abundances in the core intervals from X-ray quantitative phase analysis (Mondillo et al., 2011 modified). Drillcore Interval Mineralized zone Dol Cal Sm Cer Gp Ang Sp Gn Cha Hem Gth Kln Sau Ilt Qz 7.4 0.0 40.0 2.4 6.0 7.6 7.3 27.2 10.3 24.0 6.4 18.7 0.1 0.5 – 0.2 0.6 0.3 – 8.4 0.3 – – 3.7 – 1.0 – – – 0.4 0.1 1.2 – – 0.1 0.5 0.1 0.1 – – 0.8 0.2 0.1 0.1 – – – 0.1 – 0.3 – 0.2 0.2 0.2 1.8 1.6 4.2 3.6 3.0 2.5 0.9 0.5 7.7 0.5 0.3 1.5 0.1 – 0.1 – – – – – 0.1 – – – – – 0.1 – – – (wt.%)a (m) J109 J125 57.30–62.70 Upper zone 50.78–70.00 Upper zone 74.50–80.50 Middle zone 84.50–91.50 Lower zone J138 68.00–75.00 Upper zone Mean value on the entire sample batch 82.1 52.8 37.4 69.2 82.5 61.5 Notes: “–” not detected. Mineral abbreviations mostly after Whitney and Evans (2010). Dol; dolomite; Cal, calcite; Sm, smithsonite; Cer, cerussite; Gp, gypsum; Ang, anglesite; Sp, sphalerite; Gn, galena; Cha, chalcophanite; Hem, hematite; Gth, goethite; Kln, kaolinite; Sau, sauconite; Ilt, illite; Qz, quartz. Values derive from a weighted mean calculation referred to the length of core samples and the total length of the mineralized intervals (details in text). Statistical indicators ranges: Rp 5.11–6.40%, wRp 6.65–8.85%. one reaches from 84.5 down to 91.5 m. Also in core J125 dolomite is generally the most abundant mineral phase, though being totally absent in the best mineralized samples. Smithsonite is abundant in the upper and lower sections of the core, with average grades between 24 and 27 wt.%. In the middle section of the core, smithsonite has a mean value of 10.3 wt.%, and is associated with calcite. With an average amount of 18.7 wt.% calculated over the whole sample set, smithsonite represents the main ore mineral at Jabali. Sphalerite and galena have very low amounts (below 2 wt.% on average) in all mineralized intervals. Cerussite and anglesite were found only locally. All core samples contain goethite, with mean values ranging between 1.6 and 4.2 wt.%. Only the samples from the upper zone of the J-125 core, directly below the boundary with Unit 8, contain gypsum (~ 8.4 wt.%). Sauconite (Zn-smectite) and other clay minerals have a very limited distribution, as well as Zn–Mn-hydroxides (chalcophanite). Hydrozincite and hemimorphite are completely absent in drillcores. X-ray diffraction quantitative mineralogical analyses (Table 4) were conducted on 17 outcrop samples collected in the area of the Jabali mine site, where both high-grade ore and gossanous rocks occur. The most abundant ore mineral in the outcrop samples is again smithsonite, with a variable percentage. In some samples, smithsonite amounts to almost 100 wt.% of the rock total (JS MON 9, JS MON 10). Hydrozincite, detected only in outcrop samples, can locally reach concentrations around 50–70 wt.%. Hemimorphite is slightly less abundant. Cerussite and galena have been detected only in two specimens, sampled along a fault zone. Dolomite can be very abundant, and calcite may occur with maximum amounts of 15 wt.%. Quartz is abundant only locally (JS MON 19). Goethite is ubiquitous: together with hematite it is the main “oxide mineral” in the gossanous samples (for example in JS MON 8). Several sulfate minerals have been identified in the gossan: plumbojarosite, jarosite, dietrichite, and rozenite. All minerals detected at Jabali are listed in Table 5. 5.2. Petrography of the nonsulfide orebody In the previous published papers on the Jabali deposit (Al Ganad et al., 1994; Mondillo et al., 2011) the petrography of the nonsulfide mineral association was dealt in a limited way. Here we report the results of a comprehensive study on the secondary mineral assemblage that documents several characteristics of the deposit never highlighted before. The classic thin section microscopic observation has been integrated with cathodoluminescence and scanning electron microscopy (Figs. 6–9). Cathodoluminescence emission spectra of selected minerals are shown in Fig. 10. The Jabali host rock consists of two main generations of dolomite: an early diagenetic dolomite, which corresponds to the D1 and D2 phases of Al Ganad et al. (1994), and a saddle (hydrothermal) dolomite (D3 and D4 phases of Al Ganad et al., 1994). Saddle dolomite replaces the previous dolomite phase and occurs as well in veins (Fig. 6A). This hydrothermal dolomite contains variable amounts of Mn (up to 2 wt.% MnO) and iron (up to 6 wt.% FeO). Table 4 Mineral abundances (wt.%) from X-ray diffraction quantitative analysis of outcrop samples. JS MON 3 JS MON 6 JS MON 8 JS MON 9 JS MON 10 JS MON 13 JS MON 14 JS MON 15 JS MON 18 JS MON 19 JS MON 21 JS MON 22A JS MON 22B JS MON 28 JS MON 29 JS MON 32_2 JS MON 33 Sm Hz Hm Cer Gn Dol Cal Qz Gth Hem Prls Sau Dtrch Pb-jrs Jrs Roz Gp 64.0 96.5 – 97.3 97.5 54.9 36.4 25.3 20.5 – 33.6 69.3 0.9 39.4 19.5 8.4 – 14.6 – – – – – 21.2 9.3 – – 33.9 28.8 73.9 50.2 75.5 17.8 – – – – – – – 2.2 3.6 – – – – 7.6 0.4 2.8 4.4 – – – – – – – – – – – – – – – – 0.1 0.5 – – – – – – – – – – – – – – – – 1.0 20.2 – – – – 43.2 38.7 44.7 60.6 – – – – – – 67.6 97.5 0.2 – – – – – 0.5 1.3 15.0 0.8 – – 1.1 – 1.1 0.7 0.8 – – – – – – – – – 55.9 – – – – – – – 0.9 3.2 73.2 1.6 1.6 1.7 0.9 15.4 0.4 30.0 28.3 1.9 16.5 9.8 1.0 0.6 0.1 – – 26.8 1.1 0.9 – – – 2.7 13.2 3.8 – – – – – – – – – – – – – – – – 0.1 – – – – – – 0.1 0.3 – – – 0.2 0.1 0.4 0.8 – 0.3 – – 0.2 0.1 0.1 0.1 – – – – – – – – – 0.1 – – – – – – – – – – – – – – – tr – – – tr tr – – – – – – – – – – – tr – – – – – – – – – – – – – – tr – – – – – – – – – – – – – – – – – – – – – – – – – 0.3 – Notes: “–” not found, “tr” traces. Mineral abbreviations mostly after Whitney and Evans (2010): Sm, smithsonite; Hz, hydrozincite; Hm, hemimorphite; Cer, cerussite; Gn, galena; Dol, dolomite; Cal, calcite; Qz, quartz; Gth, goethite; Hem, hematite; Prls, pyrolusite; Sau, sauconite; Dtrch, dietrichite; Pb-jrs, plumbojarosite; Jrs, jarosite; Roz, rozenite; Gp, gypsum. Statistical indicators ranges: Rp 5.67–8.40%, wRp 8.25–9.40%. Author's personal copy 256 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Table 5 List of minerals detected at Jabali. Name Chemical formula (International Mineralogical Association) Acanthite Anglesite Calcite Cerussite Chalcophanite Covellite Dietrichite Dolomite Galena Goethite Greenockite Gypsum Hematite Hemimorphite Hydrozincite Illite Jarosite Kaolinite Plumbojarosite Pyrite/marcasite Pyrolusite Pyromorphite Quartz Rozenite Sauconite Smithsonite Sphalerite Ag2S PbSO4 CaCO3 PbCO3 ZnMn4+ 3O7·3H2O CuS (Zn,Fe2+,Mn2+)Al2(SO4)4·22H2O CaMg(CO3)2 PbS FeO(OH) CdS CaSO4·2H2O Fe2O3 Zn4(Si2O7)(OH)2·H2O Zn5(CO3)2(OH)6 (K,Ca,Na)(Mg,Al,Fe3+)2(Si,Al)4O10(OH)2·nH2O KFe3+3(SO4)2(OH)6 Al2Si2O5(OH)4 Pb0.5Fe3+3(SO4)2(OH)6 FeS2 MnO2 Pb5(PO4)3Cl SiO2 Fe2+SO4·4H2O Na0.3Zn3(Si,Al)4O10(OH)2·4H2O ZnCO3 ZnS Early diagenetic dolomite is non-luminescent under CL. Saddle dolomite has generally a bright red luminescence, but it can be locally characterized by non-luminescent crystal cores and red luminescent crystal rims (Fig. 6B). Non-luminescent crystal cores are generally associated with opaque Fe-hydroxides. Cathodoluminescence emission spectra of the saddle dolomite at Jabali have a pronounced emission maximum at 650–660 nm (Fig. 10A). As reported in the literature (Götze, 2012), this indicates that luminescence is activated by Mn2+, and that Mn2+ is hosted in the Mg-positions of the crystal structure. In fact, when Mn2 + occurs in the Ca-positions, the emission spectra show a peak (here absent) at 575 nm and yellow-orange CL colors. The non-luminescence in the crystal cores of saddle dolomite is due to their high Fe contents (measured by scanning electron microprobe and testified by the occurrence of Fe-hydroxides), which quench the Mn-related brightness. The emission spectra of the dark crystal cores of the dolomite crystals have also a maximum at 650–660 nm, but it is less pronounced and intense than in crystal rims (Fig. 10A). All dolomite phases at Jabali show two different types of alteration or replacement. The first is a dedolomitization (Bischoff et al., 1994; Coniglio, 2003), with dolomite being replaced by calcite. Manganese and Fe, previously contained in the dolomite structure, are precipitated as oxides and hydroxides in the interstices of the crystals and in small vugs and fissures (Fig. 6C, D, E). The newly formed calcite may contain Pb (up to ~ 5 wt.