Next Article in Journal
Removal of Chloride Ions from Mg(OH)2 Using NaOH with Ethanol
Previous Article in Journal
Lower Cretaceous Carbonate Sequences in the Northwestern Persian Gulf Basin: A Response to the Combined Effects of Tectonic Activity and Global Sea-Level Changes
Previous Article in Special Issue
Platinum Group Minerals Associated with Nickel-Bearing Sulfides from the Jatobá Iron Oxide-Copper-Gold Deposit, Carajás Domain, Brazil
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metasomatic Mineral Systems with IOA, IOCG, and Affiliated Critical and Precious Metal Deposits: A Review from a Field Geology Perspective

by
Louise Corriveau
1,* and
Jean-François Montreuil
2
1
Natural Resources Canada, Geological Survey of Canada, 490 Rue de la Couronne, Québec, QC G1K 9A9, Canada
2
Red Pine Exploration Inc., 145 Wellington Street West, Suite 1001, Toronto, ON M5J 1H8, Canada
*
Author to whom correspondence should be addressed.
Minerals 2025, 15(4), 365; https://doi.org/10.3390/min15040365
Submission received: 3 February 2025 / Revised: 20 March 2025 / Accepted: 25 March 2025 / Published: 31 March 2025

Abstract

:
Worldwide, a growing list of critical (Bi, Co, Cu, F, Fe, Mo, Ni, P, PGE, REE, W, U, and Zn) and precious metal (Ag and Au) resources have been identified in mineral systems forming Fe-oxide-copper-gold (IOCG) deposits; Fe-oxide-apatite (IOA); Fe-sulfide Cu-Au (ISCG); and affiliated W skarn; Fe-rich Au-Co-Bi or Ni; albitite-hosted U or Au ± Co; and five-element (Ag, As, Co, Ni, and U) vein deposits. This paper frames the genesis of this metallogenic diversity by defining the Metasomatic Iron and Alkali-Calcic (MIAC) mineral system and classifying its spectrum of Fe-rich-to-Fe-poor and alkali-calcic deposits. The metasomatic footprint of MIAC systems consists of six main alteration facies, each recording a distinct stage of mineralization as systems have evolved. The fluid flow pathways and the thermal and chemical gradients inferred from the space–time distribution of the alteration facies within a system are best explained by the ascent and lateral propagation of a voluminous hypersaline fluid plume. The primary fluid plume evolves, chemically and physically, as metasomatism progresses and through periodic ingresses of secondary fluids into the plume. Exploration strategies can take advantage of the predictability and the expanded range of exploration targets that the MIAC system framework offers, the building blocks of which are the alteration facies as mappable prospectivity criteria for the facies-specific critical and precious metal deposits the systems generate. Global case studies demonstrate that these criteria are applicable to MIAC systems worldwide.

1. Introduction

Metasomatic Iron and Alkali-Calcic (MIAC) systems can form Iron-Oxide Copper-Gold (IOCG), Fe, P, or rare-earth (REE) Iron Oxide-Apatite (IOA), as well as affiliated critical and precious metal deposits. The affiliated deposits include, but are not limited to, Iron Sulfide Copper-Gold (ISCG); Fe, W, and polymetallic Pb-Zn skarn; Fe-rich Ni and Fe-rich Au-Co-Bi; albitite-hosted U or Au ± Co; Fe-rich Au; Fe-poor Mo; Fe-poor Cu; and five-element (U, Ag, Ni, Co, As, ±Cu, and ±Bi) vein deposits (Table A1, Table A2 and Table A3). Hofstra et al. [1] have classified their deposit environment as regional metasomatic instead of the more common sedimentary, magmatic, and metamorphic-related environments. MIAC systems form along fluid-flow pathways through the upper crust, irrespective of the sedimentary, volcano–plutonic (felsic, intermediate, mafic, and ultramafic), and metamorphic host rocks (see the representative host geological environments in Table A1, Table A2 and Table A3) [1,2]. The mineral resource estimates of the deposits include the critical raw materials Bi, Co, Cu, F, Fe, Mo, Nb, Ni, P, Pd, and Pt; and the rare earths (REE), U, Ti (minor), V (minor), W, and Zn (Table A1, Table A2 and Table A3) [2,3,4,5]. The additional base and precious metal resources consist of Ag, Au, Pb, and Re. Potential by-products comprise In, Sb, Sn, Ta, and Te [3,6,7]. The district-scale metallogenic diversity of MIAC systems has been globally observed, as shown by the representative mineral resources of the deposits in Table A1, Table A2 and Table A3.
Prior to the recognition of the MIAC system (see [7] and Section 3), genetic models for IOCG deposits provided schematic cross sections of the ore and hydrothermal alteration envelopes in both magmatic and non-magmatic settings for up to ~5 km in extent [2,8]. The linkages among the extraordinary range of IOCG deposits, and the realm of Fe oxide-rich-to-Fe oxide-poor deposit types within IOCG districts have remained contentious [2,9]. The problem was most acute in poorly exposed districts, where repeated regional-scale events (e.g., orogenic or magmatic) and metal remobilization hampered the recognition of the primary metasomatic affiliations among deposits. To resolve the apparent lack of linkages, the IOCG spectrum was dramatically reduced, while the previously considered affiliated deposits were re-interpreted as formed in distinct geological environments (cf. [9]). Yet, the results from field observations, geochemical investigations, and constraints on the timing of mineralization events required a more holistic mineral system approach to interpret the observed variety of the alteration and mineralization types [4,5,7,10,11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29].
The integration of geological, geochemical, mineralogical, tectonic, and geophysical research on Canadian and global case studies has elucidated the apparent disparity in the metal associations and deposit types within MIAC systems [6,7,21,22,23,24,25,26,27,28,29,30]. Systematic and predictable metal speciation takes place through six main stages during the metasomatic evolution of the system [7]. Each stage is constrained by the formation of distinct alteration facies, each with a diagnostic range of composition to which a distinct suite of deposit types is associated [6,7,20,21,22,23,28]. The suite of metasomatic reactions is inferred to be driven by a compositionally evolving plume of hypersaline fluids derived from different sources (e.g., magmatic, basinal, and formational fluids and salt melts) depending on the host geological environment (see Section 5.1). At the scale of entire districts, in areas representing centers of maximum fluid circulation, a suite of metasomatic reactions thoroughly transforms the textures, the composition, and the mineralogy of all rock types encountered along the flow path into new rock types, i.e., metasomatites. The scientific breakthroughs on the significance of salt melts in MIAC systems help to explain this extreme reactivity of fluid plumes [31,32].
The characterization of MIAC systems in terms of alteration facies by Corriveau and colleagues was shown to be applicable globally, including in the Olympic Cu-Au Province of Australia ([13] and in [6,7,10,18,20,21,22,23,24,25,26,33]). We, here, adopt the Skirrow et al. [34] mappable criteria for IOCG deposits and expand, with a strong focus on the field geology of alteration facies, their capabilities to help interpret the mineral potential for the range of deposit types within MIAC systems. To achieve that, this review paper (1) strengthens the early definition and the concept of a MIAC system [7,22]; (2) refines the classification of MIAC-related mineral deposit types and deposit classes [1,2,5,22,35]; (3) synthesizes the shared geological attributes and space–time relationships of their alteration facies as part of a mineral system as per Wyborn et al. [36]; and, then, (4) adapts the ore deposit model for the Fe-oxide and alkali-calcic deposits (IOA and IOCG) of Corriveau et al. [7] to the broader MIAC system. The energy driver and the fluid and metal sources of the mineral systems are also discussed from a field geology perspective.

2. Case Studies

This paper draws on the field examinations and characterizations of MIAC systems with an alteration facies approach with or by colleagues worldwide (Figure 1), e.g., Canada; the Central Andes and the El Laco deposit in Chili; the Cloncurry district and the Olympic Cu-Au Province in Australia; the Norrbotten district in Sweden; the Middle–Lower Yangtze River Metallogenic Belt and the Kangdian district in China; and the YangYang deposit in the Republic of Korea (Figure 1; see Acknowledgment Section) [10,13,19,20,21,22,23,24,25,26,37,38,39]. Additional examples were drawn from key synthesis [1,2,3,4,5,8,9,33,35,40,41,42,43], special volumes [20,44], books [4,13,39], and other publications. The tables in Section 3.4 summarize the published field relationships among the alteration facies and mineralization used for this review paper.
In Canada, the Great Bear magmatic zone (e.g., the NICO, Sue-Dianne, and Terra deposits), the East Arm Basin, and the Central Mineral Belt provide insights on MIAC systems in volcano–plutonic belts [6,7,8,9,10,11,12,13,14,15,16,17,18,26]. The Romanet Horst (e.g., Delhi Pacific prospect), the Wanapitei district (e.g., Scadding deposit), and the Temagami area (e.g., Belfast prospect) are examples of MIAC systems within sedimentary basins with mafic rocks that predate the MIAC systems but where coeval regional-scale magmatism is not in the immediate vicinity of the districts [22,49,50]. The Bondy gneiss complex and the Josette deposit (Table A2) are metamorphosed examples [6,19,51]. Hundreds of field geology photographs and scans of rock slabs and cobaltinitrate-stained rock slabs have been published for these case studies. The photographs and scans are part of the extensive collections of Natural Resources Canada (e.g., 18,000 for the Great Bear magmatic zone, Canada, most of which are unpublished to date).
The Paleoproterozoic Great Bear magmatic zone in Canada is regarded as the best exposed MIAC province in the world. It consists of series of MIAC systems that were formed within volcanic centers and their subvolcanic intrusions, as well as systems in an older sedimentary basin and a metamorphic basement (Section 4.1). The glacially polished outcrops and the phaneritic grain size and distinctive patina of the metasomatic minerals due to minor micron-to-one-millimeter-scale weathering of outcrop surfaces facilitate the identification of mineral assemblages in the field [11]. The space and time distribution of the lithological units, alteration facies, and mineralization zones have been documented along semi-continuously exposed tilted or differentially uplifted MIAC systems with depth profiles reaching 2-to-3 km [10,11,12,14,15,16,17,18,52,53,54,55,56,57,58,59,60]. A lack of regional metamorphism; the abundance of pre-, syn-, and post-MIAC intrusive bodies; the extensive U-Pb geochronological data on zircon constraining the timing of geological markers; and the 3D exposures from basement and sub-volcanic intrusions to epithermal caps and underlying sedimentary basins provide constraints on the timing of MIAC activity and on the depth to paleosurface profiles of alteration zonation in MIAC systems [16,17,27,61,62].

3. Definitions

3.1. Alteration, Alteration Facies, and Metasomatism

For clarity, the word ‘alteration’ is used when describing specific products of fluid/rock reactions (alteration facies), while acknowledging that the alteration facies are the result of metasomatic reactions, as it relates better to previous work. Alteration facies are a regular set of metasomatic mineral parageneses developed under similar physicochemical conditions, the combination of which covers the entire petrogenetic process associated with the development of the system (see Section 3.4) [11,15]. Consequently, alteration facies are not simply a rock unit with specific physical (and, in this case, also) chemical features, as each facies is also a tangible record of a distinct stage in the evolution of the system. We, herein, use the alteration facies to track down the evolution of the MIAC system.
Alteration is also used as a process term to refer to weaker chemical changes following fluid/rock reactions for the late alteration types that are overprinting MIAC systems and that are not systematically leading to the formation of a new rock type because of strong metasomatism. Metasomatism or metasomatic are used herein to describe fluid–rock interactions that lead to the intense-to-megascopically complete mineralogical and chemical re-equilibration of the host rocks into metasomatites that are typically associated with the high-temperature alteration facies of MIAC systems.

3.2. Definition of MIAC Systems

From here on, we use the original definition of mineral systems of Wyborn et al. [36] for the geologically mappable footprint that records the regional distribution of ore deposition factors and the essential elements that exploration should focus on. Building on this definition, the terminology to name and describe mineral systems with IOCG, IOA, and affiliated deposits is expanded here to the Metasomatic Iron and Alkali-Calcic (MIAC) mineral system from the iron oxide and alkali deposit types of Porter [5] and the iron oxide and alkali-calcic alteration systems of Corriveau et al. [7,15]. This new umbrella name is needed as Fe silicate-, Fe sulfide-, Fe sulfarsenide-, and Fe carbonate-rich deposits and Fe-poor alkali-calcic deposits are increasingly recognized as part of the host mineral system that form polymetallic Fe-oxide deposits [7,22,23,33]. The MIAC system is defined as metasomatic to account for the thorough fluid-driven transformations of host rocks into metasomatites at the deposit and the regional scales [63]. Transformations are mineralogical, chemical, textural, and structural in nature, and they involve the pervasive dissolution of original minerals and the precipitation of new—metasomatic—mineral parageneses forming completely new rock types [7,63,64]. In the process, a wide range of new physical properties are acquired by the metasomatites [30].
The MIAC system is characterized by Fe-rich alteration facies that are typically silica-deficient and associated with Fe-rich mineralization parageneses. Both ferric and ferrous Fe can be stable, and Fe can be concentrated in oxides, silicates, sulfides, sulfarsenides, or carbonates (Section 3.4) [2,3,4,5,7,8,9,10]. The system is classified as alkali and calcic due to the alteration facies rich in alkali metals (K as biotite, K-feldspar, and the white micas of the paragonite-muscovite-phengite suite; Na as albite, oligoclase, scapolite, and the more rare Na-pyroxene or Na-amphibole) and those rich in alkaline earth metals (Ca as amphibole, calcite, dolomite and ankerite, anhydrite, clinopyroxene, epidote, fluorite, garnet and scheelite; Mg as amphibole, chlorite, clinopyroxene, dolomite, phengitic mica and talc; and Ba as barite) [2,3,4,5,8,10,11,35,65]. The outcomes of alkali-calcic metasomatism include the regional corridors of albitite that are diagnostic of these systems and the extensive zones of Na- and K-bearing alteration [2,5,8,28,50,52,54,65,66,67,68,69,70].

3.3. Classification for the Mineral Deposits in MIAC Systems

Two main groups of mineral deposits are defined to organize and group the broad range of mineral deposit types in MIAC systems: Metasomatic Iron (MI) and Metasomatic Alkali-Calcic deposits. Each group is subdivided into classes using the main commodities of economic interest of the deposit, acknowledging that most mineral deposits are polymetallic (Table A1, Table A2 and Table A3).
MI deposits include the deposit classes in which Fe-rich minerals (i.e., Fe oxides, Fe silicates, Fe carbonates, Fe sulfides and Fe sulfarsenides) prevail in the mineralization paragenesis (Table A1, Table A2 and Table A3) [33,35,41,70,71,72,73,74,75,76]. The alteration facies associated with each class of MI deposits are listed in Section 4. These deposits can also have an alkali-calcic component (e.g., Fe-rich silicates like amphibole, biotite, or clinopyroxene; and Fe-poor minerals like apatite, calcite, K-feldspar, or white micas). The MI-W, MI-Fe, MI-Ni, MI-REE, MI-Co, MI-Cu, MI-U, MI-Au, and, possibly, MI-Ag classes are currently documented, each with their own suite of deposit types (Table A1, Table A2 and Table A3; Section 3.4; Figure 2). For example, deposits of the MI-Cu class include the IOCG and ISCG deposit types. The strict definition of IOCG deposit by Williams et al. [2] is retained here to describe that deposit type.
The MI deposit types rich in Fe-oxides (magnetite and hematite) form the largest deposits in global MIAC districts [2,3,9,33,71], including those in which REE from the primary endowments are remobilized into REE mineralization [6,77,78]. Examples include IOCG; IOA (with Fe, Ni, P, REE, or V resources); Fe (magnetite) skarn; MI-Co ± (Au, Bi, and Cu); MI-Ni; and albitite-hosted MI-U deposits (Table A1, Table A2 and Table A3; Section 3.4). In MI deposits, examples of Fe silicate-rich parageneses (amphibole, biotite, clinopyroxene, chlorite, epidote, and garnet) include those of MIAC-related W and Pb-Zn skarn and some of the MI-Co deposits (e.g., Idaho Cobalt Belt [41]). Fe sulfides (pyrrhotite and pyrite) parageneses (e.g., ISCG deposits), Fe sulfarsenide (arsenopyrite) parageneses (e.g. Fe-rich Au-Co-Bi deposits) or Fe carbonates (ankerite and siderite) parageneses also occur [33,79,80].
MAC deposits comprise the deposit classes in which Fe-poor alkali-calcic minerals (e.g., carbonates like dolomite, Fe-dolomite, or calcite; silicates like Mg-chlorite, Mg-amphibole, talc, K-feldspar, albite, and white micas) prevail in mineralization parageneses [7,22,35,81]. MAC deposits can be related to a suite of the alkali-calcic alteration facies formed coevally with the Fe and alkali-calcic alteration facies [81]. Mineralization is often associated with quartz. The specific classes of MAC deposits are defined by their main commodities, with MAC-Au, MAC-Cu, MAC-Mo, and MAC-U deposits being currently identified (Figure 2).
Figure 2. Relative timing relationships of alteration facies, from 1 to 6, and transitional ones (1–2, 2–1, 2–3). Red to blue arrow: the fluid plume evolving in composition and physical conditions as it precipitates the regular sequence of alteration facies. The inflow fluid at each facies is the outflow fluid of the previous facies (primary fluid plume/inflow fluid + elements dissolved from hosts − elements precipitated = evolving fluid plume). The ingress of magmatic heat and external fluids disrupt the physicochemical conditions of the primary plume. The downward red arrow represents the ingress of magmatic heat and high-temperature fluids, and the upward yellow arrow represents the disruption that occurs due to the ingress of low-temperature fluids (see Section 5 for additional information). Note that heat ingress does not produce additional Facies 1 albitite. The mineral abbreviations are according to Warr [82], with additions for iron oxides (IO), oxides (OX), sulfides (Sulf), silicates (Sia), arsenides (As), sulfarsenides (AsS), and minerals (min.). The Great Bear magmatic zone (GBMZ), the Southern Breccia (S), the Chalco prospect (C) and NICO deposit (N) of the Lou system (Lou), and the Port Radium-Echo Bay (PR-EB) system occur in Canada (Figure 1) [15,17,56,59,60,61,62]. The Olympic Dam (O), Manxman A1 (M), Island Dam (I), Oak Dam East (D east), Oak Dam West (D west), and Bellatrix (B) are from Australia [6,10,23,71,72]. Deposits and prospects are labeled with their facies and metal associations, and the continuous to intermittent path is based on the facies observed (additional information in Section 3.4). Additional abbreviations include the following: giant quartz veins (G), hematite-group (Hem-gp), high temperature (HT), including (inc.), low temperature (LT), metasomatic alkali-calcic (MAC), metasomatic iron (MI), and magnetite-group (Mag-gp). The figure is extensively modified and updated from Fig. 12 of Corriveau et al. [7] and the published components of the original figure reproduced under GAC®’s Fair Dealing/Fair Use policy.
Figure 2. Relative timing relationships of alteration facies, from 1 to 6, and transitional ones (1–2, 2–1, 2–3). Red to blue arrow: the fluid plume evolving in composition and physical conditions as it precipitates the regular sequence of alteration facies. The inflow fluid at each facies is the outflow fluid of the previous facies (primary fluid plume/inflow fluid + elements dissolved from hosts − elements precipitated = evolving fluid plume). The ingress of magmatic heat and external fluids disrupt the physicochemical conditions of the primary plume. The downward red arrow represents the ingress of magmatic heat and high-temperature fluids, and the upward yellow arrow represents the disruption that occurs due to the ingress of low-temperature fluids (see Section 5 for additional information). Note that heat ingress does not produce additional Facies 1 albitite. The mineral abbreviations are according to Warr [82], with additions for iron oxides (IO), oxides (OX), sulfides (Sulf), silicates (Sia), arsenides (As), sulfarsenides (AsS), and minerals (min.). The Great Bear magmatic zone (GBMZ), the Southern Breccia (S), the Chalco prospect (C) and NICO deposit (N) of the Lou system (Lou), and the Port Radium-Echo Bay (PR-EB) system occur in Canada (Figure 1) [15,17,56,59,60,61,62]. The Olympic Dam (O), Manxman A1 (M), Island Dam (I), Oak Dam East (D east), Oak Dam West (D west), and Bellatrix (B) are from Australia [6,10,23,71,72]. Deposits and prospects are labeled with their facies and metal associations, and the continuous to intermittent path is based on the facies observed (additional information in Section 3.4). Additional abbreviations include the following: giant quartz veins (G), hematite-group (Hem-gp), high temperature (HT), including (inc.), low temperature (LT), metasomatic alkali-calcic (MAC), metasomatic iron (MI), and magnetite-group (Mag-gp). The figure is extensively modified and updated from Fig. 12 of Corriveau et al. [7] and the published components of the original figure reproduced under GAC®’s Fair Dealing/Fair Use policy.
Minerals 15 00365 g002

3.4. Review of Alteration Facies and Mineralization in MIAC Systems

The definition of the alteration facies that form MIAC systems [11,15] follow the definition of the metasomatic facies of Zharikov et al. [83] for the IUGS subcommission on the nomenclature of metasomatic rocks. The alteration facies are differentiated by the appearance or disappearance of specific suites of mineral assemblages, which reflect variations in one (or more) of the physicochemical parameters of the fluid-driven metasomatic reactions that formed the system (Figure 2; Section 3.3). The distinct physicochemical conditions that led to each alteration facies also led to the facies-specific deposit types as illustrated in this section (see also Figure 2).
In mature MIAC systems, metasomatism typically sequentially forms six main alteration facies, with high temperatures (HT) reaching a peak of up to about 1000 °C and low temperatures (LT) reaching about 150 °C (Figure 2; Section 5.1.1; [15] and the references therein; and [31,32]). The systems are globally going through (1) a heating stage with significant leaching of metals and volatiles, (2) a stage where iron and alkali-calcic alteration facies are associated with MI deposits, (3) a stage where alkali-calcic alteration facies are associated with MAC deposits, and (4) a stage where alteration facies are developed under epithermal conditions [55,84] or where mineralized veins are form through remobilization from or are superimposed on pre-existing MIAC systems [21,60,85].
The heating stage of the system produces barren Facies 1a Na, Facies 1b Na-Ca, and Facies 1c skarn (Table 1). At this stage, host rock elements, including metals and volatiles, are extensively leached and transferred to the fluid plume. In silicate rocks, mineral assemblages record a significant addition of Na to form albitite and other regionally extensive sodic alteration (Facies 1a), and lesser enrichment in Na, which is where Ca remains (Facies 1b) (Table 1). Carbonate rocks are transformed to skarn (Facies 1c) at high temperatures (above ~450 °C) [6] and can be albitized at lower temperatures [49], whereas the significant addition of Na is largely partitioned into silicate lithologies. Onset of Fe enrichment starts at Facies 1–2 HT Na-Ca-Fe (Table 1).
Alteration facies associated with a significant increase of Fe content and the formation of MI deposits are reflected by a prevalence of Fe-rich minerals in the alteration assemblages (through addition of Fe and leaching of other elements). For example, MI-Fe magnetite skarn and IOA deposits are associated with Facies 2–1 Fe-rich skarn and Facies 2 HT Ca-Fe (Figure 3A–D). Prior to the formation of MI-Fe deposits, the Fe enrichment observed in Facies 2–1 Fe-rich skarn can be associated with W mineralization (Table 2). In MIAC systems formed in geological environments with abundant mafic-to-ultramafic rocks, Ni mineralization can be associated with Facies 2 HT Ca-Fe with and without the development in the same MIAC system of Facies 2–3 HT Ca-K-Fe and Facies 3 HT K-Fe (Table 2; Figure 3E). Mineralization in REE is associated with the HT Ca-Fe facies in systems that evolved to form the HT K-Fe alteration facies, such as those observed at the Josette IOA-REE deposit (Table 2 and Table A3) and in the zones of IOA mineralization at the past-producing Terra Ag mine (Table 2; Figure 3D). Facies 2 is thus associated with deposits of the MI-Fe, MI-Ni, and MI-REE classes. Polymetallic Co mineralization with Fe oxide or Fe silicate parageneses with and without Fe sulfarsenides is associated with Facies 2–3 HT Ca-K-Fe and biotite-rich Facies 3 HT K-Fe, forming deposits of the MI-Co class (Figure 3F).
Polymetallic Cu mineralization of the MI-Cu class is primarily associated with Facies 3 HT K-Fe and Facies 5a LT (K, Ba, Ca, Mg, and Na)-(CO2, F, and H+)-Fe, whereas the Facies 4 formed at the transition from magnetite to hematite is commonly barren (Table 3, Table 4 and Table 5). Cu mineralization in MI-Cu deposits can be associated with abundant Fe oxides (e.g., IOCG; Figure 4A,E), Fe silicates (Figure 4B,C), or Fe sulfides (e.g., ISCG; Figure 4D). In some MIAC systems, Facies 3 and Facies 5a are also associated with U mineralization (Figure 4F), which can form deposits of the MI-U class [28]. Facies 5b LT (Ca, Mg, K, Na)-Si-Fe, overlapping with the later stage of the development of Facies 5a, can be associated with Au mineralization (Figure 4G) without significant Cu additions that can form deposits of the MI-Au class.
Facies 5c K/CO2 and Facies 5d Si/CO2 represent the alkalic-calcic alteration facies poor in Fe that are largely formed after the Fe-rich alteration facies and are associated with the diversified mineralization types associated with the formation of MAC deposits (Figure 5A–C; Table 6). The polymetallic Cu mineralization hosted in carbonate–quartz veins or in pervasive zones of silicification associated with K-feldspar replacement is associated with the Fe-poor Facies 5c (Figure 2) (Figure 5B,C). Facies 5d (Figure 2) is associated with albitite-hosted U and albitite-hosted Au and/or Co deposit types (Figure 5A).
As MIAC systems waned at Facies 6, aluminous or quartz-rich alteration were formed under epithermal conditions (Table 7) [7,15,55,71,84]. Many stages of the hydrothermal vein-, breccia-, or shear-hosted mineralization with localized alteration postdated the host MIAC systems. The metal associations are compatible with the remobilization of the primary endowment of the host MIAC system into mineralization zones hosted in discrete veins or stockworks that can reach very high grades. Repeated metal remobilization and addition during regional-scale and post-MIAC events are well documented worldwide [7,21,85].
Most of the MIAC systems exhibited, although at variable levels of maturity and intensity, the full suite of alteration facies. However, depending on local conditions and possible disruptions in the evolution of the fluid plume, some MIAC systems may not exhibit all alteration facies. This could be caused, for example, by (1) changes in the fluid circulation pathways due to active deformation; (2) the absence of rejuvenation of the voluminous fluid plume by external (e.g., basinal, magmatic, and meteoric) fluids during its evolution; (3) the distance between the MIAC system and its primary heat sources, or the lack of extensive magmatism as additional heat sources; or (4) the faulting and telescoping of albitite corridors to the field of LT alteration facies (see discussion in Section 5). In addition, alteration facies can remain non-documented within a system.
For example, extensive and intense Facies 2 HT Ca-Fe alteration and extensive weak overprints of IOCG mineralized Facies 5 LT K-Fe have been reported for the Oak Dam East deposit in Australia [132]. Yet, extensive and intense Facies 3 and 5 HT and LT K-Fe alteration were missing. Based on the typical evolution of MIAC systems, a probable cause was a knowledge gap. This gap has now been filled with the discovery of the giant Oak Dam West IOCG deposit (Table A1). The example of Oak Dam West illustrates that gaps in the sequence of alteration facies in the mapping or drilling record could be important data gaps that need to be filled in. Exploration and mapping strategies can be oriented to search for the missing alteration facies and the zones of mineralization they could host.

