Next Article in Journal
White Marble Sourcing and Regional Workshop Dynamics in Roman Thrace: An Archaeometric Study of Votive Reliefs
Previous Article in Journal
Fine Crustal Velocity Structure and Deep Mineralization in South China from Joint Inversion of Gravity and Seismic Data
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mineralogy and Fluid Inclusion Constraints on the Genesis of the Recently Discovered Ag-(Ni-Co-Sb-As-Hg ± Bi) Vein Ore Shoot Mineralization in the Aouli Pb-Zn District (Upper Moulouya, Morocco)

1
Laboratoire des Gîtes Minéraux, Hydrogéologie & Environnement, Faculté des Sciences, Oujda 60000, Morocco
2
Geology and Sustainable Mining Institute, Mohammed VI Polytechnic University, Benguerir 43150, Morocco
3
Programa de Geofluidos, Instituto de Geociencias, UNAM-Campus Juriquilla, AP 1-253, Querétaro 76230, CP, Mexico
4
Laboratory of High Energy Physics, Astrophysics and Geosciences, Faculty of Sciences Semlalia, Cadi Ayyad University, UCA, P.B. 2390, Marrakech 40000, Morocco
5
Geology Department, California State University, San Bernardino, CA 92407, USA
6
Department of Geology, Juniata College, 1700 Moore St., Huntingdon, PA 16652, USA
7
Département des Sciences de la Terre et de l’Atmosphère, Université du Québec à Montréal, CP8888, Centre-Ville, Montréal, QC H3C 3P8, Canada
8
Department of Geosciences, University of Arkansas, 340 N Campus Drive, Fayetteville, AR 72701, USA
9
Institute of Life-Earth-Environment (ILEE), University of Namur, 61 rue de Bruxelles, B-5000 Namur, Belgium
10
Helmholtz Institute Freiberg for Resource Technology, Helmholtz-Zentrum Dresden-Rossendorf, Chemnitzer Str. 40, 09599 Freiberg, Germany
11
Physical Science and Engineering Division, King Abdullah University of Science and Technology, Thuwal 23955-6900, Saudi Arabia
12
GFZ Helmholtz Centre for Geosciences, Telegrafenberg, 14473 Potsdam, Germany
13
Laboratory of Georessources, Department of Geosciences, Faculty of Sciences and Technics, Moulay Ismail University, Errachidia 52000, Morocco
14
Laboratory of Applied Geology and Geo-Environment, Faculty of Sciences, Université Ibn Zohr, Agadir 80000, Morocco
*
Authors to whom correspondence should be addressed.
Minerals 2025, 15(7), 669; https://doi.org/10.3390/min15070669 (registering DOI)
Submission received: 16 May 2025 / Revised: 14 June 2025 / Accepted: 19 June 2025 / Published: 22 June 2025
(This article belongs to the Section Mineral Deposits)

Abstract

:
Unusual Ag-(Ni-Co-Sb-As-Hg ± Bi)-bearing fault-fill vein ore shoot mineralization set in a gangue of quartz, fluorite, and barite has been identified in Morocco’s Aouli deposit. The Paleozoic host rocks consist of a succession of Cambrian to Ordovician-aged folded and low- to medium-grade metasediments and metavolcaniclastic rocks with tuff interbeds and amphibolite sills, locally intruded by late Visean calc-alkaline to alkaline granitoid intrusions. Paragenetic relationships indicate that the sequence of ore precipitation comprises a succession of Ni-Co-Fe arsenides, followed by Pb-Sb-As-Ag-Hg sulfarsenides/sulfosalts and then Zn-Pb-Fe sulfides. Results indicate that the ore shoot mineralization formed from episodic stages of fracturing and subsequent fluid migration. Precipitation of ore phases is thought to have occurred as a result of isothermal mixing and subsequent fluid–rock interactions. The timing of mineralization is thought to have occurred between Late Triassic and Late Miocene, coinciding with major crustal extension and Middle Jurassic–Upper Cretaceous alkaline magmatism. Thermal convection and seismic pumping are proposed as the main driving force for the large-scale migration of the ore-forming brines. This research bears directly upon the potential for new exploration targets in Pb-Zn ± fluorite ± barite deposits hosted in Variscan inliers throughout North Africa.

1. Introduction

Unconformity-related Pb-Zn ± fluorite ± barite vein systems are widespread throughout the European Variscan belt [1,2] and its North African counterpart [3,4]. Sulfide mineralization occurs at different stratigraphic levels of the metamorphic basement and the unconformably overlying Mesozoic-to-Cenozoic sedimentary cover. In addition to Pb-Zn ± fluorite ± barite, recent mineralogic and geochemical studies revealed that part of these veins are host to unusual native As-sulfide ± native Ag ± Ni-Co-Fe arsenide ± antimonide ± Ag-Bi sulfosalt-bearing orebodies referred to as vein ore shoots [1,2,5,6,7,8].
The Aouli fault-fill deposits hosted by Paleozoic metasediments in the Upper Moulouya metallogenic province (Figure 1) rank together with those of Tighza [9,10,11], Roc Blanc [12], and Sidi Lahcen [13] among the four major producers of Pb-Zn ± fluorite ± barite-type argentiferous veins in Morocco, with significant Ag output (locally up to 4500 g/t Ag). Between 1926 and 1983, more than 9.6 Mt of run-of-mine ore were produced, with an average extraction grade of 5% Pb and 400 g/t Ag. Apart from the well-constrained Pb-Zn ± fluorite ± barite mineralization event [14,15,16], a new Ni–Co–Fe arsenide, Sb-As-Ag-Hg ± Bi-bearing sulfide and sulfosalt paragenesis, referred to thereafter as Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore-shoot mineralization, has recently been recognized in different mineral associations from the main exploited veins.
Here, we provide detailed investigations of the mineralogical, paragenetic, and fluid characteristics of these newly discovered vein ore shoots at the Upper Moulouya metallogenic province. As such, a refined paragenetic sequence that underpins a new genetic model is proposed based on an integrated approach that combines field observations and mineralogical, textural, and paragenetic constraints, along with fluid inclusion microthermometry. The combined use of these analyses provides a suitable framework for a better understanding of the significance and origin of the Ag-(Ni-Co-Sb-As-Hg ± Bi) ore shoot mineralization.

2. Regional and District-Scale Geological Settings

The ~15–50 km wide and 100 km long Upper Moulouya massif, which hosts the Aouli deposits (this study), is part of the eastern Meseta being sandwiched between the Middle Atlas to the northwest and High Atlas to the south (Figure 1). It is an intermontane Mesozoic basin that formed as a result of subsidence during the opening of the Jurassic Maghrebian Tethys and Atlantic Ocean [18,19], and subsequent inversion and uplift during the Late Cretaceous (Senonian) compressional–transpressional event [20,21]. There, the Paleozoic basement is mainly exposed in the Aouli-Mibladen and Bou Mia inliers surrounded by Mesozoic and Cenozoic strata (Figure 1).
Sedimentation rates increased through the Jurassic as rifting continued with the breakup of Pangea, and the opening of the neo-Tethys Ocean and the North Atlantic [22]. Two rifting episodes operating from the end of Early Jurassic to Late Jurassic were documented [23,24]. While the early Jurassic rifting 1 is related to the hyper-extension of the Atlantic Ocean, the late Jurassic rifting 2 is linked to the hyper-extension of the Maghrebian Tethys. During the Mesozoic, the Upper Moulouya massif formed an emergent structural high that separates two subsiding basins, i.e., the Middle Atlas to the northwest and the High to the south [25], whereas during the Cenozoic, the massif became an intramontane basin of uplift-related sediment input sourced from the surrounding Middle and High Atlas belts [26,27]
From a metallogenic perspective, the upper Moulouya massif is unique in containing a wide range of mineral deposits including the following (Figure 2): (1) vein-type (i.e., Aouli deposit described herein; 9.6 Mt at 5% Pb), (2) stratiform diagenetic-type (i.e., Zeida deposit; 16 Mt at 3% Pb), and (3) stratabound Mississippi Valley-type (MVT) (i.e., Mibladen deposit; 6.5 Mt at 5% Pb) [15].
In the studied district, the oldest rocks are of presumed Cambrian–Ordovician age [29,30] consisting of a succession of folded and low- to medium-grade metasediments and volcaniclastic rocks with tuff interbeds, and amphibolite sills (Figure 1). Locally near Engill vein (Figure 3), the exposed sedimentary succession shows the occurrence of a dark-colored muddy rock sequence, up to 120 m thick, consisting of organic-rich siliceous shale interbedded with thin fine-grained carbonate beds. The regional stratigraphic relationships for these black shales, which could be source rocks for petroleum and natural gas, remain poorly constrained, being assigned to the Carboniferous [31]. Several late Visean calc-alkaline to alkaline metaluminous to peraluminous granitoid intrusions (ca. 330–340 Ma [32]) intruded the Cambrian–Ordovician country (Figure 1).
Structurally, the metasedimentary host rocks were affected by two main phases of Variscan deformation, the earliest of which is thought to have occurred during the Tournaisian–Visean (i.e., Eovariscan D1 phase at 360–330 Ma [30,33,34]), whereas the later Late Westphalian phase occurred under an extensional regime coinciding with the end of the D1 phase and subsequent emplacement of the late Visean granitoid intrusions. A contact metamorphic aureole of ~1500-m width formed around the Aouli granitic batholith. The peak P-T conditions (400–640 °C, <3 kb), inferred from the main metamorphic assemblages along with Ti-in-biotite and amphiboles geothermometry, correspond to batholith emplacement depths ranging from 4 to 7 km [35,36,37].
Unconformably overlying the Neoproterozoic–Paleozoic basement, the Mesozoic to Cenozoic stratigraphy consists of syn-rift, ~400–500 m thick Triassic red bed conglomerates, salt-bearing argillites and arkoses with CAMP basalt interbeds, and Jurassic (Liassic, Toarcian, Aalenian–Bajocian) limestones, dolostones, and marls. The post-rift Cretaceous–early Tertiary formations comprise Cretaceous (Lower Cenomanian, Cenomanian–Turonian) conglomerates, shallow and marginal marine limestones, calcareous shales, dolostones, and interbedded anhydrite-rich shales, Cenozoic (Oligocene) continental fluvial to lacustrine deposits, and Plio–Quaternary alkali basalts (Figure 1). The major brittle structures comprise an array of NNE-SSW to NE-SW, and WNW-ESE-trending fault systems (Figure 1) which are interpreted as pre-Variscan in age but have been repeatedly reactivated during the later Alpine orogeny and currently represent significant features of extension (thinning) or trans-tension [38].

3. Sampling and Analytical Methods

Following extensive field mapping, representative fresh and hydrothermally altered host rocks along with barren and mineralized veins across each of the five ore zones (i.e., Aouli, Sidi Ayad, Poulet, Sidi Said, and Ansegmir; Figure 3) were collected for petrographic and mineralogical characterization and geochemical analysis based on texture, mineral assemblages, and paragenetic position. The sampling sites, preferably those of high-grade ore, include surface exposures, old mining pits, and underground mining galleries (Figure 3). Specimens of massive ores (n = 50) and associated gangue phases (n = 50) with visible arsenide (n = 20) were collected mainly from Sidi Said underground mining workings due to the inaccessibility of galleries from the other ore zones.
In the Sidi Said area, sampling targeted massive to stringer to disseminated arsenide (n = 20), sulfarsenide/sulfosalt (n = 20), and sulfide (n = 50) mineralization from both the veins and the adjacent wall rocks. The sample locations are shown in Figure 3. More than 120 polished thin sections were investigated under optical microscope and scanning electron microscopy (SEM) and scanning electron microscope-backscattered electron (SEM-BSE) imaging for mineral assemblage and textural relationship characterization prior to mineral, fluid inclusions.

3.1. Optical and Cathodoluminescence Microscopy

Mineral identification along with textural analysis and cold-cathodoluminescence (CL) imaging were carried out at the Isotope Studies Laboratory, Institute of Geosciences, UNAM, Juriquilla, Mexico, using an Olympus SZ-X microscope equipped with a Qimage Micropublisher 5 Mp digital camera, a Reliotron® CL instrument, and a HITACHI TM3000 scanning electron microscope (SEM). The CL instrument was typically operated with a beam current of 500 μA and a corresponding voltage of ~15 kV. The CL imaging of the fluorite, quartz, apatite, and sphalerite was performed in complement to the classical optical petrography to better constrain textural relationships and relative age-constraints of the studied fluid inclusion assemblages (FIAs).

3.2. Electron Probe Microanalysis (EPMA)

Minerals in grain mounts and thin sections were imaged by backscattered electrons (BSE), and major and minor element data were collected at the microanalysis laboratory of Université Laval (Canada) using a CAMECA SX-100 electron microprobe. For arsenide, sulfosalt, and sulfide analyses, the instrument was operated with an accelerating voltage of 15 kV and a beam current of 20 nA with a spot diameter of 2 μm and counting times of 15–20 s for peak measurements. The element lines used for analysis included Kα lines for S, Cu, Ni, Co, Fe, Zn; Lα lines for Ba, Ag, As, Sb; Mα lines for Pb; and Mβ lines for Hg. The following in-house and published standards were used for calibration: sphalerite (for Zn), chalcopyrite (for Cu), pure Ni (for Ni), pure Co (for Co), hematite (for Fe), GaAs (for As), pure Ag (for Ag), Sb (for Sb), HgS (for Hg), galena (for Pb), and marcasite (for S).

3.3. Laser Ablation (LA)-ICP-MS Analysis

Trace element concentrations of apatite and rutile were measured at the Isotope Studies Laboratory, Institute of Geosciences, UNAM, Juriquilla, Mexico, using a Coherent Compex 102 excimer laser ablation system with an ATL excimer ArF laser source at 193 nm wavelength and ~5 ns pulse width coupled to a Thermo iCapQc inductively coupled plasma mass spectrometer instrument. Following a 15 to 20 s period of background analysis, apatite crystals were ablated with a 60 μm laser beam for 35 s at a 4 Hz repetition rate using a laser energy density of 4 J/cm2 [39], and 15 s of washout. The ICP-MS was tuned following the experimental parameters reported in [40,41]. Spot ablation was carried out using a 24 µm spot size in the epoxy-grain-mount apatite samples and a 10 µm spot size in the thin-section apatite samples to avoid abundant inclusions. Data reduction and concentration calculations were performed using the Iolite 4 software [42,43]. The errors of trace element concentrations are typically better than 5% for most elements based on repeated analyses of secondary standards.
For rutile, a modified protocol from [44] was used. The analysis was performed using a laser beam energy maintained at 6 J/cm2 with a laser repetition rate at a constant frequency of 5 Hz and a spot size of 60 µm. Analyses were conducted over 15 s of background collection, followed by 30 s of ablation and sample acquisition, with a final 15 s for cleaning. Ti measured as 49Ti was used as a standard, normalized to a concentration of 59.94% in the NIST SRM 612 standard glass [45] and rutile R10 [46]. Rutile R19 (ID-TIMS: 489.5 ± 0.9 Ma; [46,47]) was used as an unknown to evaluate the data quality. Data were processed using the Iolite 4 software [42,43].
The LA-ICP-MS trace element compositions of sphalerite, pyrite, and chalcopyrite were determined at the EleMap laboratory, GFZ Helmholtz Centre for Geosciences, Potsdam (Germany) using an Analyte Excite 193 nm ArF excimer-based laser ablation (LA) system (Teledyne Photon Machines, Bozeman, MT, USA), coupled to a quadrupole-ICP-MS iCAP RQ from Thermo Fischer Scientific. The analytical and data processing methods and the full trace element dataset are presented in [48].