% PbO) and Cd (up to ~ 2 wt.%). The second type of replacement is most evident in wide gradational bands, which mark the boundary between the host dolomite and replacive smithsonite. Under the microscope, the altered dolomite crystals appear clear and transparent along their margins, but are generally opaque and brown colored towards the center (Fig. 6F, G). These brown colored phases are mainly Zn-bearing (Figs. 7A, B, C), characterized by ZnO contents reaching 17–22 wt.% (Mondillo et al., 2011). The ZnO content is inversely correlated with that of MgO in the Zn-bearing dolomite (Zn2+ replaces Mg2 + by up to 70%), while the calcium content corresponds to the values generally measured in stoichiometric dolomites. Zinc-bearing dolomite preserves limited amounts of Fe and Mn (below 1–2 wt.%) from the saddle dolomite phase. Cadmium can be present locally (CdO around 1.5 wt.%). The Zn-bearing dolomite has a yellowish-orange CL color (Fig. 6H). The luminescence emission spectra of this dolomite are characterized by two overlapping peaks: the most prominent occurs between 585 and 600 nm, the second at 645–655 nm (Fig. 10B). No specific literature exists on the cathodoluminescence behavior of Zn-bearing dolomite. However, adapting to this case the concepts expressed by Götze (2012), we assume that the yellowish-orange color is produced by the preferred occurrence of Mn2 + in the Ca-position instead than in the Mg-position of the crystal structure. All earlier formed dolomites and Zn-bearing dolomite are replaced by smithsonite (smithsonite 1) (Fig. 8A). The boundary between the original dolomite and smithsonite can be very gradational; it generally consists of thin concentric belts with a complex assemblage of smithsonite, Zn-bearing dolomite and remnants of original dolomite (preferentially hydrothermal dolomite). Smithsonite 1, together with newly deposited gypsum, also replaces sphalerite (Fig. 8B). Replacive smithsonite has a very fine texture. Specifically, it consists of microcrystals, which form agglomerates mimicking/following the original macrocrystalline rhombohedral/saddle habitus of the replaced dolomite (Fig. 7D, E), and locally appear to fill dissolution cavities inside the dolomite macrocrystals (Fig. 7F). Smithsonite microcrystals assume a rhombohedral shape (Fig. 9A) and occasionally can occur as tiny euhedral crystals, showing combination of forms belonging to the scalenohedral class, i.e. hexagonal dipyramids, rhombohedra, scalenohedra (Fig. 9B). Smithsonite 1 microcrystals are often chemically zoned, with Mg contents locally reaching 8–10 wt.% MgO (Fig. 7D), and variable contents of Mn (up to ~3 wt.% MnO), Fe (up to 5–6 wt.% FeO), and Ca (up to 6 wt.% CaO). Smithsonite 1 has a red CL color (Fig. 8C, D). Cathodoluminescence emission spectra have a pronounced maximum at 650–660 nm, and a broader, though less intense peak around 440 nm (Fig. 10C). In analogy with dolomite, the pronounced peak indicates that cathodoluminescence is produced by variable amounts of Mn2 + hosted in the smithsonite structure. The peak at 440 nm is instead related to the presence of crystal lattice defects in the mineral. A second type of smithsonite (smithsonite 2) consists of different cement generations, occurring in vugs, cavities and veins as newly formed crystals and concretions (Fig. 8E). Crystals are characterized by a roughly rounded morphology and a size that generally increases from the earliest generation to the latest. Concretions are prevailingly botryoidal (Fig. 7G). At microscale, crystals and concretions show variably developed rhombohedral faces. Smithsonite 2 is commonly zoned. Generally, it is possible to recognize a first generation of crystals and concretions with colors variable from brown, to orange and yellow that precipitated directly over smithsonite 1. Other generations may follow, generally white colored at first sight, and clear and transparent under the microscope (Fig. 8E). A variation in color of smithsonite 2 commonly corresponds to a chemical zonation (Fig. 7H). Composition of smithsonite 2 can vary from pure end-member Zn-carbonate to terms characterized by over 20 wt.% MgO (substitution of Mg2 + for Zn2 + of about 70% in the smithsonite structure), ~ 2 wt.% MnO, ~ 2 wt.% CdO, ~ 1.5 wt.% CaO, and ~ 1.5 wt.% PbO. The concretionary phases of smithsonite 2 have different CL colors: the first phases are pinkish-red, whereas the latest are blue (Fig. 8F). Luminescence emission spectra of pinkish-red phases are similar to those of smithsonite 1. Spectra of blue smithsonite 2 are slightly different and characterized by two peaks, approximately having the same intensity, in correspondence to the same wavelengths detected before (~440 nm and ~ 650 nm) (Fig. 10C). This indicates that the Mn-related redluminescence is strongly quenched (low intensity of peak at 650 nm), and the blue color is mostly activated by crystal lattice defects. Smithsonite 2 is precipitated in veins either shortly before, or together with variable amounts of gypsum, which cut replacive smithsonite 1 (Fig. 8G). Specifically, gypsum forms fibrous, partially deformed crystals perpendicular to the veins, suggesting crystal growth contemporary to vein opening. In outcrop samples, smithsonite 2 is Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 257 Fig. 6. Thin section petrography: A) J125-2. Sphalerite associated with saddle dolomite; B) J125-2. Zoned saddle dolomite crystals under CL; C) J138-10. Dedolomitization with dolomite remnants (NII); D) the same as in panel C, under CL; E) J125-19. Dedolomitization with dolomite remnants (NII); F) J125-15. Zn-dolomite replacing dolomite in the crystal cores; G) J125-15. Zn-dolomite replacing hydrothermal dolomite in the crystal cores; smithsonite concretions at the rims (NII); H) Zn-dolomite (crystal cores) with yellow-orange CL colors; hydrothermal dolomite in red; smithsonite crusts dark blue. Cal = calcite; Dol = dolomite; Sm 2 = smithsonite 2; Sp = sphalerite; Zn-Dol = Zn-dolomite. intergrown with clays, Fe-hydroxides and oxides, forming reddish concretions. Hydrozincite and hemimorphite have been detected solely in outcrop samples. Hydrozincite generally shows a fine-grained texture and acicular crystal habitus; it may occur as smithsonite replacement, as vein and porosity filling, and in crusts. Petrographic investigation of samples collected from the mineralized outcrops shows that hydrozincite occurs as surface coatings, in vugs, as well as filaments replacing smithsonite. Hydrozincite has a blue CL color, and the emission spectra are dominated by the effects related to the defects of crystal lattice. Hemimorphite has been detected as cm-sized fan agglomerates of tabular crystals throughout the deposit and as small crystals in vugs. Several small veins with crystalline hemimorphite, locally associated with Fe-hydroxides, are quite widespread. Lead secondary minerals (cerussite and anglesite) strictly replace primary galena, and not the host rock (Fig. 11A). Cerussite is quite pure chemically and shows a green CL color (Fig. 8H). In the Jabali samples we were able to recognize not only the remnants of the primary sulfide association (sphalerite, galena and pyrite/ marcasite), but also a few secondary sulfides related to the nonsulfide mineral assemblage. A secondary galena phase precipitated between the gypsum veins and smithsonite 2 concretions. This secondary galena may be accompanied by greenockite, and by small crystals of pyromorphite. At the boundary between primary sphalerite and replacing Author's personal copy 258 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Fig. 7. Backscattered (BSE) images: A) J125-31. Zn-dolomite, patchily replacing dolomite crystals; B) J109-5. Zn-dolomite pervasively replacing the rim of a dolomite crystal; C) J125-15. Dolomite