4. Geo-Environment of MIAC Systems

4.1. Geological Context

MIAC mineral systems and their associated mineral deposits are developed within a wide range of geological environments and host rocks that include the following: (1) coeval volcanic centers; (2) above, within, or adjacent to coeval sub-volcanic intrusions or more isolated alkaline intrusions, such as carbonatites; (3) within sedimentary basins and volcano-sedimentary basins that predate the MIAC event by a few m.y. or much older ones; (4) pre-MIAC intrusive suites a few m.y. or much older, as well as metamorphic basements; and (5) major and coeval fault zones and their splays [2,5,17,43,62,138,139]. Individual MIAC systems generally have a district-scale footprint that is typically 35 km in length, 15 km in width, and 10 km in depth, with interlinked and extensive zones of intense metasomatism [17,52,55,57,140,141].
Well-exposed MIAC provinces occur in continental magmatic arcs, such as those of the Canadian Great Bear magmatic zone and Central Mineral Belt, the South American Central Andes, the USA southeast Missouri district, the Scandinavian Norrbotten district, the Chinese Middle–Lower Yangtze River Metallogenic Belt, and the Iranian Bafq district [2,3,4,5,6,7,13,25,113,142,143]. The systems are typically distributed along or in proximity to transcrustal fault systems within intra-continental arc basins (e.g., the Central Andes of Peru and Chile [40,143], and the Middle–Lower Yangtze River Metallogenic Belt in China [144]), or in a back-arc setting (e.g., southeast Missouri district, USA [145]). The array of MIAC centers forms metallogenic provinces that can be traced for up to 2000 km in Mesozoic provinces, e.g., the Central Andes [40,141,146]. Archean and Proterozoic provinces are traced continuously for up to 600 km, but tectonic reconstructions link coeval Proterozoic metallogenic provinces into more extensive belts [5,12,56,95,139,147,148].
Systems that form within large sedimentary basins are typically interlinked by extensive fault zones, which are commonly related to the re-activation of rift-related faults [149], and they either have sparse or distant coeval magmatism, such as in the Canadian Romanet Horst (Québec) and Wanapitei district (Ontario); and the Chinese Kangdian district [22,49,50,64,100]. Other MIAC systems occur along fault systems developed in supracrustal rocks among batholiths (e.g., Mount Isa Province in Australia), or they are spatially decoupled from potentially coeval batholiths (e.g., Wanapitei district in Ontario and the Romanet Horst and other systems easterly of the Labrador Trough in Québec) [49,50,66,67,89,100]. Some of the systems developed during orogenesis or were reworked during orogenesis, as illustrated by the Josette REE deposit in Québec [51,116], the Adirondack district in the United States [150], the Tennant Creek system in Australia [33], and some zones of IOCG mineralization in the Norrbotten district in Sweden [21].
Unexposed to poorly exposed MIAC systems and their IOCG and IOA deposits can include giant mining districts, such as the Olympic Cu-Au Province in Australia, the Carajás Mineral Province in Brazil, and the Norrbotten district in Sweden [24,33,42,149,150,151,152,153,154]. Once formed, MIAC systems can become buried by the deposition of younger sedimentary basins (e.g., cover sequences above the Great Bear magmatic zone and the Olympic Cu-Au Province) [42,155], or they can be metamorphosed to mid-to-high grades through orogenesis [19,21,51,152,156,157].
Figure 6 highlights some key components of the geological context of MIAC systems using the well-constrained field geology of the Great Bear magmatic zone, including the metamorphic basement rocks and overlying 1.88 Ga sedimentary basin; the 1.87 Ga volcanic centers and their coeval intrusions; and the extensive 1.866 Ga ignimbritic cover and the multi-phase batholiths ([62] and the references therein).

4.2. Pre-MIAC Sedimentary Basins

Volcanic centers and intrusions coeval with MIAC systems are deposited on or have intruded earlier sedimentary basins that consist of carbonate, siltstone, and wacke, as well as evaporite or scapolite-bearing units interpreted as evaporite [2,42,68,70,86,114,161,162,163,164]. Some of these basins only slightly predate, and underwent short-lived orogenesis before, the MIAC activity (Great Bear magmatic zone) [62]. Other sedimentary basins are nearly coeval with or postdate volcanism and underwent tectonic inversion (Central Andes) [161]. Finally, there are older metamorphosed basins, such as those of the Mount Isa and Curnamona provinces of Australia [82], Kangdian district in China [64], and Romanet Horst in Canada [50]. Such sedimentary basins are common hosts for MI-Co and albitite-hosted mineralization.

4.3. Syn-MIAC Magmatic Flare-Up

Worldwide, magmatism that is coeval with MIAC systems typically forms distinct volcanic and intrusive centers with mafic, intermediate, and felsic compositions of shoshonitic-to-calc-alkaline affinities (e.g., Figure 6) [42,62,64,76,86,114,163]. Mafic igneous rocks are limited, while intermediate-to-felsic igneous rocks are common, and their relative proportions can vary across and among metallogenic belts (Figure 6) [16,34,58,62]. In addition to volcanic and intrusive rocks, syn-magmatic and localized fluvio-lacustrine sedimentary rocks, evaporite, acidic volcanic lake, and carbonate units can also be present (e.g., Olympic Cu-Au Province in Australia, Middle–Lower Yangtze River Metallogenic Belt in China, Great Bear magmatic zone in Canada, and there are also other MIAC provinces worldwide) [10,11,53,64,78,86,162,163,164,165]. The duration of the magmatic events associated with the formation of the primary MIAC systems were in the order of about 5–10 Ma, as described below.
In the Great Bear magmatic zone in Canada (Figure 6), the onset of regional metasomatism between 1876 and 1873 Ma coincided with the emplacement of syn-tectonic granitic dykes and sparse granitic intrusions at the waning stage of a compressional event [15,16,62]. A magmatic flare-up followed between 1873 and 1868 Ma, and it was related with the development of numerous volcanic centers consisting of compositionally diverse volcanic rocks, porphyritic-to-phaneritic sub-volcanic intrusions, and localized syn-volcanic sedimentary units. This magmatism provided one of the fluid sources for the development of MIAC systems [166]. The volcanic centers were preferentially located in extensional pull-apart intra-arc basins that have been interpreted to occur along the entire exposure of the belt [62]. Hildebrand and Bowring [167] interpreted the continental arc to be of low topography. The volcanic centers were interconnected by arc-parallel major fault zones like the one largely obscured by batholith emplacement in the west and the well-preserved Wopmay fault zone in the east [62,160]. Basaltic flows occur in narrow basins along the east-bounding, orogen-parallel Wopmay fault zone (Figure 6) [15,53]. In the other 1.87 Ga basins of the Great Bear magmatic zone, mafic rocks are minor [58,61,158,165].
In the northwest part of the Great Bear magmatic zone, the volcanic centers comprise andesite stratovolcanoes and calderas, and they were formed at ca. 1873 Ma along with their co-magmatic dioritic-to-monzonitic sub-volcanic intrusions and resurgent plutons; these also host IOA and IOCG prospects and past-producing vein-type mines (Figure 6; Camsell River district hosting the past-producing Terra Ag mine) [16,57,58,62]. The subvolcanic intrusions were emplaced as laccoliths and lopoliths, which is a mode of emplacement also observed in the MIAC system of the East Arm Basin coeval with the Great Bear magmatic zone [18,51,52,57,60,165], as well as in the Central Andes [114] and in the Middle–Lower Yangtze River Metallogenic Belt [64,86].
In the southeast of the Great Bear magmatic zone, the volcanic centers evolved in composition from rhyolitic/dacitic-to-dacitic/andesitic compositions. Volcanism was concurrent with the emplacement of many generations of porphyritic-to-phaneritic granitic intrusions [16,27,62]. Mineralization took place at the transition from felsic-to-felsic-intermediate magmatism as Facies 1–2 evolved to Facies 2–3, forming MI-W and then MI-Co mineralization, as observed in the NICO Au-Co-Bi-Cu deposit. These mineralization styles predominantly occur in the sedimentary sequences forming the basement of the volcanic centers [17,62,86,122]. The subsequent phase of dacitic/andesitic magmatism was associated throughout the southeast Great Bear magmatic zone with conditions that formed the Facies 3 HT K-Fe alteration and associated IOCG mineralization, the latter of which preferentially developed in the volcanic or sub-volcanic porphyritic units. Examples include the Summit Peak IOCG prospect overlying the NICO deposit, the Sue-Dianne IOCG deposit, and several IOCG prospects in the vicinities of the Sue-Dianne deposit and in the Fab area [16,56,62,124]. This period of dacitic/andesitic magmatism was also associated with the formation of albitite-hosted MI-U mineralization associated with Facies 3 HT K-Fe, such as the showings along the Southern Breccia trend hosted in albitite that replaced the sedimentary sequences [28,64,86].
In the Olympic Cu-Au Province (Australia), minor basalts, basaltic andesite, andesite, and dacite evolved to abundant rhyolite in the Lower Gawler Range Volcanics (1595 to 1588 Ma) that host some of the IOCG deposits of the province, such as the Prominent Hill deposit [42,163,168,169,170]. At Prominent Hill, this volcanic sequence overlies an earlier clastic sedimentary basin with carbonate reef units [164]. Additionally, localized fluvio-lacustrine sedimentary rocks and acidic volcanic lakes also occur there [163,164]. At the supergiant Olympic Dam Cu-U-Au-Ag deposit, the host is a brecciated granite (Roxby Downs Granite) dated at 1593 ± 0.21 Ma [171]. The hematite associated with mineralization in three IOCG deposits of the province was dated at 1591.27 ± 0.89 Ma at Olympic Dam, 1598.9 ± 6.3 Ma at Wirrda Well, and 1590.6 ± 6.5 Ma at Acropolis [172,173].
In the Central Andes, andesite hosts are predominant and dacite is localized. Volcanism was followed by the emplacement of diorite laccoliths, local gabbro, and some tonalite, granodiorite, and monzonite with magnetite-series, as well as I-type medium-to-high K calc-alkaline affinity [75,114,174]. These units and their basement rocks have undergone extensive metasomatism and mineralization. In the Middle–Lower Yangtze River Metallogenic Belt, calc-alkaline-to-shoshonitic andesite units and their underlying diorite intrusions and sedimentary basins with carbonates, and evaporites host the MIAC centers [64,86,175]. All these case studies share strikingly similar sequences of magmatic evolution.
Some MIAC systems formed at the same time as the emplacement of carbonatite intrusions, such as those found in the Romanet Horst in Canada [49], or they were spatially associated with them, such as at Bayan Obo in China [176]. In many examples, alkaline intrusions, notably carbonatites, or potassic alkaline-to-ultrapotassic intrusions, were emplaced subsequently, e.g., the Johnnies Rewards prospect and the later Mordor complex in Australia [156]; the 1.35 Ga Bondy gneiss complex, and the 1.09–1.07 Ga Kensington-Skootamatta ultrapotassic to shoshonitic suite [19], and the Lac Cinquante U deposit, and the ultrapotassic extrusive and intrusive rocks of the Christopher Island Formation in Canada [47].

4.4. Syn- to Post-MIAC Batholith Emplacement

Batholiths are a regionally significant feature of MIAC metallogenic provinces, and they have long been interpreted as a key source of fluids for the development of MIAC systems. However, in certain MIAC provinces, the peak period of batholith development postdated the MIAC activity, as described below. This indicates that the batholiths were not necessarily involved in the development of the MIAC footprints and their primary metal endowment. However, the batholiths post-dating MIAC systems do play a role in recirculation of fluids that can remobilize and concentrate the metals deposited in pre-existing mineralization zones [85,177]. The tonalite, granodiorite, monzodiorite, monzonite, and granite intrusions that constitute batholiths (Figure 6) have continental I- to A-type geochemical affinities [62,139,174].
In the Great Bear magmatic zone, the initial magmatic flare-up associated with peak MIAC activities (1873 to 1868 Ma) transitioned to the emplacement of a batholith (1.866–1.85 Ga) that coincided with the deposition of voluminous dacitic-to-rhyolitic ignimbrite sequences [16,53,62]. Similarly, in the Olympic Cu-Au Province, there was a rapid shift from the syn-MIAC basaltic-to-rhyolitic Lower Gawler Range Volcanic, as well as from the coeval intrusions hosting the Olympic Dam and Carrapateena IOCG deposits (from 1595 to 1588 Ma) to the post-MIAC 1587 Ma dacitic Upper Gawler Range Volcanics [71,142,170]. Batholith emplacement continued for at least another eight million years following the development of the MIAC systems [170].
Granitic magmatism, invoked for the genesis of IOCG deposits in the Carajás district, has also been shown to predate the MIAC systems and their IOCG deposits in the district [149]. In the Central Andes, the batholith is, in part, coeval with the development of MIAC systems [88]. In certain geological provinces, the batholiths can also be spatially decoupled from the MIAC systems. For example, the Killarney granite is located tens of kilometers east of the MIAC systems of the Wanapitei district in Ontario [22,100].

4.5. Structural Context

At the scale of a metallogenic province, individual MIAC systems and their cluster of mineral deposits are connected via arc- or orogen-parallel fault systems and their splay faults that are active during MIAC development. Examples include (1) the Atacama Fault System in the Central Andes [40,88,141,143,161], and (2) the series of fault networks that linked the volcanic centers in the Great Bear magmatic zone, including the Wopmay fault zone [15,16,57,62,160]. Dilatational jogs along such fault systems hosted the volcanic centers, intrusions, and MIAC systems in the Great Bear magmatic zone [62] and the MIAC systems in the Central Andes [114]. Albitite corridors, formed along such faults, commonly host subsequent alteration facies and mineralization [16,28,68,88,89,97,98,104,106].
In the southern Great Bear magmatic zone, a change in the composition of the intrusions was associated with (1) a distinct change from higher-to-lower temperature alteration facies, (2) a transition from a compressional and ductile to an extensional and brittle deformation regime, and (3) the associated mineralization types [7,11,16]. The Fe-rich Au-Co-Bi NICO deposit formed in the waning stages of compressional ductile deformation and at the onset of brittle extension. In contrast, the Sue-Dianne IOCG deposit, the Southern Breccia albitite-hosted MI-U mineralization, and the Chalco and Summit Peak IOCG prospects adjacent to the NICO deposit formed in the initial stage of extension largely under a brittle regime [16]. A transition is observed from biotite-rich Facies 3, in which syn-metasomatism brittle–ductile tectonic fabrics like foliations are developed (in the Southern Breccia) to the K-feldspar rich Facies 3 and Facies 5 that were associated with brecciation in the zones of IOCG mineralization [10].

5. Discussion

In Skirrow et al. [34], Geoscience Australia addressed the parameters that affect IOCG ore formation and derived mineral potential criteria from them. Applied to the MIAC system and its entire range of deposit types, these parameters are as follows: (1) the sources of hypersaline hydrothermal fluids, salt melts, hydrosaline liquids, ligands, cations, metals, and sulfur that precipitate the metasomatites, as well as their primary metal endowment; (2) the energy and the kinetics of reactions that ‘drive’ the system; (3) the crustal and mantle lithospheric architecture, including the fault systems and major discontinuities that serve as fluid pathways; and (4) the ore depositional gradients induced by the spatial and/or temporal changes in the physicochemical parameters of the fluid plume that drive the precipitation of the distinct metal associations of MIAC systems. Field geology provides fundamental insights into all four parameters and is the focus of this discussion. The timing of the primary MIAC system and subsequent reworking is constrained through the dating of pre-, syn- and post-MIAC dykes, intrusions, and hydrothermal veins, as well as the dating of metasomatic minerals [16,17,62,172,173,177,178,179].

5.1. Metal and Fluid Sources: A Regional Field Geology Perspective

The suite of interconnected zones with extreme metasomatism—where replacement prevails over veining and brecciation, regardless of the host rock types and geological environments—indicates that the fluids driving MIAC systems are more aggressive and pervasive than common hydrothermal fluids (cf. the discussion in Tornos et al. [141]). This is acknowledged in Hofstra et al. [1] by the classification of MIAC-related deposits as ‘regional metasomatic’ rather than hydrothermal.
The next sections provide a field geology perspective on the nature of the primary metasomatizing fluid plume and the potential ingresses of external fluids based on the pre-, syn-, and post-MIAC volcano-plutonic and sedimentary events. The scientific breakthrough on the probable involvement of salt melts in the formation of mineral deposits and alteration facies in MIAC systems is also addressed as it helps to comprehend how the fluid plume can result in very intense metasomatism. The role of mantle metasomatism, the participation of mantle fluids, and the relationships between the nature of magmatism and the tectonic environment for establishing the MIAC system are beyond the scope of this paper.

5.1.1. The Primary Sources of Regional Fluid Plumes

The primary MIAC event (as opposed to subsequent remobilization) is commonly coeval with a magmatic flare-up that creates discrete volcanic centers with sub-volcanic intrusions of calc-alkaline to shoshonitic to A-type affinities (e.g., Great Bear magmatic zone, Bafq district, Middle–Lower Yangtze River Metallogenic Belt, and Olympic Cu-Au Province) [42,62,64,75,76,163,170,173,180,181]. Fertile magmas were oxidized and had very high temperatures and a lower degree of fractionation than magmas that postdate MIAC activities [145,182]. These magmas could also assimilate and potentially melt their wall rocks, including evaporite-rich units ([141] and the references therein). Magmatic systems could thus liberate a large quantity of metasomatizing fluids with a wide range of composition that would impart a magmatic signature to fluids trapped as inclusions in MIAC minerals [40,134,166,183,184,185,186].
Metasomatism proceeded broadly coevally along entire metallogenic provinces across pre-MIAC basements, sedimentary basins, and overlying syn-MIAC volcano-sedimentary basins, as well as those within syn-MIAC sub- and intra-volcanic intrusions [16,52,64]. In non tectonically disrupted components of the systems, the spatial and timing relationships among alteration Facies 1 to 6 along exposed, tilted stratigraphic sections record how systems evolved from the depth to paleosurface [15,55,56]. What elements were precipitated from the inflow fluid (i.e., metasomatites) and what was mobilized into the outflow fluids (i.e., the change in composition from hosts to metasomatites) can thus be interpreted. Such information, combined with the tight constraint on the coeval timing and province-scale extent of Facies 1 albitite corridors and the sequential formation of alteration Facies 1 to 6, indicate that (1) the metasomatizing fluids must likely collected into regional fluid plumes to trigger the system, (2) the plume must have been initially very rich in Na, and (3) single-pass infiltration of the fluid plume from the depth to the surface without complex overprinting relationships between the alteration facies, due to tectonic disruption or other regional-scale events, can take place in simple systems (cf. the discussion of single-pass infiltration in [89] and the single fluid model of [43]). The most fertile MIAC systems were commonly disrupted and are characterized by complex overprints of alteration facies as well as telescoping of alteration facies. Field indications of such disruptions could be the superimposition of high-temperature facies on lower-temperature facies or the cyclical formation of the same alteration facies. This is indicative of a dynamic environment in which the fluid plume was periodically rejuvenated by fluids from different sources and reheated from renewed magma emplacement.
The field metasomatic record testifies to the progressive dissolution of most, or even all, minerals from host rocks, as well as to the efficient transport at the local-to-regional scale of the elements dissolved by the primary fluid plume. This includes metals, Fe, S, volatiles, and, in some extreme cases, some of the normally immobile elements such as Al, REE, Ti, and Zr [6,24,27,47,187,188]. Through such intense metasomatism, a diverse range of metasomatized host rocks become sources for metals, volatiles, and fluids. Once captured by the primary plume, these external fluids will have transferred their own igneous, sedimentary, or metamorphic source signatures to the plume, including a basinal signature that is provided by evaporites and metamorphosed evaporites [121,184,185,186]. In parallel, as Ti and Zr are less mobile and—through their relative enrichment—metasomatic titanite, rutile, and zircon can be precipitated in the early formed albitite and, locally, in other alteration facies [31,61,68,177,188]. These early metasomatic minerals can then be dated to assess the timing of the primary metasomatic event [177,188], and the composition of their inclusions can be analyzed to assess the composition of the primary metasomatizing fluids [31]. Zeng et al. [31] discovered the earliest metasomatizing fluid observed to date in MIAC systems: Na-K-Fe-Ca-Cl-SO4 hydrosaline liquids with Fe-chlorides as the main Fe species within zircon grains that were precipitated during the albitization event predating the formation of IOA deposits of the Middle–Lower Yangtze River Metallogenic Belt. These hydrosaline inclusions have 12–22 wt.% Na, whereasK, Fe, Ca, and S contents are up to 10–18 wt.%. They are enriched in Ti; contain halite, sylvite, Fe-chlorides, anhydrite, fluorite, ±hematite, and pyrite; have magmatic isotopic compositions; and their homogenization temperatures reach 930 °C [31]. Multiphase inclusions in other IOA deposits also record the presence of salt melts (e.g., chloride, sulfate, carbonate, and phosphate melts) that are rich in Na, K, Ca, Fe, Cl, F, and SO4 and reach ~1000 °C in temperature [32,86,189,190].
The source of the salt melts in the fluid plumes and in IOA deposits remains unclear. As salt melts inclusions have been observed in albitite, their source must predate or be formed coevally with albitization. The exsolution of hypersaline fluids from sub-volcanic intrusions has been proposed as a possible source by Hildebrand [52] and Zhao et al. [86]. Albitite corridors are indeed common at the roof of laccoliths and lopoliths [31,52,57,86]. However, a sub-volcanic intrusion in the Great Bear magmatic zone host enclaves of pervasively albitized wall rocks while being weakly albitized in that area [14]. Hence, at least this intrusion must postdate the onset of intense albitization. In addition, at least one intra-volcanic pluton has albitite, both at its base and top (Tut pluton, [54]), and this and other sub-volcanic intrusions can be albitized themselves [14,16,17,27,52,54]. The combination of these observations indicate that the primary fluid plume must have been sourced from below the sub-volcanic intrusions and not within it, as was proposed earlier [52].
In MIAC provinces where sedimentary sequences prevail, albitite corridors have developed along regional discontinuities within sedimentary basins far from intrusions, as has been observed in the Mt Isa Province west of the Cloncurry district, as well as in the Romanet Horst and the Wanapitei district in Canada [22,50,66]. Magnetic and gravity anomalies directly associated with the metasomatic footprints of the systems extends for tens of km well beyond any single intrusion [160], and they can reach depths of 10 km, whereas the magnetotelluric footprints of the systems extend to the mantle [191]. Zeng et al. [31] interpreted the hydrosaline liquids to be formed by transcrustal magmatic systems during the migration of magmas and through exsolution from sub-volcanic intrusions, with preferential ponding of the fluids above the cupola of the intrusions at about 2 km depth. Such an interpretation is in line with the geophysical footprints of potential magmatic brine accumulation at a ca. 2 km depth in active volcanic settings [192].
The assimilation and melting of evaporite or carbonate rocks by magmas and the subsequent immiscibility of salt melts are also invoked to form IOA deposits [32,190], and they are among the many genetic models for such deposits [141]. Most models are currently irreconcilable with the observed formation of albitites prior to the IOA deposits at the regional scale [7]. However, andesite flows that host IOA deposits show evidence for Fe-rich, as well as Cu-rich immiscible sulfide, melt inclusions [193]. Such melts and their metals may have made their way into the primary fluid plume we interpret to trigger the formation of MIAC systems based on our field geology constraints.
What we envision is that fluids, including salt melts, (1) originated from magmatic sources with potential additional contributions from the assimilation, metasomatic devolatilization, and melting of evaporitic or carbonate sources, and (2) they also collected into giant hypersaline, hot, and corrosive fluid plumes in the mid-crust. Metasomatism was triggered by the fluid plume as it raised concurrently with the magmas that formed the volcanic sequences and sub-volcanic intrusions. The fluid plume and the magmas may have used similar zones of enhanced crustal permeability like crustal-scale fault systems. The plume would then flow and infiltrate along and across major discontinuities in the upper crust, including along discontinuities in which laccolith and lopoliths were emplaced and albitite corridors formed. Additionally, the faults, breccias, and damage zones related to subsidence of multiple cauldrons (e.g., the Camsell River district, Great Bear magmatic zone, Canada [52,53,165], and southeast Missouri district, USA [194]) would further facilitate fluid flow and metasomatism. Fluids that precipitate early within the alteration sequence commonly have a magmatic signature, while subsequent ones show greater interaction with host rocks or as having tapped additional non-magmatic fluid sources [86,164,195].