3.4. Fluid Inclusion Microthermometry

A total of 10 samples of ore-related fluorite and quartz were prepared as doubly polished wafers of approximately 120–200 μm thickness for fluid inclusion studies using optical and cathodoluminescence microscopy. Cathodoluminescence (CL) imaging was applied to reinforce fluid inclusion petrography by depicting different fluorite and quartz generations and growth zoning using a Reliotron® cold cathode CL system under conditions of 15 kV accelerating voltage and a current density of ca. 10 µA [49]. Cathodoluminescence photomicrographs were taken with a very high sensitivity Qimage Micropublisher 5 Mp digital camera with a Peltier-cooled CCD.
Microthermometry measurements were carried out at the Crustal Fluids Laboratory of the Instituto de Geociencias at UNAM in Querétaro, Mexico, using an Olympus BX-51® microscope equipped with a Linkam THMSG600 heating-cooling stage coupled to a QImaging Retiga 2000R CCD HD camera. For each fluid inclusion, freezing temperature (Tf), eutectic temperature (TE), final hydrohalite melting temperature (Tm-HH), freezing final ice melting temperature (Tmice), and total homogenization temperature (Th) were repeated during several runs for accuracy. Cryogenic experiments were carried out before heating to reduce the risk of decrepitating the inclusions. The volumetric fraction of phases was estimated at room temperature using the volumetric table of [50]. Total salinities expressed as equiv wt% NaCl-CaCl2 were calculated using Tm-HH and Tm-ice temperatures according to the spreadsheet provided by [50] in the ternary NaCl-CaCl2-H2O system. In some cases, Tm-HH could not be observed. In this case, the corresponding salinities were calculated based on the HOKIEFLINCS_H2O-NaCl spreadsheet of [50] in the H2O-NaCl system and reported as total wt% NaCl eq. No hydrocarbon-bearing inclusions were detected with Raman spectroscopy.

4. Results

4.1. Vein Ore Shoot Distribution, Geometry, Ore Textures, Alteration, and Paragenesis

The Aouli Pb-Zn ± fluorite ± barite mining district extends across an exposed area of ~12 × 25 km at the approximate coordinates lat. 31°47′45″ N, long. 08°00′48″ W (Figure 1). The exploited fault-fill veins, which are mostly limited to and/or close to the rocks of the granite-induced contact metamorphic aureole, are grouped into five ore zones referred to as Aouli, Sidi Ayad, Poulet, Sidi Said, and Ansegmir (Figure 2). In the five ore zones, the ENE-WSW, WNW-ESE, and E-W trans-tensional steeply dipping veins extend laterally up to 10 km along strike and over 200 m down-dip, are spaced 50 to >100 m apart, dip between 35° and 65° to the east, and vary in thickness from several cm up to 10 m (averaging 4 m). The spatial distribution of the ore zones does not correlate with stratigraphy as the fault-fill veins are either entirely hosted in the granitic intrusion (i.e., Ansegmir group), in the Cambrian–Ordovician schistose pelite (i.e., Poulet, Aouli, and Sidi Said groups) or along the schistose pelite and the granite boundary, locally extending into the unconformably overlying Triassic rocks (i.e., Sidi Ayad group) (Figure 4A). Structural analysis indicates that all these fault-fill veins developed along mostly normal faults under an extension tectonic regime consistent with a NW-SE to NNW-SSE sub-horizontal σ3 axis and a subvertical σ1 axis [38].
Individual veins consist of a succession of strongly mineralized shear segments “ore shoots” separated by lower-grade or barren segments. The resulting Ag-(Ni-Co-Sb-As-Hg ± Bi)-bearing ore shoot mineralization occurs as irregularly distributed and lenticular orebodies typically on the same quartz-fluorite fault-fill veins, but is preferentially developed in dilatational jogs, changes in strike, splays, and tensile fracture arrays. The contact between veins and wall rocks is sharp, even though alteration halos with disseminated pyrite and sulfide-quartz veinlets may extend outward from the veins to the schistose pelitic host rocks. Wall-rock alteration consists predominantly of sulfidation and silicification with subordinate sericitization, carbonitization, argillization, and hematitization.
The majority of the Aouli ore shoot mineralization occurs as quartz-fluorite polymict clast-supported breccia veins (Figure 4B) and vug infill with ribbon, colloform, and cockade (Figure 4C) textures. The breccia bodies consist of angular to moderately rounded cm-sized wall-rock clasts set in a sand-sized hydrothermal blackish matrix of hydrothermally precipitated phases. Throughout the polymict breccia bodies, the Ag-(Ni-Co-Sb-As-Hg ± Bi)-rich ore is intergrown with quartz, sericite (Figure 4D), apatite (Figure 4E), rutile (Figure 4F), and adularia (Figure 4G), with the four latter hydrothermal phases being described hereinafter for the first time. The veins display comb, cockade, breccia, and crack-and-seal textures suggesting repeated stages of dilatation, fluid circulation and subsequent ore infilling. Reopening of the same fault-fill vein structures and their subsequent ore infilling during the repeated mineralization stages explain the texturally complex and compositionally variable vein generations. Although no consistent paragenetic sequence can be defined owing to repeated and episodic pulses of mineralization, the sequence of mineral deposition shows three successive sub-stages of mineralization (Figure 5). The early Ni-Co-Fe arsenide sub-stage comprises, from early to late, rammelsbergite (NiAs2), skutterudite (CoAs3), safflorite (CoAs2), and loellingite (FeAs2). The following sub-stage consists of a succession of Ag-As-Sb sulfarsenides and sulfosalts including fahlore group minerals (tennantite-tetrahedrite series), acanthite, proustite, polybasite, imiterite (Ag2HgS2), Ag-Hg amalgams, and native Ag. The late Zn-Pb sulfide sub-stage includes a succession of sequentially crystallized sphalerite, galena, pyrite, and chalcopyrite. The synchronous gangue phases for all three sub-stages comprise, in order of decreasing abundance, quartz, fluorite, and barite along with subordinate apatite, rutile, adularia, and sericite, and to a lesser extent carbonates (mostly dolomite and ankerite).
Strikingly comparable textural features and paragenetic sequence of ore minerals have been reported from several native metal-arsenide ore districts including the Ag-Bi-Co-Ni-As ± U veins of the Erzgebirge (Germany) [1,7,51,52], the Ni-Co-As mineralization in the Midwest polymetallic U deposit of the Athabasca Basin (Canada) [53], and those world-class deposits from the Bou Azzer Co-Ni-Fe-As(±Au ± Ag) and Imiter Ag-Hg-rich districts in the Moroccan Anti-Atlas [54,55,56]. Contrary to many of the so-called five-element vein systems, especially those in the Erzgebirge, which are commonly associated with substantial amounts of U mineralization, no evidence has been found so far either for the occurrence of U-bearing veins or U-disseminated mineralization, or for carbonate gangue phases in the Aouli granitic intrusions, or the schistose pelites.
In addition to the hydrothermal ores, a supergene stage is well developed in the uppermost sulfide-rich part of the exploited orebodies. The resulting secondary mineral assemblage consists of a variety of exotic phases such Pb- and Cu-carbonates (cerussite and malachite-azurite, respectively) and hydrated arsenate (erythrite: Co3(AsO4)2•8(H2O)).

4.2. Vein Ore Shoot Mineralogy and Mineral Chemistry

The Aouli deposit have a complex ore mineralogy which consists of (1) common metal sulfides (sphalerite, galena, pyrite ± chalcopyrite), (2) uncommon metal phases including tri- and diarseniades (skutterudite/Nickelskutterudite, safflorite/clinosafflorite, rammelsbergite, and loellingite), (3) Pb-Sb-As-Ag-sulfosalts (Ag-bearing tetrahedrite, polybasite, acanthite, proustite, and stannite), and (4) Ag-Hg amalgam and native silver (Figure 6). Volumetrically, common Zn-Pb sulfides are by far the dominant phases, whereas arsenides, Pb-Sb-As-Ag-sulfosalts, and Ag-Hg amalgam and native silver are less abundant (<5 vol.%), being spatially restricted to the upper parts of the exploited veins close to the unconformity between the Triassic basal red bed series and the underlying Paleozoic basement lithologies.

4.2.1. Arsenides

Among the Ni-Co arsenides, skutterudite/Nickelskutterudite (Figure 6A) is by far the dominant triarsenide phase. It typically occurs either as clusters of euhedral to subhedral crystals or as space fillings between quartz and fluorite ± ankerite-dolomite. Skutterudite is commonly intergrown with other members of the Ni-Co-Fe-bearing arsenide series including rammelsbergite and safflorite (Figure 6B). Locally, skutterudite is replaced by native silver, acanthite, proustite, stannite, and chalcopyrite along with supergene secondary phases such as arsenolite and erythrite (Figure 6B). The EDS analyses show large variations in As, Ni, and Co contents ranging from 70.02% to 74.40%, 6.01% to 19.43%, and 6.50% to 14.46%, respectively. Nonetheless, Fe and S are present at lower concentrations reaching up to 2.85% Fe and 3.06% S.
Diarsenide minerals commonly occur as thin continuous layers of safflorite/clinosafflorite, rammelsbergite, and loellingite displaying a complex corona-like overgrowth or star-shaped texture (Figure 6C). Texturally, diarsenide minerals occur either as (i) euhedral crusts or star-shaped aggregates, ranging from 10 μm to 500 μm in thickness, growing directly on skutterudite crystals, (ii) massive aggregates (up to 2 mm) that partially replace euhedral to subhedral skutterudite, and/or (iii) crystalline aggregates displaying an arborescent growth habit or a typical fortress texture materialized by concentric banding/zoning patterns with Ni/Fe ratios decreasing from the inner zones outward. Compositionally, the diarsenide phases exhibit overlapping chemical compositions with significant substitutions among Ni, Fe, and Co (Table S1; Figure 7A,B). In this respect, EPMA analyses show that rammelsbergite has As, Ni, Co, and Fe contents ranging from 65.06 to 69.19 wt%, 10.49 to 18.59 wt%, 4.29 to 10.69 wt%, and from 6.21 to 9.98 wt%, respectively. Safflorite has As ranging from 62.83 to 70.13 wt%, Co between 9.27 and 17.17 wt%, Ni from 0 to 9.65 wt%, and Fe from 5.01 to 13.85 wt%. Similarly, loellingite has As concentrations varying from 67.75 to 70.95 wt%, Ni from 0.19 to 8.78 wt%, Fe from 11.62 to 16.20 wt%, and Co between 5.76 and 13.30 wt%. Cu contents are up to 1.32 wt%. It is noteworthy that all analyzed diarsenides contain appreciable sulfur concentrations that range widely from 0.47 wt% up to 5.77 wt.% S.
In addition to the above-mentioned diarsenide phases, an unidentified (Ni-Co) AsS amalgam is also recognized. It occurs either as thin layers, up to 50 μm, overgrowing pyrite grains or as minute grains associated with, and/or replacing, rammelsbergite, safflorite/clinosafflorite, loellingite, and skutterudite/Nickelskutterudite. The EDS analyses indicate As, Ni, Co, Fe, and S concentrations of 46.74%, 8.86%, Co 24.99%, Fe 2.26%, and S 17.15%, respectively.

4.2.2. Pb-Sb-As-Ag-Sulfosalts

The Pb-Sb–As–Ag-sulfosalts are irregularly distributed in major veins and include Ag-bearing tetrahedrite, polybasite, proustite, imiterite, Ag-Hg amalgam, acanthite, and native silver.
Polybasite is by far the most abundant silver-bearing phase. It occurs as variably sized (up to 200 μm) anhedral crystals coexisting with Ag-bearing tetrahedrite, proustite, acanthite, sphalerite, and galena (Figure 6D). Polybasite also forms massive aggregates, corroding and cementing quartz grains (Figure 6E) or infilling small fractures. The EPMA analyses reveal that polybasite has Ag contents ranging from 58.24 to 73.96 wt%, Cu from 0.29 to 8.31 wt%, Sb from 1.98 to 12.54 wt%, and As concentrations in the range of 0.53 to 5.44 wt% (Table S1). The striking compositional feature of polybasite resides in its enrichment in Hg with contents ranging from 1.31 to 5.18 wt%. Ag-rich tetrahedrite occurs either as isolated anhedral grains, up to 200 µm in size, or as fracture-infill veinlets with coexisting chalcopyrite, polybasite, and acanthite; all set in a gangue of quartz and fluorite (Figure 6E). Geochemically, argentiferous tetrahedrite shows varying Ag (1.76–10.13 wt%), Sb (16.71–21.55 wt%), Cu (30.35–36.58 wt%), and As (4.82–8.53 wt%) contents; those of Fe and Zn range from 0.77 to 3.57 wt% and 4,00 to 6.59 wt%, respectively. Minor components include Hg and Pb whose concentrations range from 2.18 to 2.75 wt% Hg and 0.01 to 0.27 wt% Pb (Table S1).
Acanthite is closely intergrown with polybasite, imiterite, and Ag-Hg amalgam (Figure 6E). The EPMA analyses show highly variable Ag and S contents, ranging from 69.60 to 83.74 wt% and 8.62 to 15.26 wt%, respectively. The Hg concentrations range from 1.11 to 3.36 wt% (Table S1). Proustite forms anhedral to euhedral crystals ranging from 100 to 600 µm. It fills intergranular spaces between quartz and fluorite. Locally, proustite is associated with galena and sphalerite typically as coarse masses. The EPMA data reveal Ag contents between 63.23 to 65.73 wt%, As from 11.36 to 12.22 wt%, and Sb from 3.04 to 3.29 wt% (Table S1).
Ag-Hg amalgam is among the most abundant Ag-bearing phases (Figure 6E). It occurs as thin rims mantling and coexisting with Ag-bearing tetrahedrite, polybasite, acanthite, and native silver. The EPMA analyses show highly variable compositions with Ag and Hg contents ranging from 61.71 to 80.64 wt% and up to 33.23 wt%, respectively. Imiterite is a rare mineral, commonly associated with acanthite and Ag-Hg amalgam. It frequently occurs either as minute inclusions within polybasite or as intergrowths with other sulfosalts, replacing polybasite and acanthite. A single EPMA analysis reveals Ag, S, and Hg contents of 56.95 wt%, 8.59 wt%, and 32.37 wt%, respectively. Native silver occurs either as flake-shaped crystals up to several cm infilling the interstices between fluorite and quartz (Figure 6F), or as filamentous silver (i.e., wire silver) lining vugs (Figure 6G). The EPMA analyses indicate the composition of wire silver close to pure silver with Ag concentrations ranging from 96.3 to 97.6 wt% and sulfur content up to 1.1 wt%. Nonetheless, the Hg contents are low ranging from 0.22 to 0.34 wt% (Table S1).
Remarkably, the Aouli arsenide, Pb-Sb-As-Ag-sulfosalt, Ag-Hg amalgam, and native silver phases are characterized by the high percentages of Hg concentrations.

