Control Estructural
Control Estructural
of Mineral Deposits
Theory and Reality
Edited by
Alain Chauvet
Printed Edition of the Special Issue Published in Minerals
www.mdpi.com/journal/minerals
Structural Control of Mineral Deposits
Structural Control of Mineral Deposits
Theory and Reality
Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland
This is a reprint of articles from the Special Issue published online in the open access journal Minerals
(ISSN 2075-163X) from 2018 to 2019 (available at: https://www.mdpi.com/journal/minerals/
special issues/structural control deposits).
For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:
LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Article Number,
Page Range.
c 2019 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents
Alain Chauvet
Editorial for Special Issue “Structural Control of Mineral Deposits: Theory and Reality”
Reprinted from: Minerals 2019, 9, 171, doi:10.3390/min9030171 . . . . . . . . . . . . . . . . . . . . 1
Alain CHAUVET
Structural Control of Ore Deposits: The Role of Pre-Existing Structures on the Formation of
Mineralised Vein Systems
Reprinted from: Minerals 2019, 9, 56, doi:10.3390/min9010056 . . . . . . . . . . . . . . . . . . . . 5
Alexandre Cugerone, Emilien Oliot, Alain Chauvet, Jordi Gavaldà Bordes, Angèle Laurent,
Elisabeth Le Goff and Bénédicte Cenki-Tok
Structural Control on the Formation of Pb-Zn Deposits: An Example from the Pyrenean
Axial Zone
Reprinted from: Minerals 2018, 8, 489, doi:10.3390/min8110489 . . . . . . . . . . . . . . . . . . . . 27
Antonio Funedda, Stefano Naitza, Cristina Buttau, Fabrizio Cocco and Andrea Dini
Structural Controls of Ore Mineralization in a Polydeformed Basement: Field Examples from
the Variscan Baccu Locci Shear Zone (SE Sardinia, Italy)
Reprinted from: Minerals 2018, 8, 456, doi:10.3390/min8100456 . . . . . . . . . . . . . . . . . . . . 47
Yang Song, Chao Yang, Shaogang Wei, Huanhuan Yang, Xiang Fang and Hongtao Lu
Tectonic Control, Reconstruction and Preservation of the Tiegelongnan Porphyry and
Epithermal Overprinting Cu (Au) Deposit, Central Tibet, China
Reprinted from: Minerals 2018, 8, 398, doi:10.3390/min8090398 . . . . . . . . . . . . . . . . . . . . 88
Johann Tuduri, Alain Chauvet, Luc Barbanson, Jean-Louis Bourdier, Mohamed Labriki,
Aomar Ennaciri, Lakhlifi Badra, Michel Dubois, Christelle Ennaciri-Leloix, Stanislas Sizaret
and Lhou Maacha
The Jbel Saghro Au(–Ag, Cu) and Ag–Hg Metallogenetic Province: Product of a Long-Lived
Ediacaran Tectono-Magmatic Evolution in the Moroccan Anti-Atlas
Reprinted from: Minerals 2018, 8, 592, doi:10.3390/min8120592 . . . . . . . . . . . . . . . . . . . . 105
Alexis Grare, Olivier Lacombe, Julien Mercadier, Antonio Benedicto, Marie Guilcher,
Anna Trave, Patrick Ledru and John Robbins
Fault Zone Evolution and Development of a Structural and Hydrological Barrier: The Quartz
Breccia in the Kiggavik Area (Nunavut, Canada) and Its Control on Uranium Mineralization
Reprinted from: Minerals 2018, 8, 319, doi:10.3390/min8080319 . . . . . . . . . . . . . . . . . . . . 153
v
Ingrid B. Maciel, Angela Dettori, Fabrizio Balsamo, Francisco H.R. Bezerra,
Marcela M. Vieira, Francisco C.C. Nogueira, Emma Salvioli-Mariani and
Jorge André B Sousa
Structural Control on Clay Mineral Authigenesis in Faulted Arkosic Sandstone of the Rio do
Peixe Basin, Brazil
Reprinted from: Minerals 2018, 8, 408, doi:10.3390/min8090408 . . . . . . . . . . . . . . . . . . . . 181
Tao Sun, Ying Xu, Xuhui Yu, Weiming Liu, Ruixue Li, Zijuan Hu and Yun Wang
Structural Controls on Copper Mineralization in the Tongling Ore District, Eastern China:
Evidence from Spatial Analysis
Reprinted from: Minerals 2018, 8, 254, doi:10.3390/min8060254 . . . . . . . . . . . . . . . . . . . . 198
Safouane Admou, Yannick Branquet, Lakhlifi Badra, Luc Barbanson, Mohamed Outhounjite,
Abdelali Khalifa, Mohamed Zouhair and Lhou Maacha
The Hajjar Regional Transpressive Shear Zone (Guemassa Massif, Morocco): Consequences on
the Deformation of the Base-Metal Massive Sulfide Ore
Reprinted from: Minerals 2018, 8, 435, doi:10.3390/min8100435 . . . . . . . . . . . . . . . . . . . . 224
vi
About the Special Issue Editor
Alain Chauvet (Dr. HDR) is a senior CNRS Researcher at the Géosciences Montpellier laboratory,
France. After a Ph.D. in Structural Geology devoted to the late-orogenic extension in Norway,
he specialized in Tectonic Control of Ore Deposits with a focus on the perigranitic mineralizations
of South America, China, Europe, . . . , and the characterization of the relationships between Large
Igneous Provinces and Mineralizations in North Africa. He is involved in several collaborative
projects and expertises with industrial mining companies.
vii
Preface to ”Structural Control of Mineral Deposits”
This compilation of publication results from more than 20 years of questioning and of applying
structural geology in mining geology by the guest editor. If it is common to place the various deposits
of the earth into large classes that allow recognizing and identifying some characters useful to detect,
explore, and find other similar deposits, experience demonstrates that each deposit is unique and
cannot answer perfectly to a generic model. This is why we suspect that there exists a gap between
theory (i.e., the classical model) and reality that needs to be estimated and taken into account in
any type expertise or study of an unknown mineral deposit. The following publications try to be
concerned by this way of working.
My knowledge and interest in the structural control of mineral deposits benefited from several
discussions, suggestions, and field trip shared with a lot of persons that are greatly acknowledged
here. An exhaustive list is impossible, but I want to particularly acknowledge the CVRD (Vale),
Buenaventura, Cedimin, Managem, Kasbah Resource, CMS, SMI, and CTT mining companies and
all geologists that took the time to discuss with me of structural problems in mining geology, with
a special mention of A.S. André, L. Badra, L. Bailly, L. Barbanson, Y. Branquet, X. Charonnat,
A. Ennaciri, C. Ennaciri, M. Faure, L. Fontboté,. A. Gaouzi, S. Gialli, E. Gloaguen, M. Iseppi;
K. Kouzmanov, J. Onezime, P. Piantone, S. Sizaret, E. Tourneur, J. Tuduri, N. Volland. All the reviewers
that significantly improved the quality of this book are also warmly acknowledged.
Alain Chauvet
Special Issue Editor
ix
minerals
Editorial
Editorial for Special Issue “Structural Control of
Mineral Deposits: Theory and Reality”
Alain Chauvet
UMR 5243, Géosciences Montpellier, University of Montpellier, cc 60, CEDEX 5, 34095 Montpellier, France;
[email protected]; Tel.: +33-(0)4-67-14-48-57
“Structural Control” remains a crucial point that is frequently absent in scientific and/or economic
analyses of ore deposits, whatever their type and class, although a selection of references illustrates its
importance [1–5]. The case of lode deposits is particularly adapted, but other types, like breccia pipes,
stockwork, massive sulphides, skarn, etc., also concern Structural Control. Works on the Structural
Control of ore deposits are not abundant in the recent literature, and, as frequently suggested, structural
geology often is not sufficiently developed in the exploration programs of many mining camp’s
strategies. A few compilations have been devoted to this theme in the last two decades, such as (i) the
special publication of the Geological Society of London, concerned with the link between fracturing,
flow, and mineralization [6], (ii) the review of the Society of Economic Geology, devoted to Structural
Control [7], (iii) a special publication of the Geological Society of London, looking to study the genetic
link that can exist between mineralization and orogenic domains [8], and finally, (iv) a special issue
of the Journal of Structural Geology, devoted to the application of Structural Geology in mineral
exploration and mining [9]. In addition to these four compilations, only a few publications have been
concerned with this theme, and most of them are dated before the year 2000. These publications mostly
concerned vein internal infilling textures [10,11], the vein formation model, with the contribution and
controversy of the crack seal, dissolution-precipitation, diffusion, and seismic-valves mechanisms
(e.g., [12–16]). In his review, Chauvet [17] discussed of some of these concepts, in order to highlight
the role and the significance of pre-existing structures in the formation of vein-style deposits.
Three publications of this volume explore the development of mineralization in the specific
context of orogenic domains. Cugerone et al. [18] offer a detailed study of a rather complex Pb–Zn
mineralisation developed within the orogenic Hercynian Pyrenees during two mineralization stages,
each of them linked with a deformational event. The syntectonic primary mineralization is remobilized
and helps the formation of the second one. The same approach is used within the two following
contributions on the same theme [19,20]. Funedda et al. [19] and Fridovsky et al. [20] also used
a detailed description of the relationships between mineralization and deformation in deformed
domains, such as the Variscan domain of Sardinia and the Verkhovansk-Kolyma folded region of
NE Russia. Funedda et al. [19] pay close attention to the opening process of structures that will
serve as traps for mineralised fluid catching, a fact that is fundamental in any tectonic understanding
of a mineralised vein system [17]. Fridovsky et al. [20] also proposed a pluri-deformational model
associated with multiple stages of mineralisation formation.
The relationship between magmatism, regional tectonic context, and mineralization remain a
question that has still been debated in several recent publications [21,22], thus demonstrating that
this question is still relevant and may help in the distinction between intrusion-related, orogenic
deposits and the Cu–Au-rich porphyry types. Two contributions explore new methods of investigation
that provide an innovative vision of the relationship between magmatism and mineralization.
Song et al. [23] examine the consequences of the telescoping of two mineralized systems (a subsequent
epithermal system affects a primary porphyric one within the Tiegelongnan Porphyry and the
epithermal overprinting Cu (Au) deposit, Central Tibet, China) with a focus on the role of the
dislocation effects on ore reserve calculations and future deposits discoveries. Tuduri et al. [24]
suggest an original way to demonstrate the genetic link between mineralization and magmatism by
establishing that both are developed in the same regional tectonic context, in the highly mineralised
Moroccan Anti-Atlas. This contribution represents an indirect but efficient way to relatively date the
emplacement of magmatism and mineralization formations, and their relationships.
In the past, the concept of a gold-bearing shear zone has not given satisfying results in terms of
our understanding of gold deposits, and has been more or less totally abandoned, except within few
specific sectors of the Canadian shield in which the role of major crustal faults is still at the centre
of the accepted models [25]. In the domain of economic geology, faults are fundamental structures
that can have two contrasting behaviours: (i) Hydrogeological barriers that help the concentration
of ore, as demonstrated by the contribution of Grare et al. [26] in the case of the Kiggavik uranium
example (Canada), and (ii) a zone of permeability that can favour fluid circulation and can serve
as a guide for the mineralisation trapping. The work of Maciel et al. [27] proposes a surprising
example in which fault occurrences have a negative role for clay authigenesis efficiency; this work also
discusses the consequence on reservoir characteristics. Sun et al. [28] end the section on relations with
brittle tectonics by presenting an innovative GIS-based spatial analysis of mineral deposit patterns
in correlation with detailed structural features, in order to propose some implications on Structural
Control. The chosen example was provided from the Copper deposit of the Tongling Ore district of
Eastern China.
Concerning other orebodies than vein-type ones, volcanic-hosted massive sulphide deposits
(VHMS) have been recently the subject of much debate, specifically with the suggestion of a significant
contribution of “replacement processes” in their modes of formation [29,30]. In addition, it has
been demonstrated that stockwork within VHMS environments can result in subsequent syntectonic
veining instead of earlier veins related to feeder zones [31]. Indeed, the observation of stockwork
within a VHMS context needs to be considered with particular attention because of the possible
coexistence of the two types of stockwork: the one related to the feeder zone and the other the result of
subsequent deformation [17]. It has been suggested that the second event and associated metal may
contribute significantly to a relative enrichment in VHMS environment. Without any reference to some
replacement process, the contribution of Admou et al. [32] ends the special issue with a very attractive
formation model of the Moroccan Guemassa VHMS deposit, strongly involving the active role of
normal faults and Structural Control, since the beginning of the volcanic activity. In fact, it appears
that most of the VHMS deposits certainly do not present the classical geometrical model exhibited
within all teaching books, but instead form by wall-rock replacement (metasomatism) strongly helped
by the re-using of pre-existing structures, such as folds, unconformities, and/or fault and deformation
features. Such a contribution is frequently underestimated.
References
1. Forde, A.; Bell, T.H. Late structural control of mesothermal vein-hosted gold deposits in Central Victoria,
Australia: Mineralization and exploration potential. Ore Geol. Rev. 1994, 9, 33–59. [CrossRef]
2. Davis, B.K.; Hippertt, J.F.M. Relationships between gold concentration and structure in quartz veins from
the Hodgkinson Province, northeastern Australia. Mineral. Depos. 1998, 33, 391–405. [CrossRef]
3. Chauvet, A.; Piantone, P.; Barbanson, L.; Nehlig, P.; Pedroletti, I. Gold deposit formation during collapse
tectonics: Structural, mineralogical, geochronological, and fluid inclusion constraints in the Ouro Preto gold
mines, Quadrilátero Ferrífero, Brazil. Econ. Geol. 2001, 96, 25–48. [CrossRef]
4. Chauvet, A.; Bailly, L.; André, A.S.; Monié, P.; Cassard, D.; Llosa Tajada, F.; Vargas, J.R.; Tuduri, J. Internal
vein texture and vein evolution of the epithermal Shila-Paula district, southern Peru. Mineral. Dep.
2006, 41, 387–410. [CrossRef]
5. Tunks, A.J.; Cooke, D.R. Geological and structural controls on gold mineralization in the Tanami District,
Northern Territory. Mineral. Dep. 2007, 42, 107–126. [CrossRef]
2
Minerals 2019, 9, 171
6. McCaffrey, K.; Lonergan, L.; Wilkinson, J. Fractures, Fluid Flow and Mineralization; Geological Society of
London Special Publication: London, UK, 1999; Volume 155, 328p.
7. Richards, J.P.; Tosdal, R.M. Structural Controls on Ore Genesis. In Reviews in Economic Geology; Society of
Economic Geologists, Inc.: Littleton, CO, USA, 2001; 181p.
8. Blundell, D.J.; Neubauer, F.; Von Quadt, A. The Timing and Location of Major Ore Deposits in an Evolving Orogen;
Geological Society of London Special Publication: London, UK, 2002; Volume 204, 358p.
9. Vearncombe, J.R.; Blenkinsop, T.G.; Reddy, S.M. Applied Structural Geology for Mineral Exploration and
Mining. J. Struct. Geol. 2004, 26, 989–994. [CrossRef]
10. Dowling, K.; Morrison, G. Application of quartz textures to the classification of gold deposits using North
Queensland examples. Econ. Geol. Mon. 1990, 6, 342–355.
11. Dong, G.; Morrison, G.; Jaireth, S. Quartz textures in epithermal veins, Queensland - classification, origin,
and implication. Econ. Geol. 1995, 90, 1841–1856. [CrossRef]
12. Ramsay, J.G. The crack-seal mechanism of rock deformation. Nature 1980, 284, 135–139. [CrossRef]
13. Cox, S.F.; Etheridge, M.A. Crack-seal fibre growth mechanism and their significance in the development of
oriented layer silicate microstructures. J. Struct. Geol. 1983, 92, 147–170. [CrossRef]
14. Bons, P.D.; Jessell, M.W. Experimental simulation of the formation of fibrous veins by localised
dissolution-precipitation creep. Mineral. Mag. 1997, 61, 53–63. [CrossRef]
15. Boullier, A.M.; Robert, F. Paleoseismic events recorded in Archean gold quartz vein networks, Val-Dor,
Abitibi, Quebec, Canada. J. Struct. Geol. 1992, 14, 161–177. [CrossRef]
16. Bons, P.D.; Elburg, M.A.; Gomez-Rivas, E. A review of the formation of tectonic veins and their
microstructures. J. Struct. Geol. 2012, 43, 33–62. [CrossRef]
17. Chauvet, A. Structural Control of Ore Deposits: The Role of Pre-existing Structures on the Formation of
Mineralised Vein Systems. Minerals 2019, 9, 56. [CrossRef]
18. Cugerone, A.; Oliot, E.; Chauvet, A.; Gavaldà Bordes, J.; Laurent, A.; Le Goff, E.; Cenki-Tok, B. Structural
Control on the Formation of Pb-Zn Deposits: An Example from the Pyrenean Axial Zone. Minerals 2018, 8, 489.
[CrossRef]
19. Funedda, A.; Naitza, S.; Buttau, C.; Cocco, F.; Dini, A. Structural Controls of Ore Mineralization in a
Polydeformed Basement: Field Examples from the Variscan Baccu Locci Shear Zone (SE Sardinia, Italy).
Minerals 2018, 8, 456. [CrossRef]
20. Fridovsky, V.Y.; Kudrin, M.V.; Polufuntikova, L.I. Multi-Stage Deformation of the Khangalas Ore Cluster
(Verkhoyansk-Kolyma Folded Region, Northeast Russia): Ore-Controlling Reverse Thrust Faults and
Post-Mineral Strike-Slip Faults. Minerals 2018, 8, 270. [CrossRef]
21. Dressel, B.C.; Chauvet, A.; Trzaskos, B.; Biondi, J.C.; Bruguier, O.; Monié, P.; Villanova, S.N.; Newton, J.B.
The Passa Tres lode gold deposit (Parana State, Brazil): An example of structurally-controlled mineralisation
formed during magmatic-hydrothermal transition and hosted within granite. Ore Geol. Rev. 2018, 102,
701–727. [CrossRef]
22. Tuduri, J.; Chauvet, A.; Barbanson, L.; Labriki, M.; Dubois, M.; Trapy, P.H.; Lahfid, A.; Poujol, M.; Melleton, J.;
Badra, L.; et al. Structural control, magmatic-hydrothermal evolution and formation of hornfels-hosted,
intrusion-related gold deposits: Insight from the Thaghassa deposit in Eastern Anti-Atlas, Morocco. Ore Geol.
Rev. 2018, 97, 171–198. [CrossRef]
23. Song, Y.; Yang, C.; Wei, S.; Yang, H.; Fang, X.; Lu, H. Tectonic Control, Reconstruction and Preservation of
the Tiegelongnan Porphyry and Epithermal Overprinting Cu (Au) Deposit, Central Tibet, China. Minerals
2018, 8, 398. [CrossRef]
24. Tuduri, J.; Chauvet, A.; Barbanson, L.; Bourdier, J.L.; Labriki, M.; Ennaciri, A.; Badra, L.; Dubois, M.;
Ennaciri-Leloix, C.; Sizaret, S.; Maacha, L. The Jbel Saghro Au(–Ag, Cu) and Ag–Hg Metallogenetic Province:
Product of a Long-Lived Ediacaran Tectono-Magmatic Evolution in the Moroccan Anti-Atlas. Minerals
2018, 8, 592. [CrossRef]
25. Poulsen, K.H.; Robert, F. Shear zones and gold: Practical examples from the southern Canadian Shield.
Geol. Assoc. Can. Short Course Notes 1989, 6, 239–266.
26. Grare, A.; Lacombe, O.; Mercadier, J.; Benedicto, A.; Guilcher, M.; Trave, A.; Ledru, P.; Robbins, J. Fault Zone
Evolution and Development of a Structural and Hydrological Barrier: The Quartz Breccia in the Kiggavik
Area (Nunavut, Canada) and Its Control on Uranium Mineralization. Minerals 2018, 8, 319. [CrossRef]
3
Minerals 2019, 9, 171
27. Maciel, I.B.; Dettori, A.; Balsamo, F.; Bezerra, F.H.R.; Vieira, M.M.; Nogueira, F.C.C.; Salvioli-Mariani, E.;
Sousa, J.A.B. Structural Control on Clay Mineral Authigenesis in Faulted Arkosic Sandstone of the Rio do
Peixe Basin, Brazil. Minerals 2018, 8, 408. [CrossRef]
28. Sun, T.; Xu, Y.; Yu, X.; Liu, W.; Li, R.; Hu, Z.; Wang, Y. Structural Controls on Copper Mineralization in the
Tongling Ore District, Eastern China: Evidence from Spatial Analysis. Minerals 2018, 8, 254. [CrossRef]
29. Aerden, D.G.A.M. Formation of Massive Sulfide Lenses by Replacement of folds: The Hercules Pb–Zn Mine,
Tasmania. Econ. Geol. 1993, 88, 377–396. [CrossRef]
30. Perkins, W.G. Mount Isa lead-zinc orebodies: Replacement lodes in a zoned syndeformational
copper–lead–zinc system? Ore Geol. Rev. 1997, 12, 61–110. [CrossRef]
31. Chauvet, A.; Onézime, J.; Charvet, J.; Barbanson, L.; Faure, M. Syn- to late-tectonic stockwork emplacement
within the Spanish section of the Iberian Pyrite Belt: Structural, textural and mineralogical constraints in the
Tharsis-La Zarza areas. Econ. Geol. 2004, 99, 1781–1792. [CrossRef]
32. Admou, S.; Branquet, Y.; Badra, L.; Barbanson, L.; Outhounjite, M.; Khalifa, A.; Zouhair, M.; Maacha, L.
The Hajjar Regional Transpressive Shear Zone (Guemassa Massif, Morocco): Consequences on the
Deformation of the Base-Metal Massive Sulfide Ore. Minerals 2018, 8, 435. [CrossRef]
© 2019 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
4
minerals
Review
Structural Control of Ore Deposits: The Role of
Pre-Existing Structures on the Formation of
Mineralised Vein Systems
Alain CHAUVET
CNRS-UMR 5243, Géosciences Montpellier, University of Montpellier, cc 60,
34095 Montpellier CEDEX 5, France; [email protected]; Tel.: +3-34-6714-4857
Abstract: The major role played by pre-existing structures in the formation of vein-style mineral
deposits is demonstrated with several examples. The control of a pre-existing decollement level on the
formation of a crustal extension-related (collapse) gold deposit is first illustrated in the Quadrilátero
Ferrífero from Brazil. Shear zone and decollement structures were also examined and shown to control
veins formation by three distinct processes: (i) re-aperture and re-using of wrench shear zones in the
case of Shila gold mines (south Peru); (ii) remobilisation of metal in volcanic-hosted massive sulphide
(VHMS) deposit by subsequent tectonic events and formation of a secondary stockwork controlled by
structures created during this event (Iberian Pyrite Belt, Spain); (iii) formation of economic stockwork
by contrasting deformation behaviours between ductile black schist versus brittle more competent
dolomite (Cu-Ifri deposit, Morocco). Two examples involve changing of rheological competence
within zones affected by deformation and/or alteration in order to receive the mineralisation (case
studies of Achmmach, Morocco, and Mina Soriana, Spain). The last case underscores the significance
of the magmatic–hydrothermal transition in the formation of mesothermal gold deposits (Bruès
mine, Spain). All these examples clearly demonstrate the crucial role played by previously formed
structures and/or texture in the development and formation of ore deposits.
Keywords: vein; structure; textures; infilling; breccia; comb quartz; pull-apart; exploration;
pre-existing structures; decollement
1. Introduction
Numerous studies have been devoted to the process of vein formation mainly because of their
significance in term of tectonics and deformation (stress and strain determination), e.g., [1–5], but
also because of their significant economic interest in the case of metal-bearing veins, e.g., [6–8].
Several works have concentrated on the external geometry of veins and their relationships with the
mode of opening and, consequently, the local or regional stress field during vein formation [9–11].
Complementary studies have also integrated information that can be deduced from vein infilling
textures, such as the classical tripartite division in syntaxial (inward growth), antitaxial (outward
growth) and stretching veins (complex pattern with no consistent growth direction) [12–14]. In
economic geology, and particularly in vein-type deposits, the study of the nature and texture of vein
infilling is particularly important because it lies at the base of the ore-forming process itself. No recent
works have been concerned by this kind of analysis. The latest contributions [15–17] only deal with the
internal texture of ore deposits without considering the (external) geometry of the veins themselves.
This paper, as an introduction to the Minerals Special Issue “Structural Control of Ore Deposits,
Theory and Reality” focuses on the relationships between the shape and internal texture of ore-bearing
veins with the objectives to better understanding vein formation processes and, consequently, to
improve mining exploration strategies. I will present and discuss seven case studies of metal-bearing
veins with different modes of formation that highlight the role of pre-existing features on their
development. This aspect seems to be frequently underestimated, at least in the case of vein deposits,
and this work aims to demonstrate its significance in the development of exploration and exploitation
programs. The re-using of some previously formed structures has, in that case, a significant but passive
role with respect to the formation of the economic feature. This concept is already exemplified in
another contribution of this volume [18]. Each of the seven cases presented herein include a brief
overview of the regional geology and deformation history, followed by a detailed geometrical and
textural analysis of ore-bearing veins and a regional-scale genetic model that integrates these data. The
relationships between neo-formed structures versus pre-existing ones will be highlighted in each case
as well as its implications for regional vein distribution and, consequently, exploration programs.
- What tectonic context is responsible for trap formation? (the geometrical analysis)
- What is the mode and condition of filling? (the internal analysis)
6
Minerals 2019, 9, 56
Figure 1. Three examples of traps and voids mode of formation. (a) Model of formation of sigmoidal
veins by antithetic bedding-controlled gliding during fold development (a1). Example from the
Bourg-d’Oisans area (a2). (b) Stockwork formation in accommodation of shearing affecting some
multi-layers rocks with contrast of competence (b1, b2 and b3). Conceptual sketch (b4) in which yellow
veins represent the stockwork formed in more competent layers. (c) Trap opening within left lateral
pull-apart (c3). The two sketches c1 and c2 illustrate the process and show the two types of filling
encountered within these structures. See text for explanation.
It may be surprising to see that Figures 1a and 2, propose that a cleavage plane, reputed to be
a plane of maximum flattening, is used as a syntectonic trap for mineralisation during its formation.
7
Minerals 2019, 9, 56
However, this solution has been adopted and demonstrated within two works on this topic [23,24].
We will see, in the following examples, what are our arguments to defend this unconventional concept.
The case of Figure 2 should not raise the same concern because, here, cleavage serves as a receptacle
for ore concentration in a late deformation event that causes dilation parallel to a pre-existing cleavage.
Figure 2. Example of re-using and re-opening of a previously formed left-lateral shear zone during
two different states of stress. Red and green arrows correspond to the shortening direction of each
stage. Note that main infilling is realised during stage 2.
- Massive or buck texture: rare examples of this texture have been interpreted to result where voids
are filled after, not during their formation. This texture is frequently characterised by euhedral
or anhedral grain of variable size throughout the vein [15] due to uniform growth rates. Grain
orientation can also be highly variable. In fact, such a texture provides limited information about
the tectonic conditions prevailing during vein formation.
- A fibrous or comb crystal shape corresponds to crystallisation coeval with vein opening and
represents the more interesting texture for the topics of this study (Figure 3). Whether comb or
fibrous textures develop depends on the rate of trap opening versus crystal growth (see below).
Comb quartz is commonly related to (Figure 3a) (i) a supersaturated fluid invading an open space
(the initial fracture) with competitive crystal growth normal to the walls [16,25,26]; and (ii) a slow
opening rate of the fracture keeping pace with the rate of crystal growth [15,27]. Veins formed
by this process only differ from crack-seal veins [28–30] by the lack of fibrous crystallisation and
evidence for incremental cracking, such as successive and parallel inclusion trails. Indeed, fibrous
textures result from the same process as comb infilling, except that crystal growth is incremental
instead of continuous. In this case, the crack is caused by fluid overpressure and crystallisation
occurs immediately after the aperture with a unique free direction for crystal growth—the vein
centre. The succession of cracking event and, consequently, of immediate filling, explains the
continuous crystallisation and, therefore, the fibres (Figure 3b) [31]. By contrast, comb texture is
supposed to form where the rate of crystallisation is lower than the opening rate. In this case,
the crystallisation only covers the vein wall, and crystals are larger as in the case of fibrous veins
and can develop during multiple growth stages (Figure 3a), sometimes associated with a change
of fluid composition and chemistry [15]. It is still uncertain, though, if all fibrous veins form by
the crack-seal mechanism, or whether they can also form by continuous fibre growth, where
diffusion keeps pace with the rate of dilation [26,32].
8
Minerals 2019, 9, 56
Figure 3. Examples of comb and fibrous textures from the Hercynian mining district of Tras-os-Montes,
Galicia, Spain. Red lines illustrate the elongated quartz and feldspar grains indicate the opening
direction. Note the difference between fibres that cross-cut the veins and comb grains that do not
traverse the vein.
- Breccia textures are witnesses of complex processes for which we have to take into account three
parameters based on the recognition between fragments and matrix in order to understand their
process of formation:
The nature of the matrix or cement (rock flour, sediment, volcanic, magmatic,
hydrothermal, . . . );
The nature and shape of the fragments (circularity, size, distribution, fabrics, monogenic
or polygenic, lithological nature);
The relationships between fragments and matrix/cement (matrix-supported or
grain-supported).