5.1.2. The Secondary Sources: Insights from Metasomatized Host Rocks and Geological Environments

The nature of the host rocks metasomatized across the upper crust, combined with the field relationships among metasomatic, igneous, metamorphic, and sedimentary rocks, provide indications of the fluids and metals added to the main fluid plume as metasomatism proceeded. The metasomatic mineral dissolution during regional-scale albitization in Facies 1 occurs in all rock types (Table 1). Such pervasive dissolution of compositionally diverse host rocks must release, into the plume, a wide variety of volatiles, metals, cations, and ligands, including those found in the more reactive carbonate and evaporite units. The self-propagating dissolution process implies that, at each stage, the elements of the minerals unstable in host rocks become available and mobilized in solution and precipitate, at a subsequent alteration stage, sustaining the metal recharge of the main fluid plume.
Pre-MIAC sedimentary basins can host SEDEX mineralization [138], which may provide a source for Pb and Zn, whereas their mafic rocks can become a source for Cu and Zn [99], leading to MIAC deposits with Zn resources such as the Mount Dore Cu-Au-Pb-Zn deposit in the Cloncurry district, Australia (Table A1) [196]. In addition, pre- and syn-MIAC sedimentary basins can contain formation waters [197]. Their evaporites, including those metamorphosed with scapolite, can provide volatiles and S once metasomatized, and they could also release sulfate melts if assimilated or partially melted by magmas [66,190,198]. Scapolite and sulfate minerals are not the sole markers of evaporite and saline host rocks. For example, scapolite is largely absent in the metasomatized sedimentary rocks of the Great Bear magmatic zone and the fluid inclusions among MIAC mineralization in the region largely recorded magmatic fluid sources [121,122,166]. Yet, albitite can have amphibole-rich components that pseudomorph the nodular textures and bedding structures typical of nodular anhydrite beds, which suggests the presence of an evaporite sequence (E in Fig. 6; Fig. 10A,B in [11]). The pseudomorphing of nodular textures has also been observed in the Norrbotten district [152]. Stromatolite-bearing units were metasomatized to magnetite and albite (Fig. 10C–G in [11]); their stratigraphic location within the Grouard Lake sequence (S in Figure 6) [53] indicates that they may have precipitated within a saline lake in a volcanic cauldron [7,10,11]. Such lakes have been interpreted as a key fluid source for IOCG deposits [164].
The metasomatism of tens of km long carbonate sedimentary units and of graphitic schists (e.g., Cloncurry district, the Great Bear magmatic zone, Romanet Horst, and Norrbotten district) [11,17,49,66,90] requires the release of CO2 to the fluid plume. Fluid signatures in deposits hosted in such settings are commonly magmatic [66,122,183,195], but later-stage alteration zones have also recorded extensive interaction between fluids and host rocks [195]. Considering the geological environments in which MIAC plumes ascend and propagate, it is likely that the plumes incorporated other fluids encountered along the fluid pathways as described above, as well as the additional magmatic fluids liberated by syn-MIAC intrusions and dyke swarms [16,31,52,124]. As the fluid plume leached metals and volatiles from a wide diversity of host rocks on a regional scale, the outflow fluids acquired a mixed and evolving source signature; hence, the commonly observed record of distinct fluid compositions at different stages of mineralization in MIAC systems [33,66,67,121,183,184,185,186,199,200,201,202]. Distinct fluid compositions at the different stages of mineralization are intrinsic to the metasomatic evolution of the system and do not imply that distinct fluid sources created the distinct stages of the MIAC system (see [203]). However, external ingresses of fluids that were not originally present in the evolving fluid plume (e.g., fluids of magmatic, basinal, metamorphic, and meteoric sources) probably did occur and changed the composition of the parental fluid plume transiently or permanently, depending on the volume of the fluid ingress. The access to external fluids may have been essential to the formation of giant mineral deposits [164,202].

5.2. Energy Driver of Fluid Flow

Coeval MIAC systems developed into 600 to 2000 km long metallogenic belts [40,62,139,146,147,152,204], and they are rooted into a metasomatized mantle through transcrustal magma systems and fault zones at the margins of thick lithospheres [9,33,34]. Consequently, the conditions that triggered, drove, and sustained, in the same geological province, the relatively coeval formation of MIAC systems over 100s of km cannot be a local phenomenon but required a significant and regional thermal event that likely originated from the mantle. This thermal event triggered the formation of a belt-scale transcrustal magmatic system (with a generation of salt melts?). In many geological provinces, the initial flare-up of the MIAC activities event was associated with oxidized calc-alkaline to shoshonitic magmatism, which formed discrete volcanic centers. At the later stages and after development of the MIAC systems, the magmatic system expanded to form belt-scale batholiths made of intrusions with a prevailing A-type signature. This is observed in the Central Andes, the Great Bear magmatic zone, the Middle–Lower Yangtze River Metallogenic Belt, and other settings globally [34,62,144,161,174,180,205]. Syn-MIAC magmas were, in part, sourced in the mantle and have high temperatures with trapped melt inclusions that indicate temperatures of up to 1145 °C in andesite [193] in systems where a very high-temperature IOA mineralization prevails and with zircon crystallization temperatures reaching ~950 °C in systems where IOCG mineralization is the most developed [182].
The magmatic systems, in addition to the fluid plumes, can transfer, at the scale of the metallogenic belt, significant heat from the mantle to the upper crust to drive (or energize, once triggered) the coeval formation of multiple MIAC systems [206]. In addition, the salt melts and hydrosaline liquids inclusions observed within albitite, skarn and IOA deposits indicate that salt melts and hydrosaline liquids probably coexisted in the fluid plume with hydrothermal fluids and considerably enhanced the fluid’s ability to carry, mobilize, and transport cations, volatiles, and metals, as well as even the most incompatible elements such as Ti. Do the salt melts form the primary fluid plume or are they added to a primary fluid plume that is hydrothermal in origin? Are the salt melts sourced in the mantle or are they only derived at the crustal level? Whatever their sources, the presence of salt melts into the hypersaline fluid plumes can sustain their extreme heat and corrosive ability to dissolve most minerals and precipitate metasomatites with mineral assemblages that reached typical of upper amphibolite-to-granulite metamorphic facies (e.g., hypersthene) [117] and textures typical of intrusive rocks [11].

5.3. Structural–Stratigraphic Architecture and Porosity Creation That Enable Fluid Flow

The plumes of fluids—and the salt melts they incorporate—are interpreted to collect and pool in the middle crust. A progressive switch from a compressional to an extensional regime that is commonly documented in MIAC provinces could have promoted the migration of the fluid plumes along transcrustal faults toward the upper crust, like the orogen-parallel Atacama Fault System along which multiple MIAC systems formed (~2000 km) [40]. Such fault systems are commonly rooted in the mantle [33,34,191]. As the plumes reached the upper crust, based on the distribution of alteration facies (and their geophysical footprints), they laterally propagated along fault systems and the interconnected crustal discontinuities that included, for example, damage, fractured, and brecciated zones [88,89,93,94,160,207]. The fluids also used other discontinuities, such as more permeable chemically reactive or brittle layers in stratified units (e.g., sedimentary rocks), major contacts between geological units (like the upper contact of sub-volcanic intrusions with their volcanic ejecta), and pre-existing zones of permeability (like the breccia zones in the walls of calderas) [52,91].
From the micro to the regional scale, the infiltration and propagation of the fluid plumes from the major discontinuities into the host rocks were also controlled by the metasomatic dissolution and reprecipitation of minerals and the creation of transient porosity [63,68]. For example, an albitite is highly porous and remains so unless that albitite was recrystallized above later sub-volcanic intrusions [14,28,52,63,68] or through orogenesis [19]. Permeability channels may also have formed through volume loss during metasomatic devolatilization (e.g., CO2 release through albitization of the carbonates in the lower temperature system of the Romanet Horst and through development of skarn or of magnetite alteration at the expense of carbonates in the Great Bear magmatic zone and Moonta-Wallaroo region of Australia [11,17,89,92]). This is a result of the thermodynamic and kinetic disequilibrium between corrosive fluids and host rocks [203].
Along fault zones, pre-MIAC fracturing can form zones of damage and brecciation in the host rocks that facilitate the formation of regional albitite corridors. Once established and aided by the intrinsic porosity of the albitite, these corridors become preferential sites for further development of damage and breccia zones as the fault zones are reactivated. The corridors of brecciated albitite become preferential pathways for subsequent Fe-rich to Fe-poor mineralization (Table 1, Table 2, Table 3, Table 4, Table 5, Table 6 and Table 7) [28,68,93,109,110,133,207].
The volcano–plutonic environment also creates extensive breccias that aid fluid flow. An example is the large breccias with fragments of the sub-volcanic intrusions spalled from the walls of a caldera during the caldera collapse in the Camsell River district in the Great Bear magmatic zone [165]. These breccias were infilled by hematite. At the Olympic Dam deposit, extensive brecciation of the host granite has been attributed to the collapse and hydraulic brecciation of a granitic intrusion cupola during the uplift that was followed by depressurization, leading to the exsolution of orthomagmatic fluids and precipitation of hematite [173,178].

5.4. Depositional Environment: Alteration Facies as Predictive Indicators of Mineralization

The confluence of the multi-scale, fluid-driven metasomatic processes, notably the infiltrative replacement, veining, and brecciation at each facies, induces the coupling and decoupling of elements and metals into distinct metal associations and the genesis of facies-specific mineralization types (Figure 7). In contrast, structural or lithological/chemical traps concentrate the deposition of metals to form mineral deposits [22,94,207], whereas, at the scale of the metallogenic belt, deposits are associated with deep, transcrustal fault zones and their splays (Section 5.3; [208,209]). Hence, a critical window for ore genesis is where fertile alteration facies encounter chemically reactive geological units or favorable structural traps. The shape of the resulting mineralization zones becomes a function of the original distribution of the host rocks in the geological sequence and of the geometry of the major discontinuities along which the fluids have circulated, such as sedimentary units and fault zones.
Field observation indicates that mineralization is varied but localized in MIAC systems that exhibit a linear and simple evolution of alteration facies, which is interpreted as indicating an absence of significant rejuvenation of the fluid plume by a voluminous ingress of external fluids. MIAC systems, with a full spectrum of alteration facies and with indications of overprinting or telescoping of alteration facies, are interpreted as evidence for cyclical rejuvenation of the fluid plume and are found to commonly host deposits with varied mineralization styles. Yet, even in a MIAC system with polyphase and complex alteration patterns, the overall evolution and the relative timing among alteration facies remain largely the same as the one depicted in Figure 7. An example of the cyclical formation of distinct facies and renewed mineralization has ben observed between Facies 2 and 2–3 at the NICO deposit, Canada [15].
The relationships described between the alteration facies and deposit types in Figure 7 are those observed from the metal associations that are associated with each alteration facies. Of note, each alteration facies can also become a preferential host to the mineralization associated with an overprinting alteration facies. During an exploration program, identifying the alteration facies associated with the main mineralization event(s) can optimize exploration as it indicates what mineralization types are most likely to be present in the tested zone of the system. For example, many IOCG exploration programs failed, as the drill targets were magnetic bodies of Facies 2 HT Ca-Fe alteration, which are not fertile for Cu mineralization. The association between alteration facies and mineralization and of the overprinting relationships between alteration facies in mineralization zones are best documented through careful macroscopic descriptions of the mineral assemblages on outcrops or drill cores in the field. Supplementing the field descriptions with techniques like cobaltinitrate staining or micro-XRF analysis of rock slabs and drill cores can considerably improve the descriptions [11,14,22]. Discriminating alteration facies using microscopic observations in thin sections is often difficult as the scale is too small to clearly define overprinting relationships between different alteration facies.
In all systems, the corridors of intensely Na-altered rocks with abundant and extensive zones of albitite are intrinsic to the development of the MIAC systems. They also serve as a diagnostic and distinctive marker of the onset of a MIAC system. Though albitization zones can be extensive in other mineral systems, such as volcanic massive sulfide mineral systems, the hosts are not as extensively and intensively albitized to pervasive albitite [6].
Prior albitization events can occur in sedimentary basins due to the internal circulation of saline basinal fluids (e.g., Cloncurry district and Mount Isa Province) [82,211]. Some mineral assemblages of Facies 5 can also contain abundant albite following the progressive recharge of the fluid plume in Na from Facies 2 to 4 [22,68]. Facies 5 albite-carbonate alteration can be associated with Cu-Au mineralization and albite-chlorite with Au mineralization [22]. Here, the emphasis on what is a diagnostic of, and a mappable prospectivity criteria for, a MIAC system is placed on the regional-scale development of early albitite, intense scapolite-albite, and intense oligoclase-rich corridors, not simply on the presence of albitized rocks.
Facies 1 Na can also act as a metallogenic preconditioning factor for a wide range of MIAC deposit types. Arrested fronts of albitization record the mobilization of dissolved cations into subsequent alteration facies and their incremental development as metasomatism intensifies [14]. Albitite also records the progression of the primary fluid flow along fractures (e.g., haloes along fractures), damage zones (e.g., coalescing fracture haloes and albitization network), intrusive contacts, preferential stratigraphic sequences, and regional-scale fault zones and splays, many stemming from a transcrustal fault network [62,91,94,140,160,208].
In carbonate rocks, skarns form coevally with the albitite zones, and the fluid signature can be similar [91], whereas an albitite can form instead of a skarn in low-temperature environments, as observed in the Romanet Horst [49,50]. Albitite (and other regional Na alteration) and associated skarn predate IOA mineralization at deposit-to-regional scales (e.g., the Great Bear magmatic zone, Middle–Lower Yangtze River Metallogenic Belt, Norrbotten district, and the Central Andes; Figure 1) [6,7,31,52,55,88,90,205]. Albitization can also slightly predate the emplacement of sub-volcanic intrusions [14].
The linkages between alteration Facies 2 to 6 and their facies-specific metal associations and deposit types provide mappable mineral potential criteria for the deposits formed in MIAC systems. The Na facies transitions from Na-Ca (albite-scapolite) to HT Na-Ca-Fe assemblages, where albite, amphibole, and magnetite prevail, and then to Facies 2 HT Ca-Fe alteration. The Na to Na-Ca-Fe facies are barren, but the Na-Ca-Fe facies, which marks the first stage of Fe ingress in a MIAC system, can indicate potential proximity to a mineralized Fe-rich alteration facies. With the progressive addition of Fe in the system, the Facies 1 skarn also transitions to the Skarn+Fe facies that can be associated with W mineralization.
Facies 2 mineral assemblages are dominated by amphibole, magnetite, and, more rarely, apatite, and they also replace host rocks or precipitate as veins and cement breccias within the zones of albitite and the least-altered host rocks. The HT Ca-Fe facies has the largest footprint in carbonate-bearing sedimentary sequences, where zones of skarn (clinopyroxene and garnet) may form early, pre- to syn-albitite but are progressively replaced by Facies 2 assemblages. Areas of interest at the regional scale are delineated by Facies 1 to 2, which also host magnetite skarn and iron oxide-apatite (IOA) deposits. The latter are commonly spatially associated with dioritic intrusions and andesite sequences, such as in the Middle–Lower Yangtze River Metallogenic belt and the Great Bear magmatic zone [31,52,64,212]. With heat ingress from high-temperature diorite, IOA mineralization at Facies 2 HT Ca-Fe alteration can thus precipitate under very high temperatures, most being between 600 and 1000 °C (Section 5.2).
Systems that evolve toward conditions forming Facies 3 within sedimentary rocks can precipitate the MI-Co deposits, with metal associations of Au, Co, and Bi, at Facies 2–3 HT Ca-K-Fe alteration, such as the NICO deposit in the Great Bear magmatic zone [17,122]. Brecciation is typically minor. In contrast, extensive brecciation is associated with Facies 3 HT K-Fe (magnetite-K-feldspar/biotite), Facies 4a K-skarn (clinopyroxene, garnet, and K-feldspar), Facies 4b K-felsite (K-feldspar), and the Facies 5 low-temperature K-Fe ± Ca, Mg, H+, CO2, Si, and Ba (sericite, K-feldspar, hematite, chlorite, carbonate, and epidote). The formation of MI-Cu deposits, including IOCG and ISCG, and of MI-U deposits is associated with Facies 3 and 5a (Figure 7). The IOCG deposits associated with Facies 3 can also be enriched in Co or include Co resources, such as at the Ernest-Henry Cu-Au deposit in the Cloncurry district with Co reaching 2 wt.% in pyrite [126] and the Kiskamavaara Cu-Co deposit in the Norrbotten district [213,214]. Biotite-rich HT K-Fe alteration is also associated with the formation of MI-Co deposits, such as the Ram deposit in the Idaho Cobalt Belt [123]. Facies 6 low-temperature K, Si, Al ± Ba, and Fe were found to be associated with the formation of vein systems, sericitic and phyllic alteration, and epithermal lithocaps [56,57,60,84].
One of the primary challenges of doing visual depictions of alteration zonation in MIAC systems is that the geometry in 3D of the zonation of the alteration facies can vary considerably among MIAC systems based on the geometries of the main discontinuities and the common tectonic disruption of the system during metasomatism. For example, the geometry of alteration zones in a MIAC system formed in predominantly layered sequences (e.g., sedimentary or volcaniclastic) will be considerably different from the geometry of alteration zones in a MIAC system formed in a brittle–ductile shear zone, or from a MIAC system forming a breccia complex at the intersection of two brittle faults in an intrusive complex. This level of geometrical complexity (e.g., discordant alteration zonation) can further be compounded by faulting and with new infiltration stages of the fluids from the plume or from the ingress of external fluids that could be circulating along crosscutting discontinuities.

5.5. Perspectives on Knowledge and Data Gaps

5.5.1. The Need to Acquire Regional Geological Data on MIAC Systems

As per Wyborn et al. [36], truly mappable criteria are, first, boots on the ground. For MIAC systems, there is a strong need to incorporate, in regional mapping programs, the routine description of observable alteration facies that record energy, fluid, and metal sources, as well as the potential geological, chemical, and structural pathways, traps, and thermal gradients, for metals and fluids. This can only be achieved by having appropriate taxonomies and lexicons to cover the realm of geological attributes that are characteristic of each alteration facies ([10,46] and in preparation). This includes lexicons to describe the extensive replacement, veining, and brecciation zones in MIAC systems and their relationships to host rock types and to coeval or later magmatism, sedimentation, metamorphism, and tectonic activities. Considering the realm of Fe oxide-rich to Fe oxide-poor to Fe-poor deposit types, exploration strategies must also move beyond an IOCG-centric approach and integrate the varied information available at district scale on prospects and alteration facies to assess the prospectivity of the MIAC system explored. The selection of geophysical methods to detect areas of mineralization needs to consider the mineralogical expression of Fe enrichments in a targeted mineralized alteration facies. For example, geophysical methods used to detect Fe oxides may not be suitable if mineralization is associated with Fe sulfides or Fe silicates.
Once formed, any MIAC system can be metamorphosed during orogenesis [157]. The Bondy gneiss complex, Grenville Province, Canada [19], and the Johnnies Rewards deposit in the Northern Territory, Australia [156], are examples of MIAC systems that were metamorphosed at granulite facies. The atypical bulk rock compositions of the alteration facies in MIAC systems, where metasomatism is intense, produce diagnostic metamorphic mineral assemblages with very atypical mineral contents in the metamorphosed metasomatites [19,157]. Notably, Facies 1 albitite are metamorphosed to plagioclase-dominant gneisses; Facies 2 HT Ca-Fe to amphibole-, garnet-, or magnetite-rich gneisses; Facies 3 and 5 HT and LT K-Fe to varied garnetite units, orthopyroxene gneisses, and semipelite-like gneisses; Facies 6 argillic and advanced argillic alteration to metapelite-like gneisses; and tourmaline alteration to rocks that were originally interpreted as being metaexhalites [19,157]. The intense leaching of Na or K in the metasomatized rocks significantly limited partial melting and enabled the preservation of some protolith textures and of the bulk-rock composition of the metasomatites [19,215]. However, the intense deformation typical of high-grade metamorphic belts can transpose the discordant alteration zones of the metamorphosed metasomatites into layered amphibole-, biotite-, cordierite-, garnet-, magnetite-, sillimanite-, or tourmaline-rich gneisses that resemble paragneisses (i.e., gneisses with a sedimentary origin) [19,215]. Nevertheless, the combination of lithological associations and mineral contents—and the variation thereof—remain diagnostic of metamorphosed metasomatites, facilitating the recognition, interpretation, and exploration of MIAC systems in high-grade metamorphic belts.
Both the Johnnies Rewards prospect and the Bondy gneiss complex have been subjected to multiple genetic interpretations. The lithogeochemical distribution of the samples from the Bondy gneiss complex in the AIOCG diagram, as well as in other alteration diagrams and their ranges of Na-Ca-Fe-K-Mg molar proportions [19], are typical of the combined iron and alkali–calcic enrichments of MIAC systems and are distinct from those of other mineral systems [19]. Of note, both case examples are within regions with potassic alkaline intrusions (the Mordor complex and the Kensington Skootamatta potassic alkaline intrusion and its minette dyke, respectively) [156,216]. Such spatial associations between metamorphosed MIAC systems and recognizable potassic and ultrapotassic intrusions provide new research avenues on settings prospective for critical minerals.

5.5.2. Beyond the Predictability of Metal Associations and Deposit Types at Each Facies

The holistic mineral system model presented in Figure 7 provides a robust framework for the regional evolution of metal associations and deposit types within MIAC systems according to the space–time distribution of their alteration facies. Deposits, prospects, and alteration facies thus serve as vectors to mineralization, whether they are rich or all the way to poor in Fe oxides and Fe, furthering the realm of mappable prospectivity criteria and the ability to assess the mineral prospectivity in geological provinces with potential for MIAC systems. What controls the metal endowment of the mineralized zones of MIAC systems remains to be resolved. Many researchers suggest that pre to syn-MIAC geodynamic environments and mantle fertility may play a crucial role on the overall metal endowment of a MIAC system [9]. The metal content of the geological sequence replaced by a MIAC system also play a crucial role as a source of metals for ore deposition [10,11,24]. Certain compositional characteristics of the geological sequence may favor the formation of certain deposit types. For example, MI-Co deposits are preferentially formed in geological sequences with abundant sedimentary rocks. The influx of additional metals and fluids from the geological sequence to the plumes as metasomatism progresses can also carry the distinctive isotopic signatures of the host rocks into the alteration zones, as demonstrated for Fe isotopes by Emproto et al. [200].
The metal endowment of the primary fluid plume is more difficult to assess, as even the source signatures detected in the fluid and melt inclusions in albitite must have undergone modification from the sodic metasomatism of their host. The salt melts cannot also directly precipitate Fe oxide magmas without the main fluid plume initiating the regional sodic metasomatism that is first based on the timing relationships between Facies 1 and Facies 2. Deposit models for IOA deposits need to account for the distribution of albitite corridors along deep-seated fault zones and other discontinuities, including above and locally below sub-volcanic intrusions. Breakthroughs in the understanding of immiscible or exsolved salt and sulfate melts, however, provide key insights on how the parental hypersaline fluid plumes can achieve the high temperatures needed to form igneous-like metasomatic textures, as well as the mineral assemblages akin to upper amphibolite through to granulite facies conditions.