4.2.3. Sulfides

Sulfide phases consist predominantly of sphalerite and galena with subordinate pyrite, and to a lesser extent chalcopyrite. The full LA-ICP-MS trace element dataset on sulfides is available in [48].
Sphalerite typically occurs as µm to cm sized greenish- to brownish-colored scattered subhedral grains set in a gangue of quartz and fluorite ± dolomite-ankerite (Figure 6H). Texturally and compositionally, the sphalerite is intimately intergrown with Ag-bearing sulfides and sulfosalts. The EPMA analyses (Table S1) along with LA-ICP-MS (Table 1) data reveal that sphalerite contains high concentrations of Ag, Hg, Sb, As, Cd, Cu, and Pb reaching up 4.3 wt% Ag, 0.36 wt% Hg, 0.8 wt% Sb, 0.17 wt% As, 0.8 wt% Cd, 0.79 wt% Cu, and 1.2 wt% Pb (Figure 8). Nonetheless, the concentrations of other elements, particularly Ga, Ge, In, and Fe, are commonly moderate to low with average value 43.22 ppm, 90.24 ppm, 16.05 ppm, and 3486 ppm, respectively.
Galena is by far the most abundant base metal sulfide in the studied deposits. Three textural varieties of galena are recognized, including (i) disseminations of varying grain size, (ii) crystalline, massive anhedral galena infills, and (iii) cm sized cubic crystals. Similar to sphalerite, galena from sub-stage II.3 is closely intergrown with sphalerite, pyrite, and chalcopyrite within a gangue of quartz and fluorite. Galena commonly hosts numerous minute inclusions of Ag-Hg amalgam and polybasite. The EPMA analyses reveal significant enrichment in Ag and Hg with contents of 4.11 wt% Ag and up to 1.53 wt% Hg (Table S1).
Pyrite is a minor sulfide. It occurs as an early fine-grained phase, forming subhedral to euhedral crystals up to 200 μm in size or larger subhedral aggregates up to 600 μm. Pyrite exhibits a distinct cataclastic texture, with the resulted brecciated crystals being cemented by tetrahedrite and Ag-bearing minerals, and locally cross-cut by thin veinlets of galena, sphalerite, and chalcopyrite. The LA-ICP-MS analyses show that pyrite has significant substitutions of Co (110.7–204.0 ppm), Ni (88.3–200.8 ppm), Cu (390.9–954.2 ppm), Sb (65.5–184.2 ppm), and Hg (75.8–171.4 ppm) whose concentrations contrast with those of Se, Te, Sn, Cd, and Bi which are low, commonly below detection limit. The Ni/Co ratio ranges from 0.74 to 1.52. Remarkably, As and Ag contents are strikingly high, reaching 0.63 wt% and 1.16 wt%, respectively (Table 1).
Chalcopyrite occurs either as massive mono- or polymineral aggregates commonly intergrown with quartz or as anhedral masses filling microfractures. The LA-ICP-MS analyses indicate a homogenous composition close to the ideal formula CuFeS2 (Table 1).

4.2.4. Ore-Related Gangue Minerals

In addition to quartz and fluorite, which have been the focus of numerous studies [14,16], hydrothermal apatite, rutile, adularia, and sericite are described therein for the first time.

Apatite

Apatite occurs as homogeneously distributed fine-grained rounded, rarely elongate, to subhedral (10–500 μm across) crystals or aggregated grains closely intergrown with Ag-(Ni-Co-Sb-As-Hg ± Bi)-rich arsenides, sulfarsenides/sulfosalts, and sulfides along with sericite, adularia, and rutile. BSE images show that apatite grains have a homogenous appearance, contain locally quartz and galena inclusions, and fluoresce yellow light under CL (Figure 4E).
Geochemically, in situ LA-ICP-MS compositions of single apatite crystals show significant variations in trace elements, such as Mn (262–618 ppm), Sr (367–899 ppm), Pb (1.84–873 ppm), U (0.62–3.55 ppm), and Th (1.04–4.26 ppm). The Cl content is 215.98 ± 94.34 ppm (Table 2). With respect to REE + Y contents, apatite grains are strongly enriched in REEs (ΣREE = 1718.5–3440.2 ppm), with average total light (La–Nd), middle (Sm–Ho), and heavy (Er–Lu) REE contents of 1455.1 ± 279.2 ppm, 838 ± 81.3 ppm, and 212.4 ± 28.7 ppm, respectively (Table 2). The chondrite-normalized REE + Y spectra show “hump”-shaped patterns (Figure 9A) with enrichment in MREE relative to LREE and heavy rare earth elements (HREE) in addition to exhibiting a prominent negative Eu anomaly with Eu/Eu* ratios of 0.6 to 0.7, and almost no Y and Ce anomalies (Ce/Ce* = 0.92–1.05). Key REE fractionation ratios are (La/Sm)N = 0.3–0.5, (La/Yb)N = 0.6–2.2, (Sm/Yb)N = 2–3.7, and (Gd/Lu)N = 2.6–4, where the subscript N indicates that the REE contents are normalized to chondrite [57]. The Y/Ho ratios are generally between 26 and 28. No consistent correlation between trace element compositions and CL colors has been noticed.
In summary, the Aouli apatite is characterized by high REE + Y (mean = 3735 ppm) and Pb contents (average = 54 ppm), but lower concentrations of U (average = 3.55 ppm). Regarding abundances of volatiles (i.e., Cl-F-OH), only Cl has been detected with concentrations showing large variations ranging from 39.96 ppm to 656 ppm (Table 2).

Rutile

As for apatite, rutile occurs as a subordinate but common phase closely intergrown with (Ni-Co)-As-S diarsenide, and tetrahedrite (Figure 4F). Under transmitted light microscopy, rutile is dark reddish-brown to opaque. It occurs as dispersed to clustered subhedral to euhedral µm-sized (10–400 μm) untwinned grains coexisting with quartz and fluorite. In BSE, the rutile grains are uniformly gray andhave no obvious chemical zoning (Figure 4F). Inclusions of galena were locally observed.
Geochemically, rutile grains are enriched in Fe (2029–75,592 ppm), V (181.32–1360.42 ppm), Zr (13–4600 ppm), Nb (207–514 ppm), W (821–10,247 ppm), and Sn (104–654 ppm), but depleted in Cr (<2–6.96 ppm), Ta (4–97 ppm), and Hf (3.79–93 ppm) (Table 3; Figure 9B). There are large variations in the concentration of U (0.3–199.1 ppm) and consistently very low Th contents (0.0–3.27 ppm). As a result, the Th/U ratio is low (mean value = 0.01). Conversely, the Pb contents are remarkably high ranging from 101 to 42,974 ppm. Key ratios such as Nb/Ta, Zr/Hf, Y/Nb, Fe/V, and Y/Ho range from 4.4 to 90.4, 0.9 to 49.1, 0.0 to 1.1, 4.1 to 113.2, and 14 to 30.4, respectively.
Regarding REE + Y elements, the rutile samples are characterized by large variations in total REE + Y concentrations ranging from 16 to 1618 ppm, display consistently chondrite-normalized flat LREE-MREE patterns (Figure 9C), and lack distinct Ce, Eu, or Y anomalies with Ce/Ce*, Eu/Eu*, and Y/Y* values of 1.06–1.16, 0.84–2.01, and 0.50–1.32, respectively (Table 3). Comparatively, these patterns are typical of rutile of hydrothermal origin [58,59,60,61].

Adularia

Texturally, adularia occurs within breccia bodies as sparsely and homogeneously distributed very fine-grained (<20 µm across) euhedral rhombic and sub-rhombic crystals or grouped in clusters of aggregated tabular-shaped crystals intimately intergrown with sulfide mineralization and recrystallized fine-grained quartz (Figure 4G). Under cross-polarized light microscopy, adularia is pale to dark gray in color, commonly exhibiting complex twinning and optic variations. Locally, adularia consistently displays a mottled texture due to the presence of numerous minute (<15 µm across) inclusions of sericite. In BSE images, the adularia crystals are uniformly gray (Figure 4G) andhave no obvious chemical zoning or heterogeneous textures. SEM and BSE images show uniformly gray platy to polyhedral adularia morphologies with longest dimension lower than 20 µm across and have no obvious chemical zoning or heterogeneous textures. Chemically, the SEM-EDS analyses indicate the composition of adularia close to pure KAlSi3O8.

Sericite

Hydrothermal sericite occurs in relatively minor amounts as an aggregate of fine- to medium-grained flakes (up to 100 μm in length) closely intergrown with sulfides (sphalerite, pyrite, and galena) along with adularia, apatite, and quartz (Figure 4D).

4.3. Fluid Inclusion Analysis

4.3.1. Petrography

Fluid inclusions in ore-related quartz and fluorite were investigated using fluid inclusion petrography and microthermometry analysis. Moreover, the fine-grained nature of most quartz samples and/or the general lack of larger fluid inclusions limited our analyses to a restricted number of quartz-hosted fluid inclusions compared to those hosted in fluorite which are more abundant and larger. Nonetheless, both ore-related phases contain visible growth zones along which several fluid inclusion assemblages (FIAs; [62,63]) occur (Figure 10). The investigated fluid inclusion assemblages were classified as isolated, cluster, primary, pseudosecondary, and secondary according to [64,65].
Three types of fluid inclusions were identified in the ore-related fluorite and quartz based on their phase assemblages at room temperature and homogenization behaviors during heating, including the following: (i) liquid monophase (L), (ii) liquid-dominated bi-phase (vapor + liquid, L + V), and (iii) sporadically liquid-dominated tri-phase (L + V + S) inclusions in which the liquid and vapor homogenize to the liquid phase, but the solid phase(s) did not dissolve during heating (Figure 10D). Accordingly, these solid crystals were interpreted as entrapped phases and not daughter minerals, and as such were excluded from microthermometric measurements.
Overall, the two-phase L-V inclusions are by far the dominant type, with the degree of filling up to 10 vol.% of vapor phases. They are abundant in fluorite and to a lesser extent in quartz, occurring as irregular, subrounded to elongated-shaped inclusions mostly 5 to 40 μm in size (Figure 10). Larger inclusions up to 120 μm were observed in host fluorite. Examination under ultraviolet micro-spectroscopy did not identify hydrocarbon-bearing fluid inclusions as single-phase inclusions or as heterogeneously trapped aqueous-petroleum inclusions. Moreover, neither liquid CO2 nor clathrates melting was observed, indicating that the CO2 content in fluid inclusions is less than 1.5 mol% [66]. Remarkably, no evidence of boiling was observed as evidenced by the lack of coexisting liquid- and vapor-rich inclusions [64].

4.3.2. Microthermometry

Microthermometric measurements were performed on the two-phase L-V fluid inclusions considered to be of primary origin according to the criteria of [62]. A total of 198 two-phase L-V inclusions from 36 FIAs and clusters with consistent vapor-to-total ratios (<10%) were investigated for microthermometry. The results and related parameters are summarized in Table 4 https://www.sciencedirect.com/science/article/pii/S0169136824005717 (accessed on 14 June 2025).
All the two-phase L + V inclusions homogenized into the liquid phase. Collectively, microthermometric data from fluid inclusions in ore-related fluorite and quartz are presented together, as both host minerals yield inclusions with roughly similar microthermometric properties. Excluding outliers, the majority of the FIAs and clusters exhibit consistent homogenization temperatures (Th) that differ by less than 30 °C. Primary and pseudosecondary inclusions homogenized at temperatures ranging from 80 to 168 °C with 85% of the measurements between 110 and 150 °C. Fluid inclusions in quartz have on average slightly higher temperatures (140–170 °C) than those measured in fluorite (120–140 °C) (Figure 11A).
The first apparent melting temperature of ice (TE) occurred mostly close to or lower than −50 °C, consistent with the eutectic of the H2O-NaCl-CaCl2 system [50]. The final melting temperature of ice (Tm(ice)) was in the range of −20.6 to −10.5 °C (mean = −14.8 °C), and that of hydrohalite between −26.2 to −21.7 °C (mean = −23.7 °C). During heating, hydrohalite exclusively melts before ice in all fluid inclusions. Based on the melting temperatures of ice and hydrohalite, the compositions of the fluid inclusions, including salinity (wt% NaCl + CaCl2) and (NaCl/(NaCl + CaCl2) mass ratio, were calculated according to [50].
Remarkably, the compositions of the fluorite- and quartz-hosted inclusions plot in the field of ice along two distinct horizontal arrays above the ice-hydrohalite and hydrohalite-halite cotectic boundaries, on both CaCl2-dominated and NaCl-dominated sides (Figure 11B). The total salinities are mostly between 14 and 23 wt% NaCl + CaCl2, though lower salinities as low as 12 wt% NaCl + CaCl2 are observed in fluorite-hosted fluid inclusions from the late base metal sulfide sub-stage NaCl + CaCl2 (Table 4).

5. Discussion

It is worth mentioning that the focus of this contribution is on the Ag-(Ni-Co-Sb-As-Hg ± Bi)-bearing vein ore shoot mineralization.

5.1. Physicochemical Evolution, P-T Conditions, and Sources of the Ore Fluids

Microthermometric measurements in ore-related fluorite- and quartz-hosted fluid inclusions indicate that ore-forming fluids correspond to evolved H2O-NaCl-CaCl2 basin-derived brines with temperatures of homogenization (Th) ranging from 80 to 168 °C and salinities between 12 and 23 wt% NaCl + CaCl2. These temperature and salinity ranges are similar to those documented for native metal-arsenide vein occurrences in Western Europe (e.g., Schwarzwald and Odenwald in Germany [1,67,68]), Morocco (e.g., Bou Azzer [54,55,69,70,71,72]), and elsewhere (e.g., Cobalt-Gowganda and Great Bear Lake/Canada [73,74,75]).
The isochore calculation and P-T estimation suggest a probable range of fluid temperature and pressure from 129 to 205 °C and 433 to 690 bars for a geothermal gradient of 30 °C/km, or from 117 to 193 °C and 308 to 480 bars for a geothermal gradient of 40 °C/km (Figure 12), assuming a surface temperature of 20 °C and a hydrostatic fluid pressure regime. A hydrostatic fluid pressure regime is considered most likely to be constrained by regional sedimentary overburden and the occurrence of breccia and ribbon and cockade textures (Figure 4C), which together indicate shallow crustal depths (i.e., 1–5 km; [16,76,77]). Assuming a depth of 3 km and a hydrostatic fluid pressure regime, the maximum fluid pressure is approximately 300 bars; therefore, a geothermal gradient around 40 °C/km is plausible, consistent with the modern thermal gradient of 30–45 °C/km that characterizes the northern part of Morocco [78]. Given these shallow depths of ore formation, the large fluctuations in calculated fluid pressures suggest that Aouli vein ore shoot mineralization formed from episodic stages of fracturing and subsequent fluid migration. Seismic pumping [79] is proposed as the likely mechanism that drove the released ore brines into the newly created and reopened veins, contributing to the vein ore shoot mineralization in the Aouli deposit. The regional high permeability faults such as those of Aouli, Georges, Ansegmir, and Amourou (Figure 3) would have served as the major pathways that focused fluid flow allowing basement-residing brines to ascend into the cover rocks.
Moreover, the Th versus salinity relationships (Figure 13A) along with the ternary H2O-NaCl-CaCl2 plot (Figure 11B) show two main saline fluid inclusion populations with total salinities in the range of ~14 to ~20 wt% NaCl + CaCl2, and between ~20 to ~24 wt% NaCl + CaCl2, respectively. In this regard, the continuous linear correlations shown on temperature vs. NaCl/(NaCl + CaCl2) (Figure 13B) and CaCl2 vs. NaCl bi-plot diagrams (Figure 13C) are interpreted as binary isothermal mixing trends between two compositionally different endmembers including a CaCl2-dominated and a NaCl-dominated brine. This compositional dichotomy indicates that the two ore-forming brines had distinct reservoirs, flow paths, and fluid–rock interactions. While the CaCl2-dominated brine is interpreted to have been derived from the interaction between the NaCl-rich brine and Ca-rich protoliths in the basement, the Na-enriched brine more likely acquired its high salinity in a subsurface setting (i.e., basin) through extensive interaction with salt-bearing rocks. In the study area, the outcropping country rocks (Figure 1) show the occurrence of halite- to gypsum-bearing evaporites in the Triassic red bed succession [82], and as such, this latter aquifer could have acted as the sedimentary endmember. The Ca-enrichment of the CaCl2-dominant ore-forming brine could have been acquired by Na/Ca and Na/K exchanges by albitization of plagioclase and K-feldspar, respectively, and Mg-loss by Mg-rich alteration of Ca-rich basaltic rocks.
In summary, the combined fluid inclusion microthermometry, isochore calculation, and P-T estimate indicate the involvement of two main fluid endmembers, including NaCl-rich and CaCl2-rich brines.