crystal, partly replaced by Zn-dolomite, at the boundary Fe-hydroxides and smithsonite 2; D) JS MON 10. Smithsonite 1 mimicking the rhombohedral habitus of a dolomite crystal; E) J125-7. Smithsonite 1, locally cut and replaced by gypsum veins; F) JS MON 2. Zoned smithsonite 1 and Zn-dolomite replacing dolomite; G) J125-15. Concretions of smithsonite 2 at the border of smithsonite 1 replacing Zn-dolomite; H) JS MON 2. Zoned concretions of Mg-bearing smithsonite 2. Dol = dolomite; Fe–Mn-Dol = Fe–Mn-dolomite; Gp = gypsum; Sm 1 = smithsonite 1; Sm 2 = smithsonite 2; Mg–Sm 1 = Mg-bearing smithsonite 1; Mg–Sm 2 = Mg-bearing smithsonite 2; Zn-Dol = Zn-dolomite. smithsonite, secondary greenockite, galena and covellite were identified (Fig. 11B). The Ag-sulfide acanthite is always associated at Jabali with the secondary ores. It occurs together with concretionary smithsonite (Fig. 11C), and also in association with gypsum (Fig. 11D) and greenockite. Hemimorphite containing small particles of Ag-sulfide has been also identified in a gypsum vein. Zn-clays are not very common but sauconite has been detected in the porosity of the host rock, locally associated with kaolinite and illite. The Fe-hydroxides do not have a typical goethite composition: they contain Zn (up to 12 wt.% ZnO), Pb (up to 7 wt.% PbO) and SiO 2 (up to 6 wt.%). Also most Mn-hydroxides consist not only of chalcophanite (which contains only Mn and Zn), but of possibly amorphous phases containing Mn, Pb and Fe in variable proportions (i.e. PbO ~ 20–30 wt.%, FeO ~ 10 wt.%). 5.3. Major and minor element geochemistry Whole rock chemical analyses by ICP-MS of the samples from the Jabali drillcores (ZincOx plc.) have been published in Mondillo et al. (2011); a short resume, from which valuable information on the average geochemistry of the Jabali ores can be inferred, is shown in Table 6. When correlating the element concentrations of the three drillcores considered, the upper section of the J125 core is particularly enriched in Zn compared with the same section of both J109 and J138 cores. The Zn grade of the upper section of core J125 (15.2 wt.%) is two times higher Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 259 Fig. 8. Thin section petrography: A) J125-2. Smithsonite 1 replacing a saddle dolomite crystal (NII); B) J125-5. Smithsonite 1 replacing dolomite and sphalerite (NII); C) J125-7. Replacive smithsonite 1 mimicking dolomite (NII); D) the same as in panel C under CL: smithsonite shows strong red colors; E) J125-15. Zoned concretions of smithsonite 2 at the border of Zn-dolomite (NII); F) the same as in panel E under CL; G) J125-3. Smithsonite 1 cut by veins bordered by concretionary smithsonite 2, and then filled by gypsum (NII); H) J125-7. Cerussite directly replacing galena and host dolomite replaced by smithsonite 1 under CL: cerussite is green. Cer = cerussite; Gn = galena; Gp = gypsum; Sm 1 = smithsonite 1; Sm 2 = smithsonite 2: Zn-Dol = Zn-dolomite. Fig. 9. Secondary electron microscopy micrographs: A) JS MON 3. Smithsonite microcrystals with rhombohedral shape; B) J125-30. Smithsonite crystals, showing a combination of forms belonging to the scalenohedral class, i.e. hexagonal dipyramids, rhombohedra, scalenohedra. Author's personal copy 260 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 A 5000 intensity (counts) 4000 3000 saddle dolomite crystal rim saddle dolomite crystal core 2000 1000 0 350 500 650 800 wavelength (nm) B 14000 Zn-dolomite sulfides and gypsum) have been measured in the topmost zone of the J125 core. Silver shows mean relevant values in the upper (141 ppm) and lower (181 ppm) sections of the J125 drillcore. Cadmium concentration is on average very high (880 ppm). The chemical composition of 20 outcrop samples is given in Table 7. The specimens from the gossan are marked by very high Fe (over 30 wt.%), and locally by anomalous vanadium values. High sulfur contents are associated with the samples containing supergene sulfates and remnant sulfides. The Zn content can be very high (over 50 wt.% locally) in some outcrop samples collected from the massive ore body. The highest Zn concentrations (e.g. JS MON 22A, JS MON 29) are associated with smithsonite, or with a smithsonite– hydrozincite–hemimorphite assemblage. Samples characterized by discrete Pb contents (~5–6 wt.% Pb) are instead low in Zn. In outcrop samples, Cd generally occurs in higher amounts than Ag, with values locally higher than 3000 ppm. Copper can reach values higher than 500 ppm in a few mineralized samples. A strategic element like Ga reaches an anomalous content in the sample JS MON 22B (68 ppm), but it is not significant in most other samples. The environmentally dangerous Tl is not abundant at Jabali, with maximum concentrations of 50 ppm measured only in one sample. intensity (counts) 12000 10000 5.4. Carbon, oxygen and sulfur isotopes 8000 6000 4000 2000 0 350 500 650 800 wavelength (nm) C 20000 intensity (counts) 16000 smithsonite 1 smithsonite 2 12000 8000 4000 0 350 500 650 800 wavelength (nm) Fig. 10. Cathodoluminescence emission spectra of various Jabali carbonates: A) dolomite; B) Zn-bearing dolomite; C) smithsonite. than in the same section of the other two drillcores (averaging both 7.3 wt.%). The middle section of the J125 drillcore is the less Znrich (mean Zn grade around 10 wt.%), whereas the lowermost section is the richest (mean Zn grade around 18 wt.%). The Fe contents are similar in the mineralized sections of all cores, whereas significant Pb amounts are detected only in the upper zones of J125 and J138. Sulfur concentrations averaging around 2–3 wt.% (associated to both unaltered In order to characterize the fluid that precipitated the secondary mineralization, we conducted stable isotope (C, O) analyses of the nonsulfide Zn-minerals and host rock. A few S-isotope analyses were also carried out, to identify the origin of gypsum detected throughout the cores. Carbon and oxygen isotope analyses were conducted on Jurassic limestone samples (bioherms from Unit 7), saddle dolomite (associated with sulfide mineralization), Zn-dolomite (drillcores), smithsonite (drillcores and outcrop), and hydrozincite (outcrop) (Table 8, Fig. 12). Two Jurassic limestone samples from Unit 7 gave different results: in a coral sample a δ13C value of −1.0‰ and a δ18O value of −6.8‰ have been measured, while the carbonate matrix between the bioherms shows δ13C and δ18O values of − 2.3 and − 7.8‰, respectively. The saddle dolomite shows δ13C and δ18O values between −1.1 and 0.2‰ and −10.7 and −9.1‰, respectively. Zinc-bearing dolomite is characterized by δ13C values ranging from − 1.8 to − 1.0‰. δ18O values are between −11.9 and −10.4‰. Smithsonite specimens from drillcore and outcrop samples have been analyzed, excluding Zn-carbonate strongly intergrown with gypsum. In some cases, it was possible to sample replacive smithsonite 1 and clear concretionary smithsonite 2 from the same sample. However, the main difference in the isotope ratios is related to their provenance from the upper, middle, or lower zones. Smithsonite from the upper zone is characterized by δ13C values between − 4.7 and − 2.9‰, and by δ18O values between − 11.4 and − 10.5‰. One sample from the middle zone has a carbon and oxygen isotope composition of − 4.6‰ and − 9.6‰, respectively. The δ13C and δ18O values of smithsonite from the lower zone are between − 5.7 and − 4.7‰, and between − 10.0 and − 9.2‰, respectively. A smithsonite from the outcrop sample JS MON 22A has a δ13C value of − 6.1‰ and a δ18O value of −8.8‰, being comparable with the values for smithsonite from subsurface samples. Hydrozincite has carbon and oxygen isotope ratios between −7.4 and −6.6‰ and −4.9 and −0.8‰, respectively. Sulfur isotope analyses have been carried out on sphalerite, galena and gypsum (Table 9). Two sphalerite samples, from interval J125-2 to J125-5 have sulfur isotope ratios of 7.1 and 6.5‰. Galena, sampled from an outcrop vein shows a composition of − 3.6‰. Gypsum from sample J125-3 has a composition of 7.3‰, which is very similar to the δ34S value of sphalerite. Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 261 Fig. 11. Backscattered (BSE) images: A) J125-7. Anglesite replacing primary galena; B) J125-3. Secondary sulfides at the border of altered sphalerite: secondary galena, greenockite, covellite (locally containing Ag); C) J125-32. Ag-sulfide and smithsonite concretions; D) J125-3. Ag-sulfides between smithsonite and gypsum. Ang = anglesite; Cv = covellite; Gn = galena; Grn = greenockite; Gp = gypsum; Sm = smithsonite; Sp = sphalerite. 6. Discussion 6.1. Relationship between host rock and mineralization The Jabali secondary deposit was formed through the replacement process of the dolomite host rock by Zn-carbonates (Zn-bearing dolomite and smithsonite). The original lithotype was a carbonate platform limestone, which underwent strong dolomitization. Differently from Al Ganad et al. (1994), we could distinguish only two (not four) dolomite types: an early diagenetic dolomite and a hydrothermal dolomite. From our observation it becomes clear that original sulfides (sphalerite, galena, pyrite/marcasite) were strictly related to hydrothermal dolomite. This primary mineral association (hydrothermal dolomite and sulfides) was then widely altered and replaced by secondary phases. During alteration, dolomite was firstly replaced by calcite, following a typical de-dolomitization process: Ca 2þ ðaqÞ 2þ þ CaMgðCO3 Þ2ðsÞ → 2CaCO3ðsÞ þ Mg dolomite ðaqÞ : calcite In this reaction, Ca2+ is usually supplied by the dissolution of a mineral source, like gypsum or anhydrite (Bischoff et al., 1994; Coniglio, 2003). In some studies (Raines and Dewers, 1997), a supergene nature of dedolomitization is invoked, possibly related to post-burial surface weathering, when meteoric waters infiltrate a fractured dolostone, or during development of karstic processes. Coniglio (2003 and references therein) evidenced that Fe-bearing dolomite is more susceptible to dedolomitization than pure dolomite, because in the surface environment Fe2+ is easily altered to Fe3+, forming hydroxides, disrupting the dolomite structure, and causing the precipitation of calcite. A possible Ca2+ supply for the dedolomitization of Jabali dolomite, can derive from the abundant gypsum intercalations occurring in the Madbi and Sab'atayn Formations, which directly overlay the dedolomitized rock. The prevailing hydrothermal nature of Jabali dolomite, and its ferromanganoan character facilitated the dedolomitization process. A second and most important process that affected the dolomite host rock consists of its partial replacement by Zn-bearing dolomite phases. Zinc-bearing dolomite is always associated with remnants of the original dolomite. It is quite common to observe the hydrothermal dolomite crystals corroded and altered along their rims, along fractures and vugs by Zn-bearing dolomite patches. Zn-bearing dolomite never occurs as isolated, newly formed crystals in connection with dissolution. Chemical analyses of Zn-bearing dolomite indicate that in this phase the Zn-enrichment is proportional with Mg-decrease. Cathodoluminescence microscopy evidences that the luminescence of Znbearing dolomite is activated by Mn2+ substituting Ca2+ in the mineral structure, whereas the CL colors of the hydrothermal dolomite associated Table 6 Major and minor element concentrations of whole rock ICP chemical analyses (Mondillo et al., 2011 modified). Drillcore Interval Mineralized zone (m) J109 J125 57.30–62.70 Upper zone 50.78–70.00 Upper zone 74.50–80.50 Middle zone 84.50–91.50 Lower zone J138 68.00–75.00 Upper zone Mean value on the entire sample batch Zn Fe Mg Pb Ca Mn S (wt.%)a 7.3 15.2 10.3 17.6 7.3 12.7 Ag Cd Cu Ni P 800 988 673 1164 537 880 34 79 25 45 39 54 10 9 23 16 18 13 96 193 400 343 257 243 (ppm)b 2.6 2.6 3.6 3.9 3.4 3.1 9.1 6.3 4.3 6.7 8.1 6.7 1.3 3.4 0.5 1.4 3.0 2.4 19.5 12.3 21.6 12.9 18.9 15.5 0.6 0.5 0.7 0.5 0.7 0.6 0.2 2.5 0.1 0.1 0.2 1.2 15 141 5 181 76 103 Notes: Values derive from a weighted mean calculation referred to the length of core samples and the total length of the mineralized intervals (details in text). a Detection limits (wt.%): Zn 0.0001, Fe 0.01, Mg 0.01, Pb 0.01, Ca 0.01, Mn 0.0001, S 0.05. b Detection limits (ppm): Ag 0.5, Cd 0.5, Cu 0.5, Ni 0.5, P 5. Author's personal copy 262 Table 7 Major (wt.%) and minor (ppm) element concentrations of outcrop samples from ICP analysis. Sample JS MON n. 4 JS MON 6 JS MON 9 JS MON 12 JS MON 13 JS MON 14 JS MON 15 JS MON 17 JS MON 18 JS MON 19 JS MON 20 JS MON 21 JS MON 22A JS MON 28 JS MON 29 JS MON 32_1 JS MON 32_2 JS MON 33 JS MON 36 41.96 12.79 0.12 1.24 1.16 0.18 0.01 0.16 0.01 b0.01 0.03 0.23 63.6 952.3 2.7 604.5 35.9 5.9 6 3.1 b0.5 29 0.9 11 1.6 5.7 14 1.06 2.8 b0.5 68 23 48.30 5.83 0.2 0.61 0.16 0.39 0 0.07 b0.01 b0.01 0.01 b0.05 13.8 2225 1.9 122.9 63.9 12 5 1.6 b0.5 7 b0.5 11 1.2 3 b5 1.41 1.6 1.1 b5 21 51.04 2.38 0.22 0.02 0.27 0.86 0.01 0.42 0.04 0.04 0.02 b0.05 21.9 603.2 3 4 53.2 12.1 b5 1.1 0.6 9 b0.5 11 9.8 3.6 158 0.22 2 3.5 b5 22 41.58 2.46 0.61 1.23 4.71 1.07 0.01 0.56 b0.01 b0.01 0.02 0.08 15.3 1224 6.2 100.7 11.3 8.2 21 4.6 0.9 26 1.2 14 4.8 7.7 9 0.25 2.8 b0.5 13 16 23.52 3.2 5.1 0.99 12.49 0.42 0 0.16 b0.01 b0.01 0.01 0.17 81.6 1639.5 2.6 86.3 6.4 2.9 6 3.5 b0.5 37 b0.5 b10 9.9 2.9 8 0.17 2.7 b0.5 26 10 1.39 2.43 8.65 7.95 19.87 0.61 b0.001 0.06 b0.01 b0.01 0.01 0.82 350.7 73.8 1.8 7.6 4.5 1.6 5 1.6 b0.5 126 b0.5 13 20 4.1 30 0.37 0.6 0.5 b5 b2 0.13 1.79 2.52 0.00 33.84 0.64 0 0.03 b0.01 b0.01 0.01 b0.05 0.9 b0.5 b0.5 1.3 b0.5 b0.5 b5 0.7 b0.5 72 b0.5 b10 3.3 1.1 9 b0.05 b0.5 b0.5 b5 b2 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Zn 3.42 38.29 42.05 2.96 29.26 36.37 19.70 1.15 2.47 1.30 23.37 34.57 47.20 Fe 2.32 13.72 8.54 18.69 12.59 2.73 14.51 34.56 32.3 24.6 35.49 19.81 5.3 Mg 6.63 0.36 0.71 0.16 2.5 2.04 1.91 0.1 0.12 0.18 0.08 0.13 0.22 Pb 0.41 0.71 0.56 7.42 1.06 0.61 1.12 1.17 5.70 0.11 0.55 1.14 0.64 Ca 24.79 0.10 0.09 0.24 3.81 6.41 11.43 11.6 6.72 8.55 0.11 0.15 0.24 Mn 0.66 0.51 0.46 0.05 0.78 0.70 0.49 0.01 0.24 0.28 0.08 0.01 0.08 Ti 0.002 0.01 0 0.04 0.01 0 0.01 0.05 0.01 0.03 0.01 0.01 0 Al 0.05 0.30 0.07 1.2 0.13 0.4 0.32 0.33 2 0.7 0.13 0.34 0.05 Na 0.01 b0.01 b0.01 0.06 0.02 b0.01 b0.01 0.03 0.08 0.18 0.07 0.01 b0.01 K b0.01 b0.01 b0.01 0.67 0.02 b0.01 b0.01 0.16 0.08 0.02 0.04 b0.01 b0.01 P 0.03 0.02 0.02 0.03 0.02 0.02 0.02 0.06 0.08 0.11 0.03 0.03 0.01 S b0.05 0.05 b0.05 5.79 b0.05 0.06 0.09 0.47 1.9 0.12 0.2 0.1 b0.05 Ag 9.4 3.4 5.1 37.9 9.6 44.4 58.4 3.9 119.6 3.4 3.6 10.9 108 Cd 319 1439.7 2307.5 22.1 3722.7 1358.2 715.3 546.9 38.5 69.3 185.9 1499.8 2952.6 Mo 1.4 3.6 2.9 3.3 4.4 5.3 6.4 29.8 5.8 8.3 7.5 2.2 0.9 Cu 13.8 49.1 30.4 76.8 25.9 14 9.2 30.7 60 378.9 7.3 79.2 308.1 Ni 9.5 128.3 152.6 23.4 240.6 38.7 43.5 12.5 52 74.4 34.8 19.5 37 Co 1.8 8.4 6.2 3.9 17.8 9.1 4.9 1.3 13.7 166.4 2.2 1.7 4.9 As 6 b5 b5 7 b5 b5 18 8 33 b5 23 34 b5 U 2.4 1.5 1.1 0.9 1.7 4.2 4 1.4 6.1 1.4 2.5 1.9 1.7 Th b0.5 b0.5 b0.5 2.7 b0.5 b0.5 0.5 0.9 1.4 1.8 b0.5 b0.5 b0.5 Sr 55 b5 b5 117 5 26 134 41 160 96 9 8 6 b0.5 b0.5 b0.5 b0.5 b0.5 b0.5 b0.5 Sb b0.5 b0.5 b0.5 b0.5 b0.5 b0.5 V 16 11 11 19 10 11 16 180 21 54 15 15 b10 La 5.7 4.9 2.5 10.3 5 3.2 3.7 1.9 14.6 15.7 2.4 1.1 0.6 Cr 9.6 5 3.8 22 8.4 3.6 4.7 48.7 20.9 47.2 6.3 6.4 2.1 Ba 41 b5 b5 400 11 22 21 56 80 75 b5 b5 b5 Hg 0.35 0.19 0.25 1.32 1.04 0.18 0.28 0.2 0.2 b0.05 0.36 0.4 0.44 Sc 1.2 1 1.2 1.9 0.9 2 2.9 2.5 7.1 3.3 4 1.3 1.1 Tl 1.1 0.5 0.7 52.8 5.2 13.2 4.1 b0.5 2.1 2.9 0.9 b0.5 b0.5 Ga b5 b5 b5 b5 b5 b5 6 16 27 28 b5 10 22 Se 2 17 20 9 9 17 9 b2 b2 2 8 14 22 Detection limits (wt.