As a function of these three parameters, a genetic classification of breccia has been proposed
by [33], which is frequently used as an indicator of the conditions of vein formation [34–39]. The most
used in lode-related economic geology are the tectonic, hydrothermal, magmatic, collapse-related,
crackle, hydraulic and dilational breccia. Their recognition is based on the following features (Figure 4):
- Tectonic breccia is easily recognisable because of grain reduction and oriented fragments (Figure 4a).
Depending of its maturity (function of the strain intensity), fragments can be in contact
(grain-supported breccia, beginning of fragmentation and subsequent comminution) or finally
flooded in a largely developed matrix (matrix-supported breccia). Tectonic breccia is more
9
Minerals 2019, 9, 56
frequently monogenic. With respect to the intensity of the deformation and the presence or lack
of clay minerals, they can be called cataclasite, ultracataclasite or gouge.
- Hydrothermal breccia is characterised by more-or-less rounded fragments of the same nature (not
always) in place within a hydrothermal matrix. Frequently, the final voids are filled by cement
that can frequently contain some metals in economic contexts (Figure 4b).
- Magmatic breccia is more or less similar to the hydrothermal ones, except that the matrix is only
magmatic and there is no cement (Figure 4c). In this case, the fragments are rounded and never
in contact (matrix supported breccia). Due to the explosive processes, magmatic breccias are
frequently polygenic. The differentiation between matrix and fragments, both magmatic, is
sometimes difficult, especially in thin sections.
- Collapse breccia is easy to recognise because they show a large variation of fragment size, the
presence of cement, and grain-supported texture (Figure 4d). Their geometry is clearly consistent
with their mode of formation: (i) collapse of the fragments in response to an underlying explosion
or void formation by dissolution and (ii) posterior cementation.
- Crackle breccia is an early stage of what is going becoming a hydrothermal, tectonic or hydraulic
breccia. Due to their mode of formation, they are monogenic, with a low matrix and they can be
assimilated to early fragmentation in response to either tectonic stress or fluid-related fracturing.
Some parts frequently exhibit the host rocks being not totally disrupted whether other parts can
be more mature with well-expressed breccia texture (Figure 4e).
- Hydraulic breccia is the result of hydraulic fracturing. It exhibits typical jigsaw geometry with
a monogenic character and a very regular pattern (Figure 4f). The matrix is well represented,
and fragments are never in contact. The mode of formation is only due to cracking due to
fluid overpressure. Dilational breccia forms within extensional relay or pull-apart (Figures 1c,
2 and 5). In this case, breccia formation is explained by the fact that void creation causes the
fragmentation of the hosted rock affected by the pull-apart formation. Fragments are weakly
transported and sometimes rotated and the occurrence of cement is common. Why some
pull-aparts are filled by fibrous/comb crystals or dilational breccia remains an open question
(Figure 2c). The outcrop in Figure 5 can help because the two types of infilling have been observed
within the same structure. Since dilational breccia has been observed on the wall of the secondary
formed comb infilling (Figure 5b), we suspect that both types of texture can be developed in the
same structural context. Field relationships demonstrate that dilational breccia texture can form
at the beginning of the process, when rates of aperture are weak and late and rapid opening can
explain the superposition of fibrous/comb infilling. Indeed, the alternative formation of dilational
breccia or comb texture in the core of pull-apart can appear as a function of opening velocity,
crystal growth rate, and fluid saturation. We guess that dilational breccia in the core of pull-apart
can be created during all main tectonic contexts (i.e., compression, extension, transtension, etc.)
and not restricted to the only case of wrench tectonics, as this has been established for the
large-scale pull-apart-related basin formation along crustal-scale faults [40].
10
Minerals 2019, 9, 56
Figure 4. Breccia texture classification frequently used in ore geology. Each case is described in detail in
the text. A conceptual sketch is indicated for each photograph in order to correctly interpret the image.
The scale of these sketches is the same as the corresponding photograph. In red, some indications
about the process responsible for breccia formation are provided.
11
Minerals 2019, 9, 56
Figure 5. Superposition of dilational breccia texture and fibre/comb ones within similar left-lateral
pull-apart structures (see text for explanation of the cause of occurrence of fibrous or breccia texture).
3. Vein Formation Process and Tectonics: Examples from Ore Deposit Study
12
Minerals 2019, 9, 56
Figure 6. (a) Formation of sigmoid vein during late-orogenic collapse tectonics (red arrows) within
the Quadrilátero Ferrífero (Minas Gerais, Brazil). Note that vein re-used an early level formed by
mica alignment related to the nappe emplacement (b). The sigmoidal shape reflects thrust-related
emplacement (black arrows). Extensional pull-apart post-dated thrust-related structures and controlled
gold-bearing quartz-sulphide veins (c).
This study demonstrates that the Ouro Preto mesothermal gold deposit was formed in context of
late-orogenic collapse, drastically different from the conventional auriferous shear-zone model, model
that has been intensively used during the 1990s [43,44]. A recent study confirms this hypothesis by the
demonstration of resetting of older zircons by ca. 496 Ma old hot fluid rock interactions in the area of
Passagem [45].
This example clearly illustrates the importance of the systematic observation of the internal vein
texture before concluding on the mechanism of vein formation based only on the geometrical analysis.
3.2. Vein Opening and Filling Controlled by Regional-Scale Structures within Volcanic Domains
The Shila-Paula district is one of the numerous Au/Ag low sulfidation epithermal one of
Southern Peru. It is characterised by numerous veins hosted by the tertiary subaerial volcanics
of the Western Cordillera. Field studies shown that most of the mineralised bodies consist of the
systematic association of main E–W veins and secondary N120–135◦ E veins (Figure 7) [46]. Two
main stages of ore deposition are identified [47]. Stage 1 consists of a quartz–adularia–pyrite–galena
–sphalerite–chalcopyrite–electrum–Mn silicates and carbonates assemblage that fills the main E–W
veins (Figure 7a,b,e). Stage 2, also called the bonanza stage, carries most of the precious mineralisation
and consists of quartz, Fe-poor sphalerite, chalcopyrite, pyrite, adularia, galena, tennantite–tetrahedrite,
polybasite–pearceite, and electrum. This stage is mainly observed within secondary veins, in final
geodic filling (Figure 7c,f) and in veinlets that cut stage 1 assemblage (Figure 7d,g). In main veins,
the ore is systematically brecciated, whereas tectonic-free environment characterised the filling of
secondary veins. The age of veins was estimated to be around 10.8 Ma using 40Ar/39Ar ages on
adularia crystals from different veins [47].
A two-stage model is proposed to explain vein formation. The first stage was assumed to
correspond to the development of E–W sinistral shear zones and associated N120◦ E cleavages under
13
Minerals 2019, 9, 56
the effects of a NE–SW trending shortening direction, which has been previously recognised at the
Andean scale (Quechua II phase) (Figure 7a). These structures serve as a receptacle for the emplacement
of stage 1 ore assemblage that was brecciated during ongoing deformation (Figure 7b). The second
event operates a re-opening of the previously formed structures under a NW–SE trending shortening
direction that allowed the re-opening of pre-existing cleavage and the formation of scarce N50◦ E
trending S2 cleavages (Figure 7c), such as in the model in Figure 2. This stage was followed by
the bonanza ore emplacement both within geodes in core of the main E–W veins and in secondary
N120–135◦ E veins (Figure 7d). The two directions of shortening, NE–SW for the first event and NW–SE
for the second one, are also recorded by the orientation of fluid inclusion planes within quartz crystals
from the host rocks.
This study represents a unique example, constrained by combined tectonic, textural, mineralogical,
geochronological, and fluid inclusion data, of the establishment of a complete model of deposit
formation in which the re-using of previously formed tectonic features as a factor of gold concentration
in epithermal environment is evidenced.
Figure 7. Example of vein formation by the re-using of pre-existing structures within the Volcanic
domain of South Peru (see text for explanation). (a) Formation and filling of sinistral shear zone
and creation of associated cleavage under the effect of NE–SW shortening direction (red arrows). (b)
Formation of Mn-rich breccia under the same shortening direction. (c) Formation of secondary veins
by re-opening of the cleavage by N120◦ E trending shortening direction (green arrows). Formation
of geodic structures filled by stage 2-related paragenesis and subsequent “bonanza” stage and richer
veins (Veta 75) (d).
14
Minerals 2019, 9, 56
15
Minerals 2019, 9, 56
Figure 8. Schematic illustration showing the relationships between first stockwork and secondary one
around the VHMS of the Iberian Pyrite Belt (south Spain). (a) Schematic distribution of the different
stockwork and mineralised features close to a VHMS body. Occurrence of second stockwork within
meter-scale shear bands (b) and within axial planar cleavage (d). (c) Stratiform pyrite-rich level cut
by secondary pyrite-rich veins. (e) Small pull-apart filled by syntectonic quartz and pyrite. (f) Pyrite
metablasts and overgrowths (2 and 3) formed close to a synkinematic second stockwork.
16
Minerals 2019, 9, 56
Figure 9. (a) Illustrations of the Ifri Cu mine model showing some images of the Cu-rich stockwork
formed in response of ductile decollement within black schist. Each photograph is associated to a
schematic cartoon in order to explain the geological process. (b) Example of vein formation within
competent dolomite rich level in response of the ductilely deformed black schist. (c) Opening of
chalcopyrite/quartz-rich veins due to NW-directed shearing. (d) Stockwork formation by contrasted
behaviour of dolomite and black schist.
3.4.1. Sn-rich Breccia Formation of the Achmmach Prospect (Moroccan Central Massif)
The Achmmach tin mineralisation occurs within the NE part of the Massif Central domain
of Morocco, hosted by Ordovician, Silurian, and Devonian low-grade meta-sediments affected by
the Hercynian deformational events, weakly represented in this area. The region concerned by the
deposit exhibits a regular, N030–045◦ E trending cleavage mainly within the Silurian calc-schists.
The mineralisation is the result of a long-lived process that includes four events that occurred in
a transtensional tectonic regime associated with the late magmatic-hydrothermal evolution of the
Hercynian orogeny (see details in [52]).
- The first event is the formation of tourmaline-rich halos in core of the calc-schist. These halos
have ellipsoidal shape resembling tension gashes and are supposed to have formed during E–W
trending shortening. Since they follow the N070◦ E trend of the cleavage, most halos are “en
echelon” and indicators of a right-lateral potential shearing. Conjugate left-lateral “en echelon”
17
Minerals 2019, 9, 56
tourmaline halos also exist but are less common. The rock shown in Figure 10 was collected in
the core of one of the alteration halos and is entirely affected by the tourmalinisation.
- The second event is link to the development of right-lateral shearing only in levels affected by
the tourmalinisation (Figure 10a,c). It is noteworthy that this deformation is consistent with the
same tectonic context and therefore probably result from ongoing transtension controlled by E–W
shortening. Main shear bands are oriented N070◦ E.
- Third, we have evidence of transformation of the previously formed shear band in tourmaline-rich
breccia levels (Figure 10a,c). Such levels can reach thickness of 2 or 3 meters. The breccia is
matrix-supported with a well-developed tourmaline-rich matrix, and can exhibit some domains
with fragment-preferred orientation thus translating to tectonic- and hydrothermal-type breccia.
- The fourth texture is the most important because it is associated with cassiterite and thus
representative of the economic stage. Transtension is transformed in extension and normal
faults developed with the formation of a clast-supported breccia with numerous voids formation
and cassiterite crystallisation (Figure 10b). These mineralised structures are systematically formed
at the core of the first breccia levels and always in association with the tourmaline halos.
To conclude, mineralisation in the case of Achmmach prospect is clearly the result of polyphase
deformation during the late orogenic evolution of this Hercynian domain and certainly associated
with some granite emplacement. Granite remains hidden except for the occurrence of some rare
outcrops. Magmatic affinity is demonstrated by the ubiquitous presence of tourmaline at each stage of
the process. It is suggested in this case that ore concentration benefitted from the change of rheology
due to tourmaline invasion (tourmaline-rich halos). The process was achieved by ongoing successive
structures until the final formation of mineralised orebodies (shearing, brecciation and cassiterite
crystallisation).
18
Minerals 2019, 9, 56
Figure 10. Deposit history and evolution of the four stages that explain the formation of the Achmmach
tin deposit (Morocco) recognised in a unique block (a). (b) Image of the “North Zone” area in which the
succession of structure can be observed. Note the limit of the tourmaline alteration halo that delimitate
the zone where mineralisation developed. (c) Parallelism between ductile shear bands and mineralised
breccia showing that breccia re-used the earlier plane of deformation to develop.
19
Minerals 2019, 9, 56
and characterised by voluminous magmatic complexes emplaced between 325 and 300 Ma (from G1 to
G4 granitic events, [57]). The area is the site of abundant ore deposits (Au, W, Sn, REE). The example
shown herein concerns the Sn–W deposit of the Mina Soriana in which mineralised veins were formed
during a tectonic context dominated by NS extension linked to EW shortening [58]. The veins formed
with E–W strike normal to a NS stretching lineation. This phenomenon is clearly exposed within the
Mina Soriana outcrop. The Mina Soriana outcrop (Figure 11a) exposes a horizontal sill of leucogranite
that was injected into mica schist during the N–S lineation-related tectonic event. Tourmaline halos
(50 cm of thickness) are developed along the upper and lower host rocks (Figure 11b). Steeply dipping
veins mainly filled by quartz and tourmaline occur normal to the magmatic sill and only within the
tourmaline-rich halo (Figure 11c). This indicates that micaschist affected by the alteration had changed
its competence and reacted differently than the surrounding unaltered micaschist. Veins are limited
to the alteration halo although some of them cross cut the magmatic sill, as in boudinage-related
structures (Figure 11a). Since tourmaline grains are aligned N–S within the alteration halo but also at
the margin of quartz vein (Figure 11d), all features, i.e., magmatic sill emplacement, tourmaline-rich
halo, and quartz vein development, are coeval and controlled by the same tectonic event.
In this case, the different behaviour, as explained in Figure 1b, lies at the origin of vein formation
and, hence, of the formation of the Mina Soriana main vein, which outcrops further north, with the
same orientation as the small-scale quartz veins and is mined as the main orebody. The difference with
Figure 1b is that, in this case, the variation of rheology is not a pre-existing lithological feature, but was
created during the same process that produced the ore concentration.
Figure 11. (a) Mina Soriana outcrop, Galicia (Spain), showing the development of mineralised quartz
veins thanks to tourmaline-rich alteration halos developed in response to granitic sill emplacement
(modified from [59]). (b) Close view of the granitic sill, tourmaline halo and vertical quartz veins. (c)
Microscopic view of quartz vein rim showing tourmaline syntectonic growing. (d) Development of
vertical quartz mineralised veins limited to the tourmaline-rich alteration halo.
20
Minerals 2019, 9, 56
Figure 12. Relationships between gold-bearing quartz vein and granitic dike within the Bruès granite
cupola, Galicia. (a) Outcrop view showing the close relationship between quartz vein and granitic dike.
(b) Sample view showing the transitional contact between granite and quartz hydrothermal vein. (c)
Thin section of the central part of the hydrothermal vein representative of the mineralised stage. Note
the occurrence of dynamic recrystallised quartz and white mica indicating a normal sense of shearing.
Red arrows indicate the sense of motion.
21
Minerals 2019, 9, 56
22
Minerals 2019, 9, 56
In the light of these results, the importance of tectonic and microtectonic analysis at different
scales in modern metallogenic studies is underlined. This work should be realised at the regional scale
down to the scale of internal vein textures. The complementarity nature of pluridisciplinary works,
even though already adopted in many previous studies, has been again demonstrated by the examples
proposed and discussed in this paper. Change of scale and integration within the regional, geological
and tectonic context are two additional conditions for a comprehensive analysis. The benefits of
such an approach are both fundamental, leading to a better understanding of the mechanism of vein
formation process, and economics, leading to better knowledge of specific orebody geometry and
distribution and hence highly recommended in any type of exploration program.
Funding: This research was partly funded by Projects CAPES-COFECUB and CNRS GDR Transmet.
Acknowledgments: The mining companies SEIMSA (Iberian Pyrite Belt, Spain), CVRD (Vale group, Brazil),
CEDIMIN and BUENAVENTURA (Peru), MANAGEM (Morocco), SMS (Seksaoua, Morocco) and KASBAH
RESSOURCES (Achmmach, Morocco) are gratefully acknowledged for their constant help, support and fruitful
discussions. L. Badra, L. Bailly, L. Barbanson, Y. Branquet, P. Chaponnière, P. Couderc, M. Dardennes, A.
Gaouzi, J.M. Georgel, E. Gloaguen, M. Majhoubi, M. Menezes, J. Onezime, J. Rosas, and J. Tuduri, are tanked
for their contribution. Two anonymous reviewers and D. Aerden are kindly acknowledged for their fruitful and
constructive review.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.
References
1. Durney, D.W.; Ramsay, J.G. Incremental strains measured by syntectonic crystal growths. In Gravity and
Tectonics; De Jong, K.A., Scholten, K., Eds.; Wiley: New York, NY, USA, 1973; pp. 67–96.
2. Durney, D.W. Pressure solution and crystallization deformation. Philos. Trans. R. Soc. Lond. 1976, A283,
229–240. [CrossRef]
3. Beach, A. The geometry of en-echelon vein arrays. Tectonophysics 1975, 28, 245–263. [CrossRef]
4. Bons, P.D. The formation of large quartz veins by rapid ascent of fluids in mobile hydrofractures.
Tectonophysics 2001, 336, 1–17. [CrossRef]
5. Bons, P.D.; Elburg, M.A.; Gomez-Rivas, E. A review of the formation of tectonic veins and their
microstructures. J. Struct. Geol. 2012, 43, 33–62. [CrossRef]
6. Fisher, D.; Byrne, T. The character and distribution of mineralized fractures in the Kodiak Formation, Alaska:
Implications for fluid flow in an underthrust sequence. J. Geophys. Res. 1990, 95, 9069–9080. [CrossRef]
7. McCaffrey, K.; Lonergan, L.; Wilkinson, J. Fractures, Fluid Flow and Mineralization; Geological Society of
London Special Publication: London, UK, 1999; 155p.
8. Richards, J.P.; Tosdal, R.M. Structural controls on ore genesis. Rev. Econ. Geol. 2001, 14, 181.
9. Smith, J.V. En echelon sigmoidal vein arrays hosted by faults. J. Struct. Geol. 1996, 18, 1173–1179. [CrossRef]
10. Smith, J.V. Geometry and kinematics of convergent conjugate vein array systems. J. Struct. Geol. 1996, 18,
1291–1300. [CrossRef]
11. Cox, S.F. Deformational controls on the dynamics of fluid flow in mesothermal gold systems. In Fractures,
Fluid Flow and Mineralization; McCaffrey, K., Lonergan, L., Wilkinson, J., Eds.; Geological Society of London
Special Publication: London, UK, 1999; Volume 155, pp. 123–140.
12. Spencer, S. The use of syntectonic fibres to determine strain estimates and deformation paths: An appraisal.
Tectonophysics 1991, 194, 13–34. [CrossRef]
13. Hilgers, C.; Urai, J.L. Microstructural observations on natural syntectonic fibrous veins: Implications for the
growth process. Tectonophysics 2002, 352, 257–274. [CrossRef]
14. Barker, S.L.L.; Cox, S.F.; Eggins, S.M.; Gagan, M.K. Microchemical evidence for episodic growth of antitaxial
veins during fracture-controlled fluid flow. Earth Planet. Sci. Lett. 2006, 250, 331–344. [CrossRef]
15. Dowling, K.; Morrison, G. Application of quartz textures to the classification of gold deposits using North
Queensland examples. Econ. Geol. Monogr. 1990, 6, 342–255.
16. Dong, G.; Morrison, G.; Jaireth, S. Quartz textures in epithermal veins, Queensland—Classification, origin,
and implication. Econ. Geol. 1995, 90, 1841–1856. [CrossRef]
23
Minerals 2019, 9, 56
17. Taylor, R. Ores Textures, Recognition and Interpretation; Economic Geology Research Unit and Springer: Berlin,
Germany, 2009; 288p.
18. Funedda, A.; Naitza, S.; Buttau, C.; Cocco, F.; Dini, A. Structural Controls of Ore Mineralization in a
Polydeformed Basement: Field Examples from the Variscan Baccu Locci Shear Zone (SE Sardinia, Italy).
Minerals 2018, 8, 456. [CrossRef]
19. De Roo, J.A. Mass transfer and preferred orientation development during extensional microcracking in
slate-belt folds, Elura Mine, Australia. J. Metamorph. Geol. 1989, 7, 311–322. [CrossRef]
20. Davis, B.K.; Hippertt, J.F.M. Relationships between gold concentration and structure in quartz veins fr’om
the Hodgkinson Province, northeastern Australia. Miner. Depos. 1998, 33, 391–405. [CrossRef]
21. Forde, A. The late orogenic timing of mineralisation in some slate belt gold deposits, Victoria, Australia.
Miner. Depos. 1991, 26, 257–266. [CrossRef]
22. Cox, S.F.; Knackstedt, M.A.; Braun, J. Principles of structural control on permeability and fluid flow in
hydrothermal systems. In Structural Controls on Ore Genesis; Richards, J.P., Tosdal, R.M., Eds.; Reviews in
Economic Geology; Society of Economic Geologists: Littleton, CO, USA, 2001; Volume 14, pp. 1–24.
23. Gratier, J.P.; Vialon, P. Deformation pattern in a heterogeneous material: Folded and cleaved sedimentary
cover immediately overlying a crystalline basement (Oisans, French Alps). Tectonophysics 1980, 65, 151–180.
[CrossRef]
24. Chauvet, A.; Onézime, J.; Charvet, J.; Barbanson, L.; Faure, M. Syn- to late-tectonic stockwork emplacement
within the Spanish section of the Iberian Pyrite Belt: Structural, textural and mineralogical constraints in the
Tharsis—La Zarza areas. Econ. Geol. 2004, 99, 1781–1792. [CrossRef]
25. Cox, S.F.; Etheridge, M.A. Crack-seal fibre growth mechanism and their significance in the development of
oriented layer silicate microstructures. J. Struct. Geol. 1983, 92, 147–170. [CrossRef]
26. Fisher, D.M.; Brantley, S.L. Models of quartz overgrowth and vein formation: Deformation and episodic
fluid flow in an ancient subduction zone. J. Geoph. Res. 1992, 97, 20043–20061. [CrossRef]
27. Hilgers, C.; Köhn, D.; Bons, P.D.; Urai, J.L. Development of crystal morphology during unitaxial growth in a
progressively widening vein: II. Numerical simulations of the evolution of antitaxial fibrous veins. J. Struct.
Geol. 2001, 23, 873–885. [CrossRef]
28. Ramsay, J.G. The crack-seal mechanism of rock deformation. Nature 1980, 284, 135–139. [CrossRef]
29. Cox, S.F. Antitaxial crack-seal vein microstructures and their relationship to displacement paths. J. Struct.
Geol. 1987, 9, 779–787. [CrossRef]
30. Hilgers, C.; Urai, J.L. On the arrangement of solid inclusions in fibrous veins and the role of the crack-seal
mechanism. J. Struct. Geol. 2005, 27, 481–494. [CrossRef]
31. Boullier, A.M.; Robert, F. Paleoseismic events recorded in Archean gold quartz vein networks, Val-Dor,
Abitibi, Quebec, Canada. J. Struct. Geol. 1992, 14, 161–177. [CrossRef]
32. Bons, P.D.; Jessell, M.W. Experimental simulation of the formation of fibrous veins by localised
dissolution-precipitation creep. Mineral. Mag. 1997, 61, 53–63. [CrossRef]
33. Jébrak, M. Hydrothermal breccias in vein-type ore deposits: A review of mechanisms, morphology and size
distribution. Ore Geol. Rev. 1997, 12, 111–134. [CrossRef]
34. Sibson, R.H. Brecciation Processes in Fault Zones: Inferences from Earthquake Rupturing. Pure Appl. Geophys.
1986, 124, 159–175. [CrossRef]
35. Taylor, R.G.; Pollard, P.J. Mineralized Breccia Systems—Methods of Recognition and Interpretation, Economic
Geology Research Unit Contribution 46; James Cook University of North Queensland: Douglas, Australia, 1993;
36p.
36. Clark, C.; James, P. Hydrothermal brecciation due to fluid pressure fluctuations: Examples from the Olary
Domain, South Australia. Tectonophysics 2003, 366, 187–206. [CrossRef]
37. Tarasewicz, J.P.T.; Woodcok, N.H.; Disckson, J.A.D. Carbonate dilation breccias: Examples from the damage
zone to the Dent Fault, northwest England. Geol. Soc. Am. Bull. 2005, 117, 736–745. [CrossRef]
38. Davies, A.G.S.; Cooke, D.R.; Gemmell, J.B.; Van Leeuwen, T.; Cesare, P.; Hartshorn, G. Hydrothermal Breccias
and Veins at the Kelian Gold Mine, Kalimantan, Indonesia: Genesis of a Large Epithermal Gold Deposit.
Econ. Geol. 2008, 103, 717–757. [CrossRef]
39. Woodcok, N.H.; Mort, K. Classification of fault breccias and related fault rocks. Geol. Mag. 2008, 145, 435–440.
[CrossRef]
24
Minerals 2019, 9, 56
40. Burchfiel, B.C.; Stewart, J.H. “Pull-apart” origin of the central segment of Death Valley, California. Geol. Soc.
Am. Bull. 1966, 77, 439–442. [CrossRef]
41. Chauvet, A.; Piantone, P.; Barbanson, L.; Nehlig, P.; Pedroletti, I. Gold deposit formation during collapse
tectonics: Structural, mineralogical, geochronological, and fluid inclusion constraints in the Ouro Preto gold
mines, Quadrilátero Ferrífero, Brazil. Econ. Geol. 2001, 96, 25–48. [CrossRef]
42. Souza Martins, B.; Lobato, L.M.; Rosière, C.A.; Hagemann, S.G.; Schneider Santos, J.O.; dos Santos Peixoto
Villanova, F.L.; Figueiredo e Silva, R.C.; de Ávila Lemos, L.H. The Archean BIF-hosted Lamego gold deposit,
Rio das Velhas greenstone belt, Quadrilátero Ferrífero: Evidence for Cambrian structural modification of an
Archean orogenic gold deposit. Ore Geol. Rev. 2016, 72, 963–988. [CrossRef]
43. Kerrich, R. Geodynamic setting and hydraulic regimes: Shear zone hosted mesothermal gold deposits. In
Mineralisation and Shear Zones; Bursnall, J.T., Ed.; Geological Association of Canada Short Course Notes;
Mineralogical Association of Canada: Quebec City, QC, Canada, 1989; Volume 6, pp. 89–128.
44. Poulsen, K.H.; Robert, F. Shear Zones and Gold: Practical Examples from the Southern Canadian Shield.
In Mineralisation and Shear Zones; Bursnall, J.T., Ed.; Geological Association of Canada Short Course Notes;
Mineralogical Association of Canada: Quebec City, QC, Canada, 1989; Volume 6, pp. 239–266.
45. Cabral, A.R.; Zeh, A. Detrital zircon without detritus: A result of 496-Ma-old fluid–rock interaction during
the gold-lode formation of Passagem, Minas Gerais, Brazil. Lithos 2015, 212–215, 415–427. [CrossRef]
46. Cassard, D.; Chauvet, A.; Bailly, L.; Llosa, F.; Rosas, J.; Marcoux, E.; Lerouge, C. Structural control and K/Ar
dating of the Au-Ag epithermal veins in the Shila Cordillera, southern Peru. C. R. Acad. Sci. Paris 2000, 330,
23–30. [CrossRef]
47. Chauvet, A.; Bailly, L.; André, A.S.; Moni#xE9;, P.; Cassard, D.; Llosa Tajada, F.; Rosas Vargas, J.; Tuduri, J.
Internal vein texture and vein evolution of the epithermal Shila-Paula district, southern Peru. Miner. Depos.
2006, 41, 387–410. [CrossRef]
48. Onézime, J.; Charvet, J.; Faure, M.; Bourdier, J.L.; Chauvet, A. A new geodynamic interpretation for the
South Portuguese Zone (SW Iberia) and the Iberian Pyrite Belt genesis. Tectonics 2003, 22, 1027. [CrossRef]
49. Barbanson, L.; Chauvet, A.; Gaouzi, A.; Badra, L.; Mechiche, M.; Touray, J.C.; Oukarou, S. Les minéralisations
Cu–(Ni–Bi–U–Au–Ag) d’Ifri (district du Haut Seksaoua, Maroc): Apport de l’étude texturale au débat
syngenèse versus épigenèse. C. R. Géosci. 2003, 335, 1021–1029. [CrossRef]
50. Gaouzi, A.; Chauvet, A.; Barbanson, L.; Badra, L.; Touray, J.C.; Oukarou, S.; El Wartiti, M. Mise en place
syntectonique des minéralisations cuprifères du gîte d’Ifri (District du Haut Seksaoua, Haut-Atlas occidental,
Maroc). C. R. Acad. Sci. Paris 2001, 333, 277–284.