5.2. Origin(s) and Fluid Evolution from the Trace Element Compositions of Syn-Ore Apatite and Rutile

In addition to being utilized for dating [47,83,84], the combined use of high-resolution in situ LA-ICP-MS trace element measurements on hydrothermal apatite and rutile have been shown to be reliable proxies for the physicochemical conditions and the oxidation state of the hydrothermal system during mineralizing events (i.e., temperature, pH, oxidation state, oxygen fugacity) [85,86,87,88,89]. The trace element variations shown by Aouli syn-ore apatite and rutile can be attributed to (i) the composition of the primary TiO2-bearing mineral phases [60,90,91], (ii) origin and composition of ore-forming fluids (Cl- or F-rich), (iii) physicochemical conditions during mineral precipitation (i.e., temperature), and (iv) oxygen (fO2) fugacity [87,88,89,92].
The middle rare earth elements’ (MREE) enrichment of apatite, except for Eu which shows a prominent negative Eu anomaly (Figure 9A), is interpreted to be related to dissolution/recrystallization processes driven by protracted fluid–rock interactions with the country rocks including igneous and metamorphic mafic and felsic protoliths along with organic-rich siliciclastic rocks. Moreover, the negative Eu anomaly shown by syn-ore apatite is interpreted as resulting from the breakdown of plagioclase, which more likely had occurred under oxidizing fluid conditions (i.e., high oxygen fugacity [93]), consistent with the rutile intragrain compositional variability of U, Th, Fe, V, and Nb (Figure 14). Altogether, the enrichment of syn-ore apatite in MREE and Cl-OH (Table 2) and rutile in Zr, Nb, and Sn-W (Figure 9B) both suggest the involvement of neutral to alkaline halogen-rich (F, Cl) hydrothermal fluids [59,94,95,96]. Additional evidence for the involvement of neutral F- and Cl-rich brines is provided by the widespread occurrence of fluorite and adularia along with the high salinity (14–23 wt.% NaCl + CaCl2) of the ore-forming fluids as inferred from fluid inclusions microthermometry.
In summary, the large compositional variations shown by syn-ore apatite and rutile suggest that the hydrothermal pulses contained fluctuating levels of trace elements driven by mixing and subsequent fluid–rock interactions.

5.3. Fluid Mixing and Subsequent Fluid-Rock Interactions as Effective Depositional Processes

The Th versus salinity relationships, and more importantly the linear correlations shown on Figure 12 and Figure 13, reveal variable degrees of isothermal mixing between two main mineralizing brines, including a deep-seated metal-rich CaCl2-dominated brine and sulfide-bearing (H2S and HS) NaCl-dominated basinal brine. Most recent genetic models for the Ni-Co-As ore shoots invoke the involvement of redox reactions between upwelling oxidized ore fluids and reducing petroleum-bearing fluids (oil and gas) or organic-rich lithologies [1,6,98,99]. Some investigators have proposed that the metals were carried by hydrocarbons derived from black shales, and it was the oxidation of the metal-rich reducing fluids that triggered the precipitation of the metals [53,100]. Although the direct evidence for the potential involvement of hydrocarbon-bearing fluids in Aouli ore system is lacking (absence of petroleum- and gas-bearing fluid inclusions), the high Hg content of the ores (up to 14 wt%; Table S1) along with the occurrence of organic–organic-rich siliceous shale within the stratigraphic succession support the concept that organically-derived ore-forming fluids would have occurred at the site of ore deposition. By comparison, hydrocarbon-bearing brines [101] were invoked as responsible for deposition of the REE-rich fluorite mineralization in the adjacent El Hammam REE-rich fluorite deposit, whose geological setting, mineralogical attributes, and the age of mineralization are similar to those of Aouli [15,16].
Overall, the geochemical characteristics discussed above are consistent with mixing between three fluid endmembers, including (i) sulfide-bearing (H2S and HS) Na-rich brine, (ii) basement-derived metal-bearing CaCl2-rich reducing brine, and (iii) oil/gas-bearing brine. Precipitation of arsenide, sulfosalt, and sulfide phases is thought to have occurred as a result of redox reactions induced by ternary mixing and subsequent fluid–rock interactions of the above-mentioned fluid endmembers.

5.4. Timing of Mineralization

Despite the discovery at Aouli deposit of hydrothermal ore-related phases reputed as representing suitable material for conventional radiometric measurements such as apatite, rutile, sericite, and adularia, our attempts to directly date hydrothermal mineralization failed to provide accurate age constraints. The U-Pb dating of apatite and rutile has been hampered by the low concentrations of U (typically <1 ppm) and moderate Th (mostly between 10 and 15 ppm), which depart from the consistently high common lead contents of both phases. Similarly, adularia and sericite were not amenable for K-Ar and/or Ar-Ar dating as their fine-grain size (<20 µm) is far lower than the ablation spot size (50 μm).
Accordingly, the relative timing of mineralization in the deposit is indirectly constrained by combining geological field observations and textural cross-cutting relationships. The Aouli vein ore shoots cut across the Triassic red-beds (Figure 4A), thereby constraining the age of mineralization to at least the Triassic. More importantly, radiometric ages along with thermochronological data [102,103,104,105,106] indicate that major development of supergene mineralization, which forms by oxidation at the expense of former sulfides, likely occurred as recently as the late Miocene. Based on these relationships, the inferred age of Aouli vein ore shoot mineralization is broadly bracketed between Late Triassic and Late Miocene time. Regionally, this timeframe coincides with the Late Jurassic–Early Cretaceous to Late Miocene time span, which represents a significant metallogenic epoch for the formation of major post-Variscan hydrothermal vein-type and MVT deposits across Europe and North Africa [1,107] and North Africa as well [108].

6. Metallogenic Model and Concluding Remarks

Previous studies at Aouli suggested that the deposit was formed by low-temperature hydrothermal fluids, but there has been no consensus as to the nature of the mineralizing fluids and the deposit classification [15,16,109]. This research reveals that the unusual Ag-(Ni-Co-Sb-As-Hg ± Bi) fluorite-barite ore shoot mineralization formed from episodic stages of fracturing and subsequent fluid migration. The mineralogy, fluid sources, and mechanisms of ore deposition at the origin of this newly discovered mineralization style are constrained therein based on an integrated approach that combines mineralogy, paragenesis, and fluid inclusion studies. The new data sets allowed us to refine preexisting genetic models [15,16], including those proposed based on structural constraints [38].
Paragenetic relationships indicate that the Aouli vein ore shoot mineralization formed as a result of three fluid pulses, including the early Ni-Co-Fe arsenides, the intermediate Pb-Sb-As-Ag sulfarsenide-sulfosalts, and the late Pb-Zn ± Cu sulfides with quartz, fluorite, and barite being precipitated in all three stages. Fluid inclusion petrography and microthermometry combined with trace element geochemistry of ore and related gangue phases along with isotopic constraints reveal that the ore shoot mineralization formed by variable degrees of isothermal mixing and subsequent fluid–rock interaction between two main ore brines, including a deep-seated metal-rich CaCl2-dominated basement brine and sulfide-bearing (H2S and HS) NaCl-dominated basinal brine. The involvement of a third methane-dominated fluid–gas fluid which triggered ore deposition is inferred from the enrichment of the Aouli ore in Hg. The development of breccia structures in the ores and the fluid pressure fluctuation revealed by fluid inclusions suggest that the deposit formed from episodic stages of fracturing and subsequent fluid migration. Seismic pumping is proposed as the likely mechanism that drove the released mineralizing brines into the newly created and reopened veins contributing to the vein ore shoot mineralization. Precipitation of ore phases is thought to have occurred as a result of isothermal mixing. Fluid–rock interactions exerted strong control over the leaching process and transport of the released metals by the ore-forming brines from which Aouli ores precipitated.
In the absence of radiometric age constraints, the timing of Aouli mineralization is indirectly bracketed between Late Triassic and Late Miocene time, predating the climax of the Atlasic compressional event in the Late Miocene (late alpine phase [110,111]). From a geodynamic point of view, the time ascribed to the Aouli deposit coincides with major geodynamic events that developed as a causal far-field effect of the drifting of the Central Atlantic Ocean and formation of the Maghrebian Tethys passive margin [23,24,111,112,113,114,115]. These events include major crustal extension and widespread Middle Jurassic–Upper Cretaceous alkaline basaltic magmatism in the Atlas system [116,117,118]. Therefore, we propose a metallogenic model which involves a genetic link among northwestern African plate geodynamic, extensional tectonics, Middle Jurassic–Upper Cretaceous alkaline magmatism, and formation of Aouli deposit. In such a setting, we suggest that thermal convection induced by the generated high heat flow and subsequent increased geothermal gradient triggered development of small-scale convection cells that controlled the large-scale migration of ore-forming brines and formation of the Aouli unconformity-related polymetallic deposit in the Upper Moulouya metallogenic province.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/min15070669/s1. Table S1: Chemical compositions of arsenides, sulfosalts and sulfides in wt % (analyzed by EPMA).

Author Contributions

K.Z.: Data curation, Formal Analysis, Investigation, Conceptualization, Writing—Original Draft, Data Curation, Validation, Visualization. M.B.: Supervision, Conceptualization, Formal Analysis, Writing—Review and Editing, Methodology, and Validation. G.L.: Writing—Review and Editing, Writing—Original Draft, Methodology, Investigation. M.I.: Writing—Review and Editing, Writing—Original Draft, Fluid Inclusion Data Acquisition, Data Curation, Software. R.M., E.M. and J.Y.: Funding Acquisition, Project Administration, Review and Editing. M.J., A.P. and L.B.: Review and Editing. M.F. and V.v.S.: LA-ICP-MS Data Acquisition and Analysis, Review and Editing. S.B.: EMPA Data Acquisition, Review and Editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Data will be made available on request.