%): Zn 0.01, Fe 0.01, Mg 0.01, Ca 0.01, Mn 0.0005, Ti 0.001, Al 0.01, Na 0.01, K 0.01, P 0.001, S 0.05. Detection limits (ppm): Ag 0.5, Cd 0.5, Mo 0.5, Cu 0.5, Ni 0.5, Co 0.5, As 5, Th 0.5, Sr 5, Sb 0.5, V 10, La 0.5, Cr 0.5, Ba 5, Hg 0.05, Sc 0.5, Tl 0.5, Ga 5, Se 2. JS MON 22B Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 263 Table 8 Carbon and oxygen isotopes of Jabali carbonates (outcrop and drillcore samples). Sample n. Mineral Description δ13C (‰ VPDB) δ18O (‰ VPDB) −6.83 −7.75 −9.03 −8.80 −0.79 −4.90 −4.82 δ18O (‰ VSMOW) JS MON 38 JS MON 38 JS MON 3 JS MON 22A JS MAR 4 JS MON 22A JS MON 22A Outcrop Cal Cal Sm Sm Hz Hz Hz Coral fossils from limestone Calcite matrix between bioherms from limestone Concretionary smithsonite Concretionary smithsonite Hydrozincite crust Hydrozincite crust Hydrozincite crust −0.97 −2.30 −2.61 −6.06 −6.75 −7.37 −6.62 Core sample n. Depth (m) Mineral Description δ13C (‰ VPDB) δ18O (‰ VPDB) δ18O (‰ VSMOW) J125-6 J125-7 J125-15 J109-5 J125-11 J125-11 J125-15 J125-9 J125-10 J125-10 J125-14 J125-15 J125-21 J125-30 J125-30 J125-30A J125-30B J125-32 J125-32 J125-32A J125-32B 59.2 60.8 69.2 62.6 65.2 65.8 69.8 63.7 64.1 64.2 68.7 69.1 77 85.6 85.65 85.7 85.7 87.6 87.7 87.75 87.75 Dol Dol Dol Zn-Dol Zn-Dol Zn-Dol Zn-Dol Sm Sm Sm Sm Sm Sm Sm Sm Sm Sm Sm Sm Sm Sm Hydrothermal dolomite crystal Hydrothermal dolomite crystal Hydrothermal dolomite crystal Dolomite crystal with Zn-dolomite core Dolomite crystal with Zn-dolomite core Dolomite crystal with Zn-dolomite core Dolomite crystal with Zn-dolomite core Vacuolar smithsonite Massive smithsonite Concretionary smithsonite Smithsonite concretion Smithsonite concretion Massive smithsonite Smithsonite concretion Smithsonite concretion Mixed massive-concretionary smithsonite Smithsonite micro-concretions Massive smithsonite Concretionary smithsonite Smithsonite concretion Smithsonite concretion −1.10 −1.14 0.23 −1.85 −0.97 −1.17 −1.19 −3.75 −3.40 −4.16 −2.88 −4.68 −4.56 −5.48 −4.85 −5.22 −5.47 −4.83 −4.92 −5.71 −4.69 −9.36 −9.12 −10.70 −11.29 −11.95 −11.16 −10.42 −11.19 −10.83 −10.50 −11.41 −11.41 −9.57 −9.85 −9.99 −9.87 −9.93 −9.20 −9.49 −9.78 −9.85 21.21 21.46 19.83 19.22 18.54 19.36 20.11 19.33 19.70 20.03 19.10 19.09 20.99 20.70 20.56 20.69 20.62 21.38 21.08 20.78 20.71 23.82 22.87 21.55 21.79 30.04 25.81 25.89 Notes: Mineral abbreviations mostly after Whitney and Evans (2010). Cal = calcite; Dol = dolomite; Hz = hydrozincite; Sm = smithsonite; Zn-Dol = Zn-dolomite. with primary sulfides were activated by Mn2 + occupying the Mgposition. This phenomenon suggests that during the secondary process Zn2 + substitutes for Mg2 + in the dolomite structure, and that this replacement causes a decrease of Mn in the same structural positions. The occurrence of Fe- and Mn-hydroxides, associated with Zn-bearing dolomite, indicates that the Zn2 + substitution for Mg2 + occurred at the same time as the oxidation of Mn2+ and Fe2+ previously contained in the dolomite structure. In our opinion the Zn-rich waters, which caused the replacement of host dolomite by Zn-bearing dolomite, were not related to the primary hydrothermal fluids depositing sulfides, but were rather original groundwaters that altered sphalerite and brought Zn2 + in solution (Boni et al., 2011). Strictly hydrothermal Zn-bearing dolomites related to the emplacement of sulfides were described in the Navan Zn–Pb deposit (Ireland) by Kucha and Wieczorek (1984), at Broken Hill (New South Wales) by Riaz Khan and Barber (1990), and in the Silesia–Cracow district (Poland) by Coppola et al. (2009 and references therein). In all these cases, Zn-bearing dolomite precipitated together or before sulfides, as newly formed zoned crystals. In the mining district of southwestern Sardinia (Italy), supergene alteration and oxidation of sulfides and formation of Zn-oxidized deposits (Boni et al., 2003) occurred after a regional scale hydrothermal dolomitization event. In this district, a Zn-dolomite of strict supergene origin, replacing the previous hydrothermal dolomite, was detected with characteristics similar to those we have described from Jabali Table 9 Sulfur isotopes of Jabali sulfides and sulfates. Sample n. Mineral Description δ34S (‰ VCDT) JS MAR 2 J125-2 J125-5 J125-3 Gn Sp Sp Gp Macrocrystalline galena agglomerate Sphalerite vein Sphalerite vein Gypsum vein within smithsonite −3.58 7.13 6.47 7.34 Abbreviations: Gn = galena; Gp = gypsum; Sp = sphalerite. (Boni et al., 2013). Another case of Zn-dolomite of supergene origin was also identified by Boni et al. (2011) and Mondillo et al. (2014) in the Yanque Zn–Pb deposit (Peru). The economically most relevant replacement of the original dolomite host rock resulted in the precipitation of the Zn-carbonate smithsonite (smithsonite 1), consisting of rhombohedral microcrystals in agglomerates mimicking the saddle shapes of hydrothermal dolomite macrocrystals, and also filling dissolution cavities in the dolomite. The red CL color indicates that smithsonite 1 inherited variable Mn amounts from original dolomite. Smithsonite 2 forms crystals and concretions occurring in veins and solution vugs of the host rock. The blue CL color of smithsonite 2, activated by lattice defects, indicates a poor crystalline state of this late phase. Gypsum, together with smithsonite 2 occurs in veins cutting smithsonite 1. These textural observations point to: 1) an increase in dolomite instability during the alteration process, 2) a partial dissolution of the dolomite, followed by smithsonite 1 precipitation, and 3) a late precipitation of smithsonite 2 and gypsum. Dolomite is generally unstable in aqueous solutions characterized by low pH, and goes into solution releasing Ca2 + and Mg2 + to carbonate and bicarbonate anionic groups (Busenberg and Plummer, 1982; Morse and Arvidson, 2002; Pokrovsky et al., 1999, 2005, 2009). After Sangameshwar and Barnes (1983), during the oxidation of a sulfide body by groundwaters, the aqueous solutions can carry metals (e.g. Zn2+) only if they have an acid pH; in this setting, the acidity of waters is mainly produced by the oxidation of pyrite, which releases sulfate anionic groups into solution. In the Jabali case, the acid Znbearing solutions altered and partially dissolved the dolomite of the host rock, releasing Ca2 + and Mg2 +. The presence of the carbonate anionic groups in solution together with Zn2+ cations favored firstly the precipitation of smithsonite 1, followed by the precipitation of smithsonite 2. Mg2 + was partially incorporated in the smithsonite structure, whereas Ca2+ precipitated mainly in gypsum from the reaction with sulfate anionic groups in solution. Author's personal copy 264 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 2.0 6.2. Fluids involved in the oxidation of sulfides and formation of secondary mineralization 13C (‰ VPDB) 0.0 -2.0 -4.0 -6.0 -8.0 -10.0 16.0 18.0 20.0 22.0 18O 24.0 26.0 28.0 30.0 32.0 (‰ VSMOW) calcite corals hydrothermal dolomite Zn-bearing dolomite J109-5 Zn-bearing dolomite J125-11 to 15 hydrozincite outcrop smithsonite outcrop smithsonite J125-9 to 15 smithsonite J125-21 smithsonite J125-30 to 32 Fig. 12. δ18O vs. δ13C compositions of Jabali carbonates (values in Table 8). This interpretation is supported by the δ34S composition of the late gypsum veins at Jabali, being very different from the published δ34S values of Mesozoic marine sulfates (Kampschulte and Strauss, 2004; Strauss, 1999), but similar to the δ34S ratio measured in the primary sphalerite. A possible scheme of the alteration processes which affected the studied deposit can be exemplified by the following chemical reactions: oxidation of sphalerite 2− 2þ 1) ZnS þ 2O2 →Zn þ SO4 sphalerite 2− 2þ þ − 2) ZnS þ 4H2 O→Zn þ SO4 þ 8H þ 8e sphalerite formation of Zn-bearing carbonates 3) CaMgðCO3 Þ2 þ Zn2þ þ SO2− 4 → dolomite →Ca Mg1−x; Znx ðCO3 Þ2 þ ð1−xÞZn2þ þ xMg2þ þ SO2− 4 Zn‐dolomite Ca Mg1−x; Znx ðCO3 Þ2 þ ð1−xÞZn2þ þ xMg2þ þ SO4 2 þ 2H2 O→ 4) Zn‐dolomite →Zn2þ þ Mg2þ þ Ca2þ þ 2CO3 2 þ SO4 2 þ 2H2 O→ →2ðZn; MgÞCO3 þ Mg‐smithsonite CaSO4 2H2 O gypsum: The association of secondary Ag- and Cd-sulfides and nonsulfide Zn phases (smithsonite 2 and hemimorphite) plus gypsum is very characteristic at Jabali. Following Sangameshwar and Barnes (1983), who traced the chemical reactions resulting in nonsulfide mineralizations at the Burgin mine (USA) and at Tynagh (Ireland), the co-precipitation of smithsonite and Ag-sulfides occurs at temperatures between 25° and 60 °C, only at neutral pH, if the Eh varies between 0 and − 2 V, because the stability