51. Chauvet, A.; Barbanson, L.; Gaouzi, A.; Badra, L.; Touray, J.C.; Oukarou, S. Example of a structurally
controlled copper deposit from ther Hercynian Western High-Atlas (Morocco): The High Seksaoua mining
district. In The Timing and Location of Major Ore Deposits in an Evolving Orogen; Blundell, D.J., Neuber, F.,
Von Quadt, A., Eds.; Geological Society of London Special Publication: London, UK, 2002; Volume 204,
pp. 247–271.
52. Mahjoubi, E.M.; Chauvet, A.; Badra, L.; Sizaret, S.; Barbanson, L.; El Maz, A.; Chen, Y.; Amman, M. Structural,
mineralogical, and paleoflow velocity constraints on Hercynian tin mineralization: The Achmmach prospect
of the Moroccan Central Massif. Miner. Depos. 2015, 51, 431–451. [CrossRef]
53. Chauvet, A.; Volland-Tuduri, N.; Lerouge, C.; Bouchot, V.; Monié, P.; Charonnat, X.; Faure, M.
Geochronological and geochemical characterization of magmatic-hydrothermal events within the southern
Variscan external domain. Intern. J. Earth Sci. 2012, 101, 69–86. [CrossRef]
54. Poulsen, K.H.; Robert, F.; Dubé, B. Geological Classification of Canadian Gold Deposits; Geological Survey of
Canada, Bulletin: Ottawa, ON, Canada, 2000; Volume 540, 113p.
55. Bouchot, V.; Milési, J.P.; Lescuyer, J.L.; Ledru, P. Les minéralisations aurifères de la France dans leur cadre
géologique autour de 300 Ma. Chron. Rech. Min. 1997, 528, 13–62.
56. Bouchot, V.; Ledru, P.; Lerouge, C.; Lescuyer, J.L.; Milési, J.P. Late Variscan mineralizing systems related to
orogenic processes: The French Massif. Ore Geol. Rev. 2005, 27, 169–197. [CrossRef]
57. Gloaguen, E. Apport D’une Étude Intégrée sur les Relations Entre Granites et Minéralisations Filoniennes
(Au et Sn-W) en Contexte Tardi Orogénique (Chaîne Hercynienne, Galice Centrale, Espagne). Ph.D. Thesis,
University of Orléans, Orléans, France, 2006.
25
Minerals 2019, 9, 56
58. Gloaguen, E.; Branquet, Y.; Chauvet, A.; Bouchot, V.; Barbanson, L.; Vigneresse, J.L. Tracing the
magmatic/hydrothermal transition in regional low-strain zones: The role of magma dynamics in strain
localization at pluton roof, implications for intrusion-related gold deposits. J. Struct. Geol. 2014, 58, 108–121.
[CrossRef]
59. Sizaret, S.; Branquet, Y.; Gloaguen, E.; Chauvet, A.; Barbanson, L.; Arbaret, L.; Chen, Y. Estimating the local
paleo-fluid flow velocity: New textural method and application to metasomatism. Earth Planet. Sci. Lett.
2009, 280, 71–82. [CrossRef]
60. Audétat, A.; Günther, D.; Heinrich, C.A. Formation of magmatic-hydrothermal ore deposits: Insights from
LA-ICP-MS analysis of fluid inclusions. Science 2008, 279, 2091–2094.
61. Pe-Piper, G.; Piper, D.J.W.; McFarlane, C.R.M.; Sangster, C.; Zhang, Y.; Boucher, B. Petrology, chronology and
sequence of vein systems: Systematic magmatic and hydrothermal history of a major intracontinental shear
zone, Canadian Appalachians. Lithos 2018, 304–307, 299–310. [CrossRef]
62. Groves, D.I.; Goldfarb, R.J.; Gebre-Mariam, M.; Hagemann, S.; Robert, F. Orogenic gold deposits: A proposed
classification in the context of their crustal distribution and relationship to other gold deposit types. Ore
Geol. Rev. 1998, 13, 7–27. [CrossRef]
63. Lang, J.R.; Baker, T. Intrusion-related gold systems: The present level of understanding. Miner. Depos. 2001,
36, 477–489. [CrossRef]
64. Everall, T.J.; Sanislav, I.V. The Influence of Pre-Existing Deformation and Alteration Textures on Rock
Strength, Failure Modes and Shear Strength Parameters. Geosciences 2018, 8, 124. [CrossRef]
© 2019 by the author. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
26
minerals
Article
Structural Control on the Formation of Pb-Zn
Deposits: An Example from the Pyrenean Axial Zone
Alexandre Cugerone 1, * , Emilien Oliot 1 , Alain Chauvet 1 , Jordi Gavaldà Bordes 2 ,
Angèle Laurent 1 , Elisabeth Le Goff 3 and Bénédicte Cenki-Tok 1
1 Géosciences Montpellier, UMR CNRS 5243, Université de Montpellier, Place E. Bataillon, CC 60,
34095 Montpellier, France; [email protected] (E.O.); [email protected] (A.C.);
[email protected] (A.L.); [email protected] (B.C.-T.)
2 Conselh Generau d’Aran, Vielha, 25530 Lleida, Spain; [email protected]
3 Bureau de Recherches Géologiques et Minières (BRGM), Territorial Direction Languedoc-Roussillon,
1039 Rue de Pinville, 34000 Montpellier, France; [email protected]
* Correspondence: [email protected]; Tel.: +33-643-983-585
Abstract: Pb-Zn deposits and specifically Sedimentary-Exhalative (SEDEX) deposits are frequently
found in deformed and/or metamorphosed geological terranes. Ore bodies structure is generally
difficult to observe and its relationships to the regional structural framework is often lacking. In the
Pyrenean Axial Zone (PAZ), the main Pb-Zn mineralizations are commonly considered as Ordovician
SEDEX deposits in the literature. New structural field analyzes focusing on the relations between
mineralization and regional structures allowed us to classify these Pb-Zn mineralizations into three
types: (I) Type 1 corresponds to minor disseminated mineralization, probably syngenetic and from
an exhalative source. (II) Type 2a is a stratabound mineralization, epigenetic and synchronous to
the Variscan D1 regional deformation event and (III) Type 2b is a vein mineralization, epigenetic
and synchronous to the late Variscan D2 regional deformation event. Structural control appears
to be a key parameter in concentrating Pb-Zn in the PAZ, as mineralizations occur associated to
fold hinges, cleavage, and/or faults. Here we show that the main exploited type 2a and type 2b
Pb-Zn mineralizations are intimately controlled by Variscan tectonics. This study demonstrates the
predominant role of structural study for unraveling the formation of Pb-Zn deposits especially in
deformed/metamorphosed terranes.
Keywords: Pb-Zn deposits; Pyrenean Axial Zone; SEDEX; remobilization; structural control; sphalerite
1. Introduction
The world’s most important Pb-Zn resources consist in Sedimentary-Exhalative (SEDEX)
mineralizations [1]. These types of ore deposits are syngenetic sedimentary to diagenetic. Occurrence
of laminated sulfides parallel to bedding associated to sedimentary features (graded beds, etc.) are the
key geological argument [2]. These important deposits occur often in ancient metamorphosed and
highly deformed terranes for example in Red Dog, Alaska [3,4]; Rampura, India [5]; or Broken Hill,
Australia [6]. In these cases, the processes of ore formation are still largely debated. In consequence,
unraveling the relationships between mineralization and orogenic remobilization(s) is essential in
order to understand the genesis of Pb-Zn deposits in deformed and metamorphosed environments.
For example, in Broken Hill [6–8] and Cannington [9] deposits in Australia some authors argued
for a metamorphogenic and epigenetic mineralization as large metasomatic zones may have refined
pre-existing Pb-Zn rich rocks. Other authors consider a pre-metamorphic and syngenetic origin with
only limited remobilization linked to tectonic events [10–12]. In the world-class Jinding Pb-Zn deposit,
the host rock has undergone a complex tectonic deformation [13]. Some authors proposed a syngenetic
origin of the deposit [14,15] whereas others argued for an epigenetic genesis of the deposit based on
field study, textural evidences [16–19], fluid inclusion [19,20], and paleomagnetic age [13]. Nowadays,
these high-tonnage Pb-Zn deposits are the preferential target of numerous academic and industrial
studies also for the presence of rare metals like Ge, Ga, In, or Cd associated with sulfides.
The Pb-Zn deposits hosted in the Pyrenean Axial Zone (PAZ) area that has suffered Variscan
tectonics [21–23] are usually considered to be SEDEX. As an example, due to their geometry and
the presence of distal volcanic rocks, Bois et al. [24] and Pouit et al. [25] considered as SEDEX
the Pb-Zn mineralizations located in the Pierrefitte anticlinorium. In Bentaillou area, Fert [26] and
Pouit [27,28] demonstrated that the stratigraphic and sedimentary controls were dominant processes
during the genesis of these mineralizations. In the Aran Valley, deposits (Liat, Victoria-Solitaria, and
Margalida) have been studied by Pujals [29] and Cardellach et al. [30,31]. These authors concluded
on a stratiform and possibly exhalative formation of Pb-Zn mineralizations associated with a poor
remobilization during Variscan deformation. Only few authors have documented the impact of
Variscan tectonics on the genesis of these mineralizations. These are Alonso [32] in Liat, Urets, and
Horcalh deposits or Nicol [33] for Pierrefitte anticlinorium deposits. In the Benasque Pass area, south
of the Bossòst anticlinorium, Garcia Sansegundo et al. [34] indicated probable Ordovician stratiform
or stratabound Pb-Zn mineralizations intensely reworked during Variscan tectonics. The Pb isotopes
study realized by Marcoux [35] showed a unique major event of Pb-Zn mineralization interpreted as
sedimentary-controlled and Ordovician or Devonian in age. Remobilization processes of Pb isotopes
seem however poorly constrained and a complete structural study related to these analyzes is lacking.
Pyrenean sulfide mineralizations are an excellent target for investigating the links between
orogenic deformation(s) and the genesis of associated mineralization(s), as well as finding key
arguments to make the distinction between strictly syngenetic or rather epigenetic mineralizations
and structurally remobilized mineralizations. In this work we will demonstrate that Pb-Zn deposits
from five districts in the PAZ, previously largely considered SEDEX, were actually formed through
processes involving a strong structural control.
2. Geological Setting
The Pyrenean Axial Zone (PAZ, Figure 1) is the result of the collision between the Iberian and Eurasian
plates since the Lower Cretaceous. Deep parts of the crust were exhumed during this orogeny. The PAZ
is composed of Paleozoic metasedimentary rocks locally intruded by Ordovician granites deformed and
metamorphosed during the Variscan orogeny, like the Aston or Canigou gneiss domes [23,36].
Figure 1. (a) Location of the Pyrenean Axial Zone (PAZ) within the Variscan belt of Western Europe.
(b) Schematic map of the Pyrenean Axial Zone (PAZ) and location of all recognized Pb-Zn deposits
(based on BRGM (French geological survey) and IGME (Spanish geological survey) databases). Note
the abundance of these deposits especially in the central and western domains of the PAZ.
28
Minerals 2018, 8, 489
The PAZ is generally divided in two domains [21,36–39]: (i) a deep-seated domain called
Infrastructure, which contains medium to high-grade metamorphic rocks and (ii) a shallow-seated
domain called the Superstructure, which is composed of low-grade metamorphic rocks. The Infrastructure
presents flat-lying foliations but highly deformed domains appear locally with steep and penetrative
crenulation foliations. Alternatively, the Superstructure presents moderate deformation associated to a
slaty cleavage [40,41] These two domains are intruded by Late-Carboniferous granites, like the Bossòst
and the Lys-Caillaouas granites [37,42,43].
In the PAZ several deformation phases essentially Variscan in age (325–290 Ma) are recognized.
The first deformation event (D1 ) is marked by a cleavage (S1 ) that is often parallel to the stratification (S0 ).
Regional M1 metamorphism is of Medium-Pressure and Low-Temperature (MP/LT) and synchronous of
this first D1 deformation [22]. The second deformation event (D2 ) is expressed by a moderate to steep axial
planar (S2 ) cleavage. M2 is a Low-Pressure and High-Temperature (LP/HT) metamorphism linked to the
Late-Variscan granitic intrusions, and it is superposed to the M1 metamorphism [44,45]. Late-Variscan
and/or Pyrenean-Alpine D3 deformations are locally expressed as fold and shear zones like the Merens
and/or probably the Bossòst faults [41,46,47].
The Pyrenean Pb-Zn regional district is the second largest in France with ~400,000 t Zn and
~180,000 t Pb extracted [48,49]. These sulfides deposits are localized in the PAZ in the Pierrefitte
and Bossòst anticlinoriums (Figure 1b). Sphalerite (ZnS) and galena (PbS) are essentially present in
Ordovician and Devonian metasediments. Few Pb-Zn deposits are hosted in granitic rocks [50].
This study focuses on Pb-Zn deposits located in the Bossòst anticlinorium (Figure 1) [42,44,51] and
includes a comparison with Pb-Zn deposits occurring in the Pierrefitte anticlinorium. The southern
part of the Bossòst anticlinorium forms the Aran Valley synclinorium. The northern part is limited by
the North Pyrenean fault (Figure 2a). It is mostly composed of Cambrian to Devonian rocks and an
intruding Late-Variscan leucocratic granite named the Bossòst granite.
Figure 2. The Bossòst Anticlinorium. (a) Geological map with positions of the three districts: (1)
Bentaillou-Liat-Urets district, see Figures 3 and 4; (2) Margalida-Victoria-Solitaria district, see Figure 5;
and (3) Pale Bidau-Argut-Pale de Rase district, see Figure 6. Pb-Zn deposits are numbered as follows:
1: Solitaria; 2: Victoria; 3: Margalida; 4: Plan del Tor; 5: Urets; 6: Horcall; 7: Mauricio-Reparadora;
8: Estrella; 9: Liat; 10: Malh de Bolard; 11: Bentaillou; 12: Crabere; 13: Uls; 14: Pale Bidau; 15: Pale de Rase;
16: Argut. Lithologies are based on geological map of BRGM (France [52–54]) and IGME (Spain, Aran
Valley; Garcia-Sansegundo et al. [55]). Metamorphic dome boundaries are related to andalousite isograd
presented by Zwart; (b) Structural map with foliation trajectories of S0 -S1 , subvertical S2 , and related F2
folds. Note preferential apparition of Pb-Zn deposits when S2 cleavage is well-expressed. (c) Schmidt
stereographic projections (lower hemisphere) of poles to S0 -S1 and S2 subvertical foliation planes.
29
Minerals 2018, 8, 489
Figure 3. (a) Structural map of the Bentaillou-Liat-Urets district based on field study and BRGM/IGME
geological maps. Location in Bossòst anticlinorium is indicated in the small sketch map (see also
location on Figure 2a); (b) Structural NNE-SSW cross-section of the Liat-Bentaillou area. Location of
the cross-section is indicated in the Figure 3a (modified from Garcia-Sansegundo and Alonso [56]).
Note presence of Pb-Zn mineralization at rock competence interface and close to F1 fold hinge in
Bentaillou mine.
Three main Pb-Zn districts are recognized in the Bossòst anticlinorium (Figure 2): (I) The Bentaillou-
Liat-Urets district is located in the eastern part of the anticlinorium and was the most productive
in the Bossòst anticlinorium, ~1.4 Mt at 9% of Zn and 2% of Pb metals [32,33]. (II) The Margalida-
Victoria-Solitaria district is located in the southern part of the anticlinorium, close to the Bossòst granite.
Production reached ~555,000 t with 11% Zn and 0.1% Pb [49]. (III) The Pale Bidau-Argut-Pale de Rase
district is located in the northern part of the anticlinorium. Pb-Zn production did not exceed ~7000 t of
Zn and ~3000 t of Pb [57].
30
Minerals 2018, 8, 489
Figure 4. Field observation and structural models in the Bentaillou-Liat-Urets deposits (see location in
Figure 3a). (a) Stratigraphic log of Bentaillou areas with position of the Pb-Zn deposits; (b) F1 fold in
Bentaillou marble with S1 cleavage marked by calcite recrystallization; (c) pull-apart geometry of Pb-Zn
mineralization in Bentaillou area; (d) oriented sample of typical mineralization in Bentaillou marble;
and (e) relationship between sphalerite mineralization and host rock structure. Note that sphalerite
is not folded by F1 folds and intersect S0 stratification; (f) 3D structural model of Bentaillou deposits
with Pb-Zn mineralization in cm to pluri-m pull apart geometry; (g) stratigraphic log of Liat-Urets area
with position of the Pb-Zn deposits; (h) stratabound mineralization in top of folded schist beds in Liat
area; (i) brecciated Pb-Zn mineralization in Liat area with clast of schist and quartz in sphalerite matrix;
(j) 3D structural model of Liat deposit with dm to m stratabound and vein mineralizations; (k) vein
Pb-Zn mineralization in Liat deposit; (l) stratabound mineralization in F2 fold hinge in Urets deposit;
(m) 3D structural model of Urets deposit with pluri-dm to m Pb-Zn mineralization in F2 fold hinge.
Mineral abbreviations: Qtz-quartz; Sp-sphalerite.
Pierrefitte anticlinorium is located north of the Cauteret granite and intersected by the
Eaux-Chaudes thrust (ECT; Figure 1). It is essentially composed of Ordovician rocks in the West
and Devonian terranes in the East. Two districts are studied: (I) Pierrefitte mines is the largest district
in the PAZ which produced ~180,000 t of Zn, ~100,000 t of Pb and ~150 t of Ag [48]. (II) Arre and
Anglas mines are located west to Pierrefitte mines. Pb-Zn production did not exceed ~6500 t of Zn [48].
31
Minerals 2018, 8, 489
32
Minerals 2018, 8, 489
Liat beds and the Silurian black-shale. The large hm-size open F2 fold is bordered to the south by a
Silurian synclinorium (Figure 3a,b). S1 cleavage is strictly parallel to S0 in the area. D2 deformation is
well expressed in the south at the contact between Silurian black-shale and Late Ordovician schists.
Mineralized stratabound bodies with pluri-dm to m-thickness appear parallel to the shallow
dipping S0 -S1 . Folds in Liat schists are present locally at the base of the mineralization (Figure 4h).
It presents a brecciated texture (Figure 4i) with clasts of quartz and schists. Sphalerite presents cm
grain sizes. At the contact with the Silurian black-shale the dip of Late Ordovician schist increases
and a normal fault is inferred. Vertical Pb-Zn vein mineralization parallel to S2 is present in this
fault. It intersects S0 stratification, S1 cleavage as well as stratabound mineralizations (Figure 4j,k).
Vein mineralization also presents a brecciated texture and sulfide grains are oriented parallel to S2 .
Sphalerite presents an infra-mm grain size.
Figure 5. Margalida-Victoria-Solitaria district (see location on Figure 2a). (a) Structural map (lithologies
are based on IGME geological map (Spain, Aran Valley; Garcia-Sansegundo et al. [1]) and location on the
Bossòst anticlinorium (pre-Silurian rocks); (b) stratigraphic log; (c) stratabound Pb-Zn mineralization
in Margalida mine hosted in Sandwich limestone level; (d) typical stratabound folded mineralization
(F2 isoclinal folds) in Victoria; (e) structural NNE-SSW cross-section of Victoria-Solitaria area; and
(f) structural model of Margalida and Victoria-Solitaria mines. Mineral abbreviations: Cal—Calcite;
Qtz—Quartz; Sp—Sphalerite.
33
Minerals 2018, 8, 489
Figure 6. Field observations and 3D structural model of Pale Bidau deposit (see location in Figure 2a).
(a) Vein Pb-Zn mineralization that occurs in pull-apart geometry; (b) 3D model presenting the relations
between stratabound and vein Pb-Zn mineralizations; and (c) vein mineralization with presence of
breccia at the base of a pull-apart mineralized structure.
34
Minerals 2018, 8, 489
In western parts km-scale Valentin NNW-SSE anticlinal is included in the Pierrefitte anticlinorium.
Compared to the Bossòst anticlinorium, the volume of late-Variscan granite or pegmatitic rocks
outcropping is smaller and there is no metamorphic dome in the core (Figure 7a).
Figure 7. The Pierrefitte anticlinorium. (a) Simplified structural map showing the location of (b,e);
(b) structural map zoomed on Pierrefitte mine (Lithologies are based on BRGM geological maps [59];
(c) photograph of typical stratabound Pb-Zn mineralization in Pierrefitte mine; (d) Schmidt stereographic
projections (lower hemisphere) of poles to S0 -S1 foliations measured in the Pierrefitte anticlinorium;
(e) structural map of the Pierrefitte-Valentin anticlinorium zoomed on Anglas-Uzious and Arre mines
(Lithologies are based on BRGM geological maps); (f) photograph of Arre vein mineralization parallel to S2
cleavage; (g) photograph of Anglas-Uzious vein mineralization; and (h) Schmidt stereographic projections
(lower hemisphere) of poles to S0 -S1 and S2 foliations measured in Anglas-Uzious and Arre areas.
35
Minerals 2018, 8, 489
The Pierrefitte anticlinorium is structured by several thrusts within Silurian levels (Figure 7b,c)
associated to D1 deformation. S2 vertical N090–100◦ E cleavage is well expressed in Devonian levels at
the rim of the anticlinorium but is less visible in the Ordovician core.
Numerous Pb-Zn mines are present in Late-Ordovician and Devonian terranes. These have
produced ~3 Mt (average 9% of Zn and 5% of Pb).
36
Minerals 2018, 8, 489
Figure 8. Paragenetic succession of ore and gangue minerals for all the eleven Pyrenean-studied Pb-Zn
deposits. Minerals in grey are common to both stratabound and vein mineralizations and minerals in
black are only present in stratabound or vein. Minerals only reported in a deposit are noted with the
deposit circle. Several minerals like apatite, ilmenite, or tourmaline are only present in stratabound
mineralization. Ge-minerals, graphite zinc carbonates or oxides, and Mg-Fe-Mn carbonates are only
observed in vein mineralization (n = 110).
37
Minerals 2018, 8, 489
38
Minerals 2018, 8, 489
Stratabound mineralization contain apatite, ilmenite, and tourmaline minerals that are only
observed in this mineralization. In the Pierrefitte mineralization, the abundance of chlorite and
muscovite associated to the mineralization is remarkable compared to other Pyrenean deposits.
Vein mineralization intersects S0 at the micron scale (Figure 9f). In the Anglas deposit, vein
mineralization is essentially composed of sphalerite, galena, quartz and calcite. The hanging wall
of the vein is parallel to S2 foliation and is marked by cordierite crystallization. Sphalerite in vein
mineralization appears highly deformed and recrystallized with mm relictual grains and recrystallized
μm-size crystals (like in Arre deposit, see Figure 9g). In Pale Bidau deposit, vein mineralization is
only present in domains where the S2 cleavage is well-marked. Note that Ge-minerals are exclusively
present in the vein mineralization (Figure 8).
6. Discussion
Figure 10. Schematic 3D sketch displaying the three main mineralization types which are typically
observed in the studied area and related to each studied deposit. Note cm to pluri-m pull apart
geometries in Bentaillou Type 2a mineralizations and in Type 2b mineralizations. Type 2b vein
mineralizations are located in intensely S2 deformed domains. Other structural traps like saddle-reef
formation in fold hinge or rock competence interfaces are represented for Victoria-Solitaria, Urets,
Pierrefitte, and Liat deposits respectively.
39
Minerals 2018, 8, 489
The second stratabound Type 2a mineralization (Figure 10) is deposited parallel to S0 -S1 .
It corresponds to the main Pb-Zn mineralization episode in the PAZ (~95% of the total exploited
ore volume). In the Bentaillou area, Type 2a mineralization intersects S0 stratification and is hosted by
S1 cleavage (Figure 4e), which is axial planar to isoclinal F1 folds. Fert [26] and Pouit [27,28] proposed
a syngenetic model for the Bentaillou deposit and described a normal stratigraphic succession that
has been later folded by F2 folds. F1 isoclinal recumbent N090◦ E folds are absent in their model.
Here we observe that Bentaillou Pb-Zn mineralization is localized essentially close to F1 fold hinges
at the interface between marble and schist or microconglomerate (Figure 4c). The source for Type 2a
sulfides may be related to layered and supposed syngenetic Type 1 sulfides that are disseminated
in the Ordovician and Devonian neighboring metasediments, or to Late-Variscan granitic intrusions,
probably at least temporally close to the Type 2a mineralizations. Opening of top to the north
cm to pluri-m pull-apart-type structures (Figure 4c) enables the formation of the large amount of
mineralizations in Bentaillou. Pb-Zn ore is not observed at the base of Bentaillou marbles due to
important karstification (Cigalere cave, Figure 3a), however it is deposed at Bularic [65] both above and
below this marble level. In the Liat area, Pujals [29] described a syngenetic or diagenetic mineralization
with apparent limited reworking. Our model shows that Type 2a stratabound mineralization is linked
to the Variscan D1 deformation. In the Victoria-Solitaria area, Type 2a stratabound mineralization
occurs where D2 -related structures are present and can be locally remobilized in fold hinges. These
thicker mineralizations in fold hinge may be linked to the saddle-reef process [66–68] associated with
formation of the dilatation zone during folding. These deposits have been studied by Pujals [29],
Cardellach et al. [30,69], Alvarez-Perez et al. [70], and Ovejero-Zappino [49,71]. These authors argued
for a SEDEX origin based on syngenetic mineralization associated to the presence of syn-sedimentary
faults. These models differ from our hypothesis: here we report that S1 cleavage is parallel to axial
plane of recumbent km-size isoclinal folds and transposes the S0 stratification. F2 folded Type 2a
stratabound mineralization is thicker in fold hinge and intersects metamorphic minerals as gahnite.
Presence of this Zn-spinel may be linked to a primary minor sulfide mineralization (Type 1, Figure 10)
or to a D1 metamorphic fluid rich in Zn. Chemistry of gahnite was analyzed by Pujals [29] and its
composition is typical to metamorphosed zinc deposits. This testifies that Type 2a Pb-Zn mineralization
is syn- to post-M1 metamorphism. Alonso [32] demonstrated a predominant role of mechanical
remobilization associated to deformation in the Bossòst anticlinorium and, especially, F2 folds in
Horcalh and fault in Liat. Our model is similar as we consider that the Variscan D2 deformation locally
remobilized Type 2a mineralization. The Margalida deposit records an additional deformational event
compared to Victoria and Solitaria. Hosted in a ductile deformed marble and close to the Bossòst
ductile fault, the Type 2a mineralization appears largely deformed with a typical durchbewegung
texture. No sedimentary structure is recognized in the marble [70]. This attests for a Late Hercynian
and/or Pyrenean deformation associated to the fault on the mineralization. Comparison with the
Pierrefitte anticlinorium shows the same syn-D1 Type 2a mineralization associated to regional thrust
tectonics. The main exploited Pb-Zn mineralization in Pierrefitte mine was pluri-m scale levels parallel
to S0 -S1 and the regional thrust (Figure 10). Our work comforts the study of Nicol [60] which has
shown an important remobilization of the mineralization in Ordovician and Devonian metasediments
linked to D1 deformation. On the contrary, Bois et al. [24] proposed a syngenetic deposition related
to the activity of Late-Ordovician syn-sedimentary faults and volcanism that may have induced
these mineralizations. In this case, remobilization is weak and sulfides crystalize prior to Variscan
metamorphism [24]. But the presence of sphalerite parallel to S1 cleavage and in pressure shadows
around magnetite clast concordant to S1 rather attests for a syn-D1 mineralization event.
The third Type 2b vein mineralization (Figure 10) is parallel to S2 cleavage. It intersects S0 -S1 cleavage
and former Type 2a stratabound mineralization. It has been recognized in the Pale Bidau-Argut-Pale de
Rase districts [57] and Arre-Uzious-Anglas districts. It appears in a limited number of deposits in the
PAZ. Type 2b mineralization is present in pluri-dm veins with restricted extension and highly differs
structurally and mineralogically. The presence of Ge-minerals and absence of apatite, tourmaline, or
40
Minerals 2018, 8, 489
ilmenite are remarkable here. Nonetheless, possible Type 2a remobilization with external contribution
is not excluded in the Type 2b vein formation. In the Uzious mine mineralization intersects magmatic
aplite. Therefore it has probably emplaced syn- or post-Cauteret granite and is certainly late-Variscan
in age (aplite from late-Variscan Cauteret granite) as supposed by Reyx [63]. Deformation of sphalerite,
which is supposed to be syn-D2 and/or syn-D3 , and the unusual sulfide paragenesis are inconsistent
with a Mesozoic age as described in Aulus-Les Argentieres undeformed sphalerite [72]. Other Pb-Zn
deposits, like the La Gela deposit [73] or Carboire deposit, could be attached to this third type as they are
characterized by vertical Pb-Zn veins and presence of Ge-minerals. These late-Variscan Pb-Zn deposits
have been recognized in Saint-Salvy (cf. M2 mineralization) even if the main Pb-Zn mineralization event
is Mesozoic [74].
6.2. Genetic Model of PAZ Pb-Zn Deposits Formation Linked to Regional Geology
The genetic model comprises four stages (Figure 11) based on the regional tectonic event model
of Mezger and Passchier [22] and Garcia-Sansegundo and Alonso [56].
Figure 11. Genetic model for the formation of the three main Pb-Zn mineralization types. Stage 1
is the disseminated Type 1 mineralization that is supposed to be syn-sedimentary. Stage 2a is the
syn-D1 Type 2a stratabound mineralization which is followed by the formation of F2 folds and local
remobilization of Pb-Zn mineralizations (saddle reef). Stage 2b represents the Type 2b late-Variscan
vein mineralizations.