Acknowledgments

This contribution is part of an ongoing Ph.D. project by the first author, conducted within the framework of an international collaborative program between Mohamed Premier University of Oujda in Morocco, in partnership with Universidad Nacional Autónoma de México (UNAM) (Querétaro, Mexico) and UNamur (Belgium). Fundings for this research were provided by a UNAM grant awarded to Gilles Levresse, and additional support from the National Science Foundation (NSF EAR) under grants No. 2233425 and 2233426, awarded to Ryan Mathur. The first author acknowledges funding from the CNRST (grant No. 18UMP2020) and the European Union’s Learning Mobility of Individuals (AC 171) grant awarded to Mohammed Bouabdellah and Johan Yans. We also acknowledge the Instituto de Geociencias laboratories at UNAM and express our gratitude to Marina Vega, Luigi Solari, and Carlos Ortega for their scientific and technical assistance, Marc Choquette (University Laval, Canada) for conducting the EPMA analyses, and Hassan Bouzahzah (University of Liège, Belgium) for his assistance with SEM-EDS analyses. Special thanks and gratitude are extended to the three anonymous reviewers for their thoughtful reviews, which significantly improved the earlier version of this paper.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Burisch, M.; Gerdes, A.; Walter, B.F.; Neumann, U.; Fettel, M.; Markl, G. Methane and the Origin of Five-Element Veins: Mineralogy, Age, Fluid Inclusion Chemistry and Ore Forming Processes in the Odenwald, SW Germany. Ore Geol. Rev. 2017, 81, 42–61. [Google Scholar] [CrossRef]
  2. Walter, B.F.; Kortenbruck, P.; Scharrer, M.; Zeitvogel, C.; Wälle, M.; Mertz-Kraus, R.; Markl, G. Chemical Evolution of Ore-Forming Brines–Basement Leaching, Metal Provenance, and the Redox Link between Barren and Ore-Bearing Hydrothermal Veins. A Case Study from the Schwarzwald Mining District in SW-Germany. Chem. Geol. 2019, 506, 126–148. [Google Scholar] [CrossRef]
  3. Bouabdellah, M.; Slack, J.F. Geologic and Metallogenic Framework of North Africa. In Mineral Deposits of North Africa; Bouabdellah, M., Slack, J.F., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 3–81. ISBN 978-3-319-31733-5. [Google Scholar]
  4. Boushaba, A.; Aabi, A.; Mouttaqi, A.; Nguidi, M.A.; Lamrani, O.; Samaoui, S.; Sadéqui, M.; Azza, A.; Ajamay, H.; Mbarki-Kadiri, B. El The Pb-Zn-Ag±Au-W-Sn Vein System in Variscan (Hercynian) Belts in North Africa. In The Geology of North Africa; Springer: Berlin/Heidelberg, Germany, 2024; pp. 393–470. [Google Scholar]
  5. Harrison, R.W. Primary structural control of epithermal mineralized shoots in southeastern Chloride mining district, New Mexico. New Mex. Geol. 1988, 10, 80–81. [Google Scholar] [CrossRef]
  6. Markl, G.; Burisch, M.; Neumann, U. Natural Fracking and the Genesis of Five-Element Veins. Miner. Depos. 2016, 51, 703–712. [Google Scholar] [CrossRef]
  7. Guilcher, M.; Schmaucks, A.; Krause, J.; Markl, G.; Gutzmer, J.; Burisch, M. Vertical Zoning in Hydrothermal U-Ag-Bi-Co-Ni-As Systems: A Case Study from the Annaberg-Buchholz District, Erzgebirge (Germany). Econ. Geol. 2021, 116, 1893–1915. [Google Scholar] [CrossRef]
  8. Scharrer, M.; Reich, R.; Fusswinkel, T.; Walter, B.F.; Markl, G. Basement Aquifer Evolution and the Formation of Unconformity-Related Hydrothermal Vein Deposits: LA-ICP-MS Analyses of Single Fluid Inclusions in Fluorite from SW Germany. Chem. Geol. 2021, 575, 120260. [Google Scholar] [CrossRef]
  9. Éric, M.; Khadija, N.; Yannick, B.; Claire, R.; Gilles, R.; Jean-Jacques, P.; Ross, S.; Michel, J. Late-Hercynian Intrusion-Related Gold Deposits: An Integrated Model on the Tighza Polymetallic District, Central Morocco. J. Afr. Earth Sci. 2015, 107, 65–88. [Google Scholar] [CrossRef]
  10. Rossi, M.; Gasquet, D.; Cheilletz, A.; Tarrieu, L.; Bounajma, H.; Mantoy, T.; Reisberg, L.; Deloule, E.; Boulvais, P.; Burnard, P. Isotopic and Geochemical Constraints on Lead and Fluid Sources of the PbZnAg Mineralization in the Polymetallic Tighza-Jbel Aouam District (Central Morocco), and Relationships with the Geodynamic Context. J. Afr. Earth Sci. 2017, 127, 194–210. [Google Scholar] [CrossRef]
  11. Rossi, M.; Tarrieu, L.; Cheilletz, A.; Gasquet, D.; Deloule, E.; Paquette, J.-L.; Bounajma, H.; Mantoy, T.; Ouazzani, L.; Ouchtouban, L. The Polymetallic (W–Au and Pb–Zn–Ag) Tighza District (Central Morocco): Ages of Magmatic and Hydrothermal Events. In Mineral Deposits of North Africa; Bouabdellah, M., Slack, J.F., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 107–131. ISBN 978-3-319-31733-5. [Google Scholar]
  12. El Arbaoui, A.; Bouabdellah, M.; Wafik, A.; Klügel, A.; Jébrak, M.; Castorina, F.; Lowry, D.; Lecumberri-Sanchez, P.; Essaifi, A.; Maacha, L. The Roc Blanc Orogenic Pb-Zn-Ag-Au Deposit (Morocco): A Product of Metamorphic Dehydration and CO2 Devolatilization during Exhumation of the Variscan Jebilet Massif. Miner. Depos. 2019, 54, 437–458. [Google Scholar] [CrossRef]
  13. Bouabdellah, M.; Talbi, F.; Benalla, M.; Boudchiche, L.; Torbi, A.; Grandia, F.; Cardellach, E.; Corbella, M. Contexte Géologique et Caractérisation Physico-Chimique Des Fluides Associés Aux Minéralisations Pb-Zn-Ag Du District de Sidi Lahcen (Maroc Nord Oriental). In Proceedings of the Deuxièmes Journées De Launay, Marrakech, Moroccan, 7–9 June 2006. [Google Scholar]
  14. Naji, M. Les Minéralisations Plombo-Barytiques Du District de La Haute Moulouya. Contexte Géologique, Contrôle Tectonique et Modèle de Mise En Place: Gisements d’Aouli-Mibladen-Zeida. Ph.D. Thesis, Mohamed V University, Rabat, Morocco, 2004. Unpublished Thesis. [Google Scholar]
  15. Bouabdellah, M.; Margoum, D. Geology, Fluid Inclusions, and Geochemistry of the Aouli Sulphide ± Fluorite ± Barite Vein Deposit (Upper Moulouya District, Morocco) and Its Relationships to Pangean Rifting and Opening of the Tethys and Central Atlantic Oceans. In Mineral Deposits of North Africa; Springer: Cham, Switzerland, 2016; pp. 291–305. [Google Scholar]
  16. Margoum, D.; Bouabdellah, M.; Klügel, A.; Banks, D.A.; Castorina, F.; Cuney, M.; Jébrak, M.; Bozkaya, G. Pangea Rifting and Onward Pre-Central Atlantic Opening as the Main Ore-Forming Processes for the Genesis of the Aouli REE-Rich Fluorite-Barite Vein System, Upper Moulouya District, Morocco. J. Afr. Earth Sci. 2015, 108, 22–39. [Google Scholar] [CrossRef]
  17. Emberger, A. Introduction à l’étude Des Minéralisations Plombifères de La Haute Moulouya. Notes Mémoires Géologique 1965, 181, 167–180.m. [Google Scholar]
  18. Gimeno-Vives, O.; Mohn, G.; Bosse, V.; Haissen, F.; Zaghloul, M.N.; Atouabat, A.; Frizon de Lamotte, D. The Mesozoic Margin of the Maghrebian Tethys in the Rif Belt (Morocco): Evidence for Polyphase Rifting and Related Magmatic Activity. Tectonics 2019, 38, 2894–2918. [Google Scholar] [CrossRef]
  19. Michard, A.; Saddiqi, O.; Chalouan, A.; Chabou, M.C.; Lach, P.; Rossi, P.; Bertrand, H.; Youbi, N. Comment on “The Mesozoic Margin of the Maghrebian Tethys in the Rif Belt (Morocco): Evidence for Polyphase Rifting and Related Magmatic Activity” by Gimeno-Vives et Al. Tectonics 2020, 39, e2020TC006165. [Google Scholar] [CrossRef]
  20. Frizon de Lamotte, D.; Saint Bezar, B.; Bracène, R.; Mercier, E. The Two Main Steps of the Atlas Building and Geodynamics of the Western Mediterranean. Tectonics 2000, 19, 740–761. [Google Scholar] [CrossRef]
  21. Beauchamp, W.; Allmendinger, R.W.; Barazangi, M.; Demnati, A.; El Alji, M.; Dahmani, M. Inversion Tectonics and the Evolution of the High Atlas Mountains, Morocco, Based on a Geological-Geophysical Transect. Tectonics 1999, 18, 163–184. [Google Scholar] [CrossRef]
  22. Ziegler, P.A. Geological Atlas of Western and Central Europe. Geol. Mag. 1982, 121, 371–372. [Google Scholar] [CrossRef]
  23. Frizon De Lamotte, D.; Zizi, M.; Missenard, Y.; Hafid, M.; El Azzouzi, M.; Maury, R.C.; Charrière, A.; Taki, Z.; Benammi, M.; Michard, A. The Atlas System. Lect. Notes Earth Sci. 2008, 116, 133–202. [Google Scholar] [CrossRef]
  24. Escosa, F.O.; Leprêtre, R.; Spina, V.; Gimeno-Vives, O.; Kergaravat, C.; Mohn, G.; Frizon de Lamotte, D. Polyphased Mesozoic Rifting from the Atlas to the North-West Africa Paleomargin. Earth Sci. Rev. 2021, 220, 103732. [Google Scholar] [CrossRef]
  25. Dresnay, R. du Extension et Développement Des Phénomènes Récifaux Jurassiques Dans Le Domaine Atlasique Marocain, Particulièrement Au Lias Moyen. Bull. Société Géologique Fr. 1971, 7, 46–56. [Google Scholar] [CrossRef]
  26. Morel, J.-L.; Zouine, E.M.; Poisson, A. Relations between the Moulouya Basins Subsidence and the Atlasic Uplift, Morocco: An Example of Tectonic Reversal since Neogene Time Relations Entre La Subsidence Des Bassins Moulouyens et La Creation Des Reliefs Atlasiques (Maroc): Un Exemple d’inversi. Bull.-Soc. Geol. Fr. 1993, 164, 79–91. [Google Scholar]
  27. Zouine, E.M. Géodynamique Récente Du Haut Atlas. Evolution de Sa Bordure Septentrionale et Du Moyen Atlas Sud-Occidental Au Cours Du Cénozoïque. Ph.D. Thesis, Mohamed V University, Rabat, Morocco, 1993. Unpublished Thesis. [Google Scholar]
  28. Emberger, A. Eléments Pour Une Synthèse Métallogénique Du District Plombifère de La Haute Moulouya. Notes Mém. Serv. Géol. Maroc. 1965, 181, 205–244. [Google Scholar]
  29. Hoepffner, C.; Houari, M.R.; Bouabdelli, M. Tectonics of the North African Variscides (Morocco, Western Algeria): An Outline. Comptes. Rendus-Geosci. 2006, 338, 25–40. [Google Scholar] [CrossRef]
  30. Hoepffner, C.; Soulaimani, A.; Piqué, A. The Moroccan Hercynides. J. Afr. Earth Sci. 2005, 43, 144–165. [Google Scholar] [CrossRef]
  31. Raddi, Y.; Aarar, M.; Michard, A. Notice Explicative de La Carte Géologique Du Maroc Au 1/50.000 (Feuille de Mibladen). Notes Mémoires Serv. Géologique Maroc. 2013, 576, 24–59. [Google Scholar]
  32. Chopin, F.; Leprêtre, R.; El Houicha, M.; Tabaud, A.-S.; Schulmann, K.; Míková, J.; Barbarand, J.; Chebli, R. U–Pb Geochronology of Variscan Granitoids from the Moroccan Meseta (Northwest Africa): Tectonic Implications. Gondwana Res. 2023, 117, 274–294. [Google Scholar] [CrossRef]
  33. Michard, A.; Soulaimani, A.; Hoepffner, C.; Ouanaimi, H.; Baidder, L.; Rjimati, E.C.; Saddiqi, O. The South-Western Branch of the Variscan Belt: Evidence from Morocco. Tectonophysics 2010, 492, 1–24. [Google Scholar] [CrossRef]
  34. Accotto, C.; Martínez Poyatos, D.; Azor, A.; Jabaloy-Sánchez, A.; Talavera, C.; Evans, N.J.; Azdimousa, A. Tectonic Evolution of the Eastern Moroccan Meseta: From Late Devonian Forearc Sedimentation to Early Carboniferous Collision of an Avalonian Promontory. Tectonics 2020, 39, e2019TC005976. [Google Scholar] [CrossRef]
  35. Dahire, M. Le Complexe Plutonique de La Haute Moulouya (Meseta Orientale, Maroc): Evolution Pétrologique et Structurale. Ph.D. Thesis, Faculty of Sciences Dhar El Mahraz, Fes, Morocco, 2004. [Google Scholar]
  36. Elabouyi, M.; Dahire, M.; Driouch, Y.; Duchêne, S.; Kriegsman, L.M.; Ntarmouchant, A.; Kahou, Z.S.; Severac, J.L.; Belkasmi, M.; Debat, P. Crustal Anatexis in the Aouli-Mibladen Granitic Complex: A Window into the Middle Crust below the Moroccan Eastern Variscan Meseta. J. Afr. Earth Sci. 2019, 154, 136–163. [Google Scholar] [CrossRef]
  37. Filali, F. Etude Pétro-Structurale de l’encaissant Métamorphique de La Boutonnière d’Aouli-Mibladen (Haute Moulouya, Maroc): Conséquences Sur La Géodynamique Hercynienne Au Maroc. Ph.D. Thesis, Muséum national d’histoire naturelle, Paris, France, 1996. [Google Scholar]
  38. Yaagoub, D.; Hinaje, S.; El Ouaragli, B.; El Fartati, M.; Gharmane, Y.; Amrani, S.; Ouhssaine, A.; Elabouyi, M. Reconstruction of Mesozoic Paleostress Associated with the Tectono-Sedimentary Evolution of the Basins Bordering the Aouli Inlier (Upper Moulouya, Morocco). Arab. J. Geosci. 2021, 14, 706. [Google Scholar] [CrossRef]
  39. Abdullin, F.; Solé, J.; de Jesús Meneses-Rocha, J.; Solari, L.; Shchepetilnikova, V.; Ortega-Obregón, C. LA-ICP-MS-Based Apatite Fission Track Dating of the Todos Santos Formation Sandstones from the Sierra de Chiapas (SE Mexico) and Its Tectonic Significance. Int. Geol. Rev. 2016, 58, 32–48. [Google Scholar] [CrossRef]
  40. Solari, L.A.; Gómez-Tuena, A.; Bernal, J.P.; Pérez-Arvizu, O.; Tanner, M. U-Pb Zircon Geochronology with an Integrated LA-ICP-MS Microanalytical Workstation: Achievements in Precision and Accuracy. Geostand. Geoanal Res. 2010, 34, 5–18. [Google Scholar] [CrossRef]
  41. Ortega-Obregón, C.; Abdullin, F.; Solari, L.; Schaaf, P.; Solís-Pichardo, G. Apatite U-Pb Dating at UNAM Laboratories: Analytical Protocols and Examples of Its Application. Rev. Mex. Cienc. Geol. 2019, 36, 27–37. [Google Scholar] [CrossRef]
  42. Paton, C.; Hellstrom, J.; Paul, B.; Woodhead, J.; Hergt, J. Iolite: Freeware for the Visualisation and Processing of Mass Spectrometric Data. J. Anal. At. Spectrom. 2011, 26, 2508–2518. [Google Scholar] [CrossRef]
  43. Paton, C.; Woodhead, J.D.; Hellstrom, J.C.; Hergt, J.M.; Greig, A.; Maas, R. Improved Laser Ablation U-Pb Zircon Geochronology through Robust Downhole Fractionation Correction. Geochem. Geophys. Geosystems 2010, 11, 3. [Google Scholar] [CrossRef]
  44. Solari, L.A.; González-León, C.M.; Ortega-Obregón, C.; Valencia-Moreno, M.; Rascón-Heimpel, M.A. The Proterozoic of NW Mexico Revisited: U–Pb Geochronology and Hf Isotopes of Sonoran Rocks and Their Tectonic Implications. Int. J. Earth Sci. 2018, 107, 845–861. [Google Scholar] [CrossRef]
  45. Jochum, K.P.; Weis, U.; Stoll, B.; Kuzmin, D.; Yang, Q.; Raczek, I.; Jacob, D.E.; Stracke, A.; Birbaum, K.; Frick, D.A.; et al. Determination of Reference Values for NIST SRM 610–617 Glasses Following ISO Guidelines. Geostand. Geoanal Res. 2011, 35, 397–429. [Google Scholar] [CrossRef]
  46. Luvizotto, G.L.; Zack, T.; Meyer, H.P.; Ludwig, T.; Triebold, S.; Kronz, A.; Münker, C.; Stockli, D.F.; Prowatke, S.; Klemme, S.; et al. Rutile Crystals as Potential Trace Element and Isotope Mineral Standards for Microanalysis. Chem. Geol. 2009, 261, 346–369. [Google Scholar] [CrossRef]
  47. Zack, T.; Stockli, D.F.; Luvizotto, G.L.; Barth, M.G.; Belousova, E.; Wolfe, M.R.; Hinton, R.W. In Situ U–Pb Rutile Dating by LA-ICP-MS: 208Pb Correction and Prospects for Geological Applications. Contrib. Mineral. Petrol. 2011, 162, 515–530. [Google Scholar] [CrossRef]
  48. Zaid, K.; Bouabdellah, M.; Frenzel, M.; van Schijndel, V.; Levresse, G.; Idbaroud, M.; Yans, J. Laser Ablation Inductively Coupled Plasma Mass Spectrometry Trace Element Chemistry of Sulfides from the Aouli Ag–Pb–Zn–(Cu) Vein System (Upper Moulouya District, Morocco). GFZ Data Serv. 2025. [Google Scholar]
  49. Neuser, R.D.; Reinecke, T.; Schertl, H.-P. Low Temperature Cathodoluminescence of Selected Minerals from High Pressure Metamorphic Rocks. Boch. Geol. Geotech. Arb. 1995, 44, 119–123. [Google Scholar]
  50. Steele-MacInnis, M.; Lecumberri-Sanchez, P.; Bodnar, R.J. HokieFlincs_H2O-NaCl: A Microsoft Excel Spreadsheet for Interpreting Microthermometric Data from Fluid Inclusions Based on the PVTX Properties of H2O–NaCl. Comput. Geosci. 2012, 49, 334–337. [Google Scholar] [CrossRef]
  51. Guilcher, M.; Albert, R.; Gerdes, A.; Gutzmer, J.; Burisch, M. Timing of Native Metal-Arsenide (Ag-Bi-Co-Ni-As±U) Veins in Continental Rift Zones–In Situ U-Pb Geochronology of Carbonates from the Erzgebirge/Krušné Hory Province. Chem. Geol. 2021, 584, 120476. [Google Scholar] [CrossRef]
  52. Burisch, M.; Frenzel, M.; Seibel, H.; Gruber, A.; Oelze, M.; Pfänder, J.A.; Sanchez-Garrido, C.; Gutzmer, J. Li-Co–Ni-Mn-(REE) Veins of the Western Erzgebirge, Germany—A Potential Source of Battery Raw Materials. Miner. Depos. 2021, 56, 1223–1238. [Google Scholar] [CrossRef]
  53. Ferguson, D.; Chi, G.; Normand, C.; Mercadier, J.; Wang, Y.; McKee, K.; Anderson, M.; Robbins, J. Relationship between U and Ni-Co-As Mineralization in the Midwest Polymetallic U Deposit, Athabasca Basin (Canada)–Constraints from Mineralogical, Geochemical, and Fluid Inclusion Studies. Miner. Depos. 2025, 60, 63–83. [Google Scholar] [CrossRef]
  54. Diallo, M.; Bouabdellah, M.; Levresse, G.; Yans, J.; Castorina, F.; Klügel, A.; Mouhagir, M.; El Mouden, S.; Maacha, L. Mineralogy, Fluid Inclusion, and CO-Sr Isotope Geochemistry to Unravel the Evolution of the Magmatic-Hydrothermal System at the Igoudrane Silver-Rich Deposit (Imiter District, Eastern Anti-Atlas, Morocco). Minerals 2021, 11, 997. [Google Scholar] [CrossRef]
  55. Bouabdellah, M.; Maacha, L.; Levresse, G.; Saddiqi, O. The Bou Azzer Co–Ni–Fe–As (±Au ±Ag) District of Central Anti-Atlas (Morocco): A Long-Lived Late Hercynian to Triassic Magmatic-Hydrothermal to Low-Sulphidation Epithermal System. In Mineral Deposits of North Africa; Springer: Cham, Switzerland, 2016; pp. 229–247. [Google Scholar]
  56. Levresse, G.; Bouabdellah, M.; Cheilletz, A.; Gasquet, D.; Maacha, L.; Tritlla, J.; Banks, D.; Moulay Rachid, A.S. Degassing as the Main Ore-Forming Process at the Giant Imiter Ag–Hg Vein Deposit in the Anti-Atlas Mountains, Morocco. In Mineral Deposits of North Africa; Springer: Cham, Switzerland, 2016; pp. 85–106. [Google Scholar]
  57. McDonough, W.F.; Sun, S.-S. The Composition of the Earth. Chem. Geol. 1995, 120, 223–253. [Google Scholar] [CrossRef]
  58. Adlakha, E.E.; Hattori, K.; Kerr, M.J.; Boucher, B.M. The Origin of Ti-Oxide Minerals below and within the Eastern Athabasca Basin, Canada. Am. Mineral. 2020, 105, 1875–1888. [Google Scholar] [CrossRef]
  59. Chen, Q.; Wang, C.; Bagas, L.; Du, B.; Shi, K.; Zhu, J. Hydrothermal Fluid Signatures of the Yulong Porphyry Cu-Mo Deposit: Clues from the Composition and U-Pb Dating of W-Bearing Rutile. Am. Mineral. 2023, 108, 1092–1108. [Google Scholar] [CrossRef]
  60. Scott, K.M. Rutile Geochemistry as a Guide to Porphyry Cu–Au Mineralization, Northparkes, New South Wales, Australia. Geochem. Explor. Environ. Anal. 2005, 5, 247–253. [Google Scholar] [CrossRef]
  61. Urban, A.J.; Hoskins, B.F.; Grey, I.E. Characterization of V-Sb-W-Bearing Rutile from the Hemlo Gold Deposit, Ontario. Can. Mineral. 1992, 30, 319–326. [Google Scholar]
  62. Goldstein, R.H.; Reynolds, T.J. Systematics of Fluid Inclusions in Diagenetic Minerals; SEPM Society for Sedimentary Geology: Claremore, OK, USA, 1994; ISBN 9781565760981. [Google Scholar]
  63. Chi, G.; Diamond, L.W.; Lu, H.; Lai, J.; Chu, H. Common Problems and Pitfalls in Fluid Inclusion Study: A Review and Discussion. Minerals 2020, 11, 7. [Google Scholar] [CrossRef]