fields of the two phases are very close, but not superimposed. The mentioned values give a constraint on the pH and Eh ranges at Jabali during precipitation of smithsonite 2, which followed the replacement of dolomite by smithsonite 1. The δ13C values measured in the reef corals from the Jabali undolomitized limestone are in the range of carbon isotope ratios of Jurassic (Kimmeridgian–Tithonian) marine carbonates; their δ18O compositions, instead, are lower than those commonly reported in literature (Jenkyns et al., 2002). The modification in oxygen isotope data can be related to a diagenetic alteration during the stabilization of the primarily aragonitic skeletons. The δ13C and δ18O ratios of the saddle dolomite are typical of other hydrothermal dolomites in the world (Boni et al., 2013; Diehl et al., 2010; Radke and Mathis, 1980). The δ13C composition of Zn-bearing dolomite at Jabali is in the same range of hydrothermal dolomite in the area, from which it is inherited most likely. The δ18O ratio probably results from a mixing between the isotopic composition of hydrothermal dolomite and that of the Zn-carrying fluid that deposited smithsonite. The δ13C values of smithsonite are always negative: they are heavier in the specimens from the upper zone of the deposit, than in those from the middle and lower zones. This range in the carbon isotope values is typical of most supergene nonsulfide Zn deposits in the world and interpreted as a result of mixing between carbonate carbon from the host rock and soil/atmospheric CO2 (Gilg et al., 2008). The negative carbon isotope ratios of the Jabali samples may be explained by a contribution of isotopically light organic carbon from weathered soils or from the black shales of Unit 8. The Jabali smithsonite shows δ18O values substantially lower and with a larger variability in comparison with other smithsonites considered as supergene in the literature (Gilg et al., 2008). As suggested by Gilg et al. (2008), this variability may indicate effects of temperaturerelated fractionation. If the δ18O value of the solution from which smithsonite was formed can be approximated, the precipitation temperature can be calculated using the following equation (Gilg et al., 2008):1000 ln αsmithsonite-water = 3.10 (106/T2) − 3.50. Modern meteoric waters in Yemen have a δ18O composition ranging from −8.0 to +10.5‰ VSMOW (Al-Ameri, 2011). In the Jabali region, several springs about 20 km south of the deposit area have an oxygen isotope composition between − 4.2 and − 3.6‰ VSMOW (Minissale et al., 2007). The δ18O values of Pleistocene to Holocene speleothems measured in various caves of Saudi Arabia and Yemen point to an oxygen isotope composition of groundwater ranging between ~ − 12 and ~0‰ VSMOW (Fleitmann et al., 2004, 2011). Thus both rainwater and groundwater in the region are characterized by a wide range in δ18O. However, δ18O compositions of groundwaters are always negative. It follows that 1) if the smithsonites from the ore deposit precipitated from a fluid with a δ18O value comparable to modern groundwater, the precipitation temperature would have been between ~ 55 and ~ 65 °C, corresponding to a low-temperature hydrothermal environment; 2) if the smithsonites precipitated from a fluid with a δ18O value comparable to Pleistocene to Holocene cave water, the precipitation temperature would have been substantially lower (30–40 °C), corresponding to a normal weathering environment at this latitude. Variability in δ18O values throughout the different mineralized zones could indicate different temperatures or processes active during smithsonite precipitation, and/or different stages of smithsonite formation. Unusually low and variable δ18O values were measured in the Sierra Mojada smithsonites by Hye In Ahn (2010) (Fig. 13). The δ18O composition of modern groundwaters in the Sierra Mojada is particularly low (−8‰ VSMOW) translating into estimated precipitation temperatures of about 33 °C for Zn-carbonate. In the Angouran Zn deposit (Iran), unusual δ18O values in smithsonite were related to a hypogene genesis of the nonsulfide ores (Fig. 13), associated with local travertine deposition (Boni et al., 2007). The mineralizing fluids were low-temperature hydrothermal, probably circulating during the waning stages of Tertiary–Quaternary volcanic activity in the area (Boni et al., 2007). Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 265 8.0 6.0 1 Supergene smithsonite Gilg et al. 2008 2 Angouran smithsonite 1 Boni et al. 2007 3 Sierra Mojada smithsonite Hye In Ahn 2010 2 4.0 13 C (‰ VPDB) 2.0 0.0 3 Jabali smithsonite this study -2.0 -4.0 1 -6.0 -8.0 -10.0 -12.0 16.0 18.0 20.0 22.0 18O 24.0 26.0 28.0 30.0 32.0 (‰ VSMOW) Fig. 13. δ18O vs. δ13C of Jabali smithsonite, compared to smithsonites from other nonsulfide deposits/districts: 1 = supergene smithsonite (Gilg et al., 2008), 2 = Angouran smithsonite 1 (Boni et al., 2007), 3 = Sierra Mojada smithsonite (Hye In Ahn, 2010). Several travertine successions associated with thermal springs have been recorded in the Marib volcanic district (about 40 km from Jabali), as well as south of Jabal as Saad (about 15 km from Jabali). Their occurrence may suggest that a deep hydrothermal circulation involving oxygenated waters could have been active also in the Jabali area. This setting, paired with the previously quoted examples, could support a model in which Jabali smithsonite precipitated from fluids characterized by different temperatures, consisting of local groundwaters variably mixed with low-temperature hydrothermal waters. The same mechanism may be assumed for precipitation of Zn-bearing dolomite. The carbon and oxygen isotope ratios of hydrozincite are both in the range of supergene Zn-carbonates (Gilg et al., 2008). This is a proof of Early diagenetic dolomite the most recent formation of hydrozincite, precipitated from meteoric waters at the surface (Takahashi, 1960). The results of our study do not confirm a possible beginning of the oxidation process during a Cretaceous weathering stage, as stated by Al Ganad et al. (1994). Instead, our observations are compatible with two different oxidation processes, probably partially combined (Fig. 14): 1) supergene weathering, mainly developed from the early Miocene (~17 Ma) when major uplift and exhumation in Yemen commenced as a result of the main phase of Red Sea extension (Menzies et al., 1992), and continued until present; 2) oxidation related to lowtemperature hydrothermal circulation in combination with magmaticinduced geothermal activity in the area (Miocene–Holocene). Late Jurassic Cretaceous (?) Final stages of host rock deposition Final Mesozoic stages of Sab'atayn basin evolution Sulfide formation Miocene Beginning of major uplift phases and geothermal activity Present Arid climate Weathering and oxidation Hydrothermal dolomite Sphalerite Galena Pyrite Calcite (dedolomitization) Zn-bearing dolomite Smithsonite 1 Anglesite - Cerussite Smithsonite 2 Acanthite - Greenockite Gypsum Hemimorphite Hydrozincite Fig. 14. Paragenesis of the main mineralogical phases observed at Jabali, framed in the geological evolution of the region. Author's personal copy 266 N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 7. Conclusions This study sheds new light on many characteristics of the Jabali secondary nonsulfide mineralization and on the related genetic processes. Smithsonite is the most abundant economic mineral in the secondary deposit, where it is associated with minor hemimorphite, hydrozincite and Ag-sulfides. The secondary mineralization evolved through different stages: 1) alteration of original sulfides (sphalerite, pyrite and galena), and release of metals in acid solutions; 2) alteration of dolomite host rock and formation of Zn-bearing dolomite; 3) partial dissolution of dolomite by metal-carrying acid fluids and replacement of dolomite and Zn-bearing dolomite by a first smithsonite phase (smithsonite 1). To this stage also belong the direct replacement of sphalerite and galena by secondary minerals (smithsonite and cerussite); 