6. Conclusions

In the last few decades, the absence of a unifying framework to explain the observed metallogenic diversity of Fe-rich but Fe oxide-poor, as well as Fe-poor alkali–calcic deposits that are spatially associated with IOCG deposits, has resulted in significant confusion on the significance of these deposits, how to differentiate them from sensu stricto IOCG deposits, how to name and classify them, and how to assess their mineral potential. This paper exemplifies that IOCG deposits are one of many deposit types that can be formed into the district-scale metasomatic mineral systems that are defined as Metasomatic Iron and Alkali-Calcic (MIAC) mineral systems.
From the extensively published field geology attributes of alteration facies, it was possible to address how, in a short time and throughout a geological province, a series of interconnected MIAC systems forming diagnostic alteration facies and mineralization types can be developed. A combination of hydrosaline liquids, salt melts, and hypersaline hydrothermal fluids are interpreted to have been derived from transcrustal magmatic systems (and mantle?) and to have accumulated and collected in discrete areas of enhanced crustal permeability in the mid crust to form a series of large fluid plumes along the metallogenic provinces. Regional field geology observations indicate that the parental fluid plume is generally magmatic in origin and can transport metals, such as Cu and Au.
Series of fluid plumes ascended toward the upper crust along a main transcrustal-, continental arc-, or orogen-parallel fault system, reaching 2000 km in length. In many MIAC systems, the ascent of the fluid plumes occurred concurrently with the syn-MIAC magmas that formed discrete volcanic centers and shallow intrusive complexes. The spatial overlap between MIAC and magmatic systems can enable the periodic transfer of heat- and magma-derived fluids to the fluid plume that evolves in the upper crust.
As the plumes reached the upper crust, they infiltrated and percolated vertically and laterally along faults, within reactive or permeable geological units, and along stratigraphic and intrusive discontinuities. Their ascent and percolation in the upper crust was found to be associated with a suite of metasomatic reactions that dissolved the original minerals in the host rocks and precipitated new (metasomatic) ones. The dissolved elements (including metals and volatiles) were added to the compositionally evolving fluid plume. As the fluid plumes percolated into the upper crust, fluids originating from upper crustal sources, like sedimentary basins, caldera lakes, and meteoric waters, could also mix with the fluid plumes and contribute metals.
The chain of metasomatic reactions that formed the sequence of diagnostic alteration facies with distinctive metal associations, gave way to a predictable series of facies-specific critical and precious metal deposit types, which are rich to poor in Fe. Deposit classes are specific to each alteration facies, whereas commodities may precipitate in more than one facies. Deposit classes are thus defined to highlight the main commodities of deposits to respond to the commonly commodity-driven exploration. Accordingly, MIAC systems host Facies 1–2 MI-W (W-skarn); Facies 2 MI-Fe (Fe-skarn, IO, and IOA), MI-REE (remobilized IOA-REE), and MI-Ni; Facies 2–3 MI-Co; Facies 3 and 5a MI-Cu (magnetite- and hematite-group IOCG, ISCG, and other MI-Cu), MI-U (including albitite-hosted IO-U), and rarer MI-Co; Facies 5b MI-Au; Facies 5c and 5d MAC-Au, MAC-Cu, MAC-Mo, or MAC-U (which are typically albitite-hosted); and, finally, Facies 6 epithermal mineralization and later-stage five-element and other mineralized veins. The distinct alteration facies and their mineralization types can have a depth to surface distribution and/or be overprinted on top of each other, but the global diachronous formation of IOA (earlier) and IOCG (later) remain.
Dynamic tectono-magmatic environments favoring episodic rejuvenation of the fluid plumes or its lateral offshoots with the addition of external fluids and heat can increase the potential of the resulting MIAC systems to host large mineral deposits. Each time new fluids circulate in older mineralized alteration facies, metals can be remobilized and reconcentrated locally remaining within their own alteration facies, or they can be channeled away along fault zones.
Once formed, any MIAC system can then be metamorphosed, with examples at granulite facies being preserved in Canada and Australia.

Author Contributions

Conceptualization, review, editing, and investigation, L.C. and J.-F.M.; funding acquisition and writing—original draft, L.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Ore system project of the Targeted Geoscience Initiative Program (Phase 6) of the Geological Survey of Canada (Natural Resources Canada).

Data Availability Statement

See the references listed. Original photographs can be obtained in the NRCan Photo Database. In this paper, the photographs were slightly adjusted—equally for the entire image for contrast, brightness, intensity, and sharpness—to best illustrate the relationships among the distinct mineral assemblages in alteration and mineralization. The photographs were also cropped on their sides to remove hammers, other object used as scales, excessive lichens, dirt, core boxes, etc., in order to best focus on the key attributes presented.

Acknowledgments

This paper is an outcome of the sub-activity Metasomatic iron and alkali–calcic systems and their IOCG and affiliated critical mineral deposits of the Targeted Geoscience Initiative program (Phase 6) of the Geological Survey of Canada (Natural Resources Canada). The authors sincerely thank the participants of the Critical Mineral Mapping Initiative, O. Blein, and the co-authors in Corriveau et al. [37] for their extensive discussions that helped translate the field geology of the Canadian MIAC systems into empirical constraints on fluid and metal sources, ore deposition environments, and tectonics. Field examination of MIAC systems worldwide was made possible by collaborators and through a series of fieldtrips led by the Society of Economic Geologists (SEG), the Society for Geology Applied to mineral deposits (SGA), and the Geological Survey of South Australia. The authors also wish to thank R. Swinscoe, and the three anonymous journal reviewers for their thorough and insightful reviews of this paper, as well as A.G. Călin for his additional insights. Permission to publish photos from the Ernest Henry mine and of the Belfast prospect were, respectively, granted by B. Miller (Ernest Henry Mining) and by T. Obradovich (Conquest). The authors also wish to thank S.-J. Pak and his colleagues from KIGAM for the opportunity to examine the Yang-Yang deposit.

Conflicts of Interest

The authors declare no conflicts of interest. J.-F.M. contributed to this research as a collaborator of the Targeted Geoscience Initiative Program of Natural Resources Canada with the permission of his employers, MacDonald Mines Exploration and Red Pine Exploration. This paper reflects his views as a scientist and not of the companies.

Appendix A

Table A1. Representative examples of the mineral resources and their deposit types within the MIAC systems of Australia. Districts and provinces are located in Figure 1.
Table A1. Representative examples of the mineral resources and their deposit types within the MIAC systems of Australia. Districts and provinces are located in Figure 1.
DepositsTotal Resources (Unless Indicated Otherwise) 1
NameClassType
Olympic Cu-Au Province (Volcano-Plutonic Environment with Additional Sedimentary Hosts)
CarrapateenaMI-CuIOCG900 Mt at 0.56% Cu, 0.24 g/t Au, 3 g/t Ag [217]
Oak Dam EastMI-FeIOA560 Mt at 41%–56% Fe, 0.2% Cu, 690 ppm U [132]
Oak Dam WestMI-CuIOCG1340 Mt at 0.66% Cu, 0.33 g/t Au [218]
Olympic Dam 2MI-CuIOCG9640 Mt at 0.58% Cu, 0.19 kg/t U3O8, 0.26 g/t Au, 1.0 g/t Ag+ 1740 Mt at 1.49% Cu, 0.44 kg/t U3O8, 0.58 g/t Au, 3 g/t Ag [217]
Prominent HillMI-CuIOCG162 Mt at 0.94% Cu, 0.81 g/t Au, 3.0 g/t Ag (UG sulfide) [217]
HillsideMI-CuFe-rich skarn-Cu337 Mt at 0.56% Cu, 0.14 g/t Au [219]
Cloncurry district and host Mount Isa Province (sedimentary basin with additional volcano-plutonic hosts)
Elaine 1MI-CuSkarn-hosted Cu-Au26.1 Mt at 0.56% Cu, 0.09 g/t Au [220]
Elaine-DorothyMAC-USkarn-hosted U-REE0.83 Mt at 280 ppm U3O8, 3200 ppm TREO [221]
EloiseMI-CuISCG4.8 Mt at 2.4% Cu, 0.6 g/t Au [222]
Ernest Henry 2MI-CuIOCG167 Mt at 1.1% Cu, 0.5 g/t Au (historic, pre-mining) [129]
Current: 101.5 Mt at 1.25% Cu, 0.73 g/t Au [223]
JerichoMI-CuISCG19.2 Mt at 2.0% Cu, 0.4 g/t Au [224]
Eva Cu 3MI-CuIOCG260.7 Mt at 0.42% Cu, 0.04 g/t Au [225]
KalmanMI-CuISCG39.2 Mt at 0.53% Cu, 0.27 g/t Au, 1.5 g/t Ag, 0.01% Mo, 2.1 g/t Re [226]
Mary KathleenMAC-USkarn-hosted U9.5 Mt at 1300 ppm U3O8 [227]
MerlinMAC-MoMAC Mo6.4 Mt at 1.5% Mo, 26 g/t Re (reserves) [95]
Mt Elliot-SwanMI-CuIOCG–ISCG353.7 Mt at 0.6% Cu, 0.35 g/t Au [228]
Mt DoreMAC-CuBreccia-hosted Cu-Au110.4 Mt at 0.55% Cu, 0.10 g/t Au, 0.05% Pb, 0.30% Zn [228]
RocklandsMI-CuIOCG56.7 Mt at 0.64% Cu, 294 ppm Co, 0.15 ppm Au, 5.3% magnetite; 178 Mt at 15% magnetite [229]
Tick HillMAC-AuAlbitite-hosted Au0.706 Mt at 22.52 g/t Au (mined) [131]
ValhallaMI-UAlbitite-hosted U34.7 Mt at 830 ppm U3O8 [230]
Other representative deposits
Johnnies RewardMI-CuIOCG (at granulite facies)2.19 Mt at 0.7 g/t Au, 0.4% Cu [231]
Peko (tailings)MI-CuIOCG3.75 Mt at 1.14 g/t Au, 0.25% Cu, 0.11% Co, 80% magnetite [232]
Savage RiverMI-FeIOA471.8 Mt at 68.3% Fe, 0.04% Ni, 0.71% TiO2, 1.43% MgO, 0.007% P, 0.35% V, 0.08% S [233]
WarregoMI-FeIOCG6.95 Mt at 2.0% Cu, 6.6 g/t Au, 0.32% Bi [5]
1 Metals in bold are critical minerals on the Canadian list [234]. 2 Case studies studied in more detail by the authors marked. 3 Eva Cu includes the Little Eva, Turkey Creek, Blackard, Scanlan, Bedford, Lady Clayre, and Ivy Ann deposits.
Table A2. Representative examples of the mineral resources and their deposit types within the MIAC systems of Canada, the United States, and of a MI-Ni in Brazil. The deposits in Idaho and the NICO deposit are representative deposits formed in a sedimentary-dominant environment. The southeast Missouri district Boss, Pea Ridge, and Pilot Knob deposits, as well as the Sue-Dianne deposit, are representative of deposits formed in a felsic-to-intermediate volcano-plutonic dominant environment. Jaguar is representative of deposits formed in an ultramafic-to-mafic volcano-plutonic dominant environment. Deposits or their districts are located in Figure 1.
Table A2. Representative examples of the mineral resources and their deposit types within the MIAC systems of Canada, the United States, and of a MI-Ni in Brazil. The deposits in Idaho and the NICO deposit are representative deposits formed in a sedimentary-dominant environment. The southeast Missouri district Boss, Pea Ridge, and Pilot Knob deposits, as well as the Sue-Dianne deposit, are representative of deposits formed in a felsic-to-intermediate volcano-plutonic dominant environment. Jaguar is representative of deposits formed in an ultramafic-to-mafic volcano-plutonic dominant environment. Deposits or their districts are located in Figure 1.
DepositsTotal Resources (Unless Indicated Otherwise) 1
NameClassType
United States
Ram/Sunshine/Sunshine East, IdahoMI-CoMI-Co5.77 Mt at 0.44% Co, 0.69% Cu, 0.53 g/t Au [123]
Iron Creek, IdahoMI-CoISi-Co4.45 Mt at 0.19% Co, 0.73% Cu (indicated), 1.23 Mt at 0.08% Co, 1.34% Cu (inferred) [235]
Boss, SE MissouriMI-CuIOCG40 Mt at 0.83% Cu, 18% Fe, 0.035% Co (historic) [236]
Pea Ridge, SE Missouri 2MI-FeIOA160.6 Mt at ~53%–55% Fe; 0.2 Mt at 12% TREE (historic) [100]
Pilot Knob, SE MissouriMI-FeIOA20 Mt at 35 to 40% Fe (produced) [113]
Pumpkin Hollow, Yerrington districtMI-CuIOCG501.7 Mt at 0.452% Cu, 0.07 g/t Au, 1.85 g/t Ag (open pit, measured and indicated), 25.4 Mt at 0.358% Cu, 0.03 g/t Au, 1.37 g/t Ag (open pit, inferred); 49.1 Mt at 1.39% Cu, 0.17 g/t Au, 3.98 g/t Ag, 17.8% Fe (underground, measured and indicated), 26.5 Mt at 1.09% Cu, 0.10 g/t Au, 2.19 g/t Ag, 12.8% Fe (underground, inferred) [237]
Moonlight-Superior project, CaliforniaMI-CuIOCG402.83 Mt at 0.31% Cu, 1.85 g/t Ag, 0.012 g/t Au (measured and indicated); 64.59 Mt at 0.31% Cu, 0.77 g/t Ag, 0.005 g/t Au (inferred) [238]
Coles Hill, VirginiaMAC-UAlbitite-hosted U119 Mt at 0.056% U3O8 (indicated resources) [239]
Buena VistaMI-FeIOA232 Mt at 18.6% Fe [240]
Canada 2
NICOMI-CoIOx-Co33 Mt at 1.02 g/t Au, 0.12% Co, 0.14% Bi, 0.04% Cu [241]
Sue-Dianne 2MI-CuIOCG8.4 Mt at 0.80% Cu, 0.07 g/t Au, 3.2 g/t Ag [242]
MichelinMI-UAlbitite-hosted U42.7 Mt at 0.098% U3O8 [243]
Upper C Moran LakeMI-UAlbitite-hosted U6.9 Mt at 0.034% U3O8, 0.078% V2O5 (indicated, historic 3) + 5.3 Mt at 0.024% U3O8, 0.058% V2O5 (inferred, historic 3) [244]
Josette 2MI-REEIOA-REE6.9 Mt at 2.7% REE2O3 (=1.83% LREE, 0.89% HREE) [245]
FostungMI-WW skarn12.4 Mt at 0.2% WO3 (historic) [246]
KiggavikMAC-UAlbitite-hosted U10.4 Mt at 0.47 U3O8 [247]
Lac CinquanteMAC-UAlbitite-hosted U2.8 Mt at 0.693% U3O8, 20.6 g/t Ag 0.167% Mo, 0.25% Cu (historic) [248]
Werner LakeMI-CoIsi-Co57.9kt at 0.51% Co, 0.25% Cu, 0.27% As, 0.22 g/t Au (indicated); 6.3kt at 0.48% Co, 0.14% Cu, 0.30% As, 0.24 g/t Au (inferred) [249]
Jaguar, Carajás, BrazilMI-NiIOA-Ni58.6 Mt at 0.96% Ni [250]
1 Metals in bold are critical minerals on the Canadian list [234]. 2 Case studies studied in more detail by the authors. 3 The past-producing five-element, and vein-type mines of the northern Great Bear magmatic zone have produced 2.113 Mt, containing 54.216 Moz of Ag, 13.401 Mlbs of U oxides, 8.4 kt of Cu, 450 g of Rd, 0.13 kt of Ni, and 0.23 kt of Co [56]. 3 Historic estimates stated not to be relied upon by [244] but serving as a case example of V enrichment in deposits of MIAC systems.
Table A3. Representative examples of the mineral resources and their deposit types within the MIAC systems of Sweden and Finland. Districts are located in Figure 1.
Table A3. Representative examples of the mineral resources and their deposit types within the MIAC systems of Sweden and Finland. Districts are located in Figure 1.
DepositsTotal Resources (Unless Indicated Otherwise) 1
NameClassType
Sweden
Per GeigerMI-FeIOA-REE734 Mt at 47.3% Fe, 2.3% P, 0.18% TREO [153]
Kiruna (+Konsuln)MI-FeIOA1437 Mt at 59.8% Fe, 0.33% P, 0.017% TREO [154]
MalmbergetMI-FeIOA1570 Mt at 53.4% Fe, 0.57% P, 0.022% TREO [154]
SvappavaaraMI-FeIOA785 Mt at 46.2% Fe, 0.47% P [154]
GrangesbergMI-FeIOA148.3 Mt at 41.3% Fe, 0.81% P [251]
Nautanen NorthMI-CuIOCG21 Mt at 1.46% Cu, 0.78 g/t Au, 6 g/t Ag, 99 g/t Mo [252]
KiskamavaaraMI-CuIOCG7.67 Mt at 0.25% Cu, 0.04% Co [214]
Finland
HannukainenMI-CuIOCG221 Mt at 32% Fe, 0.17% Cu, 0.077 g/t Au, 135 ppm Co [253]
SahavaaraMI-FeFe skarn86.8 Mt at 39.82% Fe, 1.93% S (measured + indicated) [254]
JuomasuoMAC-AuAlbitite-hosted Au-Co2.37 Mt at 0.13% Co, 4.6 g/t Au + 5.04 Mt at 0.12% Co [255]
1 Metals in bold are critical minerals on the Canadian list [234].