41
Minerals 2018, 8, 489
Stage 1 represents the syn-sedimentary layered mineralization (SEDEX deposit, Pb-Zn Type
1 disseminated mineralization). Primary sulfides were recognized in all pre-Silurian stratigraphic
succession in the Bossòst area (Figure 11) and in Devonian rocks in Anglas-Uzious-Arre district. In the
Pierrefitte area primary sphalerite is absent, which is probably linked to important hydrothermal
low-grade alteration and D1 overprint.
Stage 2a starts during the D1 Variscan deformation and induces Type 2a stratabound mineralization.
This mineralization occurs preferentially where a rheological contrast exists between two lithologies
(e.g., marble-schist; schist-microconglomerates) and in highly D1 deformed area (Figure 11). Stage 2a
continues with D2 Variscan deformation and the formation of N090–110◦ E F2 upright folds. Granitic
intrusions occur at that stage (Figure 11). This D2 deformation locally reworked mineralization like in
Victoria mines where the mineralization is remobilized in fold hinges. Horcalh mineralized fault [32] is
interpreted as synchronous to D2 deformation.
Stage 2b occurs during the doming phase and the late-Variscan Type 2b vein mineralizations
(Figure 11). This mineralization type preferentially occurs parallel to the vertical S2 cleavage and is
mostly observed in the Pierrefitte and Bossòst anticlinoriums. Pull-apart-type structures are observed
in Pale Bidau and Uzious mines. A late deformation D3 corresponds to faults like the Bossòst mylonitic
fault close to Margalida district.
We have shown that the Pb-Zn deposits in the PAZ were polyphased and closely linked to Variscan
tectonics. There are at least three Pb-Zn mineralization-forming events, and two of them are evidently
structurally controlled. Type 1 may be syngenetic, but little ore is present. The main exploited ores are
Type 2a and Type 2b which have emplaced under a marked structurally control, either associated to S1
and trapped in F1 fold hinge, at lithology interface or in highly D1 or D2 deformed areas.
7. Conclusions
Three main types of Pb-Zn mineralizations have been distinguished in the Pyrenean Axial
Zone. A minor type (Type 1) is a stratiform disseminated mineralization that presents syngenetic
42
Minerals 2018, 8, 489
characteristics. The two other mineralization types, previously described as SEDEX, are in reality
post-sedimentation and formed as a result of Variscan polyphased tectonics: Type 2a is a syn-D1
stratabound mineralization that is parallel to S1 foliation. Type 2b is a syn to post-D2 vein-type
mineralization that is parallel to subvertical S2 cleavage. Structural control is thus a key parameter
for the remobilization of Pb-Zn mineralizations in this area like in (D1 and D2 ) fold hinges (saddle
reef), high (D1 ) deformed zones, rock contrast interfaces, and S2 cleavages. A multiscale detailed
structural study is essential for unraveling the formation of Pb-Zn deposits, especially in deformed
and/or metamorphosed terranes.
Author Contributions: A.C. (Alexandre Cugerone) and B.C.-T. conceived the research within the framework of
the A.C. (Alexandre Cugerone)’s PhD project; A.C. (Alexandre Cugerone), E.O., A.C. (Alain Chauvet), B.C.-T.,
J.G.B., and A.L. participated in field work; A.C. (Alexandre Cugerone) acquired the samples and performed all
the analytical work under the guidance of A.C. (Alain Chauvet) and B.C.-T.; A.C. (Alexandre Cugerone) wrote the
paper with contributions from E.O., B.C.-T., E.L.G., and J.G.B.
Funding: This research was funded by the French Geological Survey (Bureau de Recherches Géologiques et
Minières; BRGM) through the national program “Référentiel Géologique de France” (RGF-Pyrénées).
Acknowledgments: The authors gratefully acknowledge Kalin Kouzmanov and Stefano Salvi for their involvement in
the project. We thank the ARSHAL association for Bentaillou mine access and Jean-Marc Poudevigne, Louis de Pazzis,
and Bernard Lafage for their precious knowledge of the Pyrenean Pb-Zn mines. We acknowledge Christophe Nevado
and Doriane Delmas for thin section preparation. The authors are thankful for the editorial handling of Jax Jiang and for
the constructive comments of the three anonymous reviewers.
Conflicts of Interest: The authors declare no conflicts of interest.
References
1. Wilkinson, J.J. Sediment-Hosted Zinc-Lead Mineralization: Processes and perspectives. Treatise Geochem.
2013, 13, 219–249.
2. Leach, D.L. Sediment-hosted lead-zinc deposits: A global perspective. Econ. Geol. 2005, 100, 561–607.
3. Moore, D.W.; Young, L.E.; Modene, J.S.; Plahuta, J.T. Geologic setting and genesis of the Red Dog zinc-lead-silver
deposit, western Brooks Range, Alaska. Econ. Geol. 1986, 81, 1696–1727. [CrossRef]
4. Kelley, K.D.; Jennings, S. A special issue devoted to barite and Zn-Pb-Ag deposits in the Red Dog district,
Western Brooks Range, northern Alaska. Econ. Geol. 2004, 99, 1267–1280. [CrossRef]
5. Hazarika, P.; Upadhyay, D.; Mishra, B. Contrasting geochronological evolution of the Rajpura-Dariba and
Rampura-Agucha metamorphosed Zn-Pb deposit, Aravalli-Delhi Belt, India. J. Asian Earth Sci. 2013, 73,
429–439. [CrossRef]
6. Lawrence, L.J. Polymetamorphism of the sulphide ores of Broken Hill, NSW, Australia. Miner. Depos. 1973,
8, 211–236. [CrossRef]
7. Gibson, G.M.; Nutman, A.P. Detachment faulting and bimodal magmatism in the Palaeoproterozoic
Willyama Supergroup, south-central Australia; keys to recognition of a multiply deformed Precambrian
metamorphic core complex. J. Geol. Soc. 2004, 161, 55–66. [CrossRef]
8. Hobbs, B.E.; Walshe, J.L.; Ord, A.; Zhang, Y.; Carr, G.C. The Broken Hill orebody: A high temperature, high
pressure scenario. AGSO Rec. 1998, 2, 98–103.
9. Walters, S.; Bailey, A. Geology and mineralization of the Cannington Ag-Pb-Zn deposit: An example of
Broken Hill-Type mineralization in the eastern succession, Mount Isa Inlier, Australia. Econ. Geol. 1998, 93,
1307–1329. [CrossRef]
10. Webster, A.E. The Structural Evolution of the Broken Hill Pb-Zn-Ag Deposit, New South Wales, Australia.
Ph.D. Thesis, University of Tasmania, Hobart, Australia, 2004.
11. Bodon, S.B. Paragenetic relationships and their implications for ore genesis at the Cannington Ag-Pb-Zn
deposit, Mount Isa inlier, Queensland, Australia. Econ. Geol. 1998, 93, 1463–1488. [CrossRef]
12. Haydon, R.C.; Mcconachy, G.W. The stratigraphic setting of Pb-Zn-Ag mineralization at Broken Hill.
Econ. Geol. 1987, 82, 826–856. [CrossRef]
13. Yalikun, Y.; Xue, C.; Symons, D.T.A. Paleomagnetic age and tectonic constraints on the genesis of the giant
Jinding Zn-Pb deposit, Yunnan, China. Miner. Depos. 2018, 53, 245–259. [CrossRef]
43
Minerals 2018, 8, 489
14. Shi, J.X.; Yi, F.H.; Wen, Q.D. The rock-ore characteristics and mineralisation of Jinding lead-zinc deposit,
Lanping. J. Yunnan Geol. 1983, 2, 179–195.
15. Wang, J.B.; Li, C.Y.; Chen, X. A new genetic model for the Jinding lead-zinc deposit. Geol. Explor. Non
Ferr. Met. 1992, 1, 200–206. (In Chinese)
16. Leach, D.L.; Song, Y.C.; Hou, Z.Q. The world-class Jinding Zn–Pb deposit: Ore formation in an evaporite
dome, Lanping Basin, Yunnan, China. Miner. Depos. 2017, 52, 281–296. [CrossRef]
17. Kyle, J.R.; Li, N. Jinding: A giant tertiary sandstone-hosted Zn–Pb deposit, Yunnan, China. SEG Newsl. 2002,
50, 1–9.
18. Chi, G.; Xue, C.; Qing, H.; Xue, W.; Zhang, J.; Sun, Y. Hydrodynamic analysis of clastic injection and hydraulic
fracturing structures in the Jinding Zn-Pb deposit, Yunnan, China. Geosci. Front. 2012, 3, 73–84. [CrossRef]
19. Xue, C.; Zeng, R.; Liu, S.; Chi, G.; Qing, H.; Chen, Y.; Yang, J.; Wang, D. Geologic, fluid inclusion and isotopic
characteristics of the Jinding Zn-Pb deposit, western Yunnan, South China: A review. Ore Geol. Rev. 2007, 31,
337–359. [CrossRef]
20. Chi, G.; Qing, H.; Xue, C. An overpressured fluid system associated with the giant sandstone-hosted Jinding
Zn-Pb deposit, western Yunnan, China Chapter. In Mineral Deposit Research: Meeting the Global Challenge;
Mao, J., Bierlein, F.P., Eds.; Springer: Berlin, German, 2005; pp. 93–96.
21. Zwart, H.J. The Geology of the Central Pyrenees. Leidse Geol. Meded. 1979, 50, 1–74.
22. Mezger, J.E.; Passchier, C.W. Polymetamorphism and ductile deformation of staurolite-cordierite schist of
the Bossòst dome: Indication for Variscan extension in the Axial Zone of the central Pyrenees. Geol. Mag.
2003, 140, 595–612. [CrossRef]
23. Denèle, Y.; Laumonier, B.; Paquette, J.-L.; Olivier, P.; Gleizes, G.; Barbey, P. Timing of granite emplacement,
crustal flow and gneiss dome formation in the Variscan segment of the Pyrenees. Geol. Soc. Lond. Spec. Publ.
2014, 405, 265–287. [CrossRef]
24. Bois, J.P.; Pouit, G. Les minéralisations de Zn (Pb) de l’anticlinorium de Pierrefitte: Un exemple de gisements
hydrothermaux et sédimentaires associés au volcanisme dans le Paléozoïque des Pyrénées centrales. Bureau Rech.
Geol. Min. 1976, 6, 543–567. (In French)
25. Pouit, G.; Fortuné, J.-P. Métallogénie comparée des Pyrénées et du Sud du Massif-central. In Proceedings of
the 26ème Congrès Géologique International, Paris, France, 7–17 July 1980; p. 61. (In French)
26. Fert, D. Un Aspect de la Métallogénie du Zinc et du Plomb Dans l’Ordovicien des Pyrénées Centrales: Le
District de Sentein (Ariège, Haute-Garonne). Ph.D. Thesis, University Pierre Marie Curie, Paris, France, 1976.
27. Pouit, G. Différents Modèles de Mineralisations «Hydrothermale Sédimentaire», à Zn (Pb) du Paléozoïque
des Pyrénees Centrales. Miner. Depos. 1978, 13, 411–421. (In French) [CrossRef]
28. Pouit, G. Les minéralisations Zn-Pb exhalatives sédimentaires de Bentaillou et de l’anticlinorium paléozoïque
de Bosost (Pyrénées ariégeoises, France). Chron. Rech. Min. 1986, 485, 3–16. (In French)
29. Pujals, I. Las Mineralizaciones de Sulfuros en el Cambro-Ordovicico de la Val d’Aran (Pirineo Central,
Lérida). Ph.D. Thesis, University Autónoma Barcelona, Barcelona, Spain, 1992.
30. Cardellach, E.; Phillips, R.; Ayora, C. Metamorphosed stratiform sulphides of the Liat area, Central Pyrenees,
Spain. Inst. Min. Metall. Trans. 1982, 91, 90–95. (In Spanish)
31. Cardellach, E. Estudio microscópico de las mineralizaciones de Pb-Zn de Liat, Baguergue y Montoliu.
Acta Geol. Hisp. 1977, 12, 120–122. (In Spanish)
32. Alonso, J.L. Deformaciones Sucesivas en el Area Comprendida Entre Liat y el Puerto de Orla—Control
Estructural de los Depositos de Sulfuros (Valle de Aran, Pirineos Centrales). Master’s Thesis, University
Oviedo, Oviedo, Spain, 1979. (In Spanish)
33. Nicol, N. Etude Structurale des Minéralisations Zn-Pb du Paléozoïque du Dôme de Pierrefitte (Hautes-Pyrénées).
Goniométrie de Texture Appliquée aux Minéraux Transparents et Opaques. Ph.D. Thesis, University Orléans,
Orléans, France, 1997. (In French)
34. García-Sansegundo, J.; Martin-Izard, A.; Gavaldà, J. Structural control and geological significance of the
Zn-Pb ores formed in the Benasque Pass area (Central Pyrenees) during the post-late Ordovician extensional
event of the Gondwana margin. Ore Geol. Rev. 2014, 56, 516–527. [CrossRef]
35. Marcoux, E. Isotope du plomb et paragénèses métalliques, traceurs de l’histoire des gites mineraux. Bur. Rech.
Geol. Min. 1986, 117, 1–289. (In French)
36. Zwart, H.J. Metamorphic history of the Central Pyrenees, Part II, Valle de Aran. Leidse Geol. Meded. 1963, 28,
321–376.
44
Minerals 2018, 8, 489
37. Kleinsmiede, W.F.J. Geology of the Valle de Aran (Central Pyrenees). Leidse Geol. Meded. 1960, 25, 129–245.
38. Cochelin, B.; Lemirre, B.; Denèle, Y.; De Saint Blanquat, M.; Lahfid, A.; Duchêne, S. Structural inheritance
in the Central Pyrenees: The Variscan to Alpine tectonometamorphic evolution of the Axial Zone. J. Geol.
Soc. Lond. 2017, 175, 16. [CrossRef]
39. De Sitter, L.U.; Zwart, H.J. Tectonic development in supra and infra-structures of a mountain chain.
In Structure of the Earth’s Crust and Deformation of Rocks; Det Berlingske Bogtrykkeri: Copenhagen, Denmark,
1960; Volume 18, pp. 248–256.
40. Carreras, J.; Capellà, I. Tectonic levels in the Palaeozoic basement of the Pyrenees: A review and a new
interpretation. J. Struct. Geol. 1994, 16, 1509–1524. [CrossRef]
41. Carreras, J.; Druguet, E. Framing the tectonic regime of the NE Iberian Variscan segment. Geol. Soc. Lond.
Spec. Publ. 2014, 405, 249–264. [CrossRef]
42. Cochelin, B. Champ de déformation du socle Paléozoïque des Pyrénées. Ph.D. Thesis, Université Toulouse 3
Paul Sabatier, Toulouse, France, 2016. (In French)
43. Mezger, J.E.; Gerdes, A. Early Variscan (Visean) granites in the core of central Pyrenean gneiss domes:
Implications from laser ablation U-Pb and Th-Pb studies. Gondwana Res. 2016, 29, 181–198. [CrossRef]
44. Pouget, P. Hercynian tectonometamorphic evolution of the Bosost dome (French Spanish Central Pyrenees).
J. Geol. Soc. Lond. 1991, 148, 299–314. [CrossRef]
45. Mezger, J.E. Comparison of the western Aston-Hospitalet and the Bossòst domes: Evidence for
polymetamorphism and its implications for the Variscan tectonic evolution of the Axial Zone of the Pyrenees.
J. Virtual Explor. 2005, 19, 1–19. [CrossRef]
46. Mezger, J.E.; Schnapperelle, S.; Rölke, C. Evolution of the Central Pyrenean Mérens fault controlled by near
collision of two gneiss domes. Hallesches Jahrb. 2012, 34, 11–29.
47. Carreras, J. Zooming on Northern Cap de Creus shear zones. J. Struct. Geol. 2001, 23, 1457–1486. [CrossRef]
48. BRGM International Report: Les Gisements de Pb-Zn Français (Situation en 1977); BRGM: Orléans, France, 1984;
pp. 1–278. (In French)
49. Ovejero Zappino, G. Mineralizaciones Zn-Pb ordovícicas del anticlinorio de Bossost. Yacimientos de Liat y
Victoria. Valle de Arán. Pirineo (España). Bol. Geol. Min. 1991, 102, 356–377. (In Spanish)
50. Castroviejo Bolibar, R.; Serrano, F.M. Estructura y metalogenia del campo filoniano de Cierco (Pb-Zn-Ag), en
el Pirineo de Lérida. Bol. Geol. Min. 1983, 1983, 291–320. (In Spanish)
51. Aerden, D.G.A. Kinematics of orogenic collapse in the Variscan Pyrenees deduced from microstructures in
porphyroblastic rocks from the Lys-Caillaouas massif. Tectonophysics 1994, 238, 139–160. [CrossRef]
52. Barrère, P.; Bouquet, C.; Debroas, E.J.; Pelissonnier, H.; Peybernes, B.; Soulé, J.C.; Souquet, P.; Ternet, Y.
Arreau. In BRGM Geological Map 1/50,000 with Note; BRGM: Orléans, France, 1984; p. 60. (In French)
53. Clin, M.; Taillefer, F.; Pouchan, P.; Muller, A. Bagnères de Luchon. In BRGM Geological Map 1/50,000 with
Note; BRGM: Orléans, France, 1989; p. 78. (In French)
54. Lavigne, J. Pic de Mauberme. In BRGM Geological Map 1/50,000 with Note; BRGM: Orléans, France, 1972;
p. 24. (In French)
55. García-Sansegundo, J.; Merino, J.R.; Santisteban, R.R.; Leyva, F. Canejan-Vielha Mapa geologico 1:50,000.
Inst. Geol. Min. Espana 2013, 1, 66. (In French)
56. Garcia-Sansegundo, J.; Alonso, J.L. Stratigraphy and structure of the southeastern Garona Dome. Geodin. Acta
1989, 3, 127–134. [CrossRef]
57. Cugerone, A.; Cenki-Tok, B.; Chauvet, A.; Le Goff, E.; Bailly, L.; Alard, O.; Allard, M. Relationships
between the occurrence of accessory Ge-minerals and sphalerite in Variscan Pb-Zn deposits of the Bossost
anticlinorium, French Pyrenean Axial Zone: Chemistry, microstructures and ore-deposit setting. Ore Geol. Rev.
2018, 95, 1–19. [CrossRef]
58. Dubois, C. Mangeuses d’homme: L’épopée des Mines de Bentaillou et de Bulard en Ariège; Privat Edition: Toulouse,
France, 2015; p. 320. (In French)
59. Barrère, P.; Bois, J.-P.; Soulé, J.-C.; Ternet, Y. Argelès-Gazost. In BRGM Geological Map 1/50,000 with Note;
BRGM: Orléans, France, 1980; pp. 1–48. (In French)
60. Nicol, N.; Legendre, O.; Charvet, J. Les minéralisations Zn-Pb de la série paléozoïque de Pierrefitte
(Hautes-Pyrénées) dans la succession des évènements tectoniques hercyniens. C. R. Acad. Sci. 1997,
324, 453–460. (In French)
45
Minerals 2018, 8, 489
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
46
minerals
Article
Structural Controls of Ore Mineralization in a
Polydeformed Basement: Field Examples from the
Variscan Baccu Locci Shear Zone (SE Sardinia, Italy)
Antonio Funedda 1, * , Stefano Naitza 1 , Cristina Buttau 1 , Fabrizio Cocco 1 and
Andrea Dini 2
1 Dipartimento di Scienze Chimiche e Geologiche, Università degli Studi di Cagliari,
Cittadella Universitaria (Blocco A), S.S. 554 bivio per Sestu, 09042 Monserrato (CA), Italy;
[email protected] (S.N.); [email protected] (C.B.); [email protected] (F.C.)
2 Consiglio Nazionale delle Ricerche (CNR)-Istituto di Geoscienze e Georisorse, 56100 Pisa, Italy;
[email protected]
* Correspondence: [email protected]
Abstract: The Baccu Locci mine area is located in a sector of the Variscan Nappe zone of Sardinia
(the Baccu Locci shear zone) that hosts several type of ore deposits mined until the first half of the
last century. The orebodies consist of lenses of Zn–Cu sulphides, once interpreted as stratabound,
and Qtz–As–Pb sulphide ± gold veins; the implication of structural controls in their origin were
previously misinterpreted or not considered. Detailed field mapping, structural analyses, and ore
mineralogy allowed for unraveling how different ore parageneses are superimposed each other and
to recognize different relationships with the Variscan structures. The sulphide lenses are parallel
to the mylonitic foliation, hosted in the hinges of minor order upright antiforms that acted as traps
for hydrothermal fluids. The Qtz–As–Pb sulphide veins crosscut the sulphide lenses and are hosted
in large dilatational jogs developed in the hanging wall of dextral-reverse faults, whose geometry
is influenced by the attitude of reverse limbs of late Variscan folds. The ores in the Baccu Locci
shear zone are best interpreted as Variscan orogenic gold-type; veins display mutual crosscutting
relationships with mafic dikes dated in the same district at 302 ± 0.2 Ma, a reliable age for the
mineralizing events in the area.
Keywords: sulphide lenses; hinge trap; dilational jogs; orogenic gold; mafic dikes; mineralization
chronology; arsenopyrite; late Variscan strike-slip faults
1. Introduction
Structural controls of ore deposits hosted in poly-deformed low-grade metamorphic basements
are often not easy to recognize, even more so when there are multiple generations of mineralization,
mostly because the relationships between structures and ore bodies are often not clear. The correct
understanding of the structural controls is, in fact, relevant for defining the characteristics of the
deposit, its emplacement style, its age, and therefore, its origin. In some cases, the difficulties in
unravelling the tectonic structures prevent the understanding of the ore bodies’ geometry, leading
to mistakes in mineral exploration, evaluation of ore deposits, mine planning, and even mineral
exploitation. Recognizing whether the structural controls on ore deposits are “passive”, and therefore
attributable to tectonic structures developed before mineralization, or, conversely, “active,” i.e., related
to tectonic structures progressively evolving with mineralization, can provide valuable indications
in this sense. Indeed, passive structural controls imply that there are no direct correlations between
tectonic and mineralizing events; the physical parameters that characterized the deformational
events, e.g., thermo-baric conditions, presence of fluids, state of stress, etc., were independent from
mineralization processes and they had no influence on ore formation. Conversely, in the case of active
controls, these features were critical for mineralization and shaped both the deformational context and
the genesis of the ores.
In the nappe zone of the Variscan basement of SE Sardinia, several Sb–W, As-Au, and As–Pb–(Cu,
Zn, Ag, Au) ore deposits are located along a lower green-schist facies mylonitic belt, folded by a
plurikilometric antiform (the Flumendosa Antiform) during the Variscan collisional phases and then
affected by extensional tectonics. During the postcollisional extension, early phases were characterized
by ductile-type structures that, with progressive exhumation of deeper tectonic units, further evolved
in brittle–ductile to brittle regimes [1], supporting large-scale hydrothermal fluid circulation in the crust.
In this frame, mineralization processes and ore deposition were structurally controlled, both passively
and actively, at all scales.
In this work, we present a case study of the mineralized systems occurring in the Baccu Locci
mine area, located in the core of the Flumendosa Antiform. In this area, several stacked tectonic
units are separated by thick mylonitic zones and folded together [2]. Exhumation of the antiformal
core is evidenced by the superposition of brittle–ductile and brittle structures over ductile ones;
accordingly, different kinds of structurally controlled mineral deposits exposed in the area allow a
distinction between passive and active emplacement styles, also leading to a relative chronology of
tectonic and mineralizing events. Recent studies in the Baccu Locci mine area [1,3–5] highlighted
that: (1) ore formation was structurally controlled and related to the tectonic evolution, during late
Variscan extensional phases, of the Baccu Locci regional shear zone, and (2) the relationships between
different ores materialize a superposition of tectonic events, further complicated by mafic magmatism
and diking. Apart from this general picture, however, several unsolved issues still persist: (1) type
and timing of different mineralization, (2) their relationships with tectonic structures and structural
controls, (3) ore mineral associations and mafic dikes, and (4) relationships with regional and local
stress fields.
In this paper, we intend to contribute to the first two issues and provide some new data for the
third one.
48
Minerals 2018, 8, 456
Figure 1. (a) Geological sketch map of the Variscan basement of Sardinia. (b) Tectonic sketch of the SE
Sardinia Variscan basement. (c) Geological cross-section of the Gerrei-Sarrabus region in SE Sardinia;
modified after [7], reprinted with permission.
Figure 2. Tectonic sketch map and mineral deposits of the Gerrei district.
49
Minerals 2018, 8, 456
Figure 3. Structural schematic map of the Baccu Locci shear zone and geological cross-section, after [2].
In the geological cross-section, red circles point out D2 folds that are schematically represented in
Figure 15.
50
Minerals 2018, 8, 456
51
Minerals 2018, 8, 456
Locci shear zone is characterized by widespread and penetrative mylonitic foliation with a mineral
assemblage typical of lower green-schist facies metamorphism that is parallel to the large thrust and
generally cuts at a low angle the early D1 axial plane foliation [3]. The deformation is highly partitioned;
in the core of the shear zone, it is not possible to recognize the mylonite protolith, although several slices
of less deformed rocks have been recognized and mapped [2,3,19]. (Note that, following the choice
by [2,3], in Figures 3 and 5 we distinguished the mylonite whose protolith is not recognizable from the
mylonite whose protolith is still recognizable). At a later stage during the collisional phase, a late-D1
(LD1) N-S shortening event led to development at a regional scale of large, weakly east-plunging
upright folds, with axes up to 50 km long that refolded both isoclinal folds and ductile shear zones.
Of these late folds, the main fold is the Flumendosa Antiform [1] (Figures 1 and 2), which runs roughly
from WNW to ESE for more than 50 km, folding the D1 nappe stack. In detail, the Flumendosa
Antiform consists of some minor order antiforms and synforms with km-size wavelength. One such
northern minor order folds crops out in the study area. The LD1 axis is generally east-plunging,
and at the hinge zone, a subvertical spaced crenulation cleavage discontinuously developed. Then,
the LD1 folds were in turn deformed during the D2 postcollisional extensional phase. The limbs of
the LD1 antiforms are deformed by several asymmetric recumbent folds with subhorizontal axial
planes and axes parallel to the LD1 limb attitude [1] (Figure 4). The D2 folds are overturned away
from the antiformal hinge zone [1]. They are often associated with low-angle normal shear zones that
allowed for the exhumation of deeper units and enhanced the antiformal structure. They have opposite
structural-facing direction in the fold flanks: north-facing in the northern limb and south-facing in
the southern limb. Their major order wavelength exposed in the field is about 30 to 40 m. Folds with
“outer” structural facing and low-angle normal faults are interpreted as produced by vertical shortening
of steeply inclined bedding and earlier foliations. During exhumation, rocks were progressively carried
to shallower structural levels, where brittle behavior became prevalent. Thus, the deformation style
changed, and the final stage of postcollisional extension was accommodated by high-angle normal
faults [6]. A D3 folding event, with vertical axial planes and axis trend changing from N-S to N 40◦ ,
is also recognized, but it is still not clear whether it could be related to a final stage of postcollisional
extension or to the following D4 strike slip faulting [8]. Finally, D4 strike-slip faults affected the
exhumed basement but did not involve the Permian to Eocene successions that crop out in the study
area. At the Variscan Realm scale, a late strike-slip tectonics is from far recognized. Moreover, is clearly
observable in the field that the lower Permian granitoids postdate the D1, D2, and D3 structures,
whereas there is less evidence about their relationship with strike-slip tectonics. As we will describe
below, LD1 and D2 folds, as well as late faults, played a significant role in controlling the geometry
of orebodies.
Figure 4. Schematic relationships between D1, LD1, and D2 structure. Note the development of
recumbent D2 folds in the limb of upright LD1 antiform (modified after [1]). The red dashed box
indicates the location of the scheme in Figure 15.
52
Minerals 2018, 8, 456
Figure 5. Mineralized outcrops in the Baccu Locci mine area, after [2].
53
Minerals 2018, 8, 456
Ores of (a) and (b)-types are hosted in the Palaeozoic metamorphic basement, and mostly occur
in the northern part of the district (Gerrei); (c) type ores are hosted in and/or related to the suite
of F-bearing Permian granites of San Vito and Quirra intrusions [9,29]; and, (d) type ores occur
prevalently in the southern part of the district (Sarrabus), are broadly typified by the “Sarrabus Silver
Lode” [30] and Silius [31] deposits, and they are possibly related to another suite of F-rich Permian
granites (Sette Fratelli and Monte Genis intrusions [9]). Although still very lacking and non-systematic,
some isotope and fluid inclusion data on different Gerrei-Sarrabus ore deposits are available from
several recent studies [10,25,30–32].
Only (a) and (b) type ores are present in the study area, and thus will be considered in detail in
this work. Both types are structurally controlled and are located at different structural levels.