4) precipitation of a second smithsonite phase (smithsonite 2) in veins and cavities, together with gypsum and Ag-sulfides. The Jabali smithsonite has variable δ18O values in different parts of the orebody, possibly associated with different environments or temperatures during smithsonite precipitation, and/or with different stages of smithsonite formation. The δ18O values are also generally lower than those of other smithsonites derived from pure weathering processes, whereas the carbon isotope composition is in the same range of values for supergene Zn-carbonates from other nonsulfide ores (Gilg et al., 2008). With this scenario, and considering the negative δ18O values of groundwaters and paleo-groundwaters in this area of Yemen, we argue that the Jabali smithsonite may have precipitated from a combination of fluids, possibly consisting of local groundwaters variably mixed with low-temperature hydrothermal waters. The most favorable setting for the development of supergene deposit with these geochemical characteristics could have been initiated in the early Miocene (~ 17 Ma) and continued until Recent. Lowtemperature, hydrothermal circulation at Jabali could have been possibly active through the magmatically-induced geothermal activity (Miocene–Holocene) in the area. Acknowledgments This study is part of the PhD Thesis of N. Mondillo at the University of Napoli “Federico II”. The authors would like to thank A. Woollett (ZincOx), and especially B. Grist (former ZincOx) for his help during the fieldwork and drillcore sampling. We are indebted to R. de' Gennaro of the CISAG Napoli for his support during SEM analyses. We also thank J. Götze for his help and discussion on the cathodoluminescence analyses, that were carried out at the TU Bergakademie Freiberg (Germany) in the frame of a joint Deutsche Forschungsgemeinschaft project between Freiberg and Napoli Universities. The careful comments of D. Large and of two other anonymous reviewers of Ore Geology Reviews have greatly improved the quality of the manuscript. A special thank is due to F. Pirajno for final editing. References Ahlbrandt, T.S., 2002. Madbi Amran/Qishn Total Petroleum System of the Ma'Rib-Al Jawf-Shabwah, and Masila-Jeza Basins, Yemen. USGS Bull. 2202-G, 1–28. Al Ganad, I., 1991. Etude géologique du gisement Zn-Pb-Ag de Jabali (Bordure sud du bassin du Wadi al Jawf). Unpublished doctoral thesis, Université Orleans (France), 180 pp. Al Ganad, I., Lagny, Ph., Lescuyer, J.L., Rambo, C., Touray, J.C., 1994. Jabali, a Zn–Pb–(Ag) carbonate-hosted deposit associated with Late Jurassic rifting in Yemen. Mineral. Deposita 29, 44–56. Al-Ameri, A., 2011. Regional stable isotope and hydrochemistry investigation in Yemen and in the representative area “Sana'a basin”. Unpublished doctoral thesis, University of Berlin (Germany), 152 pp. As-Saruri, M.A., Sorkhabi, R., Baraba, R., 2010. Sedimentary basins of Yemen: their tectonic development and lithostratigraphic cover. Arab. J. Geosci. 3, 515–527. Baker, J., Snee, L., Menzies, M., 1996. A brief Oligocene period of flood volcanism in Yemen: implications for the duration and rate of continental flood volcanism at the Afro-Arabian triple junction. Earth Planet. Sci. Lett. 138, 39–55. Beydoun, Z.R., 1997. Introduction to the revised Mesozoic stratigraphy and nomenclature for Yemen. Mar. Pet. Geol. 14, 617–629. Beydoun, Z.R., As-Saruri, M.L., El-Nakhal, H., Al-Ganad, I.N., Baraba, R.S., Nani, A.O., Al-Aawah, M.H., 1998. International lexicon of stratigraphy, Republic of Yemen. VIII, Asia, fasc, 10b2. IUGS Publ 34, 1–245. Bischoff, J.L., Wayne, R.J., Shanks, C., Rosenbauer, R.J., 1994. Karstification without carbonic acid: bedrock dissolution by gypsum-driven dedolomitization. Geology 22, 995–998. Bish, D.L., Howard, S.A., 1988. Quantitative phase analysis using the Rietveld method. J. Appl. Crystallogr. 21, 86–91. Bish, D.L., Post, J.E., 1993. Quantitative mineralogical analysis using the Rietveld fullpattern fitting method. Am. Mineral. 78, 932–940. Boni, M., Gilg, H.A., Aversa, G., Balassone, G., 2003. The “Calamine” of SW Sardinia (Italy): geology, mineralogy and stable isotope geochemistry of a supergene Zn-mineralization. Econ. Geol. 98, 731–748. Boni, M., Gilg, H.A., Balassone, G., Schneider, J., Allen, C.R., Moore, F., 2007. Hypogene Zn carbonate ores in the Angouran deposit, NW Iran. Mineral. Deposita 42, 799–820. Boni, M., Mondillo, N., Balassone, G., 2011. Zincian dolomite: a peculiar dedolomitization case? Geology 39, 183–186. Boni, M., Mondillo, N., Balassone, G., Joachimski, M., Colella, A., 2013. Zincian dolomite related to supergene alteration in the Iglesias mining district (SW Sardinia). Int. J. Earth Sci. (Geol. Rundsch.) 102, 61–71. Bosence, D.W.J., 1997. Mesozoic rift basins of Yemen. In: Bosence, D.W.J. (Ed.), Special issue on Mesozoic rift basins of Yemen. Mar. Pet. Geol, 14, pp. 611–730. Bosworth, W., Huchon, P., McClay, K., 2005. The Red Sea and Gulf of Aden basins. J. Afr. Earth Sci. 43, 334–378. Brannan, J., Gerdes, K.D., Newth, I.R., 1997. Tectono-stratigraphic development of the Qamar basin, Eastern Yemen. Mar. Petrol. Geol. 14, 701–730. Busenberg, E., Plummer, L.N., 1982. The kinetics of dissolution of dolomite in CO2–H2O systems at 1.5 to 65 °C and 0 to 1 atm pCO2. Am. J. Sci. 282, 45–78. Christmann, P., Lagny, P., Lescuyer, J.L., Sharaf Ad Din, A., 1983. Résultats de trois années de prospection en République arabe du Yemen. Découverte du gisement de Jabali (Pb–Zn–Ag) dans la couverture jurassique. Chron. Réch. Min. 473, 25–38. Christmann, P., Lagny, P., Lescuyer, J.L., Sharaf Ad Din, A., 1989. Discovery of the Jabali deposit (Zn–Pb–Ag) in the Jurassic cover of the Yemen Arab Republic. Chron. Réch. Min. Spec. Issue 43–52. Coniglio, M., 2003. Dedolomitization. In: Middleton, G.V. (Ed.), Encyclopedia of Sediments and Sedimentary Rocks. Kluwer Academic Encyclopedia of Earth Sciences Series, 23. Springer, Dordrecht, pp. 188–190. Coppola, V., Boni, M., Gilg, H.A., Strzelska-Smakowska, B., 2009. Nonsulfides zinc deposits in the Silesia–Cracow district, Southern Poland. Miner. Deposita 44, 559–580. Coulié, E., Quidelleur, X., Gillot, P.Y., Courtillot, V., Lefèvre, J.C., Chiesa, S., 2003. Comparative K–Ar and Ar/Ar dating of Ethiopian and Yemenite Oligocene volcanism: implications for timing and duration of the Ethiopian traps. Earth Planet. Sci. Lett. 206, 477–492. Csato, I., Habib, A., Kiss, K., Kocz, I., Kovacs, Z., Lorincz, K., Milota, K., 2001. Play concepts of oil exploration in Yemen. Oil Gas J. 99, 68–74. Diehl, S.F., Hofstra, A.H., Koenig, A.E., Emsbo, P., Christiansen, W., Johnson, C., 2010. Hydrothermal zebra dolomite in the Great Basin, Nevada — attributes and relation to Paleozoic stratigraphy, tectonics, and ore deposits. Geosphere 6, 663–690. Ellis, A.C., Kerr, H.M., Cornwell, C.P., Williams, D.O., 1996. A tectono stratigraphic framework for Yemen and its implications for hydrocarbon potential. Petr. Geosci. 2, 29–42. Fleitmann, D., Matter, A., Pint, J.J., Al-Shanti, M.A., 2004. The speleothem record of climate change in Saudi Arabia. In: Saudi Geological Survey (Ed.), Open-file Report, SGS-OF2004-8 (46 pp.). Fleitmann, D., Burns, S.J., Pekala, M., Mangini, A., Al-Subbary, A., Al-Aowah, M., Kramers, J., Matter, A., 2011. Holocene and Pleistocene pluvial periods in Yemen, southern Arabia. Quat. Sci. Rev. 30, 783–787. Gilg, H.A., Boni, M., Hochleitner, R., Struck, U., 2008. Stable isotope geochemistry of carbonate minerals in supergene oxidation zones of Zn–Pb deposits. Ore Geol. Rev. 33, 117–133. Götze, J., 2012. Application of cathodoluminescence microscopy and spectroscopy in geosciences. Microsc. Microanal. 18, 1270–1284. Grist, B., 2006. Development of Yemen's first large scale base metal mine. 9th Arab Conference for Mineral Resources (Jeddah (Saudi Arabia), 7 pp.). Hill, R.J., 1991. Expanded use of the Rietveld method in studies of phase abundance in multiphase mixtures. Powder Diffract. 