References

  1. Hofstra, A.; Lisitsin, V.; Corriveau, L.; Paradis, S.; Peter, J.; Lauzière, K.; Lawley, C.; Gadd, M.; Pilote, J.; Honsberger, I.; et al. Deposit Classification Scheme for the Critical Minerals Mapping Initiative Global Geochemical Database; Open-File Report 2021-1049; U.S. Geological Survey: Denver, CO, USA, 2021; 60 p.
  2. Williams, P.J.; Barton, M.D.; Johnson, D.A.; Fontboté, L.; de Haller, A.; Mark, G.; Oliver, N.H.S.; Marschik, R. Iron-oxide copper-gold deposits: Geology, space-time distribution, and possible modes of origin. Econ. Geol. 2005, 100, 371–405. [Google Scholar]
  3. Barton, M.D. Iron oxide(–Cu–Au–REE–P–Ag–U–Co) systems. In Treatise on Geochemistry, 2nd ed.; Holland, H.D., Turekian, K.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; Volume 13, pp. 515–541. [Google Scholar]
  4. Porter, T.M. Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 4—Advances in the Understanding of IOCG Deposits; PGC Publishing: Adelaide, Australia, 2010; p. 294. [Google Scholar]
  5. Porter, T.M. Current understanding of iron oxide associated alkali-altered mineralised systems. Part 1—An overview. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 5–32. [Google Scholar]
  6. Corriveau, L.; Montreuil, J.-F.; Blein, O.; Ehrig, K.; Potter, E.G.; Fabris, A.; Clark, J. Mineral systems with IOCG and affiliated deposits: Part 2—Geochemical footprints. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 159–204. [Google Scholar]
  7. Corriveau, L.; Montreuil, J.-F.; Potter, E.G.; Blein, O.; De Toni, A.F. Mineral systems with IOCG and affiliated deposits: Part 3—Metal pathways and ore deposit model. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 205–245. [Google Scholar]
  8. Hitzman, M.W.; Oreskes, N.; Einaudi, M.T. Geological characteristics and tectonic setting of Proterozoic iron oxide (Cu-U-Au-LREE) deposits. Precambrian Res. 1992, 58, 241–287. [Google Scholar]
  9. Groves, D.I.; Bierlein, F.P.; Meinert, L.D.; Hitzman, M.W. Iron oxide copper-gold (IOCG) deposits through Earth history. Implications for origin, lithospheric setting, and distinction from other epigenetic iron oxide deposits. Econ. Geol. 2010, 105, 641–654. [Google Scholar]
  10. Corriveau, L.; Montreuil, J.-F.; Potter, E.G.; Ehrig, K.; Clark, J.; Mumin, A.H.; Williams, P.J. Mineral systems with IOCG and affiliated deposits: Part 1—Metasomatic footprints of alteration facies. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 113–158. [Google Scholar]
  11. Corriveau, L.; Montreuil, J.-F.; De Toni, A.F.; Potter, E.G.; Percival, J.B. Mapping mineral systems with IOCG and affiliated deposits: A facies approach. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 69–111. [Google Scholar]
  12. Corriveau, L.; Mumin, A.H.; Setterfield, T. IOCG environments in Canada: Characteristics, geological vectors to ore and challenges. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 311–344. [Google Scholar]
  13. Corriveau, L.; Potter, E.G.; Mumin, A.H. Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; 424 p. [Google Scholar]
  14. Corriveau, L.; Williams, P.J.; Mumin, A.H. Alteration vectors to IOCG mineralization—From uncharted terranes to deposits. In Exploring for Iron Oxide Copper–Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 89–110. [Google Scholar]
  15. Corriveau, L.; Montreuil, J.-F.; Potter, E.G. Alteration facies linkages among IOCG, IOA and affiliated deposits in the Great Bear magmatic zone, Canada. Econ. Geol. 2016, 111, 2045–2072. [Google Scholar]
  16. Montreuil, J.-F.; Corriveau, L.; Davis, W. Tectonomagmatic evolution of the southern Great Bear magmatic zone (Northwest Territories, Canada)—Implications on the genesis of iron oxide alkali-altered hydrothermal systems. Econ. Geol. 2016, 111, 2111–2138. [Google Scholar]
  17. Montreuil, J.-F.; Corriveau, L.; Potter, E.G.; De Toni, A.F. On the relation between alteration facies and metal endowment of iron oxide–alkali-altered systems, southern Great Bear magmatic zone (Canada). Econ. Geol. 2016, 111, 2139–2168. [Google Scholar] [CrossRef]
  18. Potter, E.G.; Corriveau, L.; Kjarsgaard, B. Paleoproterozoic iron oxide apatite (IOA) and iron oxide-copper-gold (IOCG) mineralization in the East Arm Basin, Northwest Territories, Canada. Can. J. Earth Sci. 2020, 57, 167–183. [Google Scholar]
  19. Corriveau, L.; Blein, O.; Gervais, F.; Trapy, P.H.; De Souza, S.; Fafard, D. Iron-Oxide and Alkali-Calcic Alteration, Skarn and Epithermal Mineralizing Systems of the Grenville Province: The Bondy Gneiss Complex in the Central Metasedimentary Belt of Quebec as a Case Example—A Field Trip to the 14th Society for Geology Applied to Mineral Deposits (SGA) Biennial Meeting; Open File 8349; Geological Survey of Canada: Ottawa, ON, Canada, 2018; 136 p.
  20. Slack, J.; Corriveau, L.; Hitzman, M. Proterozoic Iron Oxide-Apatite (± REE) and Iron Oxide-Copper-Gold and Affiliated Deposits of Southeast Missouri, USA, and the Great Bear Magmatic Zone, Northwest Territories, Canada. Econ. Geol. 2016, 111, 1803–1814. [Google Scholar]
  21. Drejing-Carroll, D.; Hitzman, M.W.; Coller, D. Geology of the Nautanen North Cu-Au-Ag-(Mo) Deposit, Norrbotten, Sweden. Econ. Geol. 2023, 118, 1765–1794. [Google Scholar]
  22. Hamilton, M.; Montreuil, J.-F.; Adlakha, E.; Corriveau, L.; Bain, W. Base, Critical, and Precious Metals Mineralization in the Metasomatic Iron and Alkali-Calcic Systems of the Southern Province in the Sudbury Area: A Geological Guidebook; Open File Report 6391; Ontario Geological Survey: Sudbury, ON, Canada, 2023; 66p. [Google Scholar]
  23. Fabris, A.; Michaelsen, B. Reference Drillholes from IOCG and Associated Deposits in South Australia; Report Book 2024/00008; Department for Energy and Mining: Adelaide, Australia, 2024; 60p.
  24. Blein, O.; Corriveau, L.; Montreuil, J.-F.; Ehrig, K.; Fabris, A.; Reid, A.; Pal, D. Geochemical signatures of metasomatic ore systems hosting IOCG, IOA, albite-hosted uranium and affiliated deposits: A tool for process studies and mineral exploration. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 263–298. [Google Scholar]
  25. Blein, O.; Harlaux, M.; Corriveau, L.; Lynch, E.P.; Niiranen, T.; Lisitsin, V.; Ehrig, K.; Montreuil, J.-F.; Gourcerol, B. Alteration footprints of metasomatic iron and alkali-calcic systems in the Northern Norrbotten, Sweden. In Proceedings of the 17th Biennial Meeting of the SGA, Zürich, Switzerland, 28 August–1 September 2023. [Google Scholar]
  26. Acosta-Góngora, P.; Potter, E.G.; Lawley, C.J.M.; Corriveau, L.; Sparkes, G. Geochemical characterization of the Central Mineral Belt U±Cu±Mo±V mineralization, Labrador, Canada: Evaluation of IOCG potential and application of unsupervised machine-learning. J. Geochem. Expl. 2022, 237, 106995. [Google Scholar]
  27. Montreuil, J.-F.; Potter, E.; Corriveau, L.; Davis, W.J. Element mobility patterns in magnetite-group IOCG systems: The Fab IOCG system, Northwest Territories, Canada. Ore Geol. Rev. 2016, 72, 562–584. [Google Scholar]
  28. Montreuil, J.-F.; Corriveau, L.; Potter, E. Formation of albitite-hosted uranium within IOCG systems: The Southern Breccia, Great Bear magmatic zone, Northwest Territories, Canada. Mineral. Depos. 2015, 50, 293–325. [Google Scholar]
  29. Corriveau, L.; Lauzière, K.; Montreuil, J.-F.; Potter, E.G.; Hanes, R.; Prémont, S. Dataset of Geochemical Data from Iron Oxide Alkali-Altered Mineralizing Systems of the Great Bear Magmatic Zone (NWT); Open File 7643; Geological Survey of Canada: Ottawa, ON, Canada, 2015; 19p.
  30. Enkin, R.J.; Corriveau, L.; Hayward, N. Metasomatic alteration control of petrophysical properties in the Great Bear magmatic zone (Northwest Territories, Canada). Econ. Geol. 2016, 111, 2073–2085. [Google Scholar]
  31. Zeng, L.-P.; Zhao, X.-F.; Spandler, C.; Mavrogenes, J.A.; Mernagh, T.P.; Liao, W.; Fan, Y.-Z.; Hu, Y.; Fu, B.; Li, J.-W. The role of iron-rich hydrosaline liquids in the formation of Kiruna-type iron oxide–apatite deposits. Sci. Adv. 2024, 10, eadk2174. [Google Scholar]
  32. Xu, X.; Steele-MacInnis, M.; Bain, W.; Tornos, F.; Hanchar, J.M.; Lamadrid, H.M.; Lehmann, B.; Xu, X.; Steadman, J.A.; Bottrill, R.S.; et al. Magnetite-apatite ores record widespread involvement of molten salts. Geology 2024, 52, 417–422. [Google Scholar]
  33. Skirrow, R.G. Iron oxide copper-gold (IOCG) deposits—A review (part 1): Settings, mineralogy, ore geochemistry and classification. Ore Geol. Rev. 2022, 140, 104569. [Google Scholar]
  34. Skirrow, R.G.; Murr, J.; Schofield, A.; Huston, D.L.; van der Wielen, S.E.; Czarnota, K.; Coghlan, R.; Highet, L.M.; Connolly, D.; Doublier, M.; et al. Mapping iron oxide Cu-Au (IOCG) mineral potential in Australia using a knowledge-driven mineral systems-based approach. Ore Geol. Rev. 2019, 113, 103011. [Google Scholar]
  35. Williams, P.J. Classifying IOCG deposits. In Exploring for Iron Oxide Copper–Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 13–21. [Google Scholar]
  36. Wyborn, L.A.I.; Heinrich, C.A.; Jaques, A.L. Australian Proterozoic mineral systems: Essential ingredients and mappable criteria. In Proceedings of the AusIMMM Annual Conference, Transactions, Melbourne, Australia, 5–9 August 1994; pp. 109–115. [Google Scholar]
  37. Corriveau, L.; Montreuil, J.-F.; Blein, O.; Potter, E.; Ansari, M.; Craven, J.; Enkin, R.; Fortin, R.; Harvey, B.; Hayward, N.; et al. Metasomatic iron and alkali calcic (MIAC) system frameworks: A TGI-6 task force to help de-risk exploration for IOCG, IOA and affiliated primary critical metal deposits. In Scientific Presentation 127; Geological Survey of Canada: Ottawa, ON, Canada, 2021; 105p. [Google Scholar]
  38. Corriveau, L.; Potter, E.G. Advancing exploration for iron oxide-copper-gold and affiliated deposits in Canada: Context, scientific overview, outcomes and impacts. In Canada’s Northern Shield: New Perspectives from the Geo-Mapping for Energy and Minerals Program; Pehrsson, S., Wodicka, N., Rogers, N., Percival, J., Eds.; Bulletin 612; Geological Survey of Canada: Ottawa, ON, Canada, 2024; pp. 55–98. [Google Scholar]
  39. Corriveau, L.; Mumin, A.H. Exploring for iron oxide copper–gold deposits: The need for case studies, classifications and exploration vectors. In Exploring for Iron Oxide Copper–Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 1–12. [Google Scholar]
  40. Sillitoe, R.H. Iron oxide-copper-gold deposits: An Andean view. Mineral. Depos. 2003, 38, 787–812. [Google Scholar]
  41. Slack, J. Descriptive and Geoenvironmental Model for Cobalt–Copper–Gold Deposits in Metasedimentary Rocks; Scientific Investigations Report 2010–5070–G; U.S. Geological Survey: Denver, CO, USA, 2013; 218 p.
  42. Reid, A. The Olympic Cu-Au Province, Gawler Craton: A review of the lithospheric architecture, geodynamic setting, alteration systems, cover successions and prospectivity. Minerals 2019, 9, 371. [Google Scholar] [CrossRef]
  43. Skirrow, R.G. Hematite-group IOCG ± U deposits: An update on their tectonic settings, hydrothermal characteristics, and Cu-Au-U mineralizing processes. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 27–51. [Google Scholar]
  44. Schutesky, M.E.; Oliveira, C.; Hagemann, S.; Monteiro, L.V.S. A thematic issue dedicated to the 50 years of the discovery of the Carajás Mineral Province, Brazil. Ore Geol. Rev. 2021, 129, 103819. [Google Scholar]
  45. Hickin, A.S.; Ootes, L.; Orovan, E.A.; Brzozowski, M.J.; Northcote, B.K.; Rukhlov, A.S.; Bain, W.M. Critical Minerals and Mineral Systems in British Columbia; Paper 2024–01; British Columbia Geological Survey: Victoria, Canada, 2024; pp. 13–51.
  46. Rukhlov, A.S. Review of Metallic Mineralization in Alberta with Emphasis on Gold Potential; ERCB/AGS Open File Report 2011–01; Energy Resources Conservation Board: Edmonton, AB, Canada, 2011; 93 p.
  47. Avery, G.; Adlakha, E.; Powell, J.W.; Tschirhart, V. Petrography and Geochemistry of the Lac Cinquante Uranium Deposit, Nunavut; Canada; Geological Survey of Canada: Ottawa, ON, Canada, 2024.
  48. Burron, I.; Fayek, M.; Brown, J.; Quirt, D. Remnants of a 1.55 Ga hybrid between metasomatic iron alkali-calcic and unconformity-related uranium environments in the Kiggavik Region, Nunavut, Canada. Econ. Geol. 2024, 119, 1861–1888. [Google Scholar]
  49. Corriveau, L.; Nadeau, O.; Montreuil, J.-F.; Desrochers, J.-P. Report of Activities for the Core Zone: Strategic Geomapping and Geoscience to Assess the Mineral Potential of the Labrador Trough for Multiple Metals IOCG and Affiliated Deposits, Canada; Open File 7714; Geological Survey of Canada: Ottawa, ON, Canada, 2014; 12 p.
  50. Montreuil, J.-F.; Desrochers, J.-P.; Masters, J. Exploration Report (July and August 2013 program) on the Sagar Property, Romanet Horst, Labrador Trough, Québec, Canada; GM 68408; Ministère de l’Énergie et des Ressources naturelles: Québec, QC, Canada, 2014; 80 p. [Google Scholar]
  51. Clark, T.; Gobeil, A.; Chevé, S. Alterations in IOCG-type and related deposits in the Manitou Lake area, Eastern Grenville Province, Québec. In Exploring for Iron Oxide Copper–Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St John’s, NL, Canada, 2010; pp. 127–146. [Google Scholar]
  52. Hildebrand, R.S. Kiruna-type deposits: Their origin and relationship to intermediate subvolcanic plutons in the Great Bear Magmatic Zone, northwest Canada. Econ. Geol. 1986, 81, 640–659. [Google Scholar] [CrossRef]
  53. Hildebrand, R.S.; Hoffman, P.F.; Housh, T.; Bowring, S.A. The nature of volcano-plutonic relations and shapes of epizonal plutons of continental arcs as revealed in the Great Bear magmatic zone, northwestern Canada. Geosphere 2010, 6, 812–839. [Google Scholar] [CrossRef]
  54. Reardon, N.C. Altered Rocks and Magnetite—Apatite—Actinolite Deposits Associated with the Mystery Island Intrusive Suite, Echo Bay, District of Mackenzie; Open File 2507; Geological Survey of Canada: Ottawa, ON, Canada, 1992; 89 p.
  55. Mumin, A.H.; Corriveau, L.; Somarin, A.K.; Ootes, L. Iron oxide copper-gold-type polymetallic mineralisation in the Contact Lake Belt, Great Bear magmatic zone, Northwest Territories, Canada. Expl. Mining Geol. 2007, 16, 187–208. [Google Scholar]
  56. Mumin, A.H.; Somarin, A.K.; Jones, B.; Corriveau, L.; Ootes, L.; Camier, J. The IOCG-porphyry-epithermal continuum of deposit types in the Great Bear magmatic zone, Northwest Territories, Canada. In Exploring for Iron Oxide Copper–Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 59–78. [Google Scholar]
  57. Mumin, A.H. Echo Bay IOCG Thematic Map Series: Geology, Structure and Hydrothermal Alteration of a Stratovolcano Complex, Northwest Territories, Canada; Open File 7807; Geological Survey of Canada: Ottawa, ON, Canada, 2015; 19 p.
  58. Jackson, V.A.; Ootes, L.; Pierce, K.L.; Bennett, V.; Smar, L.; Mackay, D.; Sandeman, H.A. Geology of the South-Central Wopmay Orogen, Northwest Territories (Parts of NTS 86B, 86C, And 86D); Results from the South Wopmay Bedrock Mapping Project; Open File 2017-01; Northwest Territories Geological Survey: Yellowknife, NT, Canada, 2022; 103 p.
  59. Potter, E.G.; Acosta-Góngora, P.; Corriveau, L.; Montreuil, J.-F.; Yang, Z. Uranium enrichment processes in iron oxide and alkali-calcic alteration systems as revealed by trace element signatures of uraninite. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 325–345. [Google Scholar]
  60. Somarin, A.K.; Zhou, L.; Zheng, G.; Ma, X. Hydrothermal mineralization and mineral chemistry of arsenides and sulfarsenides in the Fe-Co-Ni-As-S system and introduction of three unique minerals, Port Radium deposit, Canada. Minerals 2024, 14, 85. [Google Scholar] [CrossRef]
  61. De Toni, A.F. Les Paragénèses à Magnétite des Altérations Associées aux Systèmes à Oxydes de fer et Altérations en Éléments Alcalins, Zone Magmatique du Grand lac de l’Ours. Master’s Thesis, Institut national de la Recherche Scientifique, Québec, QU, Canada, 2016; 534 p. [Google Scholar]
  62. Ootes, L.; Snyder, D.; Davis, W.J.; Acosta-Góngora, P.; Corriveau, L.; Mumin, A.H.; Montreuil, J.-F.; Gleeson, S.A.; Samson, I.A.; Jackson, V.A. A Paleoproterozoic Andean-type iron oxide copper-gold environment, the Great Bear magmatic zone, Northwest Canada. Ore Geol. Rev. 2017, 81, 123–139. [Google Scholar]
  63. Putnis, A. Fluid–mineral interactions: Controlling coupled mechanisms of reaction, mass transfer and deformation. J. Petrol. 2021, 62, egab092. [Google Scholar]
  64. Zhao, X.-F.; Chen, H.; Zhao, L.; Zhou, M.-F. Linkages among IOA, skarn, and magnetite-group IOCG deposits in China: From deposit studies to mineral potential assessment. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 83–407. [Google Scholar]
  65. Marschik, R.; Fontboté, L. The Candelaria-Punta del Cobre iron oxide Cu-Au (-Zn-Ag) deposits, Chile. Econ. Geol. 2001, 96, 1799–1826. [Google Scholar]
  66. Oliver, N.H.S.; Cleverley, J.S.; Mark, G.; Pollard, P.J.; Fu, B.; Marshall, L.C.; Rubenach, M.J.; Williams, P.J.; Baker, T. Modeling the role of sodic alteration in the genesis of iron oxide-copper-gold deposits: Eastern Mount Isa Block, Australia. Econ. Geol. 2004, 99, 1145–1176. [Google Scholar] [CrossRef]
  67. Oliver, N.H.S.; Butera, K.M.; Rubenach, M.J.; Marshall, L.J.; Cleverley, J.S.; Mark, G.; Tullemans, F.; Esser, D. The protracted hydrothermal evolution of the Mount Isa Eastern Succession: A review and tectonic implications. Precambrian Res. 2008, 163, 108–130. [Google Scholar]
  68. Engvik, A.K.; Corfu, F.; Solli, A.; Austrheim, H. Sequence and timing of mineral replacement reactions during albitisation in the high-grade Bamble lithotectonic domain, S-Norway. Precambrian Res. 2017, 291, 1–16. [Google Scholar] [CrossRef]
  69. Barton, M.D.; Dilles, J.H.; Girardi, J.D.; Haxel, G.B.; Johnson, D.A.; Kreiner, D.C.; Seedorff, E.; Zurcher, L. Jurassic igneous-related metallogeny of southwestern North America. In Great Basin Evolution and Metallogeny, Proceedings of the Geological Society of Nevada, Symposium, Reno Sparks, Reno, NV, USA, 14 May 2010; Steininger, R.C., Pennell, W.M., Eds.; Destech Publications Inc.: Lancaster, PA, USA, 2011; Volume 1, pp. 373–396. [Google Scholar]
  70. Mark, G.; Oliver, N.H.S.; Williams, P.J. Mineralogical and chemical evolution of the Ernest Henry Fe oxide-Cu-Au ore system, Cloncurry district, northwest Queensland, Australia. Mineral. Depos. 2006, 40, 769–801. [Google Scholar] [CrossRef]
  71. Warr, L.M. IMA–CNMNC approved mineral symbols. Min. Mag. 2021, 85, 291–320. [Google Scholar] [CrossRef]
  72. Ehrig, K.; McPhie, J.; Kamenetsky, V. Geology and mineralogical zonation of the Olympic Dam iron oxide Cu-U-Au-Ag deposit, South Australia. in Hedenquist, J.W., Harris, M.; Camus, F., eds., Geology and genesis of major copper deposits and districts of the world: A tribute to Richard, H. Sillitoe. Soc. Econ. Geol. Spec. Publ. 2012, 16, 237–267. [Google Scholar]
  73. Keyser, W.; Ciobanu, C.L.; Ehrig, K.; Dmitrijeva, M.; Wade, B.P.; Courtney-Davies, L.; Verdugo-Ihl, M.; Cook, N.J. Skarn-style alteration in Proterozoic metasedimentary protoliths hosting IOCG mineralization: The Island Dam Prospect, South Australia. Mineral. Depos. 2022, 57, 1227–1250. [Google Scholar] [CrossRef]
  74. Chen, H.; Clark, A.H.; Kyser, T.K.; Ullrich, T.D.; Baxter, R.; Chen, Y.; Moody, T.C. Evolution of the giant Marcona-Mina Justa iron oxide-copper-gold district, south-central Peru. Econ. Geol. 2010, 105, 155–185. [Google Scholar] [CrossRef]
  75. del Real, I.; Thompson, J.F.H.; Carriedo, J. Lithological and structural controls on the genesis of the Candelaria-Punta del Cobre iron oxide copper-gold district, Northern Chile. Ore Geol. Rev. 2018, 102, 106–153. [Google Scholar] [CrossRef]
  76. Chen, H.; Zhao, L. Mineralization, alteration, and fluid compositions in selected Andean IOCG deposits. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 365–381. [Google Scholar]
  77. Daliran, F.; Stosch, H.-G.; Williams, P.J.; Jamali, H.; Dorri, M.-B. Early Cambrian IOA-REE, U/Th and Cu(Au)-Bi-Co-Ni-Ag-As-sulphide-U/Th deposits of the Bafq district, East-Central Iran. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 409–424. [Google Scholar]
  78. Harlov, D.E.; Meighan, C.J.; Kerr, I.D.; Samson, I.M. 2016, Mineralogy, chemistry, and fluid-aided evolution of the Pea Ridge Fe oxide- (Y + REE) deposit, southeast Missouri, USA. Econ. Geol. 2016, 111, 1963–1984. [Google Scholar] [CrossRef]
  79. Bonyadi, Z.; Davidson, G.J.; Mehrabi, B.; Meffre, S.; Ghazban, F. Significance of apatite REE depletion and monazite inclusions in the brecciated Se-Chahun iron oxide–apatite deposit, Bafq district, Iran: Insights from paragenesis and geochemistry. Chem. Geol. 2011, 281, 253–269. [Google Scholar] [CrossRef]
  80. Hitzman, M.W.; Valenta, R.K. Uranium in iron oxide-copper-gold (IOCG) systems. Econ. Geol. 2005, 100, 1657–1661. [Google Scholar] [CrossRef]
  81. Kirschbaum, M.J.; Hitzman, M.W. Guelb Moghrein: An unusual carbonate-hosted iron oxide copper-gold deposit in Mauritania, northwest Africa. Econ. Geol. 2016, 111, 763–770. [Google Scholar]
  82. Cave, B.; Healy, M.; Overall, D.; Lilly, R. Genesis of the Mount Colin Cu-Au deposit, NW Queensland: Evidence from geology, multi-mineral U–Pb geochronology, S-, O- and C-isotope constraints. J. Geochem. Explor. 2024, 256, 107350. [Google Scholar] [CrossRef]
  83. Zharikov, V.A.; Pertsev, F.N.; Rusinov, V.L.; Callegari, E.; Fettes, D.J. Metasomatism and Metasomatic rocks: Recommendations by the IUGS Subcommission on the Systematics of Metamorphic Rocks, Web version 01.02.07. Available online: www.bgs.ac.uk/scmr/home.html (accessed on 3 May 2012).
  84. Kreiner, D.; Barton, M.D. Sulfur-poor intense acid hydrothermal alteration: A distinctive hydrothermal environment. Ore Geol. Rev. 2017, 88, 174–187. [Google Scholar] [CrossRef]
  85. Ehrig, K.; Kamenetsky, V.S.; McPhie, J.; Macmillan, E.; Thompson, J.; Kamenetsky, M.; Maas, R. Staged formation of the supergiant Olympic Dam uranium deposit, Australia. Geology 2021, 49, 1312–1316. [Google Scholar] [CrossRef]
  86. Zhao, X.-F.; Zeng, L.-P.; Liao, W.; Fan, Y.-Z.; Hofstra, A.H.; Emsbo, P.; Hu, H.; Wen, G.; Li, J.W. Iron oxide–apatite deposits form from hydrosaline liquids exsolved from subvolcanic intrusions. Mineral. Depos. 2024, 59, 655–669. [Google Scholar]
  87. Potter, E.G.; Montreuil, J.-F.; Corriveau, L.; Davis, W. The Southern Breccia metasomatic uranium system of the Great Bear magmatic zone, Canada: Iron oxide-copper-gold (IOCG) and albitite-hosted uranium linkages. In Ore Deposits: Origin, Exploration, and Exploitation; Decrée, S., Robb, L., Eds.; Geophysical Monograph 242, American Geophysical Union; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2019; pp. 109–130. [Google Scholar]
  88. Seymour, N.M.; Singleton, J.S.; Gomila, R.; Arancibia, G.; Ridley, J.; Gevedon, M.L.; Stockli, D.F.; Seman, S.M. Sodic-calcic alteration and transpressional shear along the Atacama fault system during IOCG mineralization, Copiapó, Chile. Mineral. Depos. 2024, 59, 1295–1323. [Google Scholar] [CrossRef]
  89. Rubenach, M. Structural controls of metasomatism on a regional scale. In Metasomatism and the Chemical Transformation of Rock: The Role of Fluids in Terrestrial and Extraterrestrial Processes; Harlov, D.E., Austrheim, H., Eds.; Lecture Notes in Earth System Sciences; Springer: New York, NY, USA, 2012; pp. 93–140. [Google Scholar]
  90. Frietsch, R.; Tuisku, P.; Martinsson, O.; Perdahl, J.-A. Early Proterozoic Cu-(Au) and Fe ore deposits associated with regional Na-Cl metasomatism in northern Fennoscandia. Ore Geol. Rev. 1997, 12, 1–34. [Google Scholar] [CrossRef]
  91. Marshall, L.J.; Oliver, N.H.S. Constraints on hydrothermal fluid pathways within Mary Kathleen Group stratigraphy of the Cloncurry iron-oxide–copper–gold District, Australia. Precambrian Res. 2008, 163, 151–158. [Google Scholar] [CrossRef]
  92. Conor, C.; Raymond, O.; Baker, T.; Teale, G.; Say, P.; Lowe, G. Alteration and mineralisation in the Moonta-Wallaroo copper-gold mining field region, Olympic domain, South Australia. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 147–170. [Google Scholar]
  93. Poulet, T.; Karrech, A.; Regenauer-Lieb, K.; Fisher, L.; Schaubs, P. Thermal–hydraulic–mechanical–chemical coupling with damage mechanics using ESCRIPTRT and ABAQUS. Tectonophysics 2012, 526–529, 124–132. [Google Scholar] [CrossRef]
  94. Austin, J.R.; Blenkinsop, T.G. Local to regional scale structural controls on mineralisation and the importance of a major lineament in the eastern Mount Isa Inlier, Australia: Review and analysis with autocorrelation and weights of evidence. Ore Geol. Rev. 2009, 35, 298–316. [Google Scholar]
  95. Babo, J.; Spandler, C.; Oliver, N.; Brown, M.; Rubenach, M.; Creaser, R.A. The high-grade Mo-Re Merlin deposit, Cloncurry District, Australia: Paragenesis and geochronology of hydrothermal alteration and ore formation. Econ. Geol. 2017, 112, 397–422. [Google Scholar]
  96. Geijer, P.; Ödman, O.H. The Emplacement of the Kiruna Iron Ores and Related Deposits; SER C NR 700; Sveriges Geologiska Undersökning: Uppsala, Sweden, 1974; 48 p.
  97. Meyer, W. Soda metasomatism. In 1986 Regional Resident Geologists Report on Activities; Kustra, C.R., Ed.; Miscellaneous Paper 134; Ontario Geological Survey: Sudbury, ON, Canada, 1987; pp. 256–257. [Google Scholar]
  98. Ray, S.K. The albitite line of northern Rajasthan—A fossil intracontinental rift zone. J. Geol. Soc. India 1990, 86, 413–423. [Google Scholar]
  99. Champion, D.C.; Huston, D.L.; Bastrakov, E.N.; Siegel, C.; Thorne, J.; Gibson, G.M.; Hauser, J. Alteration of mafic igneous rocks of the southern McArthur Basin: Comparison with the Mount Isa region and implications for basin-hosted base metal deposits. In Exploring for the Future: Extended Abstracts; Geoscience Australia: Canberra, Australia, 2020. [Google Scholar]
  100. Schandl, E.S.; Gorton, M.P.; Davis, D.W. Albitization at 1700 ± 2 Ma in the Sudbury—Wanapitei Lake area, Ontario: Implications for deep-seated alkalic magmatism in the Superior Province. Can. J. Earth Sci. 1994, 31, 597–607. [Google Scholar]
  101. Polito, P.A.; Kyser, T.K.; Stanley, C. The Proterozoic, albitite-hosted, Valhalla uranium deposit, Queensland, Australia: A description of the alteration assemblage associated with uranium mineralization in diamond drill hole V39. Mineral. Depos. 2009, 44, 11–40. [Google Scholar]
  102. Kreiner, D.C. Epithermal Style Iron Oxide(-Cu-Au) (=IOCG) Vein Systems and Related Alteration. Ph.D. Thesis, University of Arizona, Tuscon, AZ, USA, 2011; 659 p. [Google Scholar]
  103. Barton, M.D.; Johnson, D.A.; Kreiner, D.C.; Jensen, E.P. Vertical zoning and continuity in Fe oxide (-Cu-Au-Ag-Co-U-P-REE) (or IOCG) systems: Cordilleran insights. In Proceedings of the 12th SGA Biennial Meeting, Uppsala, Sweden, 12–15 August 2013; Volume 3, pp. 1348–1351. [Google Scholar]
  104. Wilde, A. Towards a model for albitite-type uranium. Minerals 2013, 3, 36–48. [Google Scholar] [CrossRef]
  105. Putnis, A. Transient porosity resulting from fluid–mineral interaction and its consequences. Rev. Mineral. Geochem. 2015, 80, 1–23. [Google Scholar]
  106. Yadav, G.S.; Muthamilselvan, A.; Shaji, T.J.; Nanda, L.K.; Rai, A.K. Recognition of a new albitite zone in northern Rajasthan: Its implications on uranium mineralization. Curr. Sci. 2015, 108, 1994–1998. [Google Scholar]
  107. Sparkes, G.W. Uranium Mineralization within the Central Mineral Belt of Labrador: A Summary of the Diverse Styles, Settings and Timing of Mineralization; Open File LAB/1684; Government of Newfoundland and Labrador, Department of Natural Resources; Geological Survey: St. John’s, NL, Canada, 2017; 198 p.
  108. Baidya, A.S.; Sen, A.; Pal, D.C. Textures and compositions of cobalt pentlandite and cobaltian mackinawite from the Madan-Kudan copper deposit, Khetri Copper Belt, Rajasthan, India. J. Earth Syst Sci 2018, 127, 56. [Google Scholar]
  109. Rasilainen, K.; Eilu, P.; Huovinen, I.; Konnunaho, J.; Niiranen, T.; Ojala, J.; Törmänen, T. Quantitative Assessment of Undiscovered Resources in Kuusamo-Type Co-Au Deposits in Finland; Bulletin 410; Geological Survey of Finland: Espoo, Finland, 2020; 32 p. [Google Scholar]
  110. Dwivedy, S.; Sahoo, P.R. Geology and trace element geochemistry of the albitite hosted iron ore mineralization around Khetri copper deposit, India: Implications for an IOA type deposit. Ore Geol. Rev. 2021, 138, 104343. [Google Scholar]
  111. Fabris, A.; Katona, L.; Gordon, G.; Reed, G.; Keeping, T.; Gouthas, G.; Swain, G. Characterising and Mapping Alteration in the Punt Hill Region: A Data Integration Project; Report Book, 2018/00010; Government of South Australia; Department of the Premier and Cabinet: Adelaide, Australia, 2018; 604 p.
  112. Magyarosi, Z.; Baker, J.; MacKenzie, J. Data Compilation and Interpretation Report, Fostung Project, Ontario, Canada. 2012. Available online: https://www.geologyontario.mines.gov.on.ca/persistent-linking?assessment=20000008530 (accessed on 27 June 2021).
  113. Day, W.C.; Slack, J.F.; Ayuso, R.A.; Seeger, C.M. Regional geologic and petrologic framework for iron oxide ± apatite ± rare earth element and iron oxide copper-gold deposits of the Mesoproterozoic St. François Mountains Terrane, Southeast Missouri, USA. Econ. Geol. 2016, 111, 1825–1858. [Google Scholar]
  114. Tornos, F.; Hanchar, J.M.; Steele-MacInnis, M.; Crespo, E.; Kamenetsky, V.S.; Casquet, C. Formation of magnetite-(apatite) systems by crystallizing ultrabasic iron-rich melts and slag separation. Mineral. Depos. 2024, 59, 189–225. [Google Scholar]
  115. Normandeau, P.X.; Harlov, D.E.; Corriveau, L.; Paquette, J.; McMartin, I. Characterization of fluorapatite within iron oxide alkali-calcic alteration systems of the Great Bear magmatic zone: A potential metasomatic process record. Can. Min. 2018, 56, 167–187. [Google Scholar]
  116. Sappin, A.-A.; Perreault, S. Drill Core Pictures and Description of Samples Collected from the REE Kwyjibo Deposit (SOQUEM Warehouse, Val d’Or, QC—October 2015); Open File 8794; Geological Survey of Canada: Ottawa, ON, Canada, 2021; 14 p.
  117. Garcia, V.B.; Schutesky, M.E.; Oliveira, C.G.; Whitehouse, M.J.; Huhn, S.R.B.; Augustin, C.T. The Neoarchean GT-34 Ni deposit, Carajás mineral Province, Brazil: An atypical IOCG-related Ni sulfide mineralization. Ore Geol. Rev. 2020, 127, 103773. [Google Scholar]
  118. Veloso, A.S.R.; Monteiro, L.V.S.; Juliani, C. The link between hydrothermal nickel mineralization and an iron oxide-copper-gold (IOCG) system: Constraints based on mineral chemistry in the Jatobá deposit, Carajás Province. Ore Geol. Rev. 2020, 121, 103555. [Google Scholar]
  119. Mansur, E.; Dare, S.; Ferreira Filho, C.; Miranda, A.C.R.; Monteiro, L.V.S. The distribution of trace elements in sulfides and magnetite from the Jaguar hydrothermal nickel deposit: Exploring the link with IOA and IOCG deposits within the Carajás Mineral Province, Brazil. Ore Geol. Rev. 2023, 152, 105256. [Google Scholar]
  120. Campo Rodriguez, Y.T.; Cook, N.J.; Ciobanu, C.L.; Schutesky, M.E.; King, S.A.; Gilbert, S.; Ehrig, K. Platinum group minerals associated with nickel-bearing sulfides from the Jatobá iron oxide-copper-gold deposit, Carajás domain, Brazil. Minerals 2024, 14, 757. [Google Scholar] [CrossRef]
  121. Acosta-Góngora, P.; Gleeson, S.; Samson, I.; Corriveau, L.; Ootes, L.; Jackson, S.E.; Taylor, B.E.; Girard, I. Origin of sulfur and crustal recycling of copper in polymetallic (Cu-Au-Co-Bi-U ± Ag) iron-oxide-dominated systems of the Great Bear magmatic zone, NWT, Canada. Mineral. Depos. 2018, 53, 353–376. [Google Scholar]
  122. Acosta-Góngora, P.; Gleeson, S.; Samson, I.; Corriveau, L.; Ootes, L.; Taylor, B.E.; Creaser, R.A.; Muehlenbachs, K. Genesis of the Paleoproterozoic NICO iron-oxide-cobalt-gold-bismuth deposit, Northwest Territories, Canada: Evidence from isotope geochemistry and fluid inclusions. Precambrian Res. 2015, 268, 168–193. [Google Scholar]
  123. Sletten, M.; Zelligan, S.; Frost, D.; Yugo, N.; Charbonneau, C.; Cameron, D.P. Idaho Cobalt Operations, Form 43-101F1 Technical Report, Feasibility Study Idaho, USA; M3 Engineering and Technology Corporation for Jervois Mining Limited: Salmon, ID, USA, 2020; 416 p. [Google Scholar]
  124. Goad, R.E.; Mumin, A.H.; Duke, N.A.; Neale, K.L.; Mulligan, D.L. Geology of the Proterozoic iron oxide-hosted NICO cobalt-gold-bismuth, and Sue-Dianne copper-silver deposits, southern Great Bear Magmatic Zone, Northwest Territories, Canada. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2000; Volume 1, pp. 249–267. [Google Scholar]
  125. Acosta-Góngora, G.P.; Gleeson, S.A.; Samson, I.; Ootes, L.; Corriveau, L. Gold refining by bismuth melts in the iron oxide-dominated NICO Au-Co-Bi (±Cu±W) deposit, NWT, Canada. Econ. Geol. 2015, 110, 291–314. [Google Scholar]
  126. Rusk, B.; Oliver, N.; Blenkinsop, T.; Zhang, D.; Williams, P.; Cleverley, J.; Habermann, H. Physical and chemical characteristics of the Ernest Henry iron oxide copper gold deposit, Cloncurry, Queensland, Australia; Implications for IOCG genesis. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 201–218. [Google Scholar]
  127. Zhao, X.-F.; Zhou, M.F.; Su, Z.K.; Li, X.C.; Chen, W.T.; Li, J.W. Geology, geochronology, and geochemistry of the Dahongshan Fe-Cu-(Au-Ag) deposit, Southwest China: Implications for the formation of iron oxide copper-gold deposits in intracratonic rift settings. Econ. Geol. 2017, 112, 603–628. [Google Scholar]
  128. Zeng, M.; Zhang, D.; Zhang, Z.; Li, T.; Li, C.; Wei, C. Structural controls on the Lala iron-copper deposit of the Kangdian metallogenic province, southwestern China: Tectonic and metallogenic implications. Ore Geol. Rev. 2018, 97, 35–54. [Google Scholar]
  129. Lilly, R.; Case, G.; Miller, B. Ernest Henry iron oxide copper-gold deposit. Publ. Australas. Inst. Min. Metall. 2017, 32, 501–506. [Google Scholar]
  130. Jébrak, M. Use of breccias in IOCG(U) exploration: An update review. In Mineral Systems with Iron Oxide Copper-Gold (IOCG) and Affiliated Deposits; Corriveau, L., Potter, E.G., Mumin, A.H., Eds.; Special Paper 52; Geological Association of Canada: St. John’s, NL, Canada, 2022; pp. 315–324. [Google Scholar]
  131. Le, T.X.; Dirks, P.H.G.M.; Sanislav, I.V.; Huizenga, J.M.; Cocker, H.; Manestar, G.N. Geological setting and mineralisation characteristics of the Tick Hill Gold Deposit, Mount Isa Inlier, Queensland, Australia. Ore Geol. Rev. 2021, 137, 104288. [Google Scholar]
  132. Davidson, G.J.; Paterson, H.; Meffre, S.; Berry, R.F. Characteristics and origin of the Oak Dam East breccia-hosted, iron oxide-Cu-U-(Au) deposit: Olympic Dam region, Gawler Craton, South Australia. Econ. Geol. 2007, 102, 1471–1498. [Google Scholar]
  133. Fox, N.; Valenta, R.; Gow, P. Chapter 7: Eloise Cu-Au deposit. NWMP Deposit Atlas; Geological Survey of Queensland: Brisbane, Australia, 2023; pp. 257–279.
  134. Somarin, A.K.; Mumin, A.H. P–T-composition and evolution of paleofluids in the Paleoproterozoic Mag Hill IOCG hydrothermal system, Contact Lake belt, Northwest Territories, Canada. Mineral. Depos. 2014, 49, 199–215. [Google Scholar]
  135. Lobo-Guerrero, S.A. Iron oxide-copper-gold mineralization in the Greater Lufilian Arc, Africa. In Exploring for Iron Oxide Copper-Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 161–175. [Google Scholar]
  136. Gandhi, S.S.; Potter, E.G.; Fayek, M. New constraints on genesis of the polymetallic veins at Port Radium, Great Bear Lake, Northwest Canadian Shield. Ore Geol. Rev. 2018, 96, 28–47. [Google Scholar]
  137. Cave, B.; Lilly, R.; Hand, M.; Varga, J.; Light, S.; Leslie, D.; North, B.; Park, J.; Klingberg, L. A temporal framework for the Carrapateena Iron Oxide Copper-Gold (IOCG) deposit, Eastern Gawler Craton, South Australia. Ore Geol. Rev. 2024, 169, 106092. [Google Scholar]
  138. Hoggard, M.J.; Czarnota, K.; Richards, F.D.; Huston, D.; Jaques, L.; Ghelichkhan, S. Global distribution of sediment-hosted metals controlled by craton edge stability. Nat. Geosci. 2020, 13, 504–510. [Google Scholar]
  139. Tiddy, C.J.; Giles, D. Suprasubduction zone model for metal endowment at 1.60–1.57 Ga in eastern Australia. Ore Geol. Rev. 2020, 122, 103483. [Google Scholar]
  140. Hayward, N.; Corriveau, L.; Craven, J.A.; Enkin, R.J. Geophysical signature of the NICO Au-Co-Bi-Cu deposit and its iron oxide-alkali alteration system, Northwest Territories, Canada. Econ. Geol. 2016, 111, 2087–2110. [Google Scholar]
  141. Tornos, F.; Hanchar, J.M.; Munizaga, R.; Velasco, F.; Galindo, C. The role of the subducting slab and melt crystallization in the formation of magnetite-(apatite) systems, Coastal Cordillera of Chile. Mineral. Depos. 2021, 56, 253–278. [Google Scholar]
  142. Porter, T.M. The Carrapateena iron oxide copper-gold deposit, Gawler Craton, South Australia: A review. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 191–200. [Google Scholar]
  143. Jara, J.J.; Barra, F.; Reich, M.; Leisen, M.; Romero, R.; Morata, D. Episodic construction of the early Andean Cordillera unravelled by zircon petrochronology. Nat. Comm. 2021, 12, 4930. [Google Scholar] [CrossRef] [PubMed]
  144. Yan, J.; Liu, X.; Wang, S.; Xie, J.; Liu, J. Metallogenic type controlled by magma source and tectonic regime: Geochemical comparisons of Mesozoic magmatism between the Middle–Lower Yangtze River Belt and the Dabie Orogen, eastern China. Ore Geol. Rev. 2021, 133, 104095. [Google Scholar] [CrossRef]
  145. Bickford, M.E.; Van Schmus, W.R.; Karlstrom, K.E.; Mueller, P.A.; Kamenov, G.D. Mesoproterozoic-trans-Laurentian magmatism: A synthesis of continent-wide age distributions, new SIMS U–Pb ages, zircon saturation temperatures, and Hf and Nd isotopic compositions. Precambrian Res. 2015, 265, 286–312. [Google Scholar] [CrossRef]
  146. Chen, H.; Cooke, D.R.; Baker, M.J. Mesozoic iron oxide copper-gold mineralization in the central Andes and the Gondwana supercontinent breakup. Econ. Geol. 2013, 108, 37–44. [Google Scholar] [CrossRef]
  147. Verbaas, J.; Thorkelson, D.J.; Crowley, J.; Davis, W.J.; Foster, D.A.; Gibson, H.D.; Marshall, D.D.; Milidragovic, D. A sedimentary overlap assemblage links Australia to northwestern Laurentia at 1.6 Ga. Precambrian Res. 2018, 305, 19–39. [Google Scholar] [CrossRef]
  148. Bockmann, M.J.; Payne, J.L.; Hand, M.; Morrissey, L.J.; Belperio, A.P. Linking the Gawler Craton and Mount Isa Province through hydrothermal systems in the Peake and Denison Domain, northeastern Gawler Craton. Geosci. Front. 2023, 14, 101596. [Google Scholar] [CrossRef]
  149. Trunfull, E.F.; Hagemann, S.G.; Xavier, R.P.; Moreto, C.P.N. Critical assessment of geochronological data from the Carajás Mineral Province, Brazil: Implications for metallogeny and tectonic evolution. Ore Geol. Rev. 2020, 121, 103556. [Google Scholar] [CrossRef]
  150. Regan, S.P.; Walsh, G.J.; Williams, M.L.; Chiarenzelli, J.R.; Toft, M.; McAleer, R. Syn-collisional exhumation of hot middle crust in the Adirondack Mountains (New York, USA): Implications for extensional orogenesis in the southern Grenville Province. Geosphere 2019, 15, 1240–1261. [Google Scholar] [CrossRef]
  151. Xavier, R.P.; Monteiro, L.V.S.; de Souza Filho, C.R.; Torresi, I.; de Resende Carvalho, E.; Dreher, A.M.; Wiedenbeck, M.; Trumbull, R.B.; Pestilho, A.L.S.; Moreto, C.P.N. The iron oxide copper-gold deposits of the Carajás mineral province, Brazil: An updated and critical review. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 285–306. [Google Scholar]
  152. Martinsson, O.; Billström, K.; Broman, C.; Weihed, P.; Wanhainen, C. Metallogeny of the Northern Norrbotten Ore Province, northern Fennoscandian Shield with emphasis on IOCG and apatite-iron ore deposits. Ore Geol. Rev. 2016, 78, 447–492. [Google Scholar] [CrossRef]
  153. LKAB. A Summary Technical Report on the Mineral Resources of LKAB, Sweden—Per Geijer Iron Ore Apatite Deposit, June 2023. Available online: https://lkab.mediaflowportal.com/documents/folder/446946/ (accessed on 3 August 2023).
  154. LKAB. LKAB 2022 Mineral Resources Summary Report, December 2022. Available online: https://lkab.mediaflowportal.com/documents/folder/446946/ (accessed on 3 August 2023).
  155. Rainbird, R.H.; Rooney, A.D.; Creaser, R.A.; Skulski, T. Shale and pyrite Re-Os ages from the Hornby Bay and Amundsen basins provide new chronological markers for Mesoproterozoic stratigraphic successions of northern Canada. Earth Planet. Sci. Lett. 2020, 548, 116492. [Google Scholar]
  156. Hussey, K.J.; Huston, D.L.; Claoué-Long, J.C. Geology and Origin of some Cu-Pb-Zn (Auag) Deposits in the Strangways Metamorphic Complex, Arunta Region, Northern Territory; Report 17; Northern Territory Geological Survey: Darwin, Australia, 2005; 96 p.
  157. Corriveau, L.; Spry, P. Metamorphosed hydrothermal ore deposits. In Treatise on Geochemistry, 2nd ed.; Holland, H.D., Turekian, K.K., Eds.; Elsevier: Amsterdam, The Netherlands, 2014; Volume 13, pp. 175–194. [Google Scholar]
  158. Hildebrand, R.S. Precambrian Geology, Leith Peninsula-Rivière Grandin Area, Northwest Territories. Canadian Geoscience Map 153; Geological Survey of Canada: Ottawa, ON, Canada, 2017.
  159. Bretzlaff, R.; Kerswill, J.A. Mineral Occurrences of the Great Bear Magmatic Zone; Open File 7959; Geological Survey of Canada: Ottawa, ON, Canada, 2016; 7p.
  160. Hayward, N.; Corriveau, L. Fault reconstructions using aeromagnetic data in the Great Bear magmatic zone, Northwest Territories, Canada. Can. J. Earth Sci. 2014, 51, 927–942. [Google Scholar]
  161. Chen, H. External sulphur in IOCG mineralization: Implications on definition and classification of the IOCG clan. Ore Geol. Rev. 2013, 51, 74–78. [Google Scholar]
  162. Badham, J.P.N.; Stanworth, C.W. Evaporites from the lower Proterozoic of the East Arm, Great Slave Lake. Nature 1977, 268, 516–518. [Google Scholar]
  163. Curtis, S.; Wade, C.; Reid, A. Sedimentary basin formation associated with a silicic large igneous province: Stratigraphy and provenance of the Mesoproterozoic Roopena Basin, Gawler Range Volcanics. Aust. J. Earth Sci. 2018, 65, 447–463. [Google Scholar]
  164. Schlegel, T.U.; Wagner, T.; Wälle, M.; Heinrich, C.A. Hematite breccia-hosted iron oxide copper-gold deposits require magmatic fluid components exposed to atmospheric oxidation: Evidence from Prominent Hill, Gawler Craton, South Australia. Econ. Geol. 2018, 113, 597–644. [Google Scholar] [CrossRef]