4. Results
54
Minerals 2018, 8, 456
Structures
The old mine exploited several lenticular sulphide orebodies that were located in the hinge zone of
a km-size open, upright antiform (a minor order structure of the largest Flumendosa Antiform) related
to the LD1 collisional phase (see cross-section in Figure 3 and the scheme in Figure 4). The antiform
axis weakly plunges toward N120 (Figures 3 and 8); the structure deformed the foliated isoclinal
folds related to the early stage of shortening (D1 Variscan phase) and the mylonitic shear zone that
separates the Riu Gruppa and Gerrei tectonic units. The exploited sulphide lens-shaped orebodies
developed parallel to the D1 mylonitic foliation, which in this area is at a low angle with the D1
axial plane foliation (Figure 9). There is no evidence of the primary bedding, completely transposed
during the D1 tectonic phase. The lenses attain a maximum thickness of 6–7 m for maximum extension
in a strike of 80–100 m. Although located in the hinge zone of an antiform, they cannot simply be
classified as typical saddle reefs, as they do not display the classical triangular shape related to hinge
collapse. The main dip of the mineralized lenses is up to 20◦ toward N 140◦ , so more or less in the axis
plunge direction. The shape of the orebodies is well detectable by studying the old room-and-pillar
mineworks, still being partially accessible.
In detail, it is possible to recognize crosscut relationships between different events of
mineralization. The large lens-shaped orebodies are mainly characterized by abundant Zn–Cu
sulphides forming veinlets 1 to 10 cm in size hosted in the mylonite (Figure 6a), being generally
parallel to the mylonitic foliation of the hinge zone (Figure 9). Galena, chalcopyrite, pyrite, and other
sulphides are arranged in cm-sized veinlets that involve the hosting rocks for about 1 m of thickness;
they cut at low angle the Zn–Cu lenses (Figure 6e), thus postdating them. In the field, these veins
became progressively steeper and more arsenopyrite-rich (Figure 7c).
55
Minerals 2018, 8, 456
Figure 6. Outcrop pictures of the Baccu Locci mineral deposits: (a) Su Spilloncargiu mineworks:
Zn–Cu–Pb sulphide lens ore (type a mineralization). The thin sulphide “beds” (dashed line) are parallel to
the Variscan mylonitic foliation (solid line). (b) Baccu Trebini outcrop (type b mineralization): Qtz–As–Pb
sulphide sheeted veins along a SW-dipping brittle shear zone. (c) Along Rio Baccu Locci, close to San
Riccardo mineworks (type b mineralization): Qtz–As–Pb dm-size vein with subhorizontal slickenlines.
(d) Baccu Trebini outcrop (type b mineralization): Fault breccia with Qtz–As–Pb sulphides wrapping
wall rock clasts along SW-dipping fault. (e) Su Spilloncargiu mineworks (type a mineralization): in a
pillar of the exploited mine a type b Qtz–As–Pb sulphide vein (white dotted line) crosscuts a sulphide
lens that is parallel to Variscan foliation (white lines).
56
Minerals 2018, 8, 456
Figure 7. Microtextural features of ores in the study area (polished sections, reflected light):
(a) Su Spilloncargiu mineworks, early mixed sulphide ore (type a mineralization): Brecciated
quartz–sphalerite layer infilled by late galena; the mineralized layer follows the Sm foliation and
exhibits sharp contact with a phyllosilicate layer in the mylonitic matrix; high-Fe sphalerite shows
a distinct chalcopyrite disease with large chalcopyrite and pyrrhotite exsolutions. (b) Su Spilloncargiu
mineworks, early mixed sulphide ore (type a mineralization cementation zone): Brecciation of
sphalerite, infilled by primary galena, is particularly highlighted by cementing fine veinlets of
secondary galena. (c) Su Spilloncargiu mineworks, quartz–arsenopyrite veins (type b mineralization
crosscutting type a mineralization): Arsenopyrite aggregates enveloping early sphalerite aggregates
(with some galena) from mixed sulphide ore. (d) Su Spilloncargiu mineworks, quartz–arsenopyrite
veins (type b mineralization): Typical brecciated arsenopyrite texture, with large aggregates of
fractured sub-idiomorphic crystals. (e) San Riccardo mineworks (type b mineralization): large galena
(chalcopyrite) enveloping arsenopyrite crystals. (f) San Riccardo mineworks (type b mineralization):
Detail of the late ore, with abundant fractured galena infilled by late chalcopyrite and inclusion poor
sphalerite; sphalerite only displays very fine chalcopyrite exsolutions along crystallographic planes.
Qtz, quartz, gl, galena, sp, sphalerite, asp, arsenopyrite.
57
Minerals 2018, 8, 456
Figure 8. Stereographic projection (equal area, lower hemisphere) of D1 tectonic foliation (black dot),
LD1 axial surface (dashed line), and calculated axis (red circle), attitude of sulphide orebodies (red dot).
Figure 9. Schematic relationships between lens-shaped sulphide orebodies, D1 fold axial plane foliation,
and D1 mylonitic foliation in the hinge zone of LD1 antiform at Spilloncargiu mine works (not to scale).
58
Minerals 2018, 8, 456
and, (4) a late stage, with cryptocrystalline quartz and pyrite. In the sulphide/sulfosalts stage, textural
evidence indicates an initial deposition of galena, which is the most abundant mineral (Figure 7);
sphalerite, chalcopyrite, and other sulphides followed. Fine disseminations of euhedral (sometimes
needle-shaped) arsenopyrite (II) in cataclased wall rocks are probably related to this stage. Sphalerite
in type b ores is distinctly different from that in type a sulphide lenses, being much less abundant,
less deformed, less dark (it shows more evident internal reflections), and much less affected by
chalcopyrite disease. Wall rock alteration occurred from the earliest stages, but it is substantially limited
to narrow zones close to the veins, commonly marked by intense sericitization, sulphidation (fine
pyrite dispersion), and silicification; in many outcrops, the footwall of the veins is strongly cataclastic
and displays a characteristic black color (black cataclasite), further indication of diffuse precipitation
of carbonaceous matter during some of the mineralizing phases. A system of E-W quartz–feldspar
cm-thick veinlets has been locally observed; it distinctly crosscuts the quartz–sulphide veins and
may be related to a different (and very late) phase of fluid circulation in the area. Gold grades in
sulphide veins are 1–12 g/t [27,38], with good persistence in the whole mineralized vein system;
silver grades are 1000–1200 g/t [27]. The Au/Ag ratio in gold grains is <1, in opposition to regional
trends [37]. According to Bakos et al. [37], 10 μm sized gold grains are particularly associated with
chalcopyrite and galena/bournonite myrmekitic intergrowths that infill microfractures in cataclased
arsenopyrite aggregates.
Structures
Qtz–As–Pb sulphide veins are generally hosted in narrow brittle shear zones, usually not thicker
than 10 m, of high-angle faults dipping about 70◦ toward N 230◦ , confirming the structural trend
recognizable at the scale of the entire district for type b mineralization [26,27]. The faults generally
involve both the quartz mylonite, whose protolith is not possible to ascribe to one of the mapped
formations, and the Ordovician rhyolitic volcanites with augen-textures (“Porfiroidi” Fm.) that
constitute some hectometer-sized tectonic slices inside the shear zone (Figure 3).
The faults clearly cut all the ductile D1, LD1, D2, and D3 structures (shear zone, folds, and
foliations) and are sealed by the lower Eocene sediments. From the structural map (Figure 3), it is
evident that the mineralized faults are parallel to the LD1 antiform axis and are mainly located in
the hinge zone. We can interpret this occurrence considering two likely types of reactivation of
previous structures, both occurring at a shallower structural level than the LD1 phase. They could
be hinge-parallel fractures developed in the fold outer arc, parallel to the bc plane according the
fold-related joints classification by Hancock [39]; or, more probably, the faults reactivated the
noncontinuous, spaced crenulation cleavage that discontinuously developed just in the hinge zone of
LD1 antiforms.
The mineralized bodies are lenticular, elongated to laminated veins that are typical of a fault-fill
vein system. They can vary from isolated veinlets no more than 1 cm thick to sheeted veinlets and
laminated veins in which the hydrothermal mineral component prevails over the host rock component
(Figure 6b). Along the fault zones hosting the veins, several kinematic indicators are found, frequently
at the contact between veins and wall rocks. In some damage zones there are fault breccia with wall
rock clasts wrapped by dominant hydrothermal quartz (Figure 6d); furthermore, some shear zone is
characterized by a foliated black cataclasite showing S-C type fabric. Slickensides and striated surfaces
occur also on the contact between the fault-fill veins (Figure 6c). Slickenlines, tension gashes and
S-C type complex foliations collected along the SW-dipping faults hosting the main quartz–sulphide
veins all indicate a dextral strike-slip kinematic with a small reverse component (a tectonic transport
direction from the top to the NW, some data are plotted in Figure 10b). A kinematic analysis was
performed also considering faults hosting Qtz–As–Pb sulphide veins but with a different orientation
allow for us to reconstruct a strain ellipsoid with a subvertical intermediate axis (λ2 , and subhorizontal
shortest (λ3 ) and longest (λ1 ) axes, respectively, oriented roughly N-S and E-W (Figure 10). Although
parallelism between strain and stress ellipsoid cannot be demonstrated and we did not perform a
59
Minerals 2018, 8, 456
paleostress inversion, the kinematics suggests a paleostress field with a subhorizontal σ1 in agreement
with strike-slip tectonics.
Figure 10. (a) Stereographic projection (equal area, lower hemisphere) of poles to Qtz–As–sulphide
main orebodies in the Baccu Locci zone; (b) Kinematic analysis of the SW dipping faults hosting quartz
Qtz–As–sulphide veins. λ1 , λ2 , and λ3 are the directions of maximum, intermediate, and minimum
strain ellipsoid axes, respectively.
Typical economic orebodies exploited in the past mine were sulphide-rich ore shoots up to 8–9 m
thick, extending along the SW-dipping faults 100–300 m along strike, and over 100 m down dip.
By using the available detailed maps of mineworks, it is possible to construct a 3D model of these
orebodies, particularly for the San Riccardo/Su Spinosu mine (Figure 11b). Along the stretched
mineralized zones, the orebody thickness increases where the fault surface is less inclined, almost
subhorizontal, and decreases where the fault is steeper, generally becoming no thicker than 2 m. At the
underground level 214.68 in the San Riccardo mineworks, which is unfortunately hardly accessible
today, this geometry has been observed exactly along section C-C’ in Figure 11a, where the fault plane
and the sulphide veins are subhorizontal (Figure 12). Although no longer accessible, from the 3D
model and from the old mine reports, the room-and-pillar exploitation between levels 201.68 and
179.68 can be considered as a less inclined sector of the fault, marked by an increase in thickness of the
orebody (section A-A’ in Figure 11a). From these observations, the more relevant economic orebodies
of type b mineralization may be associated with very large dilational jogs developed on the hanging
wall of the transpressive faults (Figure 11a), where the occurrence of less inclined segments connecting
the subvertical ones (Figure 11c) produced room for the emplacement of the orebodies during the
fault activity.
60
Minerals 2018, 8, 456
Figure 11. The Baccu Locci/San Riccardo mine, type b mineralization (numbers are elevation above
sea level, a.s.l.): (a) schematic vertical sections that point out the sigmoid shape of the sulphide veins
of the San Riccardo orebodies (see trace in b), interpreted as large dilational jogs; (b) minework plans
based on the original mine maps (different levels are identified by colors; and, (c) three-dimensional
(3D) model of the northern part of mineworks in b, in the area of section A-A’; note that the orientation
is different from the map in b.
61
Minerals 2018, 8, 456
during the late Variscan extension [12] and recently dated, in the Gerrei district, at 302 ± 0.2 Ma (U–Pb
dating on zircon [13]).
Figure 12. Outcrop relationships between Late Variscan mafic dikes and type b ores in the southern
branch of San Riccardo underground mineworks: (a) sub-horizontal Qtz–As–Pb sulphides veins parallel
to a mafic dike (mf) about 1.0 m thick (contact is highlighted by dashed white line); and, (b) Qtz–As–Pb
sulphide ore (underlined by dashed white line) that progressively became less steep toward the left
(ESE). The dashed box in the small picture indicates the location of the outcrop respect to the whole
lens-shaped orebody. Photo in a is about 2 m to the left (i.e., west) of photo in b.
62
Minerals 2018, 8, 456
5. Discussion
As previously discussed, the basic approach of this study was mainly focused on finding field
and ore microscopy elements that are able to: (1) unravel the reciprocal relationships between ore
deposit types and (2) provide indications on the controls operated by the Variscan tectonic structures
during mineralization events in the Baccu Locci mine area.
Previous works have usually considered the occurrence of type a sulphide lenses in the Baccu Locci
mine area as a single ore. This was interpreted in different ways, essentially trying to establish genetic
links with different phases of mineralization and with the type b ores. Thus, according to a proposed
“synsedimentary” model, the Zn–Cu–Pb sulphides would represent an initial (predeformation)
concentration of metals (proto-ore) that, remobilized during Variscan deformation, provided the
sulphide component to type b veins [37,40]. An interesting issue was first raised by the study of
Zucchetti [38], which evidenced a partial superposition of mineral assemblages in both ores. This was
considered as indicative of an origin of Zn–Cu–Pb sulphides by lateral infilling from Qtz–As–Pb
sulphide veins, but without clearly distinguishing the two mineralization in space and time.
Structural data and ore mineralogy observations that were carried out with this study document
a complex mineralization history, including two distinct kinds of polyphasic ores (here, type a and type
b ores in this text) that show different minerals assemblage and different spatial relationships with
tectonic structures (Figure 13).
Figure 13. Evolution of mineralizing events in the Baccu Locci mine area and their relationships with
tectonic events.
63
Minerals 2018, 8, 456
show mutual crosscutting relationships that demonstrate the progressive and polyphased deposition
of white quartz (pre-ore stage in Figure 13), arsenopyrite (As–Fe stage), and galena (Pb–Zn–Cu–Sb–Au
stage) as the most abundant mineral phases. This type b mineral assemblage is well represented in the
San Riccardo/Su Spinosu ore, where the ore paragenesis is not associated with a type a ore and linkage
between the ore and transpressive dextral faults is manifest. Furthermore, ore microscopy shows that
between the several type b mineralizing stages, there are progressive mutual relationships of cataclasis
and successive mineral infilling, showing that type b was synkinematic with a brittle deformation.
The post-ore stage of mineralization (cryptocrystalline quartz, pyrite) is widespread and crosscuts
all of the previous mineral assemblages, so it postdates the main mineralizing events. Only in type
b veins does gold occurrence assume economic relevance. On the contrary, a lack of significant gold
grades in type a mineralization is noteworthy; it could be related to lower content in arsenopyrite,
which has been recorded in other areas of the Gerrei–Sarrabus district as a probable first carrier for
gold (Brecca mine: [5,41]).
Figure 14. Sketch of the mutual relationships between type a and type b ores located at the top of LD1
antiform hinge zone.
64
Minerals 2018, 8, 456
antiform, Sa Lilla mine; Figure 2), where previously reputed “stratoid” orebodies of sulphides, similar
by mineral association and texture [42] to those of Baccu Locci/Spilloncargiu, are located at the top of
a comparable hinge zone, in which development of sulphide “beds” was once again parallel to the D1
tectonic foliations.
Some speculation can be made about the modality that allowed the brittle reactivation of
horizontal tectonic foliations to create space for type a mineral deposition. Structural data point
out that type a ores postdate D1 and LD1 structures. Further, the creation of horizontal lens-shaped
voids now filled with mineral deposits during the D2 extensional phase sounds unrealistic, because
the vertical stress σ1 (that we can image roughly higher than the lithostatic stress) operating during
the rocks’ exhumation would have prevented the opening of subhorizontal discontinuities. So, we can
argue that type a ores could have been developed after the Gerrei and Rio Gruppa tectonic units got to
shallower crustal levels, where the fluid pressure could overcome the lithostatic stress, possibly when
the vertical stress σ1 related to the extensional dynamics ceased. The role of decreased lithostatic stress
in allowing for the development of open spaces that are suitable for mineralization should be confirmed
by the absence of similar orebodies in deeper parts of the hinge zone. Anyway, this interpretation
can be demonstrated when data about the baric environment are available. Up to now, the only
available data from fluid inclusions [10] highlight that a metamorphic environment can be excluded,
thus mineralization might postdate D2 deformation.
The type b Qtz–As–Pb veins are hosted in narrow brittle–ductile shear zones, generally developed
inside the quartz–mylonitic rocks. We described their occurrence in the Spilloncargiu mine at the top
of an LD1 antiform, where they cut at low angle type a ore, but they reach their main thickness in
the San Riccardo/Su Spinosu mine. There, the availability of detailed mineworks plans permitted
the recognition of the tectonic relationship and highlighted the occurrence of large dilational jogs
(Figure 13). Interestingly, there are close relationships between jog geometry and older D1–LD1–D2
structures, although jogs clearly postdate them. In fact, the San Riccardo main fault is a generally
WSW-steepening dipping transpressive fault that mostly cuts at high angle the NE-dipping D1 foliation
in the northeast limb of LD1 antiform (Figure 15). The fault abruptly changes the dip direction to
WSW, gently dipping just where it crosses gently SW-dipping D1 foliation in the reverse limb of a D2
recumbent fold (Figure 15). So, dilational jogs developed when the D4 fault reactivated preexisting
anisotropies (in this case, D1 tectonic foliations) if they had the right attitude. Ore mineralogy
shows repeated cataclasis and mineral infilling during the several mineralizing stages of type b veins,
suggesting a progressive brittle deformation that in some cases produced large dilational jogs that
were suitable to host orebodies with economic relevance [43,44]. Actually, the most important mineral
assemblage—the galena- (and gold-) rich ore related to the Pb–Zn–Cu-Sb–Au stage of mineralization
(Figure 13)—is associated with the largest jogs. The jog structures are not perfectly cylindrical, so their
geometries can change slightly along strikes (Figure 11). This is probably due to a change in the D1
foliation attitude in the reverse limbs of D2 recumbent folds or to a change of the local stress field.
The recognition of such large dilational jogs is not very common [44] and it has been possible by
the availability of observations at different scales. Moreover, the case of Baccu Locci/San Riccardo
reveals that the occurrence of previous reactivable foliation, possibly with different attitude due to a
polyphase deformation, can be one of the situations suitable for the development of large jogs. In the
study area, the understanding of such active structural control would allow a different way to find new,
possibly more fruitful, orebodies. In the case study, the change in dip direction of D1 tectonic foliation
on the limbs of LD1 folds could be ignored, because it is not directly linkable with the mineralized
veins, but it could be a clue to identify the occurrence of large dilational jogs.
Finally, some considerations might be given to the relationships between mafic dikes and the
Qtz–As–Pb sulphide vein system, which is problematic. The definition of possible genetic links between
the mafic magmatism and the mineralization processes falls well outside the scope of this work,
requiring a wider geochemical study at the district scale. However, field mapping and explorations
into the old mineworks showed several mutual relationships between dikes and ore veins (Figure 12).
65
Minerals 2018, 8, 456
These spatial relationships suggest a coeval emplacement. Mafic dikes may have intruded in an
interval between the main mineralizing events of type b ore, in particular, between the first and
second mineralization stages (pre-ore/As–Fe sulphide stages) and the third stage (sulphide/sulfosalts
stage) of the paragenetic sequence, producing apparently contrasting timing relationships in the field.
Under this hypothesis, considering the age available for these rocks in the neighboring areas [13],
mafic dikes assume a chronological constraint, suggesting an age of mineralizing events around
302 Ma; that is, an age in which the Variscan basement of southern Europe suffered a widespread
tectonic extension. This age is, in fact, consistent with: (a) 40 Ar-39 Ar dating of hydrothermal white
mica in quartz–arsenopyrite–gold veins in the nearby Monte Ollasteddu area (307 ± 3 Ma) [10],
and (b) several geological constraints occurring in the whole Gerrei district [26], indicating that the
quartz–arsenopyrite–gold vein systems clearly predate the previously described mineralized vein
systems related to F-bearing granites, dated at 286 Ma [9,12], and they are unconformably covered by
lower Permian sediments dated at 295 Ma [13]. However, it cannot be excluded that the latest stages
of mineralization in Baccu Locci (including the type b Pb–Zn–Cu–Sb–Au sulphide/sulfosalts stage
that affects the mafic dikes) could be related to events referable to the younger part of the outlined
chronological interval. Overall, these constraints allow for us to consider the ore deposits of the
Baccu Locci mine area as part of a much wider metallogenic event that affected various massifs of the
Variscan orogen between 310 and 300 Ma [45,46]. From their geological, mineralogical, and geochemical
characteristics, they can be best classified as Variscan orogenic gold type [47]. As in other massifs
of European Variscides, in Sardinia this metallogenic event involved a crustal-scale flow of fluids
during the late Variscan extension, resulting in widespread mineralizing processes in major regional
structures, such as the Flumendosa Antiform. Large shear zones, such as the Baccu Locci shear zone,
are part of the main plumbing system through which deep fluids were focused toward the shallower
parts of the crust.
Figure 15. Baccu Locci/San Riccardo mine: scheme of the geometric relationships between foliation,
faults, mafic dikes, and Qtz–As–Pb sulphide veins. See the location of this structure in the larger LD1
antiform in the cross-section in Figures 3 and 4.
6. Conclusions
The ores in Baccu Locci are a good example of structurally controlled mineralization in a basement
characterized by the overprinting of several tectonic phases, from ductile to brittle, during both
compressive and extensional regimes. The control exerted was either a passive reactivation of older
foliations to create space for mineral deposition, or an active syn-kinematic deposition of minerals
during the progressive evolution of the hosting structure. In particular, the emplacement of the older
types of ores exposed in Baccu Locci emerges as the result of the opening of previous discontinuities
66
Minerals 2018, 8, 456
(foliations) in the hinge zone of large antiforms, where they are subhorizontal, after their exhumation
to shallow structural levels when postcollisional extension ceased. Afterward, a different stress field
produced transpressive faults that reactivated anisotropic surfaces parallel to the axial plane in the
hinge zone. Along the transpressive dextral faults, large dilational jogs developed, whose geometry
was influenced by sudden changes of the attitude of Variscan foliation in the reverse limbs of recumbent
folds. The jogs formed together with mineral deposits, exerting in this way an “active” control of
mineralization and hosting the more economically relevant ores. As a general statement, the occurrence
of older tectonic foliations and folds might be taken into account not only because they can be directly
presumed to be hosting mineralized veins, a common concept in the study of structure–orebody
relationships, but also when considering the influence that they could have in modifying hosting
structures and favoring the formation of more significant orebodies, being in this way a good tool for
prospecting new relevant ores.
Although the overprinting relationships between the different ores and the mafic dikes are now
clearer, more data are needed to better constrain the thermobaric environment in which the minerals
were deposited, the time interval between them, the source of the ore fluids, and finally the role of
mafic dikes in the large frame of the Late Variscan metallogenic epoch in Sardinia.
Author Contributions: A.F., A.D., and S.N. conceptualized the study. A.F. performed the geological mapping.
A.F., C.B, and F.C. performed the structural analysis and 3D structural modelling. A.D. and S.N. performed the
orebodies' survey. S.N. performed the macro to micro-scale ore mineralogy. A.F. and S.N. wrote the original
draft and with C.B. and F.C. reviewed and edited the draft. Funding acquisition and project administration were
performed by A.F. and S.N.
Funding: This research was funded by FdS-RAS Fondazione di Sardegna and Regione Autonoma della Sardegna
grant number F72F16003080002 and by Italian Government, project PRIN-2005 grant number 2005047008.
Acknowledgments: The authors are grateful to two anonymous reviewers for their comments that improved the
quality of the paper and thank the editors for the careful editorial management.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Conti, P.; Carmignani, L.; Oggiano, G.; Funedda, A.; Eltrudis, A. From thickening to extension in the Variscan
belt—Kinematic evidence from Sardinia (Italy). Terra Nova 1999, 11, 93–99. [CrossRef]
2. Funedda, A.; Naitza, S.; Conti, P.; Dini, A.; Buttau, C.; Tocco, S.; Carmignani, L. The geological and
metallogenic map of the Baccu Locci mine area (Sardinia, Italy). J. Maps 2011, 2011, 103–114. [CrossRef]
3. Conti, P.; Funedda, A.; Cerbai, N. Mylonite development in the Hercynian basement of Sardinia (Italy).
J. Struct. Geol. 1998, 20, 121–133. [CrossRef]
4. Funedda, A.; Naitza, S.; Conti, P.; Dini, A.; Buttau, C.; Tocco, S.; Carmignani, L. Structural control of ore
deposits: The Baccu Locci shear zone (SE Sardinia). Rend. Online SGI 2011, 15, 66–68.
5. Lerouge, C.; Bouchot, V.; Douguet, M.; Naitza, S.; Tocco, S.; Funedda, A. Variscan Gold Mineralisation of Baccu
Locci and Brecca, Southeastern Sardinia: Petrographic and Geochemical Studies; BRGM Report N RP-54431-FR;
BRGM: Orleans, France, 2007; p. 47. Available online: http://infoterre.brgm.fr/rapports/RP-54431-FR.pdf
(accessed on 13 October 2018).
6. Conti, P.; Carmignani, L.; Funedda, A. Change of nappe transport direction during the Variscan collisional
evolution of central-southern Sardinia (Italy). Tectonophysics 2001, 332, 255–273. [CrossRef]
7. Cocco, F.; Funedda, A. The Sardic Phase: Field evidence of Ordovician tectonics in SE Sardinia, Italy.
Geol. Mag. 2017, 1–14. [CrossRef]
8. Carmignani, L.; Conti, P.; Barca, S.; Cerbai, N.; Eltrudis, A.; Funedda, A.; Oggiano, G.; Patta, E.D.; Ulzega, A.;
Orrù, P.; et al. Foglio 549-Muravera. Note Illustrative; Servizio Geologico d’Italia: Roma, Italy, 2001; p. 140.
9. Conte, A.M.; Cuccuru, S.; D’Antonio, M.; Naitza, S.; Oggiano, G.; Secchi, F.; Casini, L.; Cifelli, F.
The post-collisional late Variscan ferroan granites of southern Sardinia (Italy): Inferences for inhomogeneity
of lower crust. Lithos 2017, 294–295, 263–282. [CrossRef]
67
Minerals 2018, 8, 456
10. Dini, A.; Di Vincenzo, G.; Ruggieri, G.; Rayner, J.; Lattanzi, P. Monte Ollasteddu, a new late orogenic gold
discovery in the Variscan basement of Sardinia (Italy)—Preliminary isotopic (40Ar-39Ar, Pb) and fluid
inclusion data. Miner. Depos. 2005, 40, 337–346. [CrossRef]
11. Cortesogno, L.; Cassinis, G.; Dallagiovanna, G.; Gaggero, L.; Oggiano, G.; Ronchi, A.; Seno, S.; Vanossi, M.
The Variscan post-collisional volcanism in Late Carboniferous-Permian sequences of Ligurian Alps, Southern
Alps and Sardinia (Italy): A synthesis. Lithos 1998, 45, 305–328. [CrossRef]
12. Ronca, S.; Del Moro, A.; Traversa, G. Geochronology, Sr-Nd isotope geochemistry and petrology of Late
Hercynian dike magmatism from Sarrabus (SE Sardinia). Period. Mineral. 1999, 68, 231–260.
13. Dack, A. Internal Structure And geochronology of the Gerrei Unit in the Flumendosa Area, Variscan External
Nappe Zone, Sardinia, Italy. Master’s Thesis, Boise State University, Boise, ID, USA, 2009.
14. Cassinis, G.; Ronchi, A. Upper Carboniferous to Lower Permian continental deposits in Sardinia (Italy).
Geodiversitas 1997, 19, 217–220.
15. Carmignani, L.; Carosi, R.; Di Pisa, A.; Gattiglio, M.; Musumeci, G.; Oggiano, G.; Pertusati, P.C. The Hercynian
chain in Sardinia (Italy). Geodin. Acta 1994, 7, 31–47. [CrossRef]
16. Conti, P.; Patta, E.D. Large scale W-directed tectonics in southeastern Sardinia. Geodin. Acta 1998, 11, 217–231.
[CrossRef]
17. Carosi, R.; Musumeci, G.; Pertusati, P.C. Senso di trasporto delle unità tettoniche erciniche della Sardegna
dedotto dagli indicatori cinematici nei livelli cataclastico-milonitici. Rend. Soc. Geol. Ital. 1990, 13, 103–106.
18. Casini, L.; Funedda, A. Potential of pressure solution for strain localization in the Baccu Locci Shear Zone
(Sardinia, Italy). J. Struct. Geol. 2014, 66, 188–204. [CrossRef]
19. Casini, L.; Funedda, A.; Oggiano, G. A balanced foreland–hinterland deformation model for the Southern
Variscan belt of Sardinia, Italy. Geol. J. 2010, 45, 634–649. [CrossRef]
20. Funedda, A. Foreland- and hinterland-verging structures in fold-and-thrust belt: An example from the
Variscan foreland of Sardinia. Int. J. Earth Sci. 2009, 98, 1625–1642. [CrossRef]
21. Funedda, A.; Meloni, M.A.; Loi, A. Geology of the Variscan basement of the Laconi-Asuni area
(central Sardinia, Italy): The core of a regional antiform refolding a tectonic nappe stack. J. Maps 2015,
11, 146–156. [CrossRef]
22. Montomoli, C.; Iaccarino, S.; Simonetti, M.; Lezzerini, M.; Carosi, R. Structural setting, kinematics and
metamorphism in a km-scale shear zone in the Inner Nappes of Sardinia (Italy). Ital. J. Geosci. 2018, 137,
294–310. [CrossRef]
23. Carmignani, L.; Pertusati, P.C. Analisi strutturale di un segmento della catena ercinica: Il Gerrei
(Sardegna sud-orientale). Boll. Soc. Geol. Ital. 1977, 96, 339–364.