  64. Roedder, E. Fluid Inclusions: An Introduction to Studies of All Types of Fluid Inclusions, Gas, Liquid, or Melt, Trapped in Materials from Earth and Space, and Their Application to the Understanding of Geologic Processes; Reviews in Mineralogy; Mineral in North of America: Washington, DC, USA, 1984; ISBN 978-0-939950-16-4. [Google Scholar]
  65. Goldstein, R.H. Petrographic Analysis of Fluid Inclusions. In Fluid Inclusions: Analysis and Interpretation; Mineralogical Association of Canada: Québec, QC, Canada, 2003; pp. 9–53. [Google Scholar]
  66. Diamond, L.W. Review of the Systematics of CO2-H2O Fluid Inclusions. Lithos 2001, 55, 69–99. [Google Scholar] [CrossRef]
  67. Staude, S.; Wagner, T.; Markl, G. Mineralogy, Mineral Compositions and Fluid Evolution at the Wenzel Hydrothermal Deposit, Southern Germany: Implications for the Formation of Kongsberg-Type Silver Deposits. Can. Miner. 2007, 45, 1147–1176. [Google Scholar] [CrossRef]
  68. Staude, S.; Werner, W.; Mordhorst, T.; Wemmer, K.; Jacob, D.E.; Markl, G. Multi-Stage Ag-Bi-Co-Ni-U and Cu-Bi Vein Mineralization at Wittichen, Schwarzwald, SW Germany: Geological Setting, Ore Mineralogy, and Fluid Evolution. Miner. Depos. 2012, 47, 251–276. [Google Scholar] [CrossRef]
  69. En-Naciri, A.; Barbanson, L.; Touray, J.-C. Brine Inclusions from the Co-As(Au) Bou Azzer District, Anti-Atlas Mountains, Morocco. Econ. Geol. 1997, 92, 360–367. [Google Scholar] [CrossRef]
  70. Ahmed, A.H.; Arai, S.; Ikenne, M. Mineralogy and Paragenesis of the Co-Ni Arsenide Ores of Bou Azzer, Anti-Atlas, Morocco. Econ. Geol. 2009, 104, 249–266. [Google Scholar] [CrossRef]
  71. Saintilan, N.J.; Selby, D.; Creaser, R.A.; Dewaele, S. Sulphide Re-Os Geochronology Links Orogenesis, Salt and Cu-Co Ores in the Central African Copperbelt. Sci. Rep. 2018, 8, 14946. [Google Scholar] [CrossRef]
  72. Essarraj, S.; Boiron, M.-C.; Cathelineau, M.; Banks, D.A.; Benharref, M. Penetration of Surface-Evaporated Brines into the Proterozoic Basement and Deposition of Co and Ag at Bou Azzer (Morocco): Evidence from Fluid Inclusions. J. Afr. Earth Sci. 2005, 41, 25–39. [Google Scholar] [CrossRef]
  73. Changkakoti, A.; Morton, R.D.; Gray, J. Hydrothermal Environments during the Genesis of Silver Deposits in the Northwest Territories of Canada: Evidence from Fluid Inclusions. Miner. Depos. 1986, 21, 63–69. [Google Scholar] [CrossRef]
  74. Kerrich, R.; Strong, D.F.; Andrews, A.J.; Owsiacki, L. The Silver Deposits at Cobalt and Gowganda, Ontario. III: Hydrothermal Regimes and Source Reservoirs-Evidence from H, O, C, and Sr Isotopes and Fluid Inclusions. Can. J. Earth Sci. 1986, 23, 1519–1550. [Google Scholar] [CrossRef]
  75. Marshall, D.D.; Diamond, L.W.; Skippen, G.B. Silver Transport and Deposition at Cobalt, Ontario, Canada; Fluid Inclusion Evidence. Econ. Geol. 1993, 88, 837–854. [Google Scholar] [CrossRef]
  76. Frenzel, M.; Woodcock, N.H. Cockade Breccia: Product of Mineralisation along Dilational Faults. J. Struct. Geol. 2014, 68, 194–206. [Google Scholar] [CrossRef]
  77. Moncada, D.; Mutchler, S.; Nieto, A.; Reynolds, T.J.; Rimstidt, J.D.; Bodnar, R.J. Mineral Textures and Fluid Inclusion Petrography of the Epithermal Ag-Au Deposits at Guanajuato, Mexico: Application to Exploration. J. Geochem. Explor. 2012, 114, 20–35. [Google Scholar] [CrossRef]
  78. Rimi, A. Geothermal Gradients and Heat Flow Trends in Morocco. Geothermics 1990, 19, 443–454. [Google Scholar] [CrossRef]
  79. Sibson, R.H. Fluid Involvement in Normal Faulting. J. Geodyn. 2000, 29, 469–499. [Google Scholar] [CrossRef]
  80. Bakker, R.J. Package FLUIDS 1. Computer Programs for Analysis of Fluid Inclusion Data and for Modelling Bulk Fluid Properties. Chem. Geol. 2003, 194, 3–23. [Google Scholar] [CrossRef]
  81. Zhang, Y.-G.; Frantz, J.D. Determination of the Homogenization Temperatures and Densities of Supercritical Fluids in the System NaClKClCaCl2H2O Using Synthetic Fluid Inclusions. Chem. Geol. 1987, 64, 335–350. [Google Scholar] [CrossRef]
  82. Amade, E. Les Gisements Pb-Zn de Zeida et de Bou Mia. Notes Mem. Géol Maroc Rabat 1965, 181, 175–184. [Google Scholar]
  83. Schandl, E.S.; Davis, D.W.; Krogh, T.E. Are the Alteration Halos of Massive Sulfide Deposits Syngenetic? Evidence from U-Pb Dating of Hydrothermal Rutile at the Kidd Volcanic Center, Abitibi Subprovince, Canada. Geology 1990, 18, 505–508. [Google Scholar] [CrossRef]
  84. Vry, J.K.; Baker, J.A. LA-MC-ICPMS Pb–Pb Dating of Rutile from Slowly Cooled Granulites: Confirmation of the High Closure Temperature for Pb Diffusion in Rutile. Geochim. Cosmochim. Acta 2006, 70, 1807–1820. [Google Scholar] [CrossRef]
  85. Belousova, E.A.; Griffin, W.L.; O’Reilly, S.Y.; Fisher, N.I. Apatite as an Indicator Mineral for Mineral Exploration: Trace-Element Compositions and Their Relationship to Host Rock Type. J. Geochem. Explor. 2002, 76, 45–69. [Google Scholar] [CrossRef]
  86. Bauer, M.E.; Seifert, T.; Burisch, M.; Krause, J.; Richter, N.; Gutzmer, J. Indium-Bearing Sulfides from the Hämmerlein Skarn Deposit, Erzgebirge, Germany: Evidence for Late-Stage Diffusion of Indium into Sphalerite. Miner. Depos. 2019, 54, 175–192. [Google Scholar] [CrossRef]
  87. Zack, T.; Kronz, A.; Foley, S.F.; Rivers, T. Trace Element Abundances in Rutiles from Eclogites and Associated Garnet Mica Schists. Chem. Geol. 2002, 184, 97–122. [Google Scholar] [CrossRef]
  88. Majzlan, J.; Bolanz, R.; Göttlicher, J.; Mikuš, T.; Milovská, S.; Čaplovičová, M.; Števko, M.; Rössler, C.; Matthes, C. Incorporation Mechanism of Tungsten in W-Fe-Cr-V-Bearing Rutile. Am. Mineral. 2021, 106, 609–619. [Google Scholar] [CrossRef]
  89. Zack, T.; von Eynatten, H.; Kronz, A. Rutile Geochemistry and Its Potential Use in Quantitative Provenance Studies. Sediment. Geol. 2004, 171, 37–58. [Google Scholar] [CrossRef]
  90. Czamanske, G.K.; Force, E.R.; Moore, W.J. Some Geologic and Potential Resource Aspects of Rutile in Porphyry Copper Deposits. Econ. Geol. 1981, 76, 2240–2246. [Google Scholar] [CrossRef]
  91. Rabbia, O.M.; Hernández, L.B.; French, D.H.; King, R.W.; Ayers, J.C. The El Teniente Porphyry Cu–Mo Deposit from a Hydrothermal Rutile Perspective. Miner. Depos. 2009, 44, 849–866. [Google Scholar] [CrossRef]
  92. Plavsa, D.; Reddy, S.M.; Agangi, A.; Clark, C.; Kylander-Clark, A.; Tiddy, C.J. Microstructural, Trace Element and Geochronological Characterization of TiO2 Polymorphs and Implications for Mineral Exploration. Chem. Geol. 2018, 476, 130–149. [Google Scholar] [CrossRef]
  93. Cao, G.; Chen, H.; Zhang, Y.; Sun, W.; Zhao, J.; Zhao, H.; Wang, H. Primary Controlling Factors of Apatite Trace Element Composition and Implications for Exploration in Orogenic Gold Deposits. Geochem. Geophys. Geosystems 2024, 25, e2024GC011574. [Google Scholar] [CrossRef]
  94. Salvi, S.; Fontan, F.; Monchoux, P.; Williams-Jones, A.E.; Moine, B. Hydrothermal Mobilization of High Field Strength Elements in Alkaline Igneous Systems: Evidence from the Tamazeght Complex (Morocco). Econ. Geol. 2000, 95, 559–576. [Google Scholar] [CrossRef]
  95. Samson, I.M.; Wood, S.A. The Rare Earth Elements: Behaviour in Hydrothermal Fluids and Concentration in Hydrothermal Mineral Deposits, Exclusive of Alkaline Settings. Rare-Elem. Geochem. Miner. Depos. Geol. Assoc. Can. GAC Short Course Notes 2005, 17, 269–297. [Google Scholar]
  96. Chang, Z.; Meinert, L.; Lawrence, D.; Mrozek, S.; Zhang, L. Skarns Replacing Igneous Rocks. In Proceedings of the 15th Quadrennial International Association on the Genesis of Ore Deposits, Symposium, Salta, Argentina, 28–31 August 2018; pp. 207–208. [Google Scholar]
  97. Sciuba, M.; Beaudoin, G. Texture and Trace Element Composition of Rutile in Orogenic Gold Deposits. Econ. Geol. 2021, 116, 1865–1892. [Google Scholar] [CrossRef]
  98. Kissin, S.A. Five-Element (Ni-Co-As-Ag-Bi) Veins. Geosci. Can. 1992, 19, 113–124. [Google Scholar]
  99. Scharrer, M.; Kreissl, S.; Markl, G. The Mineralogical Variability of Hydrothermal Native Element-Arsenide (Five-Element) Associations and the Role of Physicochemical and Kinetic Factors Concerning Sulfur and Arsenic. Ore Geol. Rev. 2019, 113, 103025. [Google Scholar] [CrossRef]
  100. Burke, J.S. The Origin of Polymetallic Ni-Co-As-Bi-Sb (-Ag-U) Veins in the East Arm Basin and Southern Slave Province, Northwest Territories. Master’s Thesis, Saint Mary’s University, Halifax, NS, Canada, 2019. [Google Scholar]
  101. Lecumberri-Sanchez, P.; Bouabdellah, M.; Zemri, O. Transport of Rare Earth Elements by Hydrocarbon-Bearing Brines: Implications for Ore Deposition and the Use of REEs as Fluid Source Tracers. Chem. Geol. 2018, 479, 204–215. [Google Scholar] [CrossRef]
  102. Velasco, F.; Herrero, J.M.; Suárez, S.; Yusta, I.; Alvaro, A.; Tornos, F. Supergene Features and Evolution of Gossans Capping Massive Sulphide Deposits in the Iberian Pyrite Belt. Ore Geol. Rev. 2013, 53, 181–203. [Google Scholar] [CrossRef]
  103. Charles, N.; Choulet, F.; Sizaret, S.; Chen, Y.; Barbanson, L.; Ennaciri, A.; Badra, L.; Branquet, Y. Internal Structures and Dating of Non-Sulphide Zn Deposits Using Rock Magnetism: Insights from the Moroccan High Atlas. Miner. Depos. 2016, 51, 151–175. [Google Scholar] [CrossRef]
  104. Verhaert, M.; Bernard, A.; Dekoninck, A.; Lafforgue, L.; Saddiqi, O.; Yans, J. Mineralogical and Geochemical Characterization of Supergene Cu–Pb–Zn–V Ores in the Oriental High Atlas, Morocco. Miner. Depos. 2017, 52, 1049–1068. [Google Scholar] [CrossRef]
  105. Verhaert, M.; Gautheron, C.; Dekoninck, A.; Vennemann, T.; Pinna-Jamme, R.; Mouttaqi, A.; Yans, J. Unravelling the Temporal and Chemical Evolution of a Mineralizing Fluid in Karst-Hosted Deposits: A Record from Goethite in the High Atlas Foreland (Morocco). Minerals 2022, 12, 1151. [Google Scholar] [CrossRef]
  106. Lanari, R.; Faccenna, C.; Fellin, M.G.; Essaifi, A.; Nahid, A.; Medina, F.; Youbi, N. Tectonic Evolution of the Western High Atlas of Morocco: Oblique Convergence, Reactivation, and Transpression. Tectonics 2020, 39, e2019TC005563. [Google Scholar] [CrossRef]
  107. Staude, S.; Göb, S.; Pfaff, K.; Ströbele, F.; Premo, W.R.; Markl, G. Deciphering Fluid Sources of Hydrothermal Systems: A Combined Sr-and S-Isotope Study on Barite (Schwarzwald, SW Germany). Chem. Geol. 2011, 286, 1–20. [Google Scholar] [CrossRef]
  108. Bouabdellah, M.; Levresse, G.; Boukirou, W.; Potra, A.; Samuelsen, J.R.; Caracausi, A.; Italiano, L.; Klügel, A.; Yans, J.; Castorina, F. Basin Inversion, Drifting and Hyper-Extension in the Central Atlantic Ocean and Maghrebian Tethys as Fluid Driving Mechanisms for the Superimposed Base Metal Mineralization Events in the Jbel Bou Dahar Carbonate-Hosted Pb-Zn-Ba Deposits (High Atlas, Moroc). Chem. Geol. 2024, 670, 122451. [Google Scholar] [CrossRef]
  109. Jébrak, M.; Marcoux, É. Géologie Des Ressources Minérales; Ministère des ressources naturelles et de la faune: Avenue Ouest, QC, Canada, 2008; ISBN 2551237378.
  110. Beauchamp, W.; Barazangi, M.; Demnati, A.; El Alji, M. Intracontinental Rifting and Inversion: Missour Basin and Atlas Mountains, Morocco. Am. Assoc. Pet. Geol. Bull. 1996, 80, 1459–1482. [Google Scholar]
  111. Leprêtre, R.; Missenard, Y.; Barbarand, J.; Gautheron, C.; Jouvie, I.; Saddiqi, O. Polyphased Inversions of an Intracontinental Rift: Case Study of the Marrakech High Atlas, Morocco. Tectonics 2018, 37, 818–841. [Google Scholar] [CrossRef]
  112. Stampfli, G.M.; Borel, G.D. The Transmed Transects in Space and Time: Constraints on the Paleotectonic Evolution of the Mediterranean Domain. In The Transmed Atlas. The Mediterranean Region from Crust to Mantle; Cavazza, W., Roure, F., Spakman, W., Stampfli, G.M., Ziegler, P.A., Eds.; Springer: Berlin/Heidelberg, Germany, 2004; pp. 53–80. ISBN 978-3-642-18919-7. [Google Scholar]
  113. Frizon de Lamotte, D.; Leturmy, P.; Missenard, Y.; Khomsi, S.; Ruiz, G.; Saddiqi, O.; Guillocheau, F.; Michard, A. Mesozoic and Cenozoic Vertical Movements in the Atlas System (Algeria, Morocco, Tunisia): An Overview. Tectonophysics 2009, 475, 9–28. [Google Scholar] [CrossRef]
  114. Leprêtre, R.; Missenard, Y.; Barbarand, J.; Gautheron, C.; Saddiqi, O.; Pinna-Jamme, R. Postrift History of the Eastern Central Atlantic Passive Margin: Insights from the Saharan Region of South Morocco. J. Geophys. Res. Solid Earth 2015, 120, 4645–4666. [Google Scholar] [CrossRef]
  115. Ziegler, P.A.; Dèzes, P. Crustal Evolution of Western and Central Europe. Memoirs 2006, 32, 43–56. [Google Scholar] [CrossRef]
  116. Bensalah, M.K.; Youbi, N.; Mata, J.; Madeira, J.; Martins, L.; El Hachimi, H.; Bertrand, H.; Marzoli, A.; Bellieni, G.; Doblas, M.; et al. The Jurassic-Cretaceous Basaltic Magmatism of the Oued El-Abid Syncline (High Atlas, Morocco): Physical Volcanology, Geochemistry and Geodynamic Implications. J. Afr. Earth Sci. 2013, 81, 60–81. [Google Scholar] [CrossRef]
  117. Moratti, G.; Benvenuti, M.; Santo, A.P.; Laurenzi, M.A.; Braschi, E.; Tommasini, S. New 40Ar–39Ar Dating of Lower Cretaceous Basalts at the Southern Front of the Central High Atlas, Morocco: Insights on Late Mesozoic Tectonics, Sedimentation and Magmatism. Int. J. Earth Sci. 2018, 107, 2491–2515. [Google Scholar] [CrossRef]
  118. Bouzekraoui, M.; Saadi, M.; Essalhi, M.; Karaoui, B.; Hilali, M.; Jayadi, S.; Bahaj, T. Extensional Tectonics, Structural Architecture Modeling and Geodynamic Evolution in the Cretaceous Tinghir-Errachidia-Boudenib Basin (Pre-African Trough, Morocco). J. Afr. Earth Sci. 2023, 203, 104957. [Google Scholar] [CrossRef]
Figure 1. (A) Geological map of northern Morocco showing the main structural domains. The red and white rectangles show the locations of the study deposit and the adjacent El Hammam REE-rich fluorite deposit. (B) Geological map of the Upper Moulouya metallogenic province showing the regional geology along with the locations of the historically mined Aouli, Mibladen, and Zeida Pb-Zn ± fluorite ± barite deposits (modified after [17]). (C) Representative stratigraphic column for the Aouli district, showing the main lithostratigraphic units and relative positions of known vein-type mineralization. Abbreviation: K = Cretaceous; CZ = Cenozoic.
Figure 1. (A) Geological map of northern Morocco showing the main structural domains. The red and white rectangles show the locations of the study deposit and the adjacent El Hammam REE-rich fluorite deposit. (B) Geological map of the Upper Moulouya metallogenic province showing the regional geology along with the locations of the historically mined Aouli, Mibladen, and Zeida Pb-Zn ± fluorite ± barite deposits (modified after [17]). (C) Representative stratigraphic column for the Aouli district, showing the main lithostratigraphic units and relative positions of known vein-type mineralization. Abbreviation: K = Cretaceous; CZ = Cenozoic.
Minerals 15 00669 g001
Figure 2. Longitudinal NW-SE cross section through upper Moulouya massif showing spatial distribution of historically mined Aouli, Zeida, and Mibladen Pb-Zn deposits, and their relationship to stratigraphy and major faults (modified after [28]). See Figure 1 for the position of the cross section.
Figure 2. Longitudinal NW-SE cross section through upper Moulouya massif showing spatial distribution of historically mined Aouli, Zeida, and Mibladen Pb-Zn deposits, and their relationship to stratigraphy and major faults (modified after [28]). See Figure 1 for the position of the cross section.
Minerals 15 00669 g002
Figure 3. Simplified map of the Upper Moulouya metallogenic province showing the locations of major vein ore zones and related major faults, along with the sampled sites.
Figure 3. Simplified map of the Upper Moulouya metallogenic province showing the locations of major vein ore zones and related major faults, along with the sampled sites.
Minerals 15 00669 g003
Figure 4. Representative field photographs and petrographic photomicrographs showing mode of occurrence, mineralogy, and typical ore textures and microtextures from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Panoramic view of surface exposure showing the major Henri vein cross-cutting the Paleozoic schist and the unconformably overlying Triassic salt-bearing red beds. (B) Underground mining exposure at Sidi Said deposit showing the occurrence of breccia-like orebody with clasts of schist and quartzite cemented by arsenide, sulfosalt, and sulfide ores and their supergene derivates (i.e., erythrite), all set in a gangue of quartz, fluorite, and barite. (C) Cut and polished slab of representative cockade breccia-like texture. (D) Microscopic view under cross-polarized light showing the occurrence of fine- to medium-grained flakes (up to 100 μm across) of sericite closely intergrown with hydrothermal quartz. (E) Cathodoluminescence image showing homogenous ore-related hydrothermal apatite grains fluorescing yellow under cathodoluminescence. (F) Scanning electron microscope image of ore-related hydrothermal rutile grouped in cluster of aggregated crystals intimately intergrown with apatite. Note that galena fills the open spaces between the rutile grains. (G) Scanning electron microscope image of euhedral ore-related hydrothermal adularia and apatite grains embedded in a silica matrix. Abbreviation: Ars = arsenide; Brt = barite; Qz = quartz; Fl = fluorite; Ap = apatite; Ser = Sericite; Ru = rutile; Ad = adularia; Sp = sphalerite; Gn = galena; Ccp = chalcopyrite; Slf = sulfosalt; Ery = erythrite; Sch = schist.
Figure 4. Representative field photographs and petrographic photomicrographs showing mode of occurrence, mineralogy, and typical ore textures and microtextures from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Panoramic view of surface exposure showing the major Henri vein cross-cutting the Paleozoic schist and the unconformably overlying Triassic salt-bearing red beds. (B) Underground mining exposure at Sidi Said deposit showing the occurrence of breccia-like orebody with clasts of schist and quartzite cemented by arsenide, sulfosalt, and sulfide ores and their supergene derivates (i.e., erythrite), all set in a gangue of quartz, fluorite, and barite. (C) Cut and polished slab of representative cockade breccia-like texture. (D) Microscopic view under cross-polarized light showing the occurrence of fine- to medium-grained flakes (up to 100 μm across) of sericite closely intergrown with hydrothermal quartz. (E) Cathodoluminescence image showing homogenous ore-related hydrothermal apatite grains fluorescing yellow under cathodoluminescence. (F) Scanning electron microscope image of ore-related hydrothermal rutile grouped in cluster of aggregated crystals intimately intergrown with apatite. Note that galena fills the open spaces between the rutile grains. (G) Scanning electron microscope image of euhedral ore-related hydrothermal adularia and apatite grains embedded in a silica matrix. Abbreviation: Ars = arsenide; Brt = barite; Qz = quartz; Fl = fluorite; Ap = apatite; Ser = Sericite; Ru = rutile; Ad = adularia; Sp = sphalerite; Gn = galena; Ccp = chalcopyrite; Slf = sulfosalt; Ery = erythrite; Sch = schist.