6, 74–77. Hye In Ahn, B.S., 2010. Mineralogy and geochemistry of the nonsulfide Zn deposits in the Sierra Mojada district, Coahuila, Mexico: unpublished doctoral thesis. University of Texas at Austin, USA 193 pp. Jenkyns, H.C., Jones, C.E., Gr cke, D.R., Hesselbo, S.P., Parkinson, D.N., 2002. Chemostratigraphy of the Jurassic System: application, limitations and implications for palaeoceanography. J. Geol. Soc. Lond. 159, 351–378. Kampschulte, A., Strauss, H., 2004. The sulfur isotopic evolution of Phanerozoic seawater based on the analysis of structurally substituted sulfate in carbonates. Chem. Geol. 204, 255–286. Kim, S.T., Mucci, A., Taylor, B.E., 2007. Phosphoric acid fractionation factors for calcite and aragonite between 25 and 75 °C: Revisited. Chem. Geol. 246, 135–146. Kucha, H., Wieczorek, A., 1984. Sulfide–carbonate relationships in the Navan (Tara) Zn–Pb deposit, Ireland. Miner. Deposita 19, 208–216. Larson, A.C., Von Dreele, R.B., 2000. GSAS general crystal structure analysis system. Report No. LAUR-86-748Los Alamos National Laboratory, NM, USA. Mattash, M.A., 2008. Geology and geochemistry of the sedimentary-hosted lead–zinc– vanadium–barite prospects in Yemen. Abstract volume International Geological Congress Oslo, 6–14 August 2008 (1 pp.). Author's personal copy N. Mondillo et al. / Ore Geology Reviews 61 (2014) 248–267 Mattash, M.A., Vaselli, O., Minissale, A., Ad-Dukhain, A., Hazza, M., 2005. The first geothermal resources map of Yemen, at a 1: 125 000 scale. In: Ministry of Oil and Mineral Resources (Ed.), Sana'a (Republic of Yemen). Menzies, M.A., Baker, J., Bosence, D., Dart, C., Davidson, I., Hurford, A., Al-Kadasi, M., McClay, K., Nichols, G., Al-Subbary, A., Yelland, A., 1992. The timing of magmatism, uplift and crustal extension: preliminary observations from Yemen. In: Storey, B.C., Alabaster, T., Pankhurst, R.J. (Eds.), Magmatism and the Causes of Continental Break-up. Geological Society of London: Spec. Publ., 68, pp. 293–304. Menzies, M., Al-Kadasi, M.A., Al-Khirbash, S., Al-Subbary, A.K., Baker, J., Blakey, K.R., Nichols, G., Yelland, A., 1994. Geology of Yemen. In: Geological Survey and Mineral Exploration Board, Ministry of Oil and Mineral Resources, Republic of Yemen (Ed.), Geology and Mineral Resources of Yemen, pp. 21–48. Minissale, A., Mattash, M.A., Vaselli, O., Tassi, F., Al-Ganad, I.N., Selmo, E., Shawki, N.M., Tedesco, D., Poreda, R., Ad-Dukhain, A.M., Hazzae, M.K., 2007. Thermal springs, fumaroles and gas vents of continental Yemen: their relation with active tectonics, regional hydrology and the country's geothermal potential. Appl. Geochem. 22, 799–820. Mondillo, N., Boni, M., Balassone, G., Grist, B., 2011. In search of the lost zinc: a lesson from the Jabali (Yemen) nonsulfide zinc deposit. J. Geochem. Explor. 108, 209–219. Mondillo, N., Boni, M., Balassone, G., Villa, I.M., 2014. The Yanque Prospect (Peru): from polymetallic Zn–Pb mineralization to a nonsulfide deposit. Econ. Geol. (in press). Morse, J.W., Arvidson, R.S., 2002. The dissolution kinetics of major sedimentary carbonate minerals. Earth Sci. Rev. 58, 51–84. Moseley, F., 1969. The Aden traps of Dhala, Musaymir and Radfan, South Yemen. Bull. Volcanol. 33, 889–909. Neuser, R.D., Bruhn, F., Götze, J., Habermann, D., Richter, D.K., 1995. Kathodolumineszenz: Methodik und Anwendung. Zbl. Geol. Paläontol. I 1/2, 287–306. Pokrovsky, O.S., Schott, J., Thomas, F., 1999. Dolomite surface speciation and reactivity in aquatic systems. Geochim. Cosmochim. Acta 63, 3133–3143. Pokrovsky, O.S., Golubev, S.V., Schott, J., 2005. Dissolution kinetics of calcite, dolomite and magnesite at 25 °C and 0 to 50 atm pCO2. Chem. Geol. 217, 239–255. Pokrovsky, O.S., Golubev, S.V., Schott, J., Castillo, A., 2009. Calcite, dolomite and magnesite dissolution kinetics in aqueous solutions at acid to circumneutral pH, 25 to 150 °C and 1 to 55 atm pCO2: new constraints on CO2 sequestration in sedimentary basins. Chem. Geol. 265, 20–32. Radke, B.M., Mathis, R.L., 1980. On the formation and occurrence of saddle dolomite. J. Sediment. Petrol. 50, 1149–1168. Raines, M.A., Dewers, T.A., 1997. Dedolomitization as a driving mechanism for karst generation in Permian Blaine Formation, southwestern Oklahoma. USA: Carbonates and Evaporites 12, 24–31. Reynolds, N.A., Large, D., 2010. Tethyan zinc–lead metallogeny in Europe, north Africa, and Asia. Soc. Econ. Geol. Spec. Pub. 15, 339–365. Riaz Khan, M., Barber, D.J., 1990. Composition-related microstructures in Zinc-bearing carbonate assemblages from Broken Hill, New South Wales. Miner. Petrol. 41 (2-4), 229–245. Rietveld, H.M., 1969. A profile refinement method for nuclear and magnetic structures. J. Appl. Crystallogr. 2, 67–71. 267 Roedder, E., 1976. Fluid inclusion evidence for the genesis of ores in sedimentary and volcanic rocks. In: Wolf, K.H. (Ed.), Ores in Sediments, Sedimentary and Volcanic Rocks, 2. Elsevier, Amsterdam, pp. 67–110. Rosenbaum, J., Sheppard, S.M., 1986. An isotopic study of siderites, dolomites and ankerites at high temperatures. Geochim. Cosmochim. Acta 50, 1147–1150. Sangameshwar, S.R., Barnes, H.L., 1983. Supergene processes in zinc–lead–silver sulfide ores in carbonates. Econ. Geol. 78, 1379–1397. SRK Consulting, 2005. Jabali feasibility study, geology and resources. Unpublished report, ZincOx Resources plc., 45 pp. Strauss, H., 1999. Geological evolution from isotope proxy signals. Sulfur Chem. Geol. 161, 89–101. Takahashi, T., 1960. Supergene alteration of zinc and lead deposits in limestone. Econ. Geol. 55, 1083–1115. Toby, B.H., 2001. EXPGUI, a graphical user interface for GSAS. J. Appl. Crystallogr. 34, 210–213. Ueda, A., Krouse, H.R., 1986. Direct conversion of sulphide and sulphate minerals to SO2 for isotope analysis. Geochem. J. 20, 209–212. Ukstins, I.A., Renne, P.R., Wolfenden, E., Baker, J., Ayalew, D., Menzies, M., 2002. Matching conjugate volcanic rifted margins: 40Ar/39Ar chrono-stratigraphy of pre- and syn-rift bimodal flood volcanism in Ethiopia and Yemen. Earth Planet. Sci. Lett. 198, 289–306. Watts, Griffis, and McOuat Limited, 1993. Jabali lead–zinc–silver deposit, Republic of Yemen, preliminary feasibility study. Yemen Mineral Sector Project ReportWatts, Griffis, and McOuat Limited, Toronto, Canada (15 pp.). Weiss, C., O'Neill, D., Koch, R., Gerlach, I., 2009. Petrological characterization of ‘alabaster’ from the Marib province in Yemen and its use as an ornamental stone in Sabaean culture. Arab. Archaeol. Epigr. 20, 54–63. Whitehouse, M.J., Windley, B., Ba-Bttat, M.A.O., Fanning, C.M., Rex, D.C., 1998. Crustal evolution and terrane correlation in the eastern Arabian Shield, Yemen: geochronological constraints. J. Geol. Soc. Lond. 155, 281–296. Whitehouse, M.J., Windley, B.F., Stoeser, D.B., Al-Khirbash, S., Ba-Bttat, M.A.O., Haider, A., 2001. Precambrian basement character of Yemen and correlations with Saudi Arabia and Somalia. Precambrian Res. 105, 357–369. Whitney, D.L., Evans, B.W., 2010. Abbreviations for names of rock-forming minerals. Am. Mineral. 95, 185–187. Yemen Geological Survey and Mineral Resources Board, 1994. Metallic mineral resources. In: Geological Survey and Mineral Exploration Board, Ministry of oil and Mineral Resources, Republic of Yemen (Ed.), Geology and Mineral Resources of Yemen, pp. 61–87. Yemen Geological Survey and Mineral Resources Board, 2009. Zn–Pb. In: Geological Survey and Mineral Exploration Board, Ministry of oil and Mineral Resources, Republic of Yemen (Ed.), Mineral Resources of Yemen (n. 1, 8 pp.). Youssef, E.S.A.A., 1998. Sequence stratigraphy of the Upper Jurassic evaporite–carbonate: sequence at the western area of Wadi Al-Jawj-Marib basin, Yemen. Carbonates Evaporite. 13, 168–173. Zumbo, V., Féraud, G., Bertrand, H., Chazot, G., 1995. 40Ar/39Ar chronology of Tertiary magmatic activity in Southern Yemen during the early Red Sea–Aden Rifting. J. Volcanol. Geotherm. Res. 65, 265–279.
Similar documents
In search of the lost zinc: A lesson from the Jabali (Yemen
(Mineral Resources of Yemen, 2009), and are mainly dolomite-hosted. At Jabali (Jabal Salab) the primary and secondary Zn–Pb ores are located on the southwestern flank of the NW-trending Sab'atayn Ju...
More information