  165. Hildebrand, R.S. Geology of the Rainy Lake-White Eagle Falls area district of Mackenzie: Early Proterozoic cauldrons, stratovolcanoes and subvolcanic plutons: Paper 83–20; Geological Survey of Canada: Ottawa, ON, Canada, 1984; 42 p.
  166. Kelly, C.J.; Davis, W.J.; Potter, E.G.; Corriveau, L. Geochemistry of hydrothermal tourmaline from IOCG occurrences in the Great Bear magmatic zone: Implications for fluid source(s) and fluid composition evolution. Ore Geol. Rev. 2020, 118, 103329. [Google Scholar]
  167. Hildebrand, R.S.; Bowring, S.A. Continental intra-arc depressions: A non-extensional model for their origin, with a Proterozoic example from Wopmay Orogen. Geology 1984, 12, 73–77. [Google Scholar]
  168. Huang, Q.; Kamenetsky, V.S.; Ehrig, K.; McPhie, J.; Kamenetsky, M.; Cross, K.; Meffre, S.; Agangi, A.; Chambefort, I.; Direen, N.G.; et al. Olivine-phyric basalt in the Mesoproterozoic Gawler silicic large igneous province, South Australia: Examples at the Olympic Dam iron oxide Cu–U–Au–Ag deposit and other localities. Precambrian Res. 2016, 281, 185–199. [Google Scholar] [CrossRef]
  169. Schlegel, T.U.; Heinrich, C.A. Lithology and hydrothermal alteration control the distribution of copper grade in the Prominent Hill iron oxide-copper-gold deposit (Gawler Craton, South Australia). Econ. Geol. 2015, 110, 1953–1994. [Google Scholar]
  170. Jagodzinski, E.A.; Reid, A.; Crowley, J.; Wade, C.; Curtis, S. Precise zircon U-Pb dating of the Mesoproterozoic Gawler large igneous province, South Australia. Results Geochem. 2023, 10, 100020. [Google Scholar]
  171. Cherry, A.; Ehrig, K.; Kamenetsky, V.S.; McPhie, J.; Crowley, J.; Kamenetsky, M. Precise geochronological constraints on the origin, setting and incorporation of ca. 1.59 Ga surficial facies into the Olympic Dam Breccia Complex, South Australia. Precambrian Res. 2018, 315, 162–168. [Google Scholar]
  172. Courtney-Davies, L.; Ciobanu, C.L.; Verdugo-Ihl, M.R.; Dmitrijeva, M.; Cook, N.J.; Ehrig, K.; Wade, B.P. Hematite geochemistry and geochronology resolve genetic and temporal links among iron-oxide copper gold systems, Olympic Dam district, South Australia. Precambrian Res. 2019, 335, 105480. [Google Scholar]
  173. Courtney-Davies, L.; Ciobanu, C.L.; Tapster, S.R.; Cook, N.J.; Ehrig, K.; Crowley, J.L.; Verdugo-Ihl, M.R.; Wade, B.P.; Condon, D.J. Opening the magmatic-hydrothermal window: High-precision U-Pb geochronology of the Mesoproterozoic Olympic Dam Cu-U-Au-Ag deposit, South Australia. Econ. Geol. 2020, 115, 1855–1870. [Google Scholar]
  174. Marschik, R.; Söllner, F. Early Cretaceous U-Pb zircon ages for the Copiapo plutonic complex and implications for the IOCG mineralization at Candelaria, Atacama Region, Chile. Mineral. Depos. 2006, 41, 785–801. [Google Scholar]
  175. Yu, J.; Chen, Y.; Mao, J.; Pirajno, F.; Duan, C. Review of geology, alteration and origin of iron oxide–apatite deposits in the Cretaceous Ningwu basin, Lower Yangtze River Valley, eastern China: Implications for ore genesis and geodynamic setting. Ore Geol. Rev. 2011, 43, 170–181. [Google Scholar]
  176. Wang, J.; Li, L.; Santosh, M.; Yan, G.-Y.; Shen, J.-F.; Yuan, M.-W.; Alam, M.; Li, S.-R. Multistage ore formation in the world’s largest REE-Nb-Fe deposit of Bayan Obo, North China Craton: New insights and implications. Ore Geol. Rev. 2024, 164, 105817. [Google Scholar]
  177. Davis, W.J.; Corriveau, L.; van Breemen, O.; Bleeker, W.; Montreuil, J.-F.; Potter, E.; Pelleter, E. Timing of IOCG mineralising and alteration events within the Great Bear magmatic zone [abs.]. In 39th Annual Yellowknife Geoscience Forum Abstracts; Fischer, B.J., Watson, D.M., Eds.; Northwest Territories Geological Survey: Yellowknife, NT, Canada, 2011; Abstracts Volume 2011, 97p. [Google Scholar]
  178. Verdugo-Ihl, M.R.; Ciobanu, C.L.; Cook, N.J.; Ehrig, K.; Courtney-Davies, L. Defining early stages of IOCG systems: Evidence from iron-oxides in the outer shell of the Olympic Dam deposit, South Australia. Mineral. Depos. 2020, 55, 429–452. [Google Scholar]
  179. Rodriguez-Mustafa, M.A.; Simon, A.C.; Holder, R.M.; Stein, H.; Kylander-Clark, A.R.C.; Jicha, B.R.; Blakemore, D.; Machado, E.L.B. Integrated Re-Os, Ar/Ar, and U-Pb geochronology directly dates the timing of mineralization at the Mina Justa and Marcona deposits, Peru. GSA Bull. 2024, 136, 2861–2874. [Google Scholar]
  180. Chapman, N.D.; Ferguson, M.; Meffre, S.J.; Stepanov, A.; Maas, R.; Ehrig, K.J. Pb-isotopic constraints on the source of A-type suites: Insights from the Hiltaba Suite-Gawler Range Volcanics magmatic event, Gawler craton, South Australia. Lithos 2019, 346–347, 105156. [Google Scholar]
  181. Wade, C.E.; Payne, J.L.; Hill, J.; Curtis, S.; Barovich, K.; Reid, A.J. Temporal, geochemical and isotopic constraints on plume-driven felsic and mafic components in a Mesoproterozoic flood rhyolite province. Results Geochem. 2022, 9, 100019. [Google Scholar]
  182. Wade, C.E.; Payne, J.L.; Barovich, K.M.; Crowley, J.; Jagodzinski, E.; Gilbert, S.E.; Wade, B.P.; Reid, A.J. Zircon trace element geochemistry as an indicator of magma fertility in iron oxide-copper-gold provinces. Econ. Geol. 2022, 117, 703–718. [Google Scholar]
  183. Williams, M.R.; Holwell, D.A.; Lilly, R.M.; Case, G.N.D.; McDonald, I. Mineralogical and fluid characteristics of the fluorite-rich Monakoff and E1 Cu–Au deposits, Cloncurry region, Queensland, Australia: Implications for regional F–Ba-rich IOCG mineralisation. Ore Geol. Rev. 2015, 64, 103–127. [Google Scholar]
  184. Williams, P.J.; Kendrick, M.; Xavier, R.P. Sources of ore fluid components in IOCG deposits. In Hydrothermal Iron Oxide Copper-Gold & Related Deposits: A Global Perspective, v. 3—Advances in the Understanding of IOCG Deposits; Porter, T.M., Ed.; PGC Publishing: Adelaide, Australia, 2010; pp. 107–116. [Google Scholar]
  185. Bastrakov, E.N.; Skirrow, R.G.; Davidson, G.J. Fluid evolution and origins of iron oxide Cu-Au prospects in the Olympic Dam district, Gawler Craton, South Australia. Econ Geol. 2007, 102, 1415–1440. [Google Scholar]
  186. Baker, T.; Mustard, R.; Fu, B.; Williams, P.J.; Dong, G.; Fisher, L.; Mark, G.; Ryan, C.G. Mixed messages in iron oxide–copper–gold systems of the Cloncurry district, Australia: Insights from PIXE analysis of halogens and copper in fluid inclusions. Mineral. Depos. 2008, 43, 599–608. [Google Scholar]
  187. Williams, P.J. Iron mobility during synmetamorphic alteration in the Selwyn Range area, NW Queensland: Implications for the origin of ironstone-hosted Au-Cu deposits. Mineral. Depos. 1994, 29, 250–260. [Google Scholar]
  188. Pelleter, E.; Gasquet, D.; Cheilletz, A.; Mouttaqi, A. Alteration processes and impacts on regional-scale element mobility and geochronology, Tamlalt-Menhouhou deposit, Morocco. In Exploring for Iron Oxide Copper-Gold Deposits: Canada and Global Analogues; Corriveau, L., Mumin, A.H., Eds.; Short Course Notes, No. 20; Geological Association of Canada: St. John’s, NL, Canada, 2010; pp. 177–185. [Google Scholar]
  189. Li, W.; Audétat, A.; Zhang, J. The role of evaporites in the formation of magnetite–apatite deposits along the Middle and Lower Yangtze River, China: Evidence from LA-ICP-MS analysis of fluid inclusions. Ore Geol. Rev. 2015, 67, 264–278. [Google Scholar] [CrossRef]
  190. Bain, W.M.; Steele-MacInnis, M.; Li, K.; Li, L.; Mazdad, F.K.; Marsh, E.E. A fundamental role of carbonate–sulfate melts in the formation of iron oxide–apatite deposits. Nat. Geosci. 2020, 13, 751–757. [Google Scholar]
  191. Wise, T.; Thiel, S. Proterozoic tectonothermal processes imaged with magnetotellurics and seismic reflection in southern Australia. Geosci. Front. 2020, 11, 885–893. [Google Scholar]
  192. Blundy, J.; Afanasyev, A.; Tattitch, B.; Sparks, S.; Melnik, O.; Utkin, I.; Rust, A. The economic potential of metalliferous sub-volcanic brines. Royal Soc. Open Sci. 2021, 8, 202192. [Google Scholar]
  193. Pietruszka, D.K.; Hanchar, J.M.; Tornos, F.; Wirth, R.; Graham, N.A.; Severin, K.P.; Velasco, F.; Steele-MacInnis, M.; Bain, W.M. Magmatic immiscibility and the origin of magnetite-(apatite) iron deposits. Nat. Comm. 2023, 14, 8424. [Google Scholar] [CrossRef] [PubMed]
  194. Lowell, G.R. The Butler Hill Caldera: A mid-Proterozoic ignimbrite-granite complex. Precambrian Res. 1991, 51, 245–263. [Google Scholar] [CrossRef]
  195. Marshall, L.J.; Oliver, N.H.S.; Davidson, G.J. Carbon and oxygen isotope constraints on fluid sources and fluid-wallrock interaction in regional alteration and iron-oxide–copper–gold mineralisation, eastern Mt Isa Block, Australia. Mineral. Depos. 2006, 41, 429–452. [Google Scholar] [CrossRef]
  196. Chinova Resources. Mount Dore mineral resource update summary. In Technical Report 2016; Chinova Resources: Dajarra, QLD, Australia, 2016; 15p, Available online: https://www.chinovaresources.com/images/pdfs/Mt_Dore_Resource_Estimation_Update_summary.pdf (accessed on 3 August 2023).
  197. Reid, A.J.; Fabris, A. Influence of pre-existing low metamorphic grade sedimentary successions on the distribution of iron oxide copper-gold mineralisation in the Olympic Cu-Au Province, Gawler Craton. Econ. Geol. 2015, 110, 2147–2157. [Google Scholar] [CrossRef]
  198. Morrissey, L.J.; Tomkins, A.G. Evaporite-bearing orogenic belts produce ligand-rich and diverse metamorphic fluids. Geochim. Cosmochim. Acta 2020, 275, 163–187. [Google Scholar] [CrossRef]
  199. de Melo, G.H.C.; Monteiro, L.V.S.; Hunger, R.B.; Toledo, P.I.F.; Xavier, R.P.; Zhao, X.-F.; Su, Z.-K.; Moreto, C.P.N.; De Jesus, S.d.S.G.P. Magmatic-hydrothermal fluids leaching older seafloor exhalative rocks to form the IOCG deposits of the Carajás Province, Brazil: Evidence from boron isotopes. Precambrian Res 2021, 365, 106412. [Google Scholar] [CrossRef]
  200. Emproto, C.; Mathur, R.; Simon, A.; Bindeman, I.; Godfrey, L.; Dhnaram, C.; Lisitsin, V. Integrated O, Fe, and Ti isotopic analysis elucidates multiple metal and fluid sources for magnetite from the Ernest Henry iron oxide copper gold (IOCG) Deposit, Queensland, Australia. Ore Geol. Rev. 2022, 150, 105170. [Google Scholar] [CrossRef]
  201. Melfou, M.; Richard, A.; Tarantola, A.; Villeneuve, J.; Carr, P.; Peiffert, C.; Mercadier, J.; Dean, B.; Drejing-Carroll, D. Tracking the origin of metasomatic and ore-forming fluids in IOCG deposits through apatite geochemistry (Nautanen North deposit, Norrbotten, Sweden). Lithos 2023, 438–439, 106995. [Google Scholar] [CrossRef]
  202. Li, R.; Wang, X.-L. External fluid incursion during Cu-mineralization stage of Mina Justa iron oxide copper-gold (IOCG) deposit: Evidence from triple sulfur isotope geochemistry of chalcopyrite. Ore Geol. Rev. 2022, 149, 105102. [Google Scholar] [CrossRef]
  203. Duan, G.; Ram, R.; Xing, Y.; Etschmann, B.; Brugger, J. Kinetically driven successive sodic and potassic alteration of feldspar. Nat. Commun. 2021, 12, 4435. [Google Scholar] [CrossRef]
  204. Nayebi, N.; Esmaeily, D.; D’Antonio, M.; Xia, X.-P.; Di Renzo, V.; Lehmann, B.; Shinjo, R.; Babazadeh, S.; Deevsalar, R.; Modabberi, S. Zircon U–Pb ages and Sr–Nd–Pb–Hf isotopic compositions constrain the tectono-magmatic evolution of the Anomaly 21-A iron ore region, Bafq metallogenic province, Central Iran. J. Asian Earth Sci. 2023, 250, 105646. [Google Scholar] [CrossRef]
  205. Bergman, S.; Kübler, L.; Martinsson, O. Description of Regional Geological and Geophysical Maps of Northern Norrbotten County (East of the Caledonian Orogen); BA-56; Sveriges Geologiska Undersökning: Uppsala, Sweden, 2001; 110 p.
  206. Logan, L.; Andersson, J.B.H.; Whitehouse, M.J.; Martinsson, O.; Bauer, T.E. Energy drive for the Kiruna mining district mineral system(s): Insights from U-Pb zircon geochronology. Minerals 2022, 12, 875. [Google Scholar] [CrossRef]
  207. Case, G.; Blenkinsop, T.; Chang, Z.; Huizenga, J.M.; Lilly, R.; McLellan, J. Delineating the structural controls on the genesis of iron oxide-Cu-Au deposits through implicit modelling: A case study from the E1 Group, Cloncurry District, Australia. In Characterization of Ore-forming Systems from Geological, Geochemical and Geophysical Studies; Gessner, K., Blenkinsop, T.G., Sorjonen-Ward, P., Eds.; Geological Society, London, Special Publications: London, UK, 2017; 453 p. [Google Scholar]
  208. Lu, Y.; Liu, L.; Xu, G. Constraints of deep crustal structures on large deposits in the Cloncurry district, Australia: Evidence from spatial analysis. Ore Geol. Rev. 2016, 79, 316–331. [Google Scholar]
  209. Bauer, T.E.; Andersson, J.B.H. Regional structural setting of late-orogenic IOCG mineralization along the northern Nautanen deformation zone, Norrbotten, Sweden. Ore Geol. Rev. 2023, 163, 105814. [Google Scholar]
  210. Gadd, M.G.; Lawley, C.J.; Corriveau, L.; Houlé, M.; Peter, J.M.; Plouffe, A.; Potter, E.G.; Sappin, A.-A.; Pilote, J.-L.; Marquis, G.; et al. Public geoscience solution for diversifying Canada’s critical mineral production. In Geological Society; Special Publications: London, UK, 2022; Volume 526, pp. 25–50. [Google Scholar]
  211. Cave, B.W.; Lilly, R.; Glorie, S.; Gillespie, J. Geology, apatite geochronology, and geochemistry of the Ernest Henry inter-lens: Implications for a re-examined deposit model. Minerals 2018, 8, 405. [Google Scholar] [CrossRef]
  212. Rojas, P.A.; Barra, F.; Reich, M.; Deditius, A.; Simon, A.; Uribe, F.; Romero, R.; Rojo, M. A genetic link between magnetite mineralization and diorite intrusion at the El Romeral iron oxide-apatite deposit, northern Chile. Mineral. Depos. 2018, 53, 947–966. [Google Scholar] [CrossRef]
  213. Martinsson, O. Kiskamavaara a shear zone hosted IOCG-style of Cu-Co-Au deposit in Northern Norrbotten, Sweden. In Proceedings of the 11th SGA Biennial Meeting, Antofagasta, Chile, 26–29 September 2011; Volume 2, pp. 470–472. [Google Scholar]
  214. Martinsson, O.; Wanhainen, C. Economic potential of battery metals and minerals in Sweden. In Proceedings of the Volume for the 16th SGA Biennial Meeting, Rotorua, New Zealand, 28–31 March 2022; pp. 227–230. [Google Scholar]
  215. Corriveau, L.; Bonnet, A.-L. Pinwarian (1.5 Ga) volcanism and hydrothermal activity at the eastern margin of the Wakeham Group, Grenville Province, Quebec. Can. J. Earth Sci. 2005, 42, 1749–1782. [Google Scholar]
  216. Corriveau, L.; Morin, D. Modelling 3D architecture of western Grenville from surface geology, xenoliths, styles of magma emplacement and Lithoprobe reflectors. Can. J. Earth Sci. 2000, 37, 235–251. [Google Scholar]
  217. BHP. Annual Report 2024. Available online: https://www.bhp.com/investors/annual-reporting/annual-report-2024 (accessed on 17 December 2024).
  218. BHP. Financial Results for the Year Ended 30 June 2024. Available online: https://www.bhp.com/-/media/documents/media/reports-and-presentations/2024/240827_bhpresultsfortheyearended30june2024.pdf (accessed on 13 January 2025).
  219. Rex Minerals Limited. Rex Minerals-Hillside Resources & Reserves. Available online: https://www.rexminerals.com.au/resourceshillside (accessed on 10 January 2025).
  220. Chinalco Yunnan Copper Resources. 26.1 Mt Inferred JORC Resource Estimate Elaine 1 Copper-Gold Deposit. Available online: https://www.aukingmining.com/site/pdf/efaa23c8-91c4-4c98-9019-7fc824701307/261Mt-Inferred-JORC-Resource-Elaine-1-Copper-Gold-Deposit.pdf (accessed on 29 June 2012).
  221. China Yunnan Copper Resources. Inferred Resource Estimate—Elaine-Dorothy Uranium—Rare Earth Element (REE). Available online: https://www.aukingmining.com/site/PDF/69270e44-d780-40f0-9aec-b1e695241a0e/CuCoREEUMineralisationatElaineDorothyProject (accessed on 27 January 2025).
  222. AIC Mines Limited. Eloise Copper Mine Almanac, April 2024. Available online: https://www.aicmines.com.au/downloads/eloise-almanac.pdf (accessed on 10 January 2025).
  223. Evolution Mining. Further Increase in Ernest-Henry Mineral Resource: ASX Announcement, 17 August 2023. Available online: www.evolutionmining.com.au (accessed on 15 May 2024).
  224. AIC Mines Limited. Significant Increase in Mineral Resource. 2025. Available online: https://www.listcorp.com/asx/a1m/aic-mines-limited/news/significant-increase-in-mineral-resources-3166632.html (accessed on 27 January 2025).
  225. Copper Mountain Mining Corporation. NI 43-101 Technical Report for the Eva Copper Project—Feasibility Study Update North West Queensland, Australia, 2020. Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=d47610acaa2a771985e5f54edf16b9135b71722a01bb857a83b97a332d67db57 (accessed on 18 December 2024).
  226. Hammer Metal Corporation, 2023, Annual Report 2023. Available online: www.hammermetals.com.au (accessed on 23 January 2024).
  227. McKay, A.D.; Miezitis, Y. Australia’s uranium resources, geology and development of deposits. In Mineral Resources Report 1; AGSO–Geoscience Australia: Symonston, Australia, 2001. [Google Scholar]
  228. Chinova, R. Mount Elliott Swan Resource Estimation Update Summary. 2017. Available online: www.chinovaresources.com (accessed on 10 August 2017).
  229. Cudeco. 2017 Annual Report. Available online: www.cudeco.com.au (accessed on 17 May 2023).
  230. Paladin, E. Valhalla Uranium Deposit, Mineral Resources. 2015. Available online: http://www.paladinenergy.com.au/default.aspx?MenuID=35 (accessed on 1 February 2015).
  231. Collier, J. Johnnies Reward, Au-Cu, Mineral Resource Estimation, 24th January 2018, Prepared for Conarco Consulting (Conarco). Available online: https://geoscience.nt.gov.au (accessed on 31 January 2025).
  232. Peko, B. Peko Tailings Retreatment Project Information Memorandum. 2017. Available online: https://www.99mines.com/wp-content/uploads/2017/07/345_geodir_key_documents_Peko-IM-July-2017-Executive-Summary-1.pdf (accessed on 16 December 2024).
  233. Grange Resources Limited. Annual Resource & Reserve Statement—December 2023; Maiden underground Ore Reserve Declared for Savage River Mine: ASX Announcement, 28 February 2024. Available online: https://www.listcorp.com/ (accessed on 18 December 2024).
  234. Natural Resources Canada, Canada’s Critical Minerals. 2024. Available online: https://www.canada.ca/en/campaign/critical-minerals-in-canada/critical-minerals-an-opportunity-for-canada.html (accessed on 16 January 2025).
  235. Perron, P.M.; Beauvais, M.R.; Kinnan, E.; Roy, P. NI 43-101 Technical Report and Mineral Resource Estimate for the Iron Creek Cobalt-Copper Property, Lemhi County, Idaho, USA; Electra Battery Materials Corporation: Toronto, On, Canada, 2023; 196p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=fa56487bc1f25e4e1a0bdaee8cbca3c95ffdbcbe99044a729e5a046abe01e4a4 (accessed on 9 January 2025).
  236. Hammarstrom, J.; Dicken, C.; Day, W.; Hofstra, A.; Drenth, B.; Shah, A.; McCafferty, A.; Woodruff, L.; Foley, N.; Ponce, D.; et al. Focus Areas for Data Acquisition for Potential Domestic Resources of 11 Critical Minerals in the Conterminous United States, Hawaii, and Puerto Rico—Aluminum, Cobalt, Graphite, Lithium, Niobium, Platinum-Group Elements, Rare Earth elements, tantalum, tin, titanium, and tungsten (ver. 1.1, July 2022), chap. B of U.S. Geological Survey, Focus Areas for Data Acquisition for Potential Domestic Sources of Critical Minerals; Open-File Report 2019–1023; U.S. Geological Survey: Denver, CO, USA, 2020; 67 p. [CrossRef]
  237. Nevada Copper Corporation. Pumpkin Hollow Project, Open Pit and Underground Mine Prefeasibility Study: NI 43-101 Technical Report; Nevada Copper Corporation: Yerington, NV, USA, 2019; 520p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=abf27b5892bfb4c5a9d18dd96e8474e9a58b05136024c81a3ffd693b10ca629e (accessed on 9 January 2025).
  238. Samari, H.; Lane, T.; Harvey, T.; Breckenridge, L. Preliminary Economic Assessment NI 43-101 Technical Report on the Moonlight-Superior Project; US Copper Corp: Toronto, ON, Canada, 2024; Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=0510c94aec7ae511be0f0aed8bb8b237d3b00860377b44353a278c9f527ff540 (accessed on 9 January 2025).
  239. Kyle, J.; Beahm, D. NI 43-101 Preliminary Economic Assessment Update, Coles Hill Uranium Property; Virginia Energy Resources: Pittsylvania County, VA, USA, 2012. Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=721388e103dc592c58d59cda0a57eebd97be070851512380857d1d9ce053d63d (accessed on 10 January 2025).
  240. Magnum Mining and Exploration. Maiden JORC 2012 resource for Buena Vista magnetite project. Magnum Mining and Exploration: Subiaco, WA, USA. Available online: https://www.mmel.com.au/site/projects-and-products/buena-vista-magnetite-project (accessed on 27 January 2025).
  241. Burgess, H.; Gowans, R.M.; Hennessey, B.T.; Lattanzi, C.R.; Puritch, E. Technical Report on the Feasibility Study for the NICO Gold–Cobalt–Bismuth–Copper Deposit; NI 43-101 Technical Report No. 1335; Fortune Minerals Ltd.: London, ON, Canada, 2014; 385p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=b7f689dd67e909fa62179555a672714fc561c8a7d5fcd21b5d1ee14a15edf946 (accessed on 15 September 2015).
  242. Hennessey, B.T.; Puritch, E. A Technical Report on a Mineral Resource Estimate for the Sue-Dianne Deposit, Mazenod Lake Area; NI 43-101 Technical Report; Fortune Minerals Limited: London, ON, Canada, 2008; 125p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=c64ebb5d1c5a5e3c37a62392a9209df9a59529b1c4b9a0a3a1e7b6c08f9c1a8e (accessed on 21 December 2009).
  243. Paladin Energy. Michelin Project, Labrador Canada, 2019. Available online: http://www.paladinenergy.com.au/project/michelin-canada (accessed on 1 September 2019).
  244. Kruse, F. NI 43-101 technical report, Moran Lake project, Central Mineral Belt, Newfoundland and Labrador, Canada. NI 43-101 Technical Report. Consolidated-Uranium-Inc., 2021. Available online: https://www.sedarplus.ca/csfsprod/data446/filings/03325658/00000001/z%3A%5CPUBLIC%5CConsolidated-Uranium-Inc%5C2022%5CSMG%5CCUR-MoranLk-TechRpt.pdf (accessed on 28 January 2025).
  245. Gagnon, R.; Buro, Y.A.; Ibrango, S.; Gagnon, D.; Stapinsky, M.; Del Carpio, S.; Larochelle, E. Projet de Terres Rares Kwyjibo: Rapport Technique NI 43-101 Révisé; Dra Met-Chem: Montréal, QC, Canada, 2018; 292p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=6fe7e177d4d03d91a694437e1ea496e30055103b092233ae7ba5039ea38c7265 (accessed on 15 December 2023).
  246. Chadwick, P.J.; Péloquin, A.S.; Suma-Momoh, J.; Daniels, C.M.; Hinz, S.L.K.; Kennedy, C.A.; Streit, L.; Todd, R.M. Report of Activities 2019, Resident Geologist Program, Kirkland Lake Regional Resident Geologist Report: Kirkland Lake and Sudbury Districts; Open File Report 6367; Ontario Geological Survey: Sudbury, ON, Canada, 2020; 143 p. [Google Scholar]
  247. Orano. Annual Activity Report 2023. Available online: https://cdn.orano.group/orano/docs/default-source/orano-doc/finance/publications-financieres-et-reglementees/2023/orano_annual-activity-report_2023_online.pdf (accessed on 27 January 2024).
  248. Dufresne, M.B.; Sim, R.; Davis, B. Technical Report and Resource Update for the Angilak Property; Kivalliq Energy Corporation: Kivalliq region, NU, Canada, 2013; 174p, Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=c6d42de1edac935a02f2c333558a9aac71279d3bc086be3a8c59e25d6a5d3604 (accessed on 7 January 2025).
  249. AGP Mining Consultants Inc. NI 43-101 Resource Estimate for Werner Lake Cobalt Project; Global Energy Metals Corp: Werner Lake, ON, Canada, 2017; Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=f086f72323b8bdc899fa7c7d9350404b84ccd3c00e5b25e57465b1a0b6d7255f (accessed on 7 January 2025).
  250. Centaurus Metals Limited. The Jaguar Nickel Sulphide Project Value-Add Scoping Study, Executive Summary, May 2021. Available online: www.centaurus.com.au (accessed on 28 June 2021).
  251. Anglesey Mining plc. Grängesberg PFS Highlights Post-Tax NPV8 of US$688m; 2022. Available online: https://www.angleseymining.co.uk/ (accessed on 7 January 2025).
  252. Dean, B.; McGimpsey, I.; Drejing-Carroll, D.; Årebäck, H. Boliden Summary Report, Mineral Resources and Mineral Reserves 2022, Nautanen. Available online: https://www.boliden.com/operations/exploration/mineral-reserves-and-mineral-resources (accessed on 3 August 2023).
  253. Baker, H.; MacDougall, C.; Pattinson, D. Technical Review of the Hannukainen Iron-Copper-Gold Project, Kolari District, Finland: SRK Consulting, National Instrument 43-101 Technical Report. 2014. Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=92dc83b898c7088db2d0282a1309d34fd8bc36b97671d2a6c615fa08f2f22d21 (accessed on 28 July 2015).
  254. Baker, H.; Pattinson, D.; Reardon, C. Technical Review of the Kaunisvaara Iron Project, Sweden, June 2011. Report U4067 prepared for Northland Resources AB. SRK (UK) Limited. Available online: https://www.sedarplus.ca/csa-party/records/document.html?id=47ef8c7ab490c1266285f285485325e4cb62122df76830b7cb1c79e1bd6b3d56 (accessed on 28 January 2025).
  255. Geological Survey of Finland. Juomasuo, Paleoproterozoic Kuusamo belt. Available online: https://minsysfin.gtk.fi/index.php/juomasuo-paleoproterozoic-kuusamo-belt/#:~:text=The%20host%20rocks%20of%20the,by%20pyrite%20and%20lesser%20chalcopyrite (accessed on 28 January 2025).
Figure 1. World distribution of MIAC systems and their deposits and prospects. Those with labels are discussed in this paper. Information is updated from [2,3,5,13] using the references listed in Table A1, Table A2 and Table A3 and in the tables of Section 3.4, as well as from [45,46,47,48].
Figure 1. World distribution of MIAC systems and their deposits and prospects. Those with labels are discussed in this paper. Information is updated from [2,3,5,13] using the references listed in Table A1, Table A2 and Table A3 and in the tables of Section 3.4, as well as from [45,46,47,48].
Minerals 15 00365 g001
Figure 3. Mineralization types associated with high-temperature Fe-rich alteration facies (Facies 2 and 2–3) and representative Facies 1 albitite and skarn (initially barren but key hosts for mineralization). Photographs of Facies 1 to 2–3 can be found in [7,10,11,14]. (A) Magnetite-skarn with clinopyroxene and magnetite at the transition from Fe-rich skarn to Facies 2 HT Ca-Fe in the Temagami area (Ontario, Canada). Belfast prospect, Conquest DDH BC-21-15. (B) IOA mineralization and strongly albitized host at the Yang-Yang Fe deposit (Republic of Korea). (C) IOA mineralization at Terra mine in the Great Bear magmatic zone (Canada). Geological Survey of Canada (GSC) outcrop 09CQA-0128. (D) REE mineralization associated with HT Ca-Fe alteration overprinted by biotite of Facies 3 at the JLD showing (Northwest Territories, Canada). GSC outcrop 09CQA-1170. (E) Incipient Ni mineralization at the Limestone prospect formed at Facies 2 (Ontario, Canada). (F) Relic of clinopyroxene (and scheelite) from skarn overprinted by HT Ca-K-Fe alteration fronts with Co sulfarsenide in the Co-Bi-Au ore zone of the NICO deposit (Northwest Territories, Canada). GSC outcrop CQA-07-466. Mineral abbreviations as per Figure 2. Photographs C,D,F: NRCan Photo Database 2025-037, 2025-038, and 2025-039. A,B,E: photographs by J.F. Montreuil.
Figure 3. Mineralization types associated with high-temperature Fe-rich alteration facies (Facies 2 and 2–3) and representative Facies 1 albitite and skarn (initially barren but key hosts for mineralization). Photographs of Facies 1 to 2–3 can be found in [7,10,11,14]. (A) Magnetite-skarn with clinopyroxene and magnetite at the transition from Fe-rich skarn to Facies 2 HT Ca-Fe in the Temagami area (Ontario, Canada). Belfast prospect, Conquest DDH BC-21-15. (B) IOA mineralization and strongly albitized host at the Yang-Yang Fe deposit (Republic of Korea). (C) IOA mineralization at Terra mine in the Great Bear magmatic zone (Canada). Geological Survey of Canada (GSC) outcrop 09CQA-0128. (D) REE mineralization associated with HT Ca-Fe alteration overprinted by biotite of Facies 3 at the JLD showing (Northwest Territories, Canada). GSC outcrop 09CQA-1170. (E) Incipient Ni mineralization at the Limestone prospect formed at Facies 2 (Ontario, Canada). (F) Relic of clinopyroxene (and scheelite) from skarn overprinted by HT Ca-K-Fe alteration fronts with Co sulfarsenide in the Co-Bi-Au ore zone of the NICO deposit (Northwest Territories, Canada). GSC outcrop CQA-07-466. Mineral abbreviations as per Figure 2. Photographs C,D,F: NRCan Photo Database 2025-037, 2025-038, and 2025-039. A,B,E: photographs by J.F. Montreuil.
Minerals 15 00365 g003
Figure 4. Mineralization types associated with representative high-to-low temperature Fe-rich alteration facies (Facies 3 and 5). Additional examples of all Facies can be found in [7,10,11,12,13,14,15]. (A) Magnetite-group IOCG mineralization associated with a magnetite–K-feldspar–calcite assemblage in the Ernest Henry deposit (Australia) [10,70,126]. (B,C) Iron-silicate copper–gold mineralization associated with biotite replacement fronts, localized selective K-feldspar alteration of the relicts of the albitized host, and calcite alteration in the Belfast prospect (Ontario, Canada). (D) Iron-sulfide copper–gold mineralization associated with pyrite and pyrrhotite replacement fronts with disseminated chalcopyrite and white mica alteration within an albitized basalt at the Delhi Pacific prospect (Quebec, Canada) [49,50]. (E) Facies 5 magnetite-group IOCG mineralization associated with magnetite and calcite with relicts of K-feldspar-altered fragments in the Ernest Henry deposit (Australia) [10,70,126]. (F) Albitite-hosted Fe-oxide U mineralization in the Red Hot showing of the Southern Breccia trend (Northwest Territories, Canada) [28]. (G) Iron-silicate Au mineralization at the Scadding deposit, in which Au mineralization is associated with chlorite replacement fronts containing variable amounts of disseminated pyrite and localized specks of native Au (Ontario, Canada) [22]. Mineral abbreviations as per Figure 2. Photographs A,E,F: NRCan Photo Database 2025-040, 2025-041 and 2025-042. B,C,D,G: photographs by J.F. Montreuil.
Figure 4. Mineralization types associated with representative high-to-low temperature Fe-rich alteration facies (Facies 3 and 5). Additional examples of all Facies can be found in [7,10,11,12,13,14,15]. (A) Magnetite-group IOCG mineralization associated with a magnetite–K-feldspar–calcite assemblage in the Ernest Henry deposit (Australia) [10,70,126]. (B,C) Iron-silicate copper–gold mineralization associated with biotite replacement fronts, localized selective K-feldspar alteration of the relicts of the albitized host, and calcite alteration in the Belfast prospect (Ontario, Canada). (D) Iron-sulfide copper–gold mineralization associated with pyrite and pyrrhotite replacement fronts with disseminated chalcopyrite and white mica alteration within an albitized basalt at the Delhi Pacific prospect (Quebec, Canada) [49,50]. (E) Facies 5 magnetite-group IOCG mineralization associated with magnetite and calcite with relicts of K-feldspar-altered fragments in the Ernest Henry deposit (Australia) [10,70,126]. (F) Albitite-hosted Fe-oxide U mineralization in the Red Hot showing of the Southern Breccia trend (Northwest Territories, Canada) [28]. (G) Iron-silicate Au mineralization at the Scadding deposit, in which Au mineralization is associated with chlorite replacement fronts containing variable amounts of disseminated pyrite and localized specks of native Au (Ontario, Canada) [22]. Mineral abbreviations as per Figure 2. Photographs A,E,F: NRCan Photo Database 2025-040, 2025-041 and 2025-042. B,C,D,G: photographs by J.F. Montreuil.
Minerals 15 00365 g004
Figure 5. Mineralization types associated with Fe-poor alteration facies (Facies 5 and 6). Additional examples can be found in [7,10,11,12,13,14,22]. (A) Albitite-hosted Au-U mineralization of the Mistamik Boulder Field associated with chlorite veinlets and hematization fronts (Québec, Canada) [50]. (B) Ferroan dolomite-quartz vein associated with Cu-Au mineralization at the Alwyn prospect (Ontario, Canada) [22]. (C) Zone of a strong silicification with localized K-feldspar alteration fronts and replacement veins —as defined in [11]—associated with chalcopyrite mineralization at the Alwyn prospect (Ontario, Canada) [22]. Mineral abbreviations as per Figure 2. Photographs by J.F. Montreuil.
Figure 5. Mineralization types associated with Fe-poor alteration facies (Facies 5 and 6). Additional examples can be found in [7,10,11,12,13,14,22]. (A) Albitite-hosted Au-U mineralization of the Mistamik Boulder Field associated with chlorite veinlets and hematization fronts (Québec, Canada) [50]. (B) Ferroan dolomite-quartz vein associated with Cu-Au mineralization at the Alwyn prospect (Ontario, Canada) [22]. (C) Zone of a strong silicification with localized K-feldspar alteration fronts and replacement veins —as defined in [11]—associated with chalcopyrite mineralization at the Alwyn prospect (Ontario, Canada) [22]. Mineral abbreviations as per Figure 2. Photographs by J.F. Montreuil.
Minerals 15 00365 g005
Figure 6. Simplified geology of the exposed segment of the Great Bear magmatic zone (Canada) from the Camsell River district to the north and the NICO deposit to the south (modified after [15]). The distribution of 1.87 Ga volcanic rocks were updated from Corriveau et al. [15] using Jackson et al. [58] and Hildebrand [158]. S is the location of stromatolites; E of evaporites (inferred from metasomatic pseudomorphing of the structures typical of nodular anhydrite and evaporite beds); Bas of basalt; And of andesite; Dac of dacite; and Rhy>>And of rhyolite units more abundant than andesite. Representative IOA, IOCG, and MI-Co deposits and prospects occur within the 1.87 Ga volcanic centers and the 1.88 Ga sedimentary basin, as well as along the Wopmay fault zone. Additional prospects can be found in [15,159,160].
Figure 6. Simplified geology of the exposed segment of the Great Bear magmatic zone (Canada) from the Camsell River district to the north and the NICO deposit to the south (modified after [15]). The distribution of 1.87 Ga volcanic rocks were updated from Corriveau et al. [15] using Jackson et al. [58] and Hildebrand [158]. S is the location of stromatolites; E of evaporites (inferred from metasomatic pseudomorphing of the structures typical of nodular anhydrite and evaporite beds); Bas of basalt; And of andesite; Dac of dacite; and Rhy>>And of rhyolite units more abundant than andesite. Representative IOA, IOCG, and MI-Co deposits and prospects occur within the 1.87 Ga volcanic centers and the 1.88 Ga sedimentary basin, as well as along the Wopmay fault zone. Additional prospects can be found in [15,159,160].
Minerals 15 00365 g006
Figure 7. A mineral deposit model for the MIAC system that is framed according to the relative timing among alteration facies, as regularly observed across MIAC systems (modified from Corriveau et al. [7] and Gadd et al. [210]). The representative mineral resources that support the metal associations for each facies are listed in Table A1, Table A2 and Table A3. Arrows grade in color from dark red (hottest) to grey (coolest).
Figure 7. A mineral deposit model for the MIAC system that is framed according to the relative timing among alteration facies, as regularly observed across MIAC systems (modified from Corriveau et al. [7] and Gadd et al. [210]). The representative mineral resources that support the metal associations for each facies are listed in Table A1, Table A2 and Table A3. Arrows grade in color from dark red (hottest) to grey (coolest).
Minerals 15 00365 g007
Table 1. Alteration facies at the heating stage of MIAC systems that serve as a ground preparation for subsequent mineralization. Additional examples of deposits are listed in Table A1, Table A2 and Table A3. Deposits listed or their districts are located in Figure 1.
Table 1. Alteration facies at the heating stage of MIAC systems that serve as a ground preparation for subsequent mineralization. Additional examples of deposits are listed in Table A1, Table A2 and Table A3. Deposits listed or their districts are located in Figure 1.
Alteration FaciesMinerals 1General CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-PoorAccessory
Facies 1a
Na
Ab–Olc;
res. Qz
Mnz, Ru, Ttn, ZrnBarren
  • Preferential distribution along regional-to-local discontinuities (e.g., lithological contacts, roofs of laccoliths, faults, and shear zones)
  • A regional extent along albitite corridors (≤100 s of km in length and a few km in width) in zones of enhanced crustal permeability
  • Ab replaces most minerals in host rocks, not only feldspars
  • Facies replaces any host rocks
  • Transient porosity and damage zones in albitite favour brecciation and serve as ground preparation for mineralization
  • Fine-grained to locally very coarse-grained where adjacent to coeval intrusions
  • Albitite corridors are diagnostic of a MIAC system
  • Regularly overprinted by all fertile alteration facies, with Fe-poor alkali-calcic alteration components of Facies 5 being most common
Barren when formed, e.g., Southern Province, Great Bear, CAN; Kethri Cu Belt and its Albitite Line, IND; Central Andes, CHL-PER; Cloncurry district and Mt Isa Province, AUS; Norrbotten district, SWE; Kuusamo belt, FIN; and Middle–Lower Yangtze Metallogenic Belt, CHN.
Host to subsequent albitite-hosted mineralization.
Sources: [10,11,12,13,14,15,22,26,52,54,55,66,67,68,69,70,76,86,87,88,89,90,91,92,93,94,95,96,97,98,99,100,101,102,103,104,105,106,107,108,109,110].
Facies 1b
Na-Ca
Ab–Olc, Anh, Scp; res. QzAmp, Cpx, Mnz, Ru, Ttn, ZrnBarren
Facies 1c
Skarn
Mg-Fe Amp (Tr), Mg-Fe Chl, Cpx, Grt, Phl, Tlc, res. Cb Barren
  • Localized in carbonate-bearing units
  • Spatially associated with, and pre, syn, or post Facies 1 Na alteration; can cut albitite
  • No causative intrusions; driven by the heat of fluids
  • Fine-grained but can be medium-grained
  • Distal-to-proximal to other MIAC alteration facies
Barren when formed, e.g., Great Bear, CAN; Punt Hill (Olympic Cu-Au Province), AUS; and Middle–Lower Yangtze Metallogenic Belt, CHN.
Host to subsequent skarn-hosted mineralization
Sources: [14,64,92,111].
1 The abbreviations follow [82], except for oligoclase (Olc), carbonates (Cb), sulfides (Sulf), silicates (Sia), As (arsenides), AsS (sulfarsenides), minerals (min.), remobilized (rem.), and residual (res.). Abbreviations for countries: AUS—Australia, BRA—Brazil, CAN—Canada, CHL—Chile, CHN—China, FIN—Finland, IND—India, MRT—Mauritania, PER—Peru, SWE—Sweden, VNW—Vietnam, and USA—United States.
Table 2. The field geology attributes of Facies 2 and transitional Facies 2–3, and the associated deposit types of the Metasomatic Iron (MI) class. Abbreviations for minerals and countries as per Table 1.
Table 2. The field geology attributes of Facies 2 and transitional Facies 2–3, and the associated deposit types of the Metasomatic Iron (MI) class. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-Rich
Fe-Poor
AccessoryMain
Accessory
Facies 1–2
HT Na-Ca-Fe
Amp (Act, Hbl), Cpx, Mag
Ab–Olc
Ap, Mnz, Ru, Qz, Ttn, ZrnMag, Py
  • Close to thermal core of MIAC systems
  • Overprint/spatially above and evolving from albitite
  • Proximal/evolving to HT Ca-Fe facies
  • Medium-grained-to-coarse-grained with pseudo-pegmatitic textures
  • Onset of Fe-rich alteration types in certain MIAC systems
  • Bt-Mag-Ab can precipitate and be associated with Py and Pyh, with no or only minor Ccp as it transitions to Facies 3
Proximal to iron oxide (IO) and IOA deposits.
Incipient REE mineralization possible.
Examples: Great Bear magmatic zone, CAN and Moonta-Wallaroo district (Olympic Cu-Au Province), AUS. Sources: [12,14,55,61,92].
Facies 2–1
Fe-rich skarn
Figure 3A
Fe-rich Cpx and Grt, Mag, Pyh
res. Cb
Amp, Aln, Ap, Qz, TtnMag, Sch
  • Close to the thermal core of a MIAC system
  • No causative intrusion; driven by heat of fluids
  • Significant increase of Fe in minerals
  • Fine-to-very-coarse-grained
  • Commonly overprinted by HT Ca-Fe alteration
MI-W, e.g., Fostung (Wanapitei district), CAN and Punt Hill, AUS.
Sources: [6,7,111,112].
Facies 2a
HT Ca-Fe (Fe, P, or REE endowed)
Figure 3B–D
Fe-rich Amp (Act, Hbl), Ep, Mag or Pyh
Ap, Qz
res. Cpx and Grt; TtnAp, Mag, REE-min. (e.g., Aln, Adr, Bri, Mnz Flr)
Ccp, Py, Thr, Urn;
rem. Sch
  • Thermal core of MIAC systems; peak temperatures recorded in hydrosaline liquid and salt melt inclusions
  • In zones of enhanced crustal permeability proximal to or within local-to-regional structural corridors or above sub-volcanic intrusions (generally of intermediate composition)
  • Ductile behavior is common with the development of a mineral foliation and local folding
  • Can occur as fluidized IOA breccias, including with dyke-like morphologies
  • Fine-to-very-coarse-grained
  • Amp-dominant to Mag-dominant to Ap-dominant
  • Veins, replaces, or overprints and/or is spatially above albitite
  • Infills albitite breccias and can replace its fragments; Mag-selective to pervasive alteration of fragments is common
  • Pyh precipitates instead of Mag with Amp, and it can be the main Fe-rich minerals in geological environments that are rich in graphitic sedimentary units
  • Commonly overprinted by Facies 2–3, 3, and 5a
MI-Fe (IO, IOA) deposits, e.g., El Laco, CHL; Kiruna, Per Geiger, Malberget (Norrbotten district), SWE; Pilot Knob, Pea Ridge (southeast Missouri district), USA; Middle–Lower Yangtze Metallogenic Belt, CHN.
Magnetite-skarn as magnetite-dominant HT Ca-Fe overprint on skarn.
REE remobilization from IOA apatite forms MI-REE deposits, e.g., Josette, CAN and Pea Ridge (southeast Missouri district), USA.
Sources: [6,11,14,27,61,70,77,78,110,113,114,115,116].
Facies 2b
HT Ca-(Mg)-Fe (Ni-endowed)
Figure 3E
Fe-rich Amp (Act, Hbl), Mag, Pyh, Py
Ap, Qz
Bt
res. Cpx and Grt
Mlr, Pn
Ccp
  • In or very close to the thermal core of a MIAC system
  • In provinces with abundant mafic–ultramafic units
  • Formed in discrete zones of enhanced crustal permeability
  • Can occur as fluidized breccias
  • Precedes Cu-rich mineralization types
MI-Ni deposits, e.g., Jaguar, Jatobá, GT-34 (Carajás), BRA
Sources: [117,118,119,120].
Facies 2–3
HT Ca-K-Fe to Bt-rich HT K-Fe (Bt > Kfs)
 