24. Carmignani, L.; Oggiano, G.; Barca, S.; Conti, P.; Salvadori, I.; Eltrudis, A.; Funedda, A.; Pasci, S. Geologia
della Sardegna. Note Illustrative della Carta Geologica in Scala 1:200.000; Servizio Geologico d’Italia: Roma, Italy,
2001.
25. Carmignani, L.; Cortecci, G.; Dessau, G.; Duchi, G.; Oggiano, G.; Pertusati, P.; Saitta, M. The antimony
and tungsten deposit of Villasalto in South-Eastern Sardinia and its relationship with Hercynian tectonics.
Schweiz. Mineral. Petrogr. Mitt. 1978, 58, 163–188.
26. Funedda, A.; Naitza, S.; Tocco, S. Caratteri giacimentologici e controlli strutturali nelle mineralizzazioni
idrotermali tardo-erciniche ad As-Sb-W-Au del basamento metamorfico paleozoico della Sardegna
Sud-orientale. Resoconti dell’Associazione Mineraria Sarda 2005, 110, 25–46.
27. Garbarino, C.; Naitza, S.; Tocco, S.; Farci, A.; Rayner, J. Orogenic Gold in the Paleozoic Basement of SE
Sardinia. In Mineral Exploration an Sustainable Development; Eliopoulos, D.G., Ed.; Mill Press: Rotterdam,
The Netherlands, 2003; pp. 767–770.
28. Naitza, S.; Oggiano, G.; Cuccuru, S.; Casini, L.; Puccini, A.; Secchi, F.; Funedda, A.; Tocco, S. Structural and
magmatic controls on Late Variscan Metallogenesis: Evidences from Southern Sardinia (Italy). In Mineral
Resources in a Sustainable World, Proceedings of the 13th Biennial SGA Meeting, Nancy, France, 24–27 August 2015;
André-Mayer, A.S., Cathelineau, M., Muchez, P.H., Pirard, E., Sindern, S., Eds.; The Society for Geology
Applied to Mineral Deposits (SGA): Nancy, France, 2015; pp. 161–164.
29. Naitza, S.; Conte, A.M.; Cuccuru, S.; Oggiano, G.; Secchi, F.; Tecce, F. A Late Variscan tin province associated
to the ilmenite-series granites of the Sardinian Batholith (Italy): The Sn and Mo mineralisation around the
Monte Linas ferroan granite. Ore Geol. Rev. 2017, 80, 1259–1278. [CrossRef]
68
Minerals 2018, 8, 456
30. Belkin, H.E.; De Vivo, B.; Valera, R. Fluid inclusion study of some Sarrabus fluorite deposits, Sardinia, Italy.
Econ. Geol. 1984, 79, 409–414. [CrossRef]
31. Boni, M.; Balassone, G.; Fedele, L.; Mondillo, N. Post-Variscan hydrothermal activity and ore deposits in
southern Sardinia (Italy): Selected examples from Gerrei (Silius Vein System) and the Iglesiente district.
Period. Mineral. 2009, 78, 19–35.
32. Giamello, M.; Protano, G.; Riccobono, F.; Sabatini, G. The W-Mo deposit of Perda Majori (SE Sardinia, Italy):
A fluid inclusion study of ore and gangue minerals. Eur. J. Mineral. 1992, 4, 1079–1084. [CrossRef]
33. Allmendinger, R.W.; Cardozo, N.; Fisher, D. Structural Geology Algorithms: Vectors and Tensors; Cambridge
University Press: Cambridge, UK, 2012; p. 302.
34. Lena, G.; Barchi, M.R.; Alvarez, W.; Felici, F.; Minelli, G. Mesostructural analysis of S-C fabrics in a shallow
shear zone of the Umbria–Marche Apennines (Central Italy). Geol. Soc. Lond. Spec. Publ. 2018, 409, 149–166.
[CrossRef]
35. Rutter, E.H.; Maddock, R.H.; Hall, S.H.; White, S.H. Comparative microstructures of natural and
experimentally produced clay-bearing fault gouges. Pure Appl. Geophys. 1986, 124, 3–30. [CrossRef]
36. Barton, P.B.; Bethke, P.M. Chalcopyrite disease in sphalerite: Pathology and epidemiology. Am. Mineral.
1987, 72, 451–467.
37. Bakos, F.; Carcangiu, G.; Fadda, S.; Mazzella, A.; Valera, R. The gold mineralization of Baccu Locci
(Sardinia, Italy): Origin, evolution and concentration processes. Terra Nova 1990, 2, 232–237. [CrossRef]
38. Zucchetti, S. The lead-arsenic-sulfide ore deposit of Bacu Locci (Sardinia-Italy). Econ. Geol. 1958, 53, 867–876.
[CrossRef]
39. Hancock, P.L. Brittle microtectonics: Principles and practice. J. Struct. Geol. 1985, 7, 437–457. [CrossRef]
40. Schneider, H.-J. Schichtgebundene NE-Metall- und F-Ba-Lagerstätten im Sarrabus-Gerrei-Gebiet,
SE-Sardinien. I. Bericht: Zur Lagerstättenkunde und Geologie. Neues Jahrb. Mineral. Monatshefte 1972,
1972, 529–541.
41. Lerouge, C.; Naitza, S.; Bouchot, V.; Funedda, A.; Tocco, S. Invisible gold in arsenopyrite of the Variscan
Au-Sb Brecca mineralization (Gerrei district, Southeastern Sardinia). Geol. Fr. 2007, 2, 124.
42. Violo, M. Contributo alla conoscenza dei giacimenti stratoidi polimetallici, in area metamorfica. Il giacimento
di Sa Lilla (San Vito, Cagliari-Sardegna). Resoconti dell’Associazione Mineraria Sarda 1966, 71, 5–110.
43. Cox, S.F.; Braun, J.; Knackstedt, M.A. Principles of structural control on permeability and fluid flow in
hydrothermal systems. Rev. Econ. Geol. 2001, 14, 1–24.
44. Robert, F.; Poulsen, K.H. Vein formation and deformation in greenstone gold deposits. Rev. Econ. Geol. 2001,
14, 111–155.
45. Bouchot, V.; Ledru, P.; Lerouge, C.; Lescuyer, J.L.; Milesi, J.P. Late Variscan mineralizing systems related to
orogenic processes: The French Massif Central. Ore Geol. Rev. 2005, 27, 169–197. [CrossRef]
46. De Boorder, H. Spatial and temporal distribution of the orogenic gold deposits in the Late Palaeozoic
Variscides and Southern Tianshan: How orogenic are they? Ore Geol. Rev. 2012, 46, 1–31. [CrossRef]
47. Bouchot, V.; Milesi, J.P.; Ledru, P. Crustal Scale Hydrothermal Palaeofield and Related Au, Sb, W Orogenic
Deposits at 310–305 Ma (French Massif Central, Variscan Belt). SGA News 2000, 10, 6–12.
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
69
minerals
Article
Multi-Stage Deformation of the Khangalas Ore
Cluster (Verkhoyansk-Kolyma Folded Region,
Northeast Russia): Ore-Controlling Reverse Thrust
Faults and Post-Mineral Strike-Slip Faults
Valery Y. Fridovsky 1, *, Maxim V. Kudrin 1 and Lena I. Polufuntikova 2
1 Diamond and Precious Metal Geology Institute, SB RAS, Yakutsk 677000, Russia; [email protected]
2 M.K. Ammosov North-Eastern Federal University, Yakutsk 677000, Russia; [email protected]
* Correspondence: [email protected]; Tel.: +7-4112-33-58-72
Abstract: This study reports the results of the analysis of multi-stage deformation structures of
the Khangalas gold ore cluster, northeast Russia. Four Late Mesozoic-Early Eocene deformation
stages were identified. The first deformation event (D1) was characterized by the development
of NW-striking tight to isoclinal folds of the first generation (F1) and interstratal detachment
thrusts. Major folds, extensive thrusts, boudinage, cleavage, auriferous mineralized fault zones
and quartz-vein gold mineralization were formed in the reverse and thrust fault stress field during
the progressive deformation stage (D1), with NE-SW-oriented σ1. Post-ore deformation is widely
manifested in the region. Structures D2 and D3 are coaxial. Sinistral strike-slip motions (D2 and D3)
occurred along NW-trending faults under prevailing W-E compression. They were accompanied
by the formation of NS- and NE-striking F2–3 folds with steep hinges and by bending of the earlier
formed structures, among them ore-controlling ones. The last deformation event (D4) was represented
by normal-dextral strike-slip faulting, refolding of rocks, pre-existing structures and ore bodies
and by the development of folds with steep hinges. Key structural elements of varying age are
described, the chronology of deformation events and mineralization reconstructed and their relation
to geodynamic events in northeast Asia established.
Keywords: Verkhoyansk-Kolyma folded region; Khangalas ore cluster; orogenic gold mineralization;
deformation structure; thrust fault; strike-slip fault
1. Introduction
The Khangalas ore cluster (KOC) is located in the southeastern part of the Kular-Nera slate belt of
the Verkhoyansk-Kolyma folded region in northeast Russia. It was discovered in the 1940s. Rich placer
deposits with large gold nuggets are known there. Commercial exploitation of the KOC commenced
in the latter half of the 20th Century and continues to present day. At an early stage of geologic
investigation (1940–1980), considerable attention was given to concordant ore bodies confined to the
limbs of brachyanticlinal folds [1]. Then, in the late 1980s to the early 2000s, mineralized fault zones of
complex structure with diverse mineralization were identified and investigated, which considerably
enlarged the mineral resource potential of the KOC [2,3]. Southeasterly, in the Upper Kolyma gold
district with a similar geological-structural setting, several small- and medium-sized gold deposits
are found, such as Vetrenskoye, Chay-Yuruye, Svetloye, etc. [4,5]. Gold mineralization of the KOC
is of the orogenic type, which is characterized by a close relationship with Late Jurassic-Neocomian
tectonomagmatic events in the Verkhoyansk-Kolyma folded region [6–10]. The paper presents new
data on the KOC geology obtained by the authors in the last few years, which provide a better
understanding of the relations between folds, faults and mineralization in the region.
3. Geology of the Southeastern Part of the Kular-Nera Slate Belt and the Khangalas Ore Cluster
The Kular-Nera slate belt (KNSB) is situated in the central part of the Verkhoyansk-Kolyma folded
region [6]. It is mainly composed of Upper Permian, Triassic and Lower Jurassic terrigenous rocks.
Extensive faults separate the belt from adjacent tectonic structures. In the northeast, it is separated from
the In yali-Debin synclinorium by the Charky-Indigirka and Chai-Yureye faults, and in the southwest,
the Adycha-Taryn fault separates it from structures of the passive margin of the North Asian craton.
The structural pattern of KNSB is defined by linear folds and faults of NW strike that developed over
several deformation stages. Within the KOC, NW-striking faults represent branches of the Nera Fault,
which manifests itself as 4 km-wide zones of intensive deformations and subvertical foliation of the
rocks. Dextral strike-slip motions have been reported along the Nera Fault [20].
Magmatism is poorly manifested in KNSB. It is mainly represented by granitoid massifs,
subvolcanic magma of dacite composition and dikes belonging to the NNW-striking Tas-Kystabyt
magmatic belt. They were formed in Late Jurassic-Albian times [21]. Various tectonomagmatic
events characteristic of the Late Jurassic-Late Cretaceous history of the eastern margin of the North
Asian craton are manifested within KNSB [6,22,23]. The Late Jurassic-Early Cretaceous was marked
by accretion and collision of the Kolyma-Omolon microcontinent against the craton margin and
by subduction processes in the Uda-Murgal island arc. These events produced different-age fold
and thrust structures, S- and I-type granitoids and orogenic gold deposits. In the Late Neocomian,
the direction of the Kolyma-Omolon microcontinent motion and of subduction in the Uda-Murgal arc
changed [6]. At that time, left-lateral strike slip motions first occurred in KNSB along NW-trending
faults. Post-accretionary tectonic events and Au-Sb and Ag mineralization events were related to Late
Cretaceous subduction within the Okhotsk-Chukotka arc [24].
The Khangalas ore cluster is located in the arch of the Nera anticlinorium that is represented
in the study area by the NW-striking Dvoinaya anticline composed of dislocated Upper Permian
and Lower-Middle Triassic terrigenous rocks (Figure 1). The Upper Permian (P2 ) deposits make up
the core of the Dvoinaya anticline. The lower part of the section consists of massive brownish-grey
71
Minerals 2018, 8, 270
and grey greywacke sandstones with thin siltstone interbeds. The upper part is dominated by an
800 m thick sequence of dark-grey and black siltstones with inclusions of pebbles of sedimentary,
magmatic and metamorphic rocks. The limbs of the Dvoinaya anticline are made of 680–750 m
thick Lower-Middle Triassic deposits (T1 ), mainly dark-grey shales, mudstones and siltstones with
rare interbeds of light-grey sandstones. The Middle Triassic deposits of the Anisian stage (T2 a) are
represented by a 700–800 m thick sequence of alternating sandy siltstones and siltstones with rare
fine-grained sandstone interbeds. The Ladinian strata (T2 l) are chiefly made of interbedded siltstones
and sandstones with a total thickness of 850–950 m.
The main ore-controlling rupture dislocations are the Khangalas, Dvoinoy and Granitny faults
represented by zones of breccia and fracture, low sulfidation of rocks and quartz-carbonate vein
mineralization (Figure 1). The Khangalas Fault crosscuts the Khangalas ore cluster in a northwest
direction. It controls localization of the Khangalas deposit and Ampir and Klich-Kontrolnoye
occurrences. Within the study area, the exposed fault changes its strike from NW-SE to E-W and has a
dip direction to S-W and S. The bedding of rocks exposed in the S-W wall strikes N-W, and rocks of
the N-E wall strike NE-SW and E-W. The Dvoinoy Fault strikes E-W, and its fault plane is subvertical.
In the central part of the KOC, northward of the Klich-Kontrolnoye occurrence, the Dvoinoy Fault
adjoins the Khangalas Fault. The northeastern branch of the Dvoinoy Fault controls mineralization at
the Nagornoye deposit. The rocks of the S wall of the fault have a N-E strike, while those of the N one
strike E-W. The Granitny Fault is located in the southwestern part of the KOC. Outside the Khangalas
ore cluster, the Ala-Chubuk massif of biotite granites is confined to it.
Magmatic activity is manifested by rare mafic and intermediate dikes of the normal and
subalkaline series of Late Jurassic (Nera Complex (J3 n)) and Late Cretaceous (Khulamrinsk Complex
(K2 ch)) age (Figure 1). The Nera magmatic complex includes basalt, gabbro and diorite porphyry dikes
that extend for a distance from a few tens of meters to 2 km and have NE strike and a thickness of
1–20 m. The dikes underwent alteration. They contain quartz-carbonate veinlets. The Khulamrinsk
magmatic complex consists of rare trachybasalt dikes extending for 200–500 m. They have a NW strike
and are 1–10 m thick.
At 7 km to the northwest of the Khangalas ore cluster is the exposed Ala-Chubuk massif of
biotite granites. The K-Ar age of the massif determined on orthoclase from porphyry phenocrysts is
145.0 ± 3.0 Ma and on biotite from the groundmass 149.0 ± 3.0 Ma [25]. The available geophysical
data imply the presence of unexposed intrusions of similar composition at the Nagornoye and
Khangalas deposits [25]. The Khangalas, Dvoinoye and Duk ore fields are identified within the
KOC. The first field occurs in the southeast of the ore cluster and includes the Khangalas deposit
and the Ozhidaniye occurrence. To the northwest of them are the Klich-Kontrolnoye, Dvoinoye and
Ampir occurrences belonging to the Dvoinoye ore field. The Duk ore field includes the Nagornoye
deposit. The ore bodies consist of extensive mineralized fault zones and concordant and cross-cutting
gold-quartz veins and veinlets with simple mineral composition. The amount of ore minerals does
not exceed 1–3%. These are arsenopyrite, pyrite, galena, sphalerite, chalcopyrite and native gold with
820–830‰ fineness and rare antimonite and Pb-sulfosalts. Quartz is the main gangue mineral, with less
abundant carbonates (calcite and siderite) and chlorite. A series of successive mineral assemblages
are identified in the ores of the deposits. These are pyrite-arsenopyrite-quartz metasomatic,
quartz-pyrite-arsenopyrite vein, chalcopyrite-sphalerite-galena and sulfosalt-carbonate assemblages.
The early pyrite-arsenopyrite-quartz mineralization is developed in wall-rock metasomatites. It is
represented by irregular disseminations of pyrite and arsenopyrite metacrysts and by thin quartz
streaks. Pyrite prevails over arsenopyrite. Minerals of the metasomatic assemblage are characterized by
euhedral and subhedral crystals and a streaky-disseminated structure. Pyrite and arsenopyrite grains
show evidence of deformation and corrosion. Pyrite and arsenopyrite of the early vein assemblage
occur as disseminated euhedral grains and intergrowths. Pyrites of the vein assemblage contain Co,
Sb, As, Ni, Cu and Zn trace contaminants. Minerals of the productive chalcopyrite-sphalerite-galena
assemblage sporadically occur as disseminations and small aggregates in milk-white quartz and
72
Minerals 2018, 8, 270
Figure 1. Geological sketch map, sections and location of gold deposits and occurrences of the
Khangalas ore cluster, modified and supplemented from [2]. The inset map shows the position
of the Khangalas ore cluster. Faults: Ch-I, Charky-Indigirka; Ch-Yu, Chai-Yureye; N, Nera;
A-T, Adycha-Taryn.
Microthermometric studies of the fluid were conducted at the laboratory of the M.K. Ammosov
North-Eastern Federal University on an optical microscope AxioScope.Al fitted with a motorized
heating stage (up to 600 ◦ C) and a liquid nitrogen sample cooling system (down to −196 ◦ C) (LNP95).
Analyses were made on milk-white quartz samples from the ore veins. The results of thermo-
73
Minerals 2018, 8, 270
and cryo-metric studies showed that ore-forming fluids of the KOC originated at temperatures of
310–330 ◦ C and a pressure of 0.9 kbar. The data obtained suggest that gold mineralization was formed
at a depth of about 3.5 km.
Age determinations are scarce for the KOC mineralization. The K-Ar sericite age of the
Nagornoye deposit is 130.0 ± 4.0 Ma [23]. The authors of the given paper conducted Re-Os
dating of gold (Sample Number X-45-14) from a quartz vein from the Yuzhnaya ore zone of the
Khangalas deposit at the Center of Isotope Research of the Karpinsky All-Russian Scientific-Research
Geological Institute (St. Petersburg, Russia). The isochron Re-Os age of gold was 137.0 ± 7.6 Ma.
This indicates that productive orogenic gold-quartz mineralization of the region was formed in
Valanginian-Hauterivian times.
The mineralized brittle fault zones consist of breccias and blocks of quartzose sandstones and
siltstones and are often accompanied by concordant (a few cm to 1–2 m thick, in swells up to 5 m) and
cross-cutting quartz veins and veinlets and disseminated sulfide mineralization (Figure 2). They are
mainly localized in sandstones and at their contacts with siltstones and have conformable and crossing
relations with the host rocks. The ore zones underwent strong supergene alteration as indicated by
the presence of Fe oxides, sulfates, clay minerals, etc. Nesterov N.V. [26] has reported on a secondary
gold enrichment at the Khangalas deposit. The quartz veins are often deformed in the mineralized
fault zones, which is indicative of post-ore deformation. The host siltstones and sandstones contain
disseminated sulfide mineralization (Figure 2D) represented by fine to coarse crystalline arsenopyrite
and pyrite occurring as crystals, nests and veinlets.
Figure 2. Khangalas Fault (A) and types of mineralization in the Khangalas ore cluster (KOC):
(B) concordant veins, Centralnaya zone of the Khangalas deposit; (C) vein-veinlet mineralization;
(D) veinlet-disseminated mineralization.
4. Deformation Structures of Key Deposits and Localities of the Khangalas Ore Cluster
This section presents the results of the analysis of the deformation structures of
accretionary-collisional and post-accretionary stages in the formation history of the Khangalas and
Nagornoye deposits, Dvoinoye occurrence and Mudeken locality.
74
Minerals 2018, 8, 270
Figure 3. Geological sketch map (A), cross-section (B) and stereograms of the quartz veins poles (C–F),
Khangalas deposit, (G) Schematic block-model of Khangalas deposit. In (A), the I-I line shows the
position of the cross-section (B). Mineralized fault zones: S, Severnaya; P, Promezhutochnaya; C,
Centralnaya; Yu, Yuzhnaya; Z, Zimnyaya. Symbols in stereograms and figures hereafter are: S, position
of fault or ore zone; l, calculated direction of rock motion; n, number of measurements; dashed line,
belt of vein poles. Contours of poles to planes (per 2% area).
Various deformation structures are manifested at the Khangalas deposit (Figures 3 and 4).
Early isoclinal and tight concentric folds (F) with N-W strike and subhorizontal hinges (b) occur
in narrow (up to a few tens of meters) zones (Figure 4D). In the study area, such folds were first
mapped on the northeast limb of the Nera anticlinorium, in the zone influenced by the Chay-Yureye
75
Minerals 2018, 8, 270
Fault [8]. Early folds are draped during progressive deformation into late folds, so that crests of late
folds can often be seen on the limbs of early folds (Figure 4C). F1 folds are the most widespread in the
KOC area; they have NW-SE strike and gentle hinges (b1) (Figure 4A). These are for the most part open
folds that pass into tight ones nearby the fault zones. On the right side of Uzkiy Creek, folds and ore
zones have E-W and, less frequently, N-E strike due to superimposed strike-slip deformation. F1 folds
are accompanied by Cl cleavage (Figure 4). It is platy, rarely shelly-platy, and its intensity depends
on the rock composition. The most intense cleavage is observed in siltstones, whereas in sandstones,
it becomes coarse-platy. Its regional NW strike changes to NE-SW and E-W in areas of superposed
strike-slip deformation.
The Severnaya mineralized zone is the most extensive one. On the western side of the Khangalas
deposit, the Promezhutochnaya, Centralnaya and Yuzhnaya zones branch off from the Severnaya zone
forming a horse tail termination structure, and on the eastern side, the Zimnyaya zone diverges from it
in the E-W direction. The ore-controlling structures are confined to the core of the Dvoinaya anticlinal
fold. The strike of the ore zones varies from NW-SE to S-W and, locally, to NE-SW.
Analysis of the attitude of quartz veins and veinlets revealed five variously-oriented systems
(Figures 2 and 3). Veins of the first system have persistent parameters; they are conformable with
the host rocks (Figures 2B,C and 3). Quartz veins of the second system follow the orientation of the
bedding plane and ore zones, but they dip in the opposite direction. Low-angle veins of the third
system localized in tension fractures in sandstones are rather common. The orientation of the fourth
vein system is normal to the strike of mineralized faults. In some areas, all five systems of veins and
veinlets are present, which form stockworks. Such systems of quartz vein mineralization, which are
related to the reverse and thrust fault stress field, are also found at other gold deposits in the central
part of the Kular-Nera slate belt (Bazovskoye, Malo-Tarynskoye, Levoberezhnoye and Sana) [7,28–32].
76
Minerals 2018, 8, 270
77
Figure 4. Bing Maps-satellite image (B), panoramic photo (C) and folds (A,D) of the Khangalas deposit. (E–I) Stereograms of bedding poles; Cl, cleavage. Mineralized
fault zones: S, Severnaya; P, Promezhutochnaya; C, Centralnaya; Yu, Yuzhnaya; Z, Zimnyaya.
Minerals 2018, 8, 270
Figure 5. Geological sketch map and section of the Nagornoye occurrence. In (A), the I-I line shows
the position of the cross-section (B). The diagrams show: (C,D) bedding poles; (E) quartz vein poles in
Ore Zone 1; (F) quartz vein poles in Ore Zone 2. Contours of poles to planes (per 2% area).
78
Minerals 2018, 8, 270
The most extensively studied is Ore Zone 2 (Figures 5 and 6A). It is exposed over much of its
length in a mining trench, where it parallels a subvertical sequence of quartzy sandstones interlayered
with siltstones (Figure 6A). Statistical analysis of the attitude of quartz veins showed that on the
diagrams, fields of poles of veins are grouped, in spite of their significant scatter, along the great
circle arcs corresponding to the projection of mineralized ore zones (S) (190/85) conformable with the
rock bedding (S0) (Figure 5E,F). This indicates that feathering veins are mainly localized in extension
fractures oriented at an obtuse angle to ore-bearing structures. Concordant veins of the first system are
common (Figure 5E,F).
Variously-oriented slickenlines are established (Figure 6B). One can observe subvertical
slickenlines (l-178/81) of the early thrust-faulting stage of deformation on E-W fault planes (Figure 6B).
Strike slip accretionary slickenlines are manifested at the contacts of ore bodies (Figure 6C).
These structural elements are associated with low-amplitude zones of warping observed on the
northern wall of a trench that exposed Ore Zone 1. The axes of the warping zones plunge to SE
(120/59) and are orthogonal to (l-290/30).
Figure 6. Ore Zone 2 (A), reverse-faulting (B) (red lines on the diagram) and strike-slip faulting (C)
(blue lines on the diagram) slickenlines, Nagornoye deposit. Arrows show the direction of displacement
of the faults’ hanging walls.
Figure 7. Dextral strike-slip fault (A) and boudinage-structures (B,C), Dvoinoy Creek.
79
Minerals 2018, 8, 270
80
Minerals 2018, 8, 270
81
Figure 8. Dextral strike-slip deformations in Upper Triassic sandy siltstones, Mudeken Creek. (A,B) Fold and fault structures; (C) ladder veins in sandstone beds
(plan view); (D) shaped overturned fold (plan view); (E) dextral strike-slip fault (plan view); (F–I) diagrams: (F) bedding poles, (G) cleavage poles, (H) projection of
quartz-carbonate veins shown in (C,I) projections of faults and hinges of folds; l, direction of tectonic transportation. Contours of poles to planes (per 2% area).
Minerals 2018, 8, 270
5. Discussion
Structural-kinematic analysis of deformation elements within the KOC revealed specific structures
of deposits and occurrences. Figure 9 shows stereograms of major structural elements of the KOC
(bedding, cleavage, quartz veins and veinlets, as well as mineralized fault zones). Fold structures of
the Khangalas deposit have NW-SE to E-W strike. Fold hinges (b1) dip to WSW at angles varying
from 4–28◦ . Steep dip angles are due to superposed strike-slip deformations. Hinges of the third order
F1 folds smoothly undulate in the direction of NW regional folding within the Nera anticlinorium.
Cl1 cleavage at the Khangalas deposit has a NW-SE strike. In the fault zones, cleavage is deformed
by late strike-slip faulting, as well as bedding. In stereographic projections, poles of quartz veins
and veinlets are arranged along subvertical belts. From the aforesaid, it appears that the formation
of auriferous quartz veins and veinlets is related to major fold and thrust deformations of D1 stage
(J3 -Knc). Faults and ore zones at the Khangalas deposit have sublatitudinal and, more rarely, northeast
and northwest strike. The majority of them dip S at 30–60◦ .
F1 compressed folds of sublatitudinal strike are deformed, like Cl1 cleavage at the Nagornoye
deposit, by late strike-slip faults. This led to the formation on the limbs of F1 folds of F4 open folds with
steep (b4) hinges plunging to NNE and SSW. On the stereogram, vein bodies form a steeply-dipping
belt of poles. Faults and ore zones are, for the most part, interstratal and have latitudinal to NE-SW,
rarely NW-SE orientation.
Bedding of rocks in the Dvoinoy ore field is characterized by NW-SE and NE-SW strike related to
two different deformation stages: D1 and D4, respectively. Cleavages of NW-SE and E-W orientation
are recognized. The first cleavage Cl1 is associated with the D1 stage. It was formed in relation to
early fold-and-thrust dislocations. Cleavage Cl4 is related to the right-lateral strike-slip stage (D4).
Fault zones within the Dvoinoye ore field have mostly WNW-ESE strike.
Figure 9. Stereograms of poles of bedding, cleavage, veins and faults, Khangalas ore cluster.
Dashed line, belt of vein poles. Contours of poles to planes (per 2% area).
The results of studying other deposits within the Kular-Nera slate belt [22,31–36] in
combination with the available information on the general tectonic and metallogenic evolution of the
Verkhoyansk-Kolyma folded region [6,7,9,37] and the data on the relationships between the mapped
82
Minerals 2018, 8, 270
structural elements obtained in this study indicate that deformation occurred in four stages: D1, D2,
D3 and D4 (Table 1).
Table 1. Evolution of tectonic events and associated mineralization in the southwestern part of the
Kular-Nera slate belt.
Deformation Event
Characteristic
D1 D2 D3 D4
Geodynamic
Frontal accretion Oblique accretion Post-accretionary Post-accretionary
setting
Kinematics of NW
Thrust Sinistral strike-slip Sinistral strike-slip Dextral strike-slip
faults
Interstratal detachment thrust,
interstratal ramps, thrusts, Dextral strike-slip,
Sinistral strike-slip, Sinistral strike-slip,
cross and oblique ramps, NW W-E and NW-SE
NE-SW and NE-SW and
Structural tight and isoclinal folds, fold (F4),
NW-SE folds (F2), NW-SE folds (F3),
paragenesis NW-SE open and tight folds sublatitudinal
horizontal horizontal
(F1) with horizontal hinges cleavage horizontal
slickenlines slickenlines
(b1), boudinage, fault cleavage, slickenlines
downdip slickenlines
Veins V1 - V3 V4
Mineralization Au - Sb Ag
Attitude of σ1 Subhorizontal, NE-SW Subhorizontal, E-W Subhorizontal, E-W Subhorizontal, N-S
Graphic model
83
Minerals 2018, 8, 270
6. Conclusions
Studies of deformation structures of the Khangalas gold-ore cluster showed that they were
forming over a long time during the course of the Late Jurassic-Neocomian accretionary and Late
Cretaceous-Early Paleocene post-accretionary events in the Verkhoyansk-Kolyma folded region.
The first deformation event (D1) was characterized by the development of NW-striking tight to
isoclinal folds of the first generation (F1) and interstratal detachment thrusts. Major folds, extensive
thrusts, boudinage, cleavage, Au-bearing mineralized fault zones and quartz-vein mineralization
were formed in the conditions of the tectonic stress field characteristic of reverse and thrust faulting,
with the horizontal σ1 and vertical σ3. The D1 stage was progressive deformation under a contractional
regime. In the zones of regional faults, where deformations are most intensely manifested, inter- and
intra-stratal reverse and thrust faults developed in sandstones and at their contacts with siltstones,
which were accompanied by intense, small-scale folding. These were favorable structural conditions
for localization of mineralized fault zones with concordant and cross Au-quartz veins.
Post-ore deformations are widely manifested within the KOC. The D2 and D3 structures are
co-axial. Sinistral strike-slip motions (D2–3) occurred along NW-striking faults. Associated with them
were submeridional and NE-trending folds (F2–3) with steep hinges, as well as warping of the earlier,
including ore-controlling, structures. The sinistral strike-slip stage is poorly manifested within the
84
Minerals 2018, 8, 270
KOC, but its presence is established from the analysis of fractures, slickenlines and fault-line folds.
Faults of NE strike can also be assigned to this structural paragenesis. It is likely that sinistral strike-slip
deformations changed to dextral ones that are strongly manifested within the KOC.
The fourth event (D4) is represented by normal-dextral strike-slip motions, refolding of rocks,
earlier structural elements and ore bodies. At this stage, latitudinal structures (F4 folds, Cl4 cleavage
and S4 faults) were formed, which are more widespread here than in other metallogenic zones of
the Upper Indigirka district (Adycha-Taryn, Mugurdakh-Selerikan). It is assumed that large-scale
dextral strike-slip faults modified the structure of deposits and occurrences in the KOC. In the most
strongly-deformed areas, the strike of the structures changed to sublatitudinal and, more rarely, to NE
(Khangalas deposit). Ore zones of the Khangalas deposit were previously considered as “horse tail”
structures [2], but detailed analysis of the relationships between auriferous quartz veins, ore zones and
rock bedding permitted assigning them to the first-stage paragenesis (D1). The large scale and long
duration of post-ore strike-slip motions can be inferred from the observation that early Au-bearing
quartz veins are ground to “quartz flour” in the zones of later strike-slip faults. Also observed are
quartz breccias in which early milk-white quartz is cemented by later chalcedony-like grey quartz
typical for Ag-Sb mineralization known from the Verkhoyansk-Kolyma folded region [39,40].
Thus, it can be concluded that tectogenesis within the Verkhoyansk-Kolyma folded region
followed a regular change from the Late Jurassic-Neocomian frontal accretionary regime to the
Aptian-Early Eocene strike-slip regime and that gold mineralization was related to orogenic processes,
as is exemplified by the Khangalas ore cluster described in this article.
Author Contributions: Idea of the study conceived by V.Y.F. Collection of field materials by V.Y.F., M.V.K. and
L.I.P. Treatment of data and writing the text of the paper by V.Y.F. and M.V.K. Figure drawing by M.V.K.
Funding: This research was funded by Diamond and Precious Metal Geology Institute, Siberian Branch of the
Russian Academy of Sciences, project number [No. 381-2016-004] and by Russian Foundation for Basic Research,
grant number [No. 18-35-00336].
Conflicts of Interest: The authors declare no conflicts of interest.
References
1. Rozhkov, I.S.; Grinberg, G.A.; Gamyanin, G.N.; Kukhtinskiy, Y.G.; Solovyev, V.I. Late Mesozoic Magmatism
and Gold Mineralization of the Upper Indigirka District; Nauka: Moscow, Russia, 1971; p. 238. (In Russian)
2. Oxman, V.S.; Suzdalova, N.I.; Kraev, A.A. Deformation Structures and Dynamic Conditions for the Formation of
Rocks in Upper Indigirka District; Yakut Scientific Center Siberian Branch of the Russian Academy of Sciences:
Yakutsk, Russia, 2005; p. 200, ISBN 5-463-00128-6. (In Russian)
3. Fridovsky, V.Y.; Kudrin, M.V. Deformation structures of the Khangalas ore cluster. In Proceedings of the
All-Russian Scientific-Practical Conference Geology and Mineral Resources of Northeast Russia, Yakutsk,
Russia, 31 March–2 April 2015; pp. 537–540. (In Russian)
4. Voroshin, S.V.; Tyukova, E.E.; Newberry, R.J.; Layer, P.W. Orogenic gold and rare metal deposits of the
Upper Kolyma District, Northeastern Russia: Relation to igneous rocks, timing, and metal assemblages.
Ore Geol. Rev. 2014, 62, 1–24. [CrossRef]
5. Petrov, O.V.; Morozov, A.F.; Mikhailov, B.K.; Orlov, V.P.; Militenko, N.V.; Mezhelovsky, N.V.; Feoktistov, V.P.;
Shatov, V.V.; Molchanov, A.V.; Migachev, I.F.; et al. Mineral Potential of the Russian Federation; Petrov, O.V.,
Ed.; FSBI A.P. Karpinsky Russian Geological Research Institute: St. Petersburg, Russia, 2009; p. 223,
ISBN 978-5-93761-156-7. (In Russian)
6. Tectonics, Geodynamics, and Metallogeny of the Sakha Republic (Yakutia) Territory; Parfenov, L.M.;
Kuzmin, M.I. (Eds.) MAIK Nauka/Interperiodika: Moscow, Russia, 2001; p. 571, ISBN 5-7846-0046-X.
(In Russian)
7. Fridovsky, V.Y.; Prokopiev, A.B. Tectonics, geodynamics and gold mineralization of the eastern margin of
the North Asia Craton. In The Timing and Location of Major Ore Deposits in an Evolving Orogen; Blundel, D.J.,
Neuber, F., von Quadt, A., Eds.; Geological Society: London, UK, 2002; Volume 204, pp. 299–317.
85
Minerals 2018, 8, 270
8. Fridovsky, V.Y.; Polufuntikova, L.I.; Solovyev, E.E. Dynamics of the formation and structures of the
southeastern sector of the Adycha-Nera metallogenic zone (northeast Yakutia). Russ. J. Domest. Geol.
2003, 3, 16–21. (In Russian)
9. Fridovsky, V.Y. Structural control of orogenic gold deposits of the Verkhoyansk-Kolyma folded region,
northeast Russia. Ore Geol. Rev. 2017, in press. [CrossRef]
10. Goryachev, N.A.; Pirajno, F. Gold deposits and gold metallogeny of Far East Russia. Ore Geol. Rev. 2014, 59,
123–151. [CrossRef]
11. Groves, D.I.; Goldfarb, R.J.; Gebre-Mariam, M.; Hagemann, S.G.; Robert, F. Orogenic gold deposits:
A proposed classification in the context of their crustal distribution and relationship to other gold deposit
types. Ore Geol. Rev. 1998, 13, 7–27. [CrossRef]
12. Groves, D.I.; Condie, K.C.; Goldfarb, R.J.; Hronsky, J.M.A.; Vielreicher, R.M. Secular changes in global tectonic
processes and their influence on the temporal distribution of gold-bearing mineral deposits. Econ. Geol. 2005,
100, 203–224. [CrossRef]
13. Ramsay, J.G.; Huber, M.I. The Techniques of Modern Structural Geology; Academic press: London, UK, 1987;
Volume 2, p. 704, ISBN 0-12-576922-9.
14. Fossen, H. Structural Geology; Cambridge University Press: Cambridge, UK, 2010; p. 463, ISBN 978-0-521-51664-8.
15. Prokopiev, A.V.; Fridovsky, V.Y.; Gaiduk, V.V. Faults (Morphology, Geometry, Kinematics); Parfenov, L.M., Ed.;
Yakut Scientific Center Siberian Branch of the Russian Academy of Sciences: Yakutsk, Russia, 2004; p. 148,
ISBN 5-463-00016-6. (In Russian)
16. Price, N.J.; Cosgrove, J.W. Analysis of Geological Structures; Cambridge University Press: Cambridge, UK,
2005; p. 502, ISBN 0521319587.
17. Fossen, H.; Cavalcante, G.C.G.; Pinheiro, R.V.L.; Archanjo, C.J. Deformation—Progressive or multiphase.
J. Struct. Geol. 2018, in press. [CrossRef]
18. Sherman, S.I.; Dneprovsky, Y.I. Stress Fields of the Earth’s Crust and Geological Structural Methods of Their Study;
Nauka: Novosibirsk, Russia, 1989; p. 158. (In Russian)
19. Spencer, E.W. Introduction to the Structure of the Earth; McGraw-Hill Book Company: New York, NY,
USA, 1977.
20. Gusev, G.S. Folded Structures and Faults of the Verkhoyansk-Kolyma System of the Mesozoic; Nauka: Moscow,
Russia, 1979; p. 208. (In Russian)
21. Bakharev, A.G.; Zaitsev, A.I. The Tas-Kystabyt magmatic belt. In Tectonics, Geodynamics and Metallogeny of
the Rupublic of Sakha (Yakutia); Parfenov, L.M., Kuzmin, M.I., Eds.; MAIK Nauka/Interperiodika: Moscow,
Russia, 2001; pp. 263–269, ISBN 5-7846-0046-X. (In Russian)
22. Fridovsky, V.Y. Structures of gold ore fields and deposits of the Yana-Kolyma ore belt. In Metallogeny of
Collisional Geodynamic Settings; Mezhelovsky, N.V., Gusev, G.S., Eds.; GEOS: Moscow, Russia, 2002; Volume 1,
pp. 6–241. (In Russian)
23. Sokolov, S.D. Tectonics of northeast Asia: An overview. Geotectonics 2010, 44, 493–509. (In Russian) [CrossRef]
24. Bortnikov, N.S.; Gamynin, G.N.; Vikent’eva, O.V.; Prokof’ev, V.Y.; Prokop’ev, A.V. The Sarylakh and Sentachan
gold-antimony deposits, Sakha-Yakutia: A case of combined mesothermal gold-quartz and epithermal
stibnite ores. Geol. Ore Depos. 2010, 52, 339–372. (In Russian) [CrossRef]
25. Akimov, G.Y. New data on the age of Au-quartz mineralization in the Upper-Indigirka district.
Dokl. Acad. Nauk. 2004, 398, 80–83. (In Russian)
26. Nesterov, N.V. Secondary zonation of gold ore deposits in Yakutia. Izvestiya Tomsk Polytech. Univ. 1970, 239,
242–247. (In Russian)
27. GeoInfoComLLC. Available online: http://mestor.geoinfocom.ru/publ/1-1-0-55 (accessed on 5 April 2017).
28. Fridovsky, V.Y. Collisional metallogeny of gold deposits of the Verkhoyansk-Kolyma orogenic region.
Izvestiya VUZOV Geol. Explor. 2000, 4, 53–67. (In Russian)
29. Fridovsky, V.Y.; Gamyanin, G.N.; Polufuntikova, L.I. Dora-Pil ore field: Structure, mineralogy, and
geochemistry of the ore-formation environment. Ores Met. 2012, 5, 7–21. (In Russian)
30. Fridovsky, V.Y.; Gamyanin, G.N.; Polufuntikova, L.I. The Sana Au-quartz deposit within the Taryn ore cluster.
Raz. I Okhrana Nedr. 2013, 2, 3–7. (In Russian)
31. Fridovsky, V.Y.; Gamyanin, G.N.; Polufuntikova, L.I. The structure, mineralogy, and fluid regime of ore
formation in the polygenic Malo-Taryn gold field, northeast Russia. Russ. J. Pac. Geol. 2015, 9, 274–286.
[CrossRef]
86
Minerals 2018, 8, 270
32. Fridovsky, V.Y.; Polufuntikova, L.I.; Goryachev, N.A.; Kudrin, M.V. Ore-controlling thrusts of the Bazovskoe
gold deposit (East Yakutia). Dokl. Earth Sci. 2017, 474, 617–619. [CrossRef]
33. Akimov, G.Y. Lithological-structural control of Au-quartz ores of the Nagornoye deposit, East Yakutia.
Ores Met. 2000, 4, 42–46. (In Russian)
34. Fridovsky, V.Y. Strike-slip duplexes on the Badran deposit. Izvestiya VUZOV Geol. Explor. 1999, 1, 60–65.
(In Russian)
35. Fridovsky, V.Y. Analysis of deformational structures of the El’gi ore cluster (East Yakutia). Russ. J.
Domest. Geol. 2010, 4, 39–45. (In Russian)
36. Fridovsky, V.Y.; Gamyanin, G.N.; Polufuntikova, L.I. Gold quartz and antimony mineralization in the Maltan
deposit in northeast Russia. Russ. J. Pac. Geol. 2014, 8, 276–287. [CrossRef]
37. Prokopiev, A.V.; Tronin, A.V. Structural and sedimentation characteristics of the zone of junction of the
Kular-Nera slate belt and Inyali-Debin synclinorium. Russ. J. Domest. Geol. 2004, 5, 44–48. (In Russian)
38. Khanchuk, A.I.; Ivanov, V.V. Meso-Cenozoic geodynamic settings and gold mineralization of the Russian Far
East. Russ. Geol. Geophys C/C Geol. Geofiz. 1999, 40, 1607–1617.
39. Gamyanin, G.N.; Goryachev, N.A. Subsurface mineralization of eastern Yakutia. Tikhook. Geol. 1988, 2, 82–89.
(In Russian)
40. Goryachev, N.A.; Gamyanin, G.N.; Prokofiev, V.Y.; Velivetskaya, A.V.; Ignatiev, A.V.; Leskova, N.V.
Silver-antimony mineralization of the Yana-Kolyma belt (Northeast Russian). Tikhook. Geol. 2011, 30,
12–26. (In Russian)
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
87
minerals
Article
Tectonic Control, Reconstruction and Preservation of
the Tiegelongnan Porphyry and Epithermal
Overprinting Cu (Au) Deposit, Central Tibet, China
Yang Song 1, *, Chao Yang 2 , Shaogang Wei 3 , Huanhuan Yang 1,4 , Xiang Fang 1,2 and Hongtao Lu 1
1 Key Laboratory of Metallogeny and Mineral Assessment, Institute of Mineral Resources,
Chinese Academy of Geological Sciences, Beijing 100037, China; [email protected] (H.Y.);
[email protected] (X.F.); [email protected] (H.L.)
2 Department of Geology and Geological Engineering, University Laval, QC G1V 0A6, Canada;
[email protected]
3 First Crust Monitoring and Application Center, China Earthquake Administration, Tianjin 300180, China;
[email protected]
4 College of Earth, Ocean, and Atmospheric Sciences, Oregon State University, Corvallis, OR 97331, USA
* Correspondence: [email protected]; Tel.: +86-010-6899-9087
Abstract: The newly discovered Tiegelongnan Cu (Au) deposit is a giant porphyry deposit
overprinted by a high-sulfidation epithermal deposit in the western part of the Bangong–Nujiang
metallogenic belt, Duolong district, central Tibet. It is mainly controlled by the tectonic movement
of the Bangong–Nujiang Oceanic Plate (post-subduction extension). After the closure of the
Bangong–Nujiang Ocean, porphyry intrusions emplaced at around 121 Ma in the Tiegelongnan
area, which might be the result of continental crust thickening and the collision of Qiangtang and
Lhasa terranes, based on the crustal radiogenic isotopic signature. Epithermal overprinting on
porphyry alteration and mineralization is characterized by veins and fracture filling, and replacement
textures between two episodes of alteration and sulfide minerals. Alunite and kaolinite replaced
sericite, accompanied with covellite, digenite, enargite, and tennantite replacing chalcopyrite and
bornite. This may result from extension after the Qiangtang–Lhasa collision from 116 to 112 Ma,
according to the reopened quartz veins filled with later epithermal alteration minerals and sulfides.
The Tiegelongnan deposit was preserved by the volcanism at ~110 Ma with volcanic rocks covering
on the top before the orebody being fully weathered and eroded. The Tiegelongnan deposit was
then probably partly dislocated to further west and deeper level by later structures. The widespread
post-mineral volcanic rocks may conceal and preserve some unexposed deposits in this area.
Thus, there is a great potential to explore porphyry and epithermal deposit in the Duolong district,
and also in the entire Bangong–Nujiang metallogenic belt.
1. Introduction
In the past two decades, some large porphyry deposits have been found in Tibet, China, such as
Yulong deposit (6.22 Mt at 0.99% Cu) [1], Qulong deposit (7.1 Mt at 0.5% Cu) [2], Jiama deposit
(7.4 Mt at 0.5% Cu) [3], Duobuza deposit (2.9 Mt at 0.46% Cu) [4], and Bolong deposit (3.8 Mt at
0.5% Cu) [5]. This indicates that Tibet can be considered one of the most significant potential porphyry
Cu systems in the world. Recently, epithermal deposits have also been discovered and reported in
Tibet. Epithermal deposits are genetically associated with porphyry Cu deposits, especially high and
intermediate sulfidation epithermal ones, which could be discovered at upper or lateral locations of
porphyry deposits in some cases [6]. However, the epithermal deposits do not always occur close
to porphyry deposits, because epithermal deposits are normally at shallow crustal levels (surface to
1–2 km depth), therefore they could be easily eroded by later orogenesis [6].
The Duolong porphyry Cu-Au district is located in the Bangong–Nujiang metallogenic belt
(BNMB), central Tibetan Plateau, which was discovered in 2007 and hosts several large porphyry and
epithermal deposits and ore prospects (Figure 1). The Tiegelongnan deposit was discovered in 2013,
containing the largest scale Cu resource within this district, and it was documented as a porphyry Cu
(Au) deposit overprinted by high-sulfidation mineralization [7]. The total Cu content persevered in
the Tiegelongnan deposit is around 1600 Mt at 0.51% Cu. The Au content is small at about 280 Mt with
a low grade of 0.13g/t Au on average.
Despite numerous studies on the metallogeny of the Tiegelongnan deposit, the tectonic control of
this type deposit has rarely been demonstrated. Formation of the porphyry and epithermal deposits in
the Duolong district was indicated to be associated with the magma arising from the closure of the
Bangong–Nujiang Ocean (BNO) in the Early Cretaceous [8,9]. However, how the tectonic activities
control the formation of the porphyry Cu system is controversial. The Tiegelongnan deposit is the
first high-sulfidation epithermal deposit being discovered in the Tibetan Plateau. Previous studies
suggested that the limited number of epithermal deposits found in the Tibetan Plateau is due to
the dramatic uplift and deep level erosion. The Tiegelongnan deposit is an example to study the
tectonic control, reconstruction and preservation process of porphyry Cu systems in the Tibetan
Plateau. In this study, we reviewed history of the tectonic setting, magma emplacement, multiple
episodes’ mineralization, exhumation, and preservation of the Tiegelongan deposit, based on published
literatures and the detailed drill core logging and deposit 1:500 scale mapping. Besides, we discussed
the implications of this study on exploration of porphyry and epithermal deposits in the Duolong
district and other places in Tibet.
89
Minerals 2018, 8, 398
These sequences are composed of the Upper Triassic Riganpeicuo Formation (T3 r), the Lower Jurassic
Quse Formation (J1 q), the Lower to Middle Jurassic Sewa Formation (J1-2 s), the Upper Cretaceous
Abushan Formation (K2 a), and the Upper Oligocene Kangtuo Formation (E3 k). The Riganpeicuo
Formation dominated by limestone is unconformably overlain by the Quse and Sewa formations.
The Quse Formation mainly occurred in the center and southwestern part of the ore district as
the main host formation of the Duobuza and Bolong porphyry deposits. It conformably contacts
with the overlying Sewa Formation, which is the predominant host formation of the Tiegelongnan
deposit. These two Jurassic formations are thought to be part of the metamorphosed accretionary
complex formed by north-dipping subduction of the BNO plate under the Qiangtang terrane [18].
They were also interpreted to be bathyal to abyssal flysch succession, implying a stable shallow-marine
continental-shelf sedimentary environment along the southern continental margin of the South
Qiangtang terrane [10]. Furthermore, Wei et al. (2017) [9] proposed that a continental margin arc
setting in the southern Qiangtang terrane during the Early Cretaceous.
Figure 1. Regional geological map of the Duolong ore cluster, modified after [8], ages are from [17].
There are three main faults in the Duolong district striking at NE–SW, E–W, and NW–SE,
respectively (Figure 1). The NE–SW fault is a major ore-controlling structure. A number of ore-bearing
granodioritic porphyry intrusions emplaced along this fault, and therefore, many large porphyry
copper deposits such as the Bolong, Duobuza, Tiegelongnan, and Naruo deposits occurred.
Most E–W thrust faults are large scale and traverse across the entire Duolong district, dipping to
the south with an angle between 49◦ and 16◦ [19]. There are some granodiorite porphyries (125–120 Ma,
unpublished data) beaded along this NE–SW fault. A mylonite sample obtained from the fault zone
was well constrained with a 40 Ar/39 Ar plateau age at 127.8 ± 1.1 Ma [19], which represents an early
period of thrusting. The NW–SE faults are normal slip faults dipping to the south, which might be
related to the neo-tectonic movements. These faults are characterized by sunken landform and valleys
with fault breccia exposed.
90
Minerals 2018, 8, 398
at the eastern and southern margins of this area [8]. Several phases of granodiorite porphyries are
syn-mineral intrusions with ages ranging from 121 Ma to 116 Ma [8,20,21]. They are indistinguishable
from petrology and crosscutting relationships, because subsequent strong alteration weakened their
differences and boundaries. Therefore, the geochemistry data, especially the mobile elements, could not
be used to discriminate their geochemical features.
3.1. Alteration
Drill core logging reveals concealed features of the Tiegelongnan deposit (Figure 2b). Five phases
of hydrothermal alteration were identified in the Tiegelongnan deposit, according to the dominant
alteration mineral assemblage, including: biotite alteration, sericite-pyrite-quartz (phyllic) alteration,
chlorite alteration, alunite alteration, and kaolinite-dickite alteration [22]. Alunite-kaolinite-dickite
assemblages are also named as advanced argillic alteration in high-sulfidation epithermal deposits [23].
Figure 2. Ground surface and cross-section map of the Tiegelongnan porphyry Cu (Au) deposit.
91
Minerals 2018, 8, 398
Figure 3. The epithermal alteration overprinted on the porphyry alteration. (a) Alunite breccia break
phyllic altered and mineralized host rocks, (b) kaolinite vein crosscutting biotite altered host rocks
and filling inbiotite-molybdenite vein, (c) kaolinite filling in cavity of quartz-pyrite vein, (d) sericite
replaced by kaolinite grains. Alu: alunite, Bio: biotite, Ser: sercite, Kao: kaolinite, Mol: molybdenite.
92
Minerals 2018, 8, 398
3.2. Mineralization
Chalcopyrite, bornite, and pyrite are sulfide assemblage precipitating in biotite and phyllic
alteration zone with minor molybdenite, whereas the Cu (Fe)-As-S minerals enargite, tennantite,
and Cu-S covellite, digenite are the dominant sulfides in advanced argillic alteration zone [26,27].
Chalcopyrite and bornite in the biotite and phyllic alteration zone are the main Cu mineralization
of the porphyry stage, including quartz-chalcopyrite ± bornite veins and disseminated chalcopyrite
and bornite. This is the main porphyry Cu oreody, hosted as quartz-sulfides veins, and disseminated
sulfides in wall rocks. The enargite-tennatite-covellite-digenite assemblage mostly occurs in the alunite
and kaolinite alteration zone, which are the products of the high-sulfidation epithermal Cu orebody [8].
Epithermal stage Cu sulfides are mainly presented as alunite-kaolinite-sulfides veins.
These two stages of sulfide mineral assemblages display a complicated overprinting and
cross-cutting relationship. Kaolinite-sulfide veins crosscut quartz veins (Figure 4a), those sulfides are
mostly Cu (Fe)-(As)-S minerals, also some chalcopyrite and bornite were reported as result of solid
solution from those minerals [25]. Under the microscope, we find some enargite filling in the fractures
of quartz veins along with kaolinite. The pyrite occurs as early phyllic alteration product, because
it is the most easily being replaced by the enargite. Replacement textures of chalcopyrite, bornite,
and pyrite affected by Cu (Fe)-As-S and Cu-S minerals are common in the Tiegelongnan deposit.
The pyrite is replaced from the edge firstly by bornite, and then the bornite is replaced by digenite
and covellite (Figure 4b). Enargite and tennantite replace chalcopyrite (Figure 4c). There are some
arguments that replacement relationship between sulfides is supergene replacement textures, because
covellite and digenite are typical supergene sulfides also. However, δ65 Cu of covellite and digentie
in the Tiegelongnan are averaging at 0.25‰ [28], which is similar to the hypogene copper sulfides
δ65 Cu value [29]. In some cases, the Cu (Fe)-As-S and Cu-S minerals are filled in the fractures instead
of replacing Fe-bearing minerals (Figure 4d), which might indicate a brittle force condition before the
epithermal mineralization. This corresponds with the alunite and kaolinite breccia in Figure 3a. The Cu
(Fe)-As-S and Cu-S sulfides even cut through post-mineral porphyry and breccia rocks. Generally,
overprinting of the Cu (Fe)-As-S and Cu-S on chalcopyrite-bornite-pyrite assemblage is common in
the Tiegelongnan deposit, and it was demonstrated in different occurrences, including the former
replacing the latter minerals, the former filling in fractures of the latter sulfides, and the former cutting
the chalcopyrite-bornite mineralized rocks or veins.
93
Minerals 2018, 8, 398
4. Structures
The Tiegelongnan deposit and most of other deposits in the Duolong district are along the
NE–SW faults. It is widely accepted that these faults mainly controlled the emplacement of magma
and hydrothermal fluids in the Duolong district [30]. However, there were few convincing studies
clarifying the overlying volcanic rocks which conceal the whole porphyry and epithermal ore bodies.
The NW–SE faults, named as Rongna Fault, are characterized by geomorphologically linear sunken
terrain, valleys, with fault springs seen at the ground level. In the deposits area, the fault occurred as
a river valley, which is called the Rongna Valley (Figure 5). The fault divides the Meiriqiecuo Formation
andesite into two parts, suggesting the structural movement took place after andesite eruption, which
is dated at ~110 Ma [8].
Figure 5. Rongna Valley, topography of the Rongna fault in the Tiegelongnan deposit.
The audio-frequency magneto-telluric method (AMT) was applied to understand the fault features,
and further to predict the occurrences of the orebody in a deep level on the south side. From the ATM
tests, electrical properties of different rocks obviously vary from each other in the Tiegelongnan deposit.
The Cretaceous volcanic cap-rocks have low polarizability, whereas extremely high resistivity is shown
in the paleo-weathering crust. The Jurassic sandstone showed low resistivity and high polarization,
while the advanced argillic altered sandstone has high resistivity. We found the >0.5% grade Cu whole
porphyry and epithermal orebodies correspondent with the low resistivity zone (Figure 6). There are
two low-resistivity anomalies (C1 and C2) in the E103 AMT cross-section and the C1 anomaly coincides
with the explored orebody. Therefore, the C2 low-resistivity anomaly could be another part of the
whole orebody. The fault plane shown in the AMT cross-section is dipping to the south with an angle
of 70◦ to 80◦ (Figure 6).
94
Minerals 2018, 8, 398
Figure 6. 2D inversion resistivity section of the audio-frequency magneto-telluric method (AMT) test
(a) north–south cross-section; (b) east–west cross-section.
Figure 7. (a) High-resolution remote sensing image, and (b) a 3-D map of the Meiriqiecuo Formation
andesite [31]. Red dotted lines: faults; the yellow dotted line: the boundary of the mineralized body.
95
Minerals 2018, 8, 398
6. Discussion
96
Minerals 2018, 8, 398
transposed, intruded by granitoids, and were uplifted above sea level before around 118 Ma [40].
The extensive magmatism in the Duolong district is associated with the Qiangtang–Lhasa collision
event [9,37].
Numerous igneous rocks such as gabbro, basalt, basaltic andesite, andesites, rhyolite,
and intermediate to felsic porphyries are widely distributed in vicinity of the Tiegelongnan deposit.
Zircon U-Pb ages of the porphyry intrusions in the Tiegelongnan deposit range from 115.9 Ma to
123.1 Ma, which is consistent with the mineralization (molybdenite Re-Os) ages (119.0 ± 1.4 Ma [20];
121.2 ± 1.2 Ma [8]. This is also consistent with porphyry intrusions and mineralization ages of
other deposits in the Duolong district, such as Bolong and Duobuza deposits [17]. They are
temporally associated with this younger generation of magmatic emplacement which is related
with the Qiangtang–Lhasa collision. Intrusion rocks geochemistry and isotope studies have been
conducted to understand the genetic association between those porphyry deposits and tectonic settings.
Geochemistry data mostly obtained from the ore-bearing porphyritic intrusions in the Duolong district
indicate they are magmatic rocks and adakite-like rocks [11,14,44,45]. They have relatively high oxygen
fugacity (f O2 ) and high H2 O contents that are critical to the formation of porphyry and epithermal
deposits [46]. During the process of magma upwelling, adakite-like melts might get mixed with large
amount of copper and other metals and sulfur from either interaction with hot peridotite in the mantle
wedge region [47] or mixing with mantle-derived melts [48]. It eventually resulted in mantle-derived
juvenile materials, which are thought to bring heat and materials to generate juvenile mafic lower crust.
The magmas experience various degrees of fractional crystallization and crustal contamination during
its emplacement, when it is derived from the remelting of the juvenile mafic lower crust as a result of
previous arc magmatism. Some of these hybrid magmas formed calc-alkaline ore bearing porphyries
via the shallow magma emplacement, leading to the formation of giant porphyry and epithermal Cu
(Au) deposits [49–52].
The Jurassic (170–145 Ma) intermediate–felsic intrusive rocks of the southern Qiangtang terrane
primarily exhibit negative whole-rock εNd (t) and zircon εHf (t) and old Hf isotope crustal model ages,
indicating that those Jurassic rocks were largely derived from mature or recycled continental crust
materials [35,36,53]. This is compatible with what been observed in the Early Cretaceous Fuye pluton,
Caima pluton and Qingcaoshan pluton in the Qiangtang Terrane [53]. In contrast, εNd (t) and εHf (t)
value and Hf isotope model ages of the Early Cretaceous (~126–116 Ma) magmatic rocks from the
Duolong district indicate they were probably derived from magma as a mixture of the juvenile lower
crust and mature crustal materials [9]. In addition, previous studies on Pb isotopic compositions of
the porphyry intrusions, sulfides, and sulfate in the Tiegelongnan deposit suggest that the Pb of this
deposit is mainly derived from a crust–mantle mixed subduction zone [26,54,55].
Although there is no specific research on the geochemistry of the porphyry intrusions in the
Tiegelongnan deposit, owing to their strong alteration and leaching erasing its geochemical signature,
the features of the intrusions in the Duolong district could represent that in the Tiegelongnan
deposit. Thus, it suggests that plenty of juvenile crust materials are involved in the intrusions in the
Tiegelongnan porphyry Cu (Au) deposit. The juvenile crust materials have been becoming gradually
dominated during the Late Mesozoic since the vertical growth and thickening of the continental crust
of the southern Qiangtang terrane during the Early Cretaceous [9,56]. It is commonly accepted that
variable sources conjunctly contributed to the formation of magma in this district [53]. The dominant
crust signature from radiogenic isotopes is reported as features of the post-subduction products, which
is well explained by Richards (2009) [57]. All of these suggest that magmatism and mineralization
of the Tiegelongnan porphyry Cu (Au) deposit probably occurred in an active continental margin
environment after the subduction of BNO plate, as result of continental crust thickening and terranes
collision (Figure 9).
97
Minerals 2018, 8, 398
Figure 9. Tectonic setting model for the formation of the Duolong deposit, Tibet.
98
Minerals 2018, 8, 398
magmatic–epithermal environment study suggest that the ductile-brittle transition commonly occurs
about 370–400 ◦ C [59]. This is higher than epithermal hydrothermal fluid temperature 160–270 ◦ C [6].
Therefore, before epithermal events taking place, the wall rocks of the Tiegelongnan deposit is brittle
and could be easily broken by the accumulation of hydrothermal fluid or by fault events in the Duolong
district. This gives access to epithermal fluid arriving at shallow sites, and overprints porphyry system
along those faults, fractures, and other opened space.
99
Minerals 2018, 8, 398
epithermal precious metal deposit existing in theory. The reason of not being detected on the deposit
is that it either was not well preserved because of severe erosion or has not been found yet. Deep in
the porphyry Cu (Au) system, the bottom of the porphyry Cu orebody has not been detected yet.
We suggest there could be economic potential at depth, because biotite alteration is shown at depth,
if it could represent the typical potassic alteration, which usually is the core of mineralization orebody
in the porphyry Cu system [68]. Obviously, we did not reveal the whole biotite alteration, which might
be concealed at a deeper level.
Figure 10. (a) Simplified geology map of the A-B sections in Tiegelongnan [31]. (b) Anatomy of
a condensed porphyry Cu system showing the spatial interrelationships of porphyry Cu orebody,
high-sulfidation epithermal Cu ± Au orebody, and late-mineralization andesitic volcanic rocks [31].
100
Minerals 2018, 8, 398
to a deeper domain. The nearly 11 Mt of Cu resource currently being explored might be part of
the entire Tiegelongnan porphyry and epithermal Cu (Au) orebody, which is similar to the San
Manuel-Kalamazoo porphyry copper deposit in South America [69]. Comparably, we are confident
on the great potential of the Tiegelongnan deposit. Further understanding of the dislocation of the
Rongna Fault should be conducted later, which would contribute to increasing the ore reserves of the
deposit at depth.
The volcanic rocks unconformably overlie on the whole porphyry and epithermal Cu (Au)
orebody, which prevents the orebody from being subject to further erosion. This might be the reason
for only epithermal copper orebody being found at the top the Tiegelongnan deposit so far, but not
anywhere else. There might be epithermal mineralization on the Duobuza and Bolong deposits,
but they might be fully eroded away due to the lack of overlying protection. Furthermore, due to the
large range of volcanic rocks in the Duolong district, more porphyry and epithermal copper and gold
deposits are of great potential to be preserved, and that could be the future exploration direction in the
Duolong district.
Author Contributions: Y.S. carried out the project and coordinated this study. All authors took part in the field
work, analyzed the results and wrote the manuscript; Y.S. and C.Y mainly revised and edited the paper.
Funding: This research was partly funded by the National Key R & D Program of China (No. 2018YFC0604106),
the National Natural Science Foundation of China (No. 41402178), and the Chinese Geological Survey
Program (DD20160026).
Acknowledgments: We thank three anonymous reviewers, the Editor Shi, and the academic editor of Minerals
Alain Chauvet for their positive and constructive comments that are helpful to improve the manuscript
significantly. Qing Zhang from the University of Wollongong provided major contributions to the English
editing of the paper.
Conflicts of Interest: The authors declare no conflict of interest.
101
Minerals 2018, 8, 398
11. Li, J.X.; Qin, K.; Li, G.; Xiao, B.; Zhao, J.; Chen, L. Petrogenesis of Cretaceous igneous rocks from the Duolong
porphyry Cu–Au deposit, central Tibet: Evidence from zircon U–Pb geochronology, petrochemistry and
Sr–Nd–Pb–Hf isotope characteristics. Geol. J. 2016, 51, 285–307. [CrossRef]
12. Pan, G.; Wang, L.; Li, R.; Yuan, S.; Ji, W.; Yin, F.; Zhang, W.; Wang, B. Tectonic evolution of the Qinghai-Tibet
plateau. J. Asian Earth Sci. 2012, 53, 3–14. [CrossRef]
13. Zhu, D.; Zhao, Z.; Niu, Y.; Dilek, Y.; Hou, Z.; Mo, X. The origin and pre-Cenozoic evolution of the Tibetan
Plateau. Gondwana Res. 2013, 23, 1429–1454. [CrossRef]
14. Ding, S.; Chen, Y.; Tang, J.; Zheng, W.; Lin, B.; Yang, C. Petrogenesis and Tectonics of the Naruo Porphyry
Cu (Au) Deposit Related Intrusion in the Duolong Area, Central Tibet. Acta Geol. Sin. 2017, 91, 581–601.
[CrossRef]
15. Li, J.; Qin, K.; Li, G.; Noreen, J.; Zhao, J.; Cao, M.; Huang, F. The Nadun Cu–Au mineralization, central Tibet:
Root of a high sulfidation epithermal deposit. Ore Geol. Rev. 2016, 78, 371–387. [CrossRef]
16. Li, G.; Li, J.; Qin, K.; Duo, J.; Zhang, T.; Xiao, B.; Zhao, J. Geology and Hydrothermal Alteration of the
Duobuza Gold-Rich Porphyry Copper District in the Bangongco Metallogenetic Belt, Northwestern Tibet.
Resour. Geol. 2012, 62, 99–118. [CrossRef]
17. Lin, B.; Chen, Y.; Tang, J.; Wang, Q.; Song, Y.; Yang, C.; Wang, L.; He, W.; Zhang, L. 40 Ar/39 Ar and Rb-Sr
Ages of the Tiegelongnan Porphyry Cu-(Au) Deposit in the Bangong Co-Nujiang Metallogenic Belt of Tibet,
China: Implication for Generation of Super-Large Deposit. Acta Geol. Sin. 2017, 91, 602–616. [CrossRef]
18. Li, G.; Duan, Z.; Liu, B.; Zhang, H.; Dong, S.; Zhang, L. The discovery of Jurassic accretionary
complexes in Duolong area, northern Bangong Co-Nujiang suture zone, Tibet, and its geologic significance.
Geol. Bull. China 2011, 30, 1256–1260. (In Chinese)
19. Liu, Y.; Wang, M.; Li, C.; Xie, C.; Chen, H.; Li, Y.; Fan, J.; Li, X.; Xu, W.; Sun, Z. Cretaceous structures in the
Duolong region of central Tibet: Evidence for an accretionary wedge and closure of the Bangong–Nujiang
Neo-Tethys Ocean. Gondwana Res. 2017, 48, 110–123. [CrossRef]
20. Fang, X.; Tang, J.; Song, Y.; Yang, C.; Ding, S.; Wang, Y.; Wang, Q.; Sun, X.; Li, Y.; Wei, L.; et al. Formation epoch
of the South Tiegelong superlarge epithermal Cu (Au-Ag) deposit in Tibet and its geological implications.
Acta Geosci. Sin. 2015, 36, 168–176. (In Chinese)
21. Yang, C.; Beaudoin, G.; Tang, J.; Song, Y. An extreme long life span of porphyry and epithermal Cu deposit:
The Tiegelongnan deposit, Tibet, China. 2018; in preparation.
22. Yang, C.; Beaudoin, G.; Tang, J.; Song, Y. Geology and genesis of Tiegelongnan porphyry and epithermal
base metal deposit in Duolong district, Tibet, China: From stable isotope and fluid inclusions constrains.
2018; in preparation.
23. Sillitoe, R.H.; Hedenquist, J.W. Linkages between volcanotectonic settings, ore-fluid compositions, and
epithermal precious metal deposits. Spec. Publ. Soc. Econ. Geol. 2003, 10, 315–343.
24. Seedorff, E. Porphyry deposits: Characteristics and origin of hypogene features. Econ. Geol. 2005, 29,
251–298.
25. Stoffregen, R. Genesis of Acid-Sulfate Alteration and Au-Cu-Ag Mineralization at Summitville, Colorado.
Econ. Geol. 1987, 82, 1575–1591. [CrossRef]
26. Wang, Y.; Tang, J. The First Discovery of Colusite in the Tiegelongnan Supper-large Cu (Au, Ag) Deposit and
Significance for the Genesis of the Deposit. Acta Geol. Sin. 2018, 92, 400–401. [CrossRef]
27. Yang, C.; Tang, J.; Wang, Y.; Yang, H.; Wang, Q.; Sun, X.; Feng, J.; Yin, X.; Ding, S.; Fang, X.; et al. Fluid and
geological characteristics researches of Southern Tiegelong epithermal porphyry Cu-Au deposit in Tibet.
Miner. Depos. 2014, 33, 1287–1305. (In Chinese)
28. Duan, J.; Tang, J.; Li, Y.; Liu, S.; Wang, Q.; Yang, C.; Wang, Y. Copper isotopic signature of the Tiegelongnan
high-sulfidation copper deposit, Tibet: Implications for its origin and mineral exploration. Miner. Depos.
2016, 51, 591–602. [CrossRef]
29. Mathur, R.; Munk, L.; Nguyen, M.; Gregory, M.; Annell, H.; Lang, J. Modern and paleofluid pathways
revealed by Cu isotope compositions in surface waters and ores of the Pebble porphyry Cu-Au-Mo deposit,
Alaska. Econ. Geol. 2013, 108, 529–541. [CrossRef]
30. Chen, H.Q.; Qu, X.M.; Fan, S.F. Geological characteristics and metallogenic prospecting model of Duolong
porphyry copper gold ore concentration area in Gerze County, Tibet. Miner. Depos. 2015, 34, 321–332.
102
Minerals 2018, 8, 398
31. Song, Y.; Yang, H.H.; Lin, B.; Liu, Z.B.; Qin, W.; Ke, G.; Chao, Y.; Xiang, F. The Preservation System of
Epithermal Deposits in South Qiangtang Terrane of Central Tibetan Plateau and Its Significance: A Case
Study of the Tiegelongnan Superlarge Deposit. Acta Geosci. Sin. 2017, 38, 659–669. (In Chinese)
32. Kapp, P.; Yin, A.; Manning, C.; Harrison, T.; Taylor, M.; Ding, L. Tectonic evolution of the early Mesozoic
blueschist-bearing Qiangtang metamorphic belt, central Tibet. Tectonics 2003, 22. [CrossRef]
33. Kapp, P.; Yin, A.; Harrison, T.; Ding, L. Cretaceous-Tertiary shortening, basin development, and volcanism
in central Tibet. Geol. Soc. Am. Bull. 2005, 117, 865–878. [CrossRef]
34. Pullen, A.; Kapp, P.; Gehrels, G.; Ding, L.; Zhang, Q. Metamorphic rocks in central Tibet: Lateral variations
and implications for crustal structure. Geol. Soc. Am. Bull. 2011, 123, 585–600. [CrossRef]
35. Liu, D.; Huang, Q.; Fan, S.; Zhang, L.; Shi, R.; Ding, L. Subduction of the Bangong–Nujiang Ocean:
Constraints from granites in the Bangong Co area, Tibet. Geol. J. 2014, 49, 188–206. [CrossRef]
36. Hao, L.; Wang, Q.; Wyman, D.A.; Ou, Q.; Dan, W.; Jiang, Z.; Wu, F.; Yang, J.; Long, X.; Li, J. Underplating of
basaltic magmas and crustal growth in a continental arc: Evidence from Late Mesozoic intermediate–felsic
intrusive rocks in southern Qiangtang, central Tibet. Lithos 2016, 245, 223–242. [CrossRef]
37. Zhu, D.; Li, S.; Cawood, P.; Wang, Q.; Zhao, Z.; Liu, S.; Wang, L. Assembly of the Lhasa and Qiangtang
terranes in central Tibet by divergent double subduction. Lithos 2016, 245, 7–17. [CrossRef]
38. Allmendinger, R.; Jordan, T.; And, S.; Isacks, B. The evolution of the Altiplano-Puna plateau of the Central
Andes. Annu. Rev. Earth Planet. Sci. 1997, 25, 139–174. [CrossRef]
39. Zhu, D.; Pan, G.; Wang, L.; Mo, X.; Zhao, Z.; Zhou, C.; Liao, Z.; Dong, G.; Yuan, S. Tempo-spatial variations of
Mesozoic magmatic rocks in the Gangdese belt, Tibet, China, with a discussion of geodynamic setting-related
issues. Geol. Bull. China 2008, 27, 1535–1550. (In Chinese)
40. Kapp, P.; Decelles, P.; Gehrels, G.; Heizler, M.; Lin, D. Geological records of the Lhasa-Qiangtang and
Indo-Asian collisions in the Nima area of central Tibet. Geol. Soc. Am. Bull. 2007, 119, 917–933. [CrossRef]
41. Yin, A.; Harrison, T.M. Geologic evolution of the Himalayan-Tibetan orogen. Annu. Rev. Earth Planet. Sci.
2000, 28, 211–280. [CrossRef]
42. Qu, X.-M.; Wang, R.; Xin, H.; Jiang, J.; Chen, H. Age and petrogenesis of A-type granites in the middle
segment of the Bangonghu–Nujiang suture, Tibetan plateau. Lithos 2012, 146, 264–275. [CrossRef]
43. Li, J.-X.; Qin, K.; Li, G.; Xiao, B.; Zhao, J.; Cao, M.; Chen, L. Petrogenesis of ore-bearing porphyries from the
Duolong porphyry Cu–Au deposit, central Tibet: Evidence from U–Pb geochronology, petrochemistry and
Sr–Nd–Hf–O isotope characteristics. Lithos 2013, 160, 216–227. [CrossRef]
44. Li, X.; Li, C.; Sun, Z.; Wang, M. Origin and tectonic setting of the giant Duolong Cu–Au deposit,
South Qiangtang Terrane, Tibet: Evidence from geochronology and geochemistry of Early Cretaceous
intrusive rocks. Ore Geol. Rev. 2017, 80, 61–78. [CrossRef]
45. Li, G.; Qin, K.; Li, J.; Evans, N.; Zhao, J.; Cao, M.; Zhang, X. Cretaceous magmatism and metallogeny in the
Bangong–Nujiang metallogenic belt, central Tibet: Evidence from petrogeochemistry, zircon U–Pb ages, and
Hf–O isotopic compositions. Gondwana Res. 2017, 41, 110–127. [CrossRef]
46. Hou, Z.; Mo, X.; Gao, Y.; Qu, X.; Meng, X. Adakite, a possible host rock for porphyry copper deposits:
Case studies of porphyry copper belts in Tibetan Plateau and in Northern Chile. Miner. Depos. 2003, 22, 1–12.
(In Chinese)
47. Defant, M.J.; Drummond, M.S. Derivation of some modern arc magmas by melting of young subducted
lithosphere. Nature 1990, 347, 662. [CrossRef]
48. Sillitoe, R. Epochs of intrusion-related copper mineralization in the Andes. J. S. Am. Earth Sci. 1988, 1, 89–108.
[CrossRef]
49. Sillitoe, R.H. A plate tectonic model for the origin of porphyry copper deposits. Econ. Geol. 1972, 67, 184–197.
[CrossRef]
50. Hou, Z.; Gao, Y.; Qu, X.; Rui, Z.; Mo, X. Origin of adakitic intrusives generated during mid-Miocene east–west
extension in southern Tibet. Earth Planet. Sci. Lett. 2004, 220, 139–155. [CrossRef]
51. Hou, Z.; Yang, Z.; Lu, Y.; Kemp, A.; Zheng, Y.; Li, Q.; Tang, J.; Yang, Z.; Duan, L. A genetic linkage between
subduction-and collision-related porphyry Cu deposits in continental collision zones. Geology 2015, 43,
247–250. [CrossRef]
103
Minerals 2018, 8, 398
52. Oyarzun, R.; Márquez, A.; Lillo, J.; López, I.; Rivera, S. Reply to Discussion on “Giant versus small porphyry
copper deposits of Cenozoic age in northern Chile: Adakitic versus normal calc-alkaline magmatism”
by Oyarzun R, Márquez A, Lillo J, López I, Rivera S (Mineralium Deposita 36: 794–798, 2001). Miner. Depos.
2002, 37, 795–799. [CrossRef]
53. Li, J.; Qin, K.; Li, G.; Richards, J.; Zhao, J.; Cao, M. Geochronology, geochemistry, and zircon Hf
isotopic compositions of Mesozoic intermediate–felsic intrusions in central Tibet: Petrogenetic and tectonic
implications. Lithos 2014, 198–199, 77–91. [CrossRef]
54. Lv, L.; Zhao, Y.; Song, L.; Tian, Y.; Xin, H. Characteristics of C, Si, O, S and Pb isotopes of the Fe-rich and Cu
(Au) deposits in the western Bangong–Nujiang metallogenic belt, Tibet, and their geological significance.
Acta Geol. Sin. 2011, 85, 1291–1304. (In Chinese)
55. Xin, H.; Qu, X.; Wang, R.; Liu, H.; Zhao, Y.; Wei, H. Geochemistry and Pb, Sr, Nd isotopic features of
ore-bearing porphyries in Bangong Lake porphyry copper belt, western Tibet. Miner. Depos. 2009, 28,
785–792. (In Chinese)
56. Hawkesworth, C. The generation and evolution of the continental crust. J. Geol. Soc. 2010, 167, 229–248.
[CrossRef]
57. Richards, J.P. Postsubduction porphyry Cu-Au and epithermal Au deposits: Products of remelting of
subduction-modified lithosphere. Geology 2009, 37, 247–250. [CrossRef]
58. Sillitoe, R. Styles of high-sulphidation gold, silver and copper mineralisation in porphyry and epithermal
environments. In Proceedings of the Australasian Institute of Mining and Metallurgy, Melbourne, Australia,
11–13 September 2000; Volume 305, pp. 19–34.
59. Fournier, R.O. Hydrothermal processes related to movement of fluid from plastic into brittle rock in the
magmatic-epithermal environment. Econ. Geol. 1999, 94, 1193–1211. [CrossRef]
60. Zhang, K.; Zhang, Y.; Tang, X.; Xia, B. Late Mesozoic tectonic evolution and growth of the Tibetan plateau
prior to the Indo-Asian collision. Earth Sci. Rev. 2012, 114, 236–249. [CrossRef]
61. Li, Y.; Wang, C.; Li, Y.; Ma, C.; Wang, L.; Peng, S. The Cretaceous tectonic event in the Qiangtang Basin and
its implications for hydrocarbon accumulation. Pet. Sci. 2010, 7, 466–471. [CrossRef]
62. Yang, K.; Ma, C. Some advances in the rates of continental erosion and mountain uplift. Geol. Sci. Technol. Inf.
1996, 15, 89–96.
63. Kesler, S.E.; Wilkinson, B.H. The role of exhumation in the temporal distribution of ore deposits. Econ. Geol.
2006, 101, 919–922. [CrossRef]
64. Murphy, M.; Yin, A.; Harrison, T.; Dürr, S.; Chen, Z.; Ryerson, J.; Kidd, F.; Wang, X.; Zhou, X. Did the
Indo-Asian collision alone create the Tibetan plateau? Geology 1997, 25, 719–722. [CrossRef]
65. Ketcham, R.A. Forward and inverse modeling of low-temperature thermochronometry data.
Rev. Mineral. Geochem. 2005, 58, 275–314. [CrossRef]
66. Yang, H.H.; Tang, J.; Dilles, J.; Song, Y. Temperature Study of the Duolong Porphyry Cu-Au District and its
implications for the Evolution of the Qiangtang Terrane in Tibet, China. Int. Geol. Rev. 2018. submitted.
67. Li, G. High temperature, salinity and strong oxidation ore-forming fluid at Duobuza gold-rich porphyry
copper in the Bangonghu tectonic belt, Tibet: Evidence from fluid inclusions study. Acta Petrol. Sin. 2007, 23,
935–952.
68. Sillitoe, R.H. Porphyry Copper Systems. Econ. Geol. 2010, 105, 3–41. [CrossRef]
69. Lowell, J.D.; Guilbert, J.M. Lateral and vertical alteration-mineralization zoning in porphyry ore deposits.
Econ. Geol. 1970, 65, 373–408. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
104
minerals
Review
The Jbel Saghro Au(–Ag, Cu) and Ag–Hg
Metallogenetic Province: Product of a Long-Lived
Ediacaran Tectono-Magmatic Evolution in the
Moroccan Anti-Atlas
Johann Tuduri 1,2, *, Alain Chauvet 3 , Luc Barbanson 2 , Jean-Louis Bourdier 2 ,
Mohamed Labriki 4 , Aomar Ennaciri 4 , Lakhlifi Badra 5 , Michel Dubois 6 ,
Christelle Ennaciri-Leloix 4 , Stanislas Sizaret 2 and Lhou Maacha 4
1 BRGM, F-45060 Orléans, France
2 ISTO, UMR7327, Université d’Orléans, CNRS, BRGM, F-45071 Orléans, France;
[email protected] (L.B.); [email protected] (J.-L.B.);
[email protected] (S.S.)
3 Géosciences Montpellier, cc. 060, Université de Montpellier 2, CEDEX 5, 34095 Montpellier, France;
[email protected]
4 MANAGEM, Twin Center, BP 5199, Casablanca 20100, Morocco; [email protected] (M.L.);
[email protected] (A.E.); [email protected] (C.E.-L.); [email protected] (L.M.)
5 Faculté des Sciences, Université Moulay Ismaïl, BP 11201 Zitoune, Meknes 50 000, Morocco;
badra_lakhlifi@yahoo.fr
6 Laboratoire de Génie Civil et géo-Environnement–Lille Nord de France, EA 4515, Département des Sciences
de la Terre, Université de Lille, Bât. SN5, 59655 Villeneuve d’Ascq, France; [email protected]
* Correspondence: [email protected]; Tel.: +33-238-644-790
Abstract: The Jbel Saghro is interpreted as part of a long-lived silicic large igneous province. The area
comprises two lithostructural complexes. The Lower Complex consists of folded metagreywackes
and N070–090◦ E dextral shear zones, which roughly results from a NW–SE to NNW–SSE shortening
direction related to a D1 transpressive tectonic stage. D1 is also combined with syntectonic plutons
emplaced between ca. 615 and 575 Ma. The Upper Complex is defined by ash-flow caldera
emplacements, thick and widespread ignimbrites, lavas and volcaniclastic sedimentary rocks with
related intrusives that were emplaced in three main magmatic flare ups at ca. 575, 565 and 555 Ma.
It lies unconformably on the Lower Complex units and was affected by a D2 trantensive tectonic stage.
Between 550 and 540 Ma, the magmatic activity became slightly alkaline and of lower extent. Ore
deposits show specific features, but remain controlled by the same structural setting: a NNW–SSE
shortening direction related to both D1 and D2 stages. Porphyry Au(–Cu–Mo) and intrusion-related
gold deposits were emplaced in an earlier stage between 580 and 565 Ma. Intermediate sulfidation
epithermal deposits may have been emplaced during lull periods after the second and (or) the third
flare-ups (560–550 Ma). Low sulfidation epithermal deposits were emplaced late during the felsic
alkaline magmatic stage (550–520 Ma). The D2 stage, therefore, provided extensional structures that
enabled fluid circulations and magmatic-hydrothermal ore forming processes.
Keywords: structural control; silicic large igneous province; ignimbrite flare-ups; ash-flow caldera;
epithermal; porphyry; IRGD; Anti-Atlas
1. Introduction
In northwest Africa, the Anti-Atlas, Ougarta and Hoggar domains consist of pericratonic terranes
located at the margin of the West African Craton (WAC, Figure 1a) and that were mostly amalgamated
from Palaeoproterozoic to Phanerozoic times [1–5]. These terranes were the site of recurring tectonic
activity and periods of intense magmatic activity. The most important magmatic pulses occurred
during the Mesoproterozoic [6–8] and at the end of Triassic when the Central Atlantic Magmatic
Province (CAMP) was emplaced [9–11]. According to Ernst [12] and Ernst and Bleeker [13], both
events have been related to large igneous provinces (LIPs). Indeed LIPs consist of large volumes of
mainly mafic magma (>0.1 Mkm3 ) in provinces whose areal extent might exceed 0.1 Mkm2 . LIPs
are thought to emplace in a short duration pulse or multiple pulses (less than 1–10 Myr each) with
a whole maximum duration of ca. 50 Myr. Intense felsic magmatism may also occur as silicic
large igneous provinces (SLIPs) [12,14]. Dacite–rhyolite pyroclastic rocks (ignimbrites), along with
transitional calc-alkaline I-type [15] to A-type granites, mainly characterise these SLIPs [16]. Further,
LIPs are commonly related to a wide variety of metal deposits [17] including world-class deposits
such as magmatic sulfide ore deposits associated with mafic and ultramafic magmatism (Ni, Cu,
PGE, Cr, Ti, Fe [18,19]), with carbonatite and peralkaline complexes (Nb, Ti, REE, Zr [20,21]), or with
diamondiferous kimberlites [22]. Iron oxide copper gold (IOCG) deposit types [23] and epithermal
deposits of mostly low and intermediate sulfidation gold-based metal types [24–26] may be also
related to more silicic LIPs. Another magmatic event described in peri-Gondwanan terranes of, e.g.,
Avalonian and Cadomian types, is mostly characterised by huge volumes of pyroclastic flows and was
emplaced at the end of the Neoproterozoic era [27–30]. Conditions and geodynamical environment
at the origin of such a silicic province remain insufficiently understood, although Moume et al. [31]
recently proposed that this event might be related to the Central Iapetus Magmatic Province event
(CIMP) of Ediacaran-Cambrian age [13,32]. In the Moroccan Anti Atlas, world-class deposits occur in
an area mostly dominated by rhyolitic ignimbrites such as the giant Ag–Hg Imiter deposit, the Ag-Hg
Zgounder deposit and the Co–Ni–Fe–As(–Au–Ag) Bou Azzer district (Figure 1b).
In fact, the Moroccan Anti-Atlas hosts several precious and base-metal deposits affected by at
least four major tectonic phases: i.e., the Palaeoproterozoic, the Neoproterozoic, the Variscan and the
Alpine cycles [29,33–35]. Until recently, most of the ore deposits from the Anti-Atlas were considered
as Neoproterozoic in age due to their occurrence within Proterozoic inliers. One exception was the vein
and stratabound copper deposits hosted within the early Palaeozoic cover, which were assumed to