Minerals 15 00669 g004
Figure 5. Summary of the paragenetic sequence of the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. Horizontal lines indicate the relative timing of mineral formation. Dotted lines indicate uncertainties.
Figure 5. Summary of the paragenetic sequence of the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. Horizontal lines indicate the relative timing of mineral formation. Dotted lines indicate uncertainties.
Minerals 15 00669 g005
Figure 6. Representative reflected plane-polarized light microphotographs (AE), photographs of cut and polished slab (F) and hand specimen (G) along with transmitted light microphotograph (H) of typical ore textures and microtextures from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Back-scattered electron image of cataclastic skutterudite/Ni-skutterudite with the resulting fractures being filled with polybasite, galena, erythrite, and arsenolite. (B) Reflected light microphotograph showing a succession of star-like textured rammelsbergite (Rmb), safflorite/clinosafflorite (Saf/Csaf), and loellingite (Lo), all of which set in gangue of fluorite and quartz ± dolomite-ankerite. (C) Elemental maps (SEM-EDS images) showing the distribution of As, Ni, Co, and Fe in star-shaped aggregates of safflorite/clinosafflorite, rammelsbergite, and loellingite solid solutions in quartz gangue. (D) Reflected plane-polarized light microphotograph showing infillings of polybasite and Ag-Hg amalgam along with galena and sphalerite within a gangue of quartz. (E) Back-scattered electron image showing corroded quartz grains cemented by polybasite, Ag-bearing tetrahedrite, acanthite, and Ag-Hg amalgam. (F) Cut and polished slab showing the occurrence of flakes of native silver along with arsenide and sulfosalt phases as intergranular infills within fluorite. (G) Close view of hand sample showing filamentous silver “wire silver” lining a vug in quartz. (H) Microscopic view under transmitted light showing sphalerite (honey brown) intimately intergrown with arsenide and sulfosalt within a gangue of quartz, and to a lesser extent dolomite-ankerite. Abbreviation: Ac = acanthite; Ad = adularia; Ag-Hg = Ag-Hg amalgam; Ank = ankerite; Ap = apatite; Aso = arsenolite; Ccp = chalcopyrite; Dol = dolomite; Ery = erythrite; Fl = fluorite; Qz = quartz; Gn = galena; Imi = imiterite; Lo = loellingite; Plb = polybasite; Rmb = rammelsbergite; Ru = rutile; Saf/Csaf = safflorite/clinosafflorite; Ser = sericite; Skt/Nskt = skutterudite/nickelskutterudite; Sp = sphalerite; Ttr = tetrahedrite; Ars = arsenide; Slf = sulfosalt.
Figure 6. Representative reflected plane-polarized light microphotographs (AE), photographs of cut and polished slab (F) and hand specimen (G) along with transmitted light microphotograph (H) of typical ore textures and microtextures from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Back-scattered electron image of cataclastic skutterudite/Ni-skutterudite with the resulting fractures being filled with polybasite, galena, erythrite, and arsenolite. (B) Reflected light microphotograph showing a succession of star-like textured rammelsbergite (Rmb), safflorite/clinosafflorite (Saf/Csaf), and loellingite (Lo), all of which set in gangue of fluorite and quartz ± dolomite-ankerite. (C) Elemental maps (SEM-EDS images) showing the distribution of As, Ni, Co, and Fe in star-shaped aggregates of safflorite/clinosafflorite, rammelsbergite, and loellingite solid solutions in quartz gangue. (D) Reflected plane-polarized light microphotograph showing infillings of polybasite and Ag-Hg amalgam along with galena and sphalerite within a gangue of quartz. (E) Back-scattered electron image showing corroded quartz grains cemented by polybasite, Ag-bearing tetrahedrite, acanthite, and Ag-Hg amalgam. (F) Cut and polished slab showing the occurrence of flakes of native silver along with arsenide and sulfosalt phases as intergranular infills within fluorite. (G) Close view of hand sample showing filamentous silver “wire silver” lining a vug in quartz. (H) Microscopic view under transmitted light showing sphalerite (honey brown) intimately intergrown with arsenide and sulfosalt within a gangue of quartz, and to a lesser extent dolomite-ankerite. Abbreviation: Ac = acanthite; Ad = adularia; Ag-Hg = Ag-Hg amalgam; Ank = ankerite; Ap = apatite; Aso = arsenolite; Ccp = chalcopyrite; Dol = dolomite; Ery = erythrite; Fl = fluorite; Qz = quartz; Gn = galena; Imi = imiterite; Lo = loellingite; Plb = polybasite; Rmb = rammelsbergite; Ru = rutile; Saf/Csaf = safflorite/clinosafflorite; Ser = sericite; Skt/Nskt = skutterudite/nickelskutterudite; Sp = sphalerite; Ttr = tetrahedrite; Ars = arsenide; Slf = sulfosalt.
Minerals 15 00669 g006
Figure 7. Arsenide compositions analyzed by electron probe microanalysis (EPMA) from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Ternary Fe-Co-Ni diagram showing the composition of diarsenides in mol% %. (B) Binary Cu-Ag diagram showing the compositions of tetrahedrite in wt.%. Abbreviation: Lo = loellingite; Rmb = rammelsbergite; Saf/Csaf = safflorite/clinosafflorite.
Figure 7. Arsenide compositions analyzed by electron probe microanalysis (EPMA) from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Ternary Fe-Co-Ni diagram showing the composition of diarsenides in mol% %. (B) Binary Cu-Ag diagram showing the compositions of tetrahedrite in wt.%. Abbreviation: Lo = loellingite; Rmb = rammelsbergite; Saf/Csaf = safflorite/clinosafflorite.
Minerals 15 00669 g007
Figure 8. Selected Hg-Ag (A), Ag-Sb (B), Sb-As (C), and Cu-Sb (D) binary diagrams showing LA-ICP-MS trace elements compositions of sphalerite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Figure 8. Selected Hg-Ag (A), Ag-Sb (B), Sb-As (C), and Cu-Sb (D) binary diagrams showing LA-ICP-MS trace elements compositions of sphalerite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Minerals 15 00669 g008
Figure 9. Ore-related apatite and rutile compositions analyzed by LA-ICP-MS from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Chondrite-normalized REE + Y patterns of ore-related hydrothermal apatite. (B) Selected trace element contents of ore-related hydrothermal rutile. (C) Chondrite-normalized REE + Y patterns of ore-related hydrothermal rutile, the red line refers to the mean composition. Chondrite normalizing values are from [57].
Figure 9. Ore-related apatite and rutile compositions analyzed by LA-ICP-MS from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) Chondrite-normalized REE + Y patterns of ore-related hydrothermal apatite. (B) Selected trace element contents of ore-related hydrothermal rutile. (C) Chondrite-normalized REE + Y patterns of ore-related hydrothermal rutile, the red line refers to the mean composition. Chondrite normalizing values are from [57].
Minerals 15 00669 g009
Figure 10. Transmitted-light photomicrographs showing types and occurrences of fluid inclusions in ore-related fluorite and quartz from the Aouli main vein ore zones. (A) Distribution of primary fluid inclusion assemblage in growth bands of fluorite. (B) Distribution of primary FIA in growth bands of quartz. (C) Detail of a liquid-rich, two-phase primary fluid inclusion. (D) Detail of a solid bearing-fluid inclusion. Abbreviation: L = aqueous liquid; S = entrapped solid; V = vapor; Fl = fluorite; Qz = quartz.
Figure 10. Transmitted-light photomicrographs showing types and occurrences of fluid inclusions in ore-related fluorite and quartz from the Aouli main vein ore zones. (A) Distribution of primary fluid inclusion assemblage in growth bands of fluorite. (B) Distribution of primary FIA in growth bands of quartz. (C) Detail of a liquid-rich, two-phase primary fluid inclusion. (D) Detail of a solid bearing-fluid inclusion. Abbreviation: L = aqueous liquid; S = entrapped solid; V = vapor; Fl = fluorite; Qz = quartz.
Minerals 15 00669 g010
Figure 11. (A) Frequency distribution of homogenization temperatures measured from fluid inclusions in fluorite and quartz. (B) Compositions of liquid-dominated fluid inclusions plotted in the ternary H2O-NaCl-CaCl2 diagram, calculated from the melting temperature of ice and hydrohalite using the program of [50]. Abbreviation: L = aqueous liquid; S = entrapped solid; V = vapor.
Figure 11. (A) Frequency distribution of homogenization temperatures measured from fluid inclusions in fluorite and quartz. (B) Compositions of liquid-dominated fluid inclusions plotted in the ternary H2O-NaCl-CaCl2 diagram, calculated from the melting temperature of ice and hydrohalite using the program of [50]. Abbreviation: L = aqueous liquid; S = entrapped solid; V = vapor.
Minerals 15 00669 g011
Figure 12. A temperature–pressure diagram showing the isochores of the fluid inclusions from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization and estimation of the P–T trapping conditions. The blue area is the potential range of ambient temperature and pressure constrained by the intersections of the isochores and the geothermal profiles. Isochore was calculated based on the mean salt content and the Th mode using the FLUIDS 1 package [80] and the equation state of [81].
Figure 12. A temperature–pressure diagram showing the isochores of the fluid inclusions from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization and estimation of the P–T trapping conditions. The blue area is the potential range of ambient temperature and pressure constrained by the intersections of the isochores and the geothermal profiles. Isochore was calculated based on the mean salt content and the Th mode using the FLUIDS 1 package [80] and the equation state of [81].
Minerals 15 00669 g012
Figure 13. Microthermometric data from the Aouli main vein ore zones. (A) Homogenization temperature-Tmice (bulk salinity) diagram for the fluid inclusions trapped in ore-related fluorite and quartz from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) ore mineralization. Individual fluid inclusions are plotted. Two elongated fluid inclusion populations are distinguished based on their calculated bulk salinities. (B) Binary plot showing the variations of homogenization temperature vs. NaCl/(NaCl + CaCl2) ratio (in molar) diagram. Considering an individual fluid inclusion assemblage, both the melting temperature of hydrohalite and ice varied to some degree. As stressed by [8], this results in a variation in both NaCl and CaCl2 while the total salinity remains roughly constant. (C) Binary diagram showing the variation of NaCl (wt%) as a function of CaCl2 (wt%).
Figure 13. Microthermometric data from the Aouli main vein ore zones. (A) Homogenization temperature-Tmice (bulk salinity) diagram for the fluid inclusions trapped in ore-related fluorite and quartz from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) ore mineralization. Individual fluid inclusions are plotted. Two elongated fluid inclusion populations are distinguished based on their calculated bulk salinities. (B) Binary plot showing the variations of homogenization temperature vs. NaCl/(NaCl + CaCl2) ratio (in molar) diagram. Considering an individual fluid inclusion assemblage, both the melting temperature of hydrohalite and ice varied to some degree. As stressed by [8], this results in a variation in both NaCl and CaCl2 while the total salinity remains roughly constant. (C) Binary diagram showing the variation of NaCl (wt%) as a function of CaCl2 (wt%).
Minerals 15 00669 g013
Figure 14. Biplot diagrams showing LA-ICP-MS data (in ppm) for ore-related hydrothermal rutile from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) W vs. Nb/V. (B) Fe/V vs. V/Nb. (C) U vs. Th. The metamorphic-hydrothermal and magmatic-hydrothermal fields shown in diagram (A) are from [97].
Figure 14. Biplot diagrams showing LA-ICP-MS data (in ppm) for ore-related hydrothermal rutile from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization. (A) W vs. Nb/V. (B) Fe/V vs. V/Nb. (C) U vs. Th. The metamorphic-hydrothermal and magmatic-hydrothermal fields shown in diagram (A) are from [97].
Minerals 15 00669 g014
Table 1. Summary LA-ICP-MS trace element compositions of sphalerite, pyrite, and chalcopyrite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Table 1. Summary LA-ICP-MS trace element compositions of sphalerite, pyrite, and chalcopyrite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Sphalerite
Sample ID.n MnFeCoCuGaGeAsAgCdInSnSbHgPb
S4722Min0.137.46<0.01165.850.21<0.0624.8966.86987.340.000.0290.629.9832.80
Max4.393486.0022.967934.8713.6290.24976.468322.068749.1716.052.098084.063683.782680.84
Mean1.571489.74-3048.482.67-393.262009.413227.921.590.302895.02250.71974.94
Std. Dev.1.551136.90-2503.153.46-326.741948.302329.353.610.432591.99771.53803.73
S38.157Min0.1865.950.0293.080.60<0.0319.28183.3291.54<0.001<0.0175.3623.88584.60
Max3480.342452.997.942054.4343.2243.351791.0043,790.031495.930.8657.934078.821311.9112,988.14
Mean126.721024.702.76953.689.39-201.614129.06718.03--1615.84406.883239.50
Std. Dev.507.72695.011.80399.1510.21-247.046709.34287.01--975.99332.222543.81
Pyrite
CoNiCuZnAsSeAgCdSnSbHgTlPbBi
S38.124Min110.7788.39390.921.792318.10<0.165297.300.11<0.0165.5675.858.96932.11<0.001
Max204.05200.81954.2262.069769.390.5519,954.920.830.18184.29171.4891.976251.460.01
Mean144.23136.70703.7813.636305.85-11,661.360.35-139.54108.1221.691558.19-
Std. Dev.22.5630.80121.9113.702461.34-4608.320.17-29.9227.3923.641149.65-
Chalcopyrite
CoNiZnAsSeAgCdSnSbTeHgTlPbBi
S924Min<0.02<0.079.4825.00<0.5216.950.242.9440.81<0.074.160.8641.98<0.01
Max0.310.3134.30158.921.4592.921.4431.39190.440.19117.956.79246.280.14
Mean--17.5474.24-43.770.7311.6996.13-23.973.16140.64-
Std. Dev.--5.8737.82-15.960.296.8037.04-22.611.4156.17-
Abbreviation: n = number of analyses; Min = minimum; Max = maximum; Std. Dev. = Standard Deviation.
Table 2. Summary of LA-ICP-MS trace element compositions of ore-related hydrothermal apatite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Table 2. Summary of LA-ICP-MS trace element compositions of ore-related hydrothermal apatite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
MineralApatite (n = 53)
MinMaxAvg.Std. Dev.
Cl39.96656.58215.9894.34
Mn262.96618.09417.6094.46
Sr367.91899.79525.56122.93
Y930.051605.611229.51158.76
La89.05259.22166.8037.30
Ce320.83820.37560.43125.10
Pr62.74149.37103.4721.04
Nd416.86874.28624.34100.41
Sm145.48276.18203.0725.02
Eu37.0964.7950.055.32
Gd191.97338.40262.1925.17
Tb29.0050.1739.704.16
Dy177.24298.96237.0127.73
Ho34.6660.0646.065.82
Er88.75152.69115.4215.52
Tm9.9617.2813.241.82
Yb55.3994.8174.1910.17
Lu7.2512.399.521.30
∑REE1718.493440.162505.49339.80
∑REY2648.544908.393735.00404.05
∑LREE928.042103.231455.04279.24
∑MREE615.431079.51838.0981.25
∑HREE161.57277.05212.3628.69
Pb1.84873.9753.60141.43
Th1.044.261.730.50
U0.623.551.080.54
Y/Ho25.6327.7726.690.51
Sr/Y0.280.940.440.14
Ce/Ce*Chondrite0.921.051.000.03
Eu/Eu*Chondrite0.580.710.660.03
Y/Y*Chondrite0.830.910.870.02
Abbreviation: n = number of analyses; Min = minimum; Max = maximum; Avg. = Average; Std. Dev. = Standard Deviation.
Table 3. Summary of LA-ICP-MS trace elements compositions of ore-related hydrothermal rutile from the the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Table 3. Summary of LA-ICP-MS trace elements compositions of ore-related hydrothermal rutile from the the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
MineralRutile (n = 17)
MinMaxAvgStd. Dev.
Sc64.60240.15144.8953.92
V181.321360.42552.31316.47
Cr<26.96--
Fe2029.7575,592.559227.6317,586.14
Sr10.571236.32167.73301.95
Y5.43314.2872.9983.44
Zr13.034600.56680.221084.11
Nb207.44514.80343.0591.97
Mo2.51663.6473.10171.76
Sn104.25654.02354.39163.87
Sb5.771091.10169.62315.94
Ba21.65586.37169.85171.71
La0.72154.8029.2643.97
Ce2.44426.5679.43120.64
Pr0.3654.7910.0715.41
Nd1.72211.6939.9859.91
Sm0.6044.449.2112.71
Eu0.2122.764.426.21
Gd0.9443.929.7212.29
Tb0.199.562.062.61
Dy1.2083.6317.3622.54
Ho0.2322.324.466.03
Er0.7889.2017.6823.67
Tm0.1214.642.923.91
Yb0.87109.8522.4729.31
Lu0.1215.853.404.30
∑REE10.901303.99252.46361.57
∑REY16.331618.28325.45443.51
∑LREE5.24847.84158.75239.85
∑MREE3.75226.6247.2462.27
∑HREE1.90229.5346.4760.99
Hf3.7993.5225.8320.77
Ta4.0297.4126.9021.75
W821.3710,247.184664.372414.41
Pb101.1242,974.237694.6114,082.39
Th<0.0013.27--
U0.30199.1831.8053.66
Nb/Ta4.4190.3620.7719.58
Zr/Hf0.8849.1919.9512.40
Y/Nb0.021.060.230.28
Fe/V4.09113.1715.5525.49
Y/Ho14.0830.3719.865.13
Nb/V0.201.700.830.50
Ce/Ce*Chondrite1.061.161.120.03
Eu/Eu*Chondrite0.842.011.320.26
Y/Y*Chondrite0.531.320.720.22
Abbreviation: n = number of analyses; Min = minimum; Max = maximum; Avg. = Average; Std. Dev. = Standard Deviation.
Table 4. Summary of microthermometric data of fluid inclusions hosted in ore-related quartz and fluorite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
Table 4. Summary of microthermometric data of fluid inclusions hosted in ore-related quartz and fluorite from the Aouli Ag-(Ni-Co-Sb-As-Hg ± Bi) vein ore shoot mineralization.
MineralMicrothermometryCalculated Salinity
Th (°C)Tice-melting (°C)Tm(HH) (°C)Salinity NaCl/(NaCl + CaCl2)
n Min Max Avg. Std. Dev. n Min Max Avg. Std. Dev. n Min Max Avg. Std. Dev. n Min Max Mean Std. Dev. n Min Max Avg. Std. Dev.
Ore-related quartz 2748.7168.1128.529.832−20.6−10.5−14.82.022−26.2−21.7−23.81.33214.522.818.41.6220.61.00.80.1
Ore-related fluorite 17490.9141.1122.18.4174−22.0−8.2−16.03.2151−32.1−18.6−23.61.817412.023.719.32.61460.41.00.80.1
Abbreviation: n = number of analyses; Min = minimum; Max = maximum; Avg. = Average; Std. Dev. = Standard Deviation.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zaid, K.; Bouabdellah, M.; Levresse, G.; Idbaroud, M.; Melchiorre, E.; Mathur, R.; Jébrak, M.; Potra, A.; Yans, J.; Frenzel, M.; et al. Mineralogy and Fluid Inclusion Constraints on the Genesis of the Recently Discovered Ag-(Ni-Co-Sb-As-Hg ± Bi) Vein Ore Shoot Mineralization in the Aouli Pb-Zn District (Upper Moulouya, Morocco). Minerals 2025, 15, 669. https://doi.org/10.3390/min15070669