Figure 3F
Fe-rich Amp (Act, Hbl), Bt, Mag, Pyh, Sid
Ap, Kfs, Qz
res. Cpx and Grt, Flr, Ilm, TtnApy, Co-As/AsS, Bin, Au in As-min. or native
Overprint: Bin, Emp, Mdo, native Au/Bi
Ccp, Mlr, Mol, Py, Sch
  • In sedimentary sequences where magmatic events occur
  • In ultramafic-to-mafic environments
  • In zones of enhanced crustal permeability proximal to, or within, local-to-regional structures
  • Commonly stratabound and replacing earlier stratabound Facies 2 Amp-dominant HT Ca-Fe alteration
  • Replacement by Facies 3 biotite-rich K-Fe alteration is common where Cu mineralization occurs
  • Ductile behavior with mineral foliation and local folding
  • Mineralization hosted in, or outside of, the shoulders of chemically reactive rocks (carbonates)
  • Minor-to-absent brecciation associated with mineralization
  • Fine-to-medium-grained
MI-Co deposits, e.g., Guelb Morghein, MRT; NICO, CAN; Sin Quyen, VNM; Blackbird, Ram, Sunshine, Sunshine East, Merle (Idaho Cobalt Belt), USA; and components of Ernest-Henry, AUS.
Sources: [7,10,15,17,41,61,78,80,121,122,123,124,125,126].
Facies 2–3
HT Ca-K-Fe (Kfs>Bt)
Fe-rich Amp (Act), Bt, Mag, Pyh, KfsAp, Ttn, ZrnCcp, Py, Urn
  • Evolves from HT Ca-Fe and transitions to and is proximal to a HT K-Fe alteration in felsic and intermediate volcanic and plutonic rocks
  • Incipient brecciation with Amp- or Amp-Mag infill and Kfs-altered fragments and K-feldspar haloes to breccias
  • Amp and Amp-Mag veins with Kfs haloes
  • Fine-to-medium-grained
Halo of MI-Cu deposits.
Breccia infill and Kfs-altered ore breccia, e.g., Ernest Henry, AUS
Sources: [7,10,15,17,70,126].
Table 3. Facies 3 HT K-Fe alteration and the onset of Cu mineralization in the MI deposit class. Abbreviations for minerals and countries as per Table 1.
Table 3. Facies 3 HT K-Fe alteration and the onset of Cu mineralization in the MI deposit class. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-rich
Fe-poor
AccessoryMain
Accessory
Facies 3
HT K-Fe (Bt-rich)
 
Figure 4B,C
Bt, Mag and/or Pyh
Kfs, Qz
Ap, Amp, GrtBrn, Ccp, Py, native Au, Urn
  • In sedimentary-, volcano-sedimentary-, volcaniclastic-, or mafic and ultramafic-dominant environments
  • Stratabound alteration and mineralization
  • Mineralization in areas of enhanced crustal permeability along a local-to-regional discontinuity
  • Weak-to-absent brecciation associated with mineralization
  • Fine-to-medium-grained
  • Commonly replaces Facies 2–3 or Facies 1 albitite; breccias are rare
  • Pyh and Py can be the main Fe-rich minerals
MI-Cu, including Mag-group IOCG and ISCG, e.g., components of Candelaria, CHL; Dahongshan (Kangdian district), CHN; and components of Ernest-Henry (Cloncurry district), AUS.
ISCG, which are typical of geological sequences with graphitic units.
Sources: [64,70,74,75,123,126,127,128,129].
Facies 3
HT K-Fe (Kfs-rich)
 
Figure 4A,F
Bt, Mag
Kfs, Qz
Ap, Brt, Flr, Ilm, Mnz, Thr, Ttn, ZrnCcp, native Au, Py, Urn
Apy, Bn, Mol, Pyh
  • In felsic-to-intermediate volcanic and plutonic environments
  • Breccia-hosted mineralization marking a transition to brittle conditions
  • In breccias, Kfs replaces fragments and Mag and Sulf precipitate in matrix; with increasing intensity, matrix assemblages replace Kfs-altered fragments; can develop in albitite
  • Mag veins with Kfs haloes cut Facies 1 albitite; fragment in albitite can be Mag-altered in their cores and Kfs-altered at their margin
  • Structurally controlled mineralization in zones of enhanced crustal permeability
  • Fine-to-medium-grained
MI-Cu, including Mag-group IOCG, e.g., components of Candelaria, CHL; Ernest Henry (Cloncurry district), AUS; and Sue-Dianne, CAN.
MI-U, including albite-hosted U, e.g., Southern Breccia (Great Bear magmatic zone), CAN.
Sources: [7,10,11,14,28,35,51,70,72,87,126,130].
Table 4. Facies 4 barren K-felsite and barren-to-mineralized K-skarn. Abbreviations for minerals and countries as per Table 1.
Table 4. Facies 4 barren K-felsite and barren-to-mineralized K-skarn. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-rich
Fe-poor
AccessoryMain
Accessory
Facies 4a
K-felsite
Kfs
  • At the Mag-to-Hem transition
  • Commonly brecciated and forming haloes to Fe-oxide breccias
  • Can replace albitite and K-skarn
  • Fine-grained
Haloes to zones of mineralization, e.g., Ernest-Henry (Cloncurry district), AUS.
Sources: [14,70].
Facies 4b
K-skarn
Ank, Mag
Aln, Cpx (Adr-Grs), Ep, Kfs
Cal, DolCcp, Gn, Mol, Py, Sp
  • In carbonate sedimentary units or carbonate-altered rocks
  • Commonly brecciated
  • Fine-to-medium-grained
  • Can be replaced by K-felsite
  • Overprinted by and host to Facies 5 mineralization
Example: Mile Lake prospect (Great Bear magmatic zone), CAN.
Source: [56].
Table 5. Facies 5 alteration and associated mineralization in the MI deposit class. Abbreviations for minerals and countries as per Table 1.
Table 5. Facies 5 alteration and associated mineralization in the MI deposit class. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-rich
Fe-poor
AccessoryMain
Accessory
Facies 5a
LT (K, Ba, Ca, Na)-(CO2, F, H+)-Fe
 
Figure 4D,E
Chl, Ep, Hem, Sid
Ab, Ank, Brt, Cal, Dol, Flr, Kfs, Msc, Qz
Anh, Ank, Ap, Dol, TurBrn, Ccp, Cct, LREE-min., native Au, Py, Urn
Apy, Mol, Sch
  • In felsic-to-intermediate volcanic and plutonic environments
  • Breccia-hosted mineralization
  • Structurally controlled zonation of Cu-Sulf in deposits
  • Fine-to-medium-grained
  • Commonly overprints Facies 2 IOA mineralization and albitite and can form distally from them.
MI-Cu, e.g., Hem-group IOCG Olympic Dam, AUS.
MI-U, e.g., albitite-hosted U Michelin (Central Mineral Belt), CAN.
Sources: [2,10,33,43,71,107,131,132].
Aln, Amp (Act, Hst, Stp), Chl, Ep, Grt, Hem, Mag
Ab, Ank, Brt, Cal, Dol, Flr, Kfs, Msc, Qz
ApBrn, Ccp, Cct, LREE-min., native Au, Py, Urn
Gn, Mol, Sp
  • In environments rich in carbonate units or in mafic-ultramafic units
  • Breccia-, shear-hosted, K-skarn-hosted, or albitite-hosted mineralization
  • Fine-to-medium-grained
MI-Cu (e.g., Hem-group IOCG and K-skarn hosted; Sossego (Carajás), BRA; and Hillside, Punt Hill (Olympic Cu-Au Province), AUS.
MI-U, e.g., Moran Lake (Central Mineral Belt), CAN.
Sources: [6,7,56,92,107,111].
Act, Bt, Chl, Pyh, Py
Msc, Qz
Ap, Cb, LREE-min., TurBrn, Ccp
Gn, Mol, Sp
  • In geological environments with abundant graphitic units
  • Structurally controlled in fault or shear zones
  • Sulfide replacement lodes
  • Fine-to-medium-grained
MI-Cu, e.g., Eloise, Jericho (Cloncurry district), AUS; Delhi Pacific (Romanet Horst), CAN.
Sources: [50,133].
Facies 5b
Si-rich LT (Ca, Mg)-(K, Na)-Si-Fe; Figure 4G
Amp, Chl, Ep, Hem, Mag, Pyh
Ab, Ank, Brt, Dol, Flr, Qz
LREE-min., Kfs, MscNative Au, Py
Ccp
  • Breccia, shear-hosted, or replacement zones
  • Can be albitite- or skarn-hosted
  • At the transition from aluminosilicates + Fe min. to Qz + Fe min.
  • Fine-to-medium-grained
MI-Au, e.g., Scadding (Wanapitei district), CAN.
MI-Ag.
Sources: [7,22,33,80].
Table 6. Fe-poor alkali-calcic alteration facies forming deposits of the MAC class. Abbreviations for minerals and countries as per Table 1.
Table 6. Fe-poor alkali-calcic alteration facies forming deposits of the MAC class. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-rich
Fe-poor
AccessoryMain
Accessory
Facies 5c
K/CO2
 
Figure 5B,C
Ank
Chl, Cal, Dol (ferroan), Kfs, Msc, Qz, Tur
Ccp, Mol, Native Au, Pyh, Py
Brn
  • Contemporaneous with Fe-rich metasomatism
  • Typically overprints or formed close to earlier corridors of albitite
  • Structurally controlled and generally hosted in fault zones
  • Occur as breccias, replacements, or networks of veins
  • Probable remobilization and reconcentration of metals from earlier mineralized alteration facies
  • Fine-grained-to-pegmatitic with hydrothermal textures
MAC-Cu, e.g., Mount Dore (Cloncurry district), AUS; Alwyn (Wanapitei district), CAN; Pahtohavare (Norrbotten district), SWE.
MAC-Mo, e.g., Merlin (Cloncurry district), AUS.
Hanging wall of MI-REE, e.g., Per Geijer (Norrbotten district), SWE.
Sources: [22,95,96].
Facies 5d
Si/CO2
 
Figure 5A
Cal, Chl, Dol (ferroan), QzAb, Kfs, MscNative Au, Pyh, Py, Urn
Ccp, LREE-min.
MAC-Au, e.g., Tick Hill (Cloncurry district), AUS; Kish showing (Romanet Horst), CAN.
Sources: [50,131].
Table 7. Waning stage of MIAC systems at the transition to epithermal conditions and metal remobilization from pre-existing alteration facies of a MIAC system with post-MIAC fluid circulation. Abbreviations for minerals and countries as per Table 1.
Table 7. Waning stage of MIAC systems at the transition to epithermal conditions and metal remobilization from pre-existing alteration facies of a MIAC system with post-MIAC fluid circulation. Abbreviations for minerals and countries as per Table 1.
Alteration FaciesMineralsGeneral CharacteristicsAssociation with Mineralization and Case Examples
AlterationMineralization
Fe-rich
Fe-poor
Main
Facies 6
K-Si-Al ± Fe-Ba epithermal caps
Hem
Phyllic, sericitic, silicic, advanced argillic,
silicification
Py
  • Apex of system near or at paleosurface
  • Peak acidic conditions and thermal low of a MIAC system
  • Occurs as veins, vein networks, or replacement zones
  • Postdates MIAC and MAC alteration
  • Metal reconcentration from earlier alteration facies and mineralization
Epithermal mineralization, e.g., Gossan Island, Echo Bay Gossan (Great Bear magmatic zone), CAN; Central Andes, CHL
[10,55,56,64,84].
Facies 6
LT Fe-(Si,CO2) ± Ba
Hem
Brt, Cal, Dol, Qz
As–AsS, Ccp, Gn, Sp, Urn
  • Postdates MIAC and MAC alteration
  • Veins, vein networks, and breccias in brittle structures
  • Metal signature of the host facies
  • Giant quartz veins (≤ tens of km); more local silica flooding
  • Varied time gap between primary MIAC system and recirculation or addition of fluids and metals
Vein-type mineralization, five-element veins, e.g., Camsell River and Port Radium-Echo Bay districts (Great Bear magmatic zone), CAN; Olympic Cu-Au Province, AUS; Lufilian arc, Africa
[21,56,60,85,134,135,136,137].
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

Corriveau, L.; Montreuil, J.-F. Metasomatic Mineral Systems with IOA, IOCG, and Affiliated Critical and Precious Metal Deposits: A Review from a Field Geology Perspective. Minerals 2025, 15, 365. https://doi.org/10.3390/min15040365

AMA Style

Corriveau L, Montreuil J-F. Metasomatic Mineral Systems with IOA, IOCG, and Affiliated Critical and Precious Metal Deposits: A Review from a Field Geology Perspective. Minerals. 2025; 15(4):365. https://doi.org/10.3390/min15040365

Chicago/Turabian Style

Corriveau, Louise, and Jean-François Montreuil. 2025. "Metasomatic Mineral Systems with IOA, IOCG, and Affiliated Critical and Precious Metal Deposits: A Review from a Field Geology Perspective" Minerals 15, no. 4: 365. https://doi.org/10.3390/min15040365

APA Style

Corriveau, L., & Montreuil, J.-F. (2025). Metasomatic Mineral Systems with IOA, IOCG, and Affiliated Critical and Precious Metal Deposits: A Review from a Field Geology Perspective. Minerals, 15(4), 365. https://doi.org/10.3390/min15040365

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

Article Metrics

Back to TopTop