AMA Style

Zaid K, Bouabdellah M, Levresse G, Idbaroud M, Melchiorre E, Mathur R, Jébrak M, Potra A, Yans J, Frenzel M, et al. Mineralogy and Fluid Inclusion Constraints on the Genesis of the Recently Discovered Ag-(Ni-Co-Sb-As-Hg ± Bi) Vein Ore Shoot Mineralization in the Aouli Pb-Zn District (Upper Moulouya, Morocco). Minerals. 2025; 15(7):669. https://doi.org/10.3390/min15070669

Chicago/Turabian Style

Zaid, Khadra, Mohammed Bouabdellah, Gilles Levresse, Mohamed Idbaroud, Erik Melchiorre, Ryan Mathur, Michel Jébrak, Adriana Potra, Johan Yans, Max Frenzel, and et al. 2025. "Mineralogy and Fluid Inclusion Constraints on the Genesis of the Recently Discovered Ag-(Ni-Co-Sb-As-Hg ± Bi) Vein Ore Shoot Mineralization in the Aouli Pb-Zn District (Upper Moulouya, Morocco)" Minerals 15, no. 7: 669. https://doi.org/10.3390/min15070669

APA Style

Zaid, K., Bouabdellah, M., Levresse, G., Idbaroud, M., Melchiorre, E., Mathur, R., Jébrak, M., Potra, A., Yans, J., Frenzel, M., Schijndel, V. v., Benaissi, L., & Belkacim, S. (2025). Mineralogy and Fluid Inclusion Constraints on the Genesis of the Recently Discovered Ag-(Ni-Co-Sb-As-Hg ± Bi) Vein Ore Shoot Mineralization in the Aouli Pb-Zn District (Upper Moulouya, Morocco). Minerals, 15(7), 669. https://doi.org/10.3390/min15070669

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Article metric data becomes available approximately 24 hours after publication online.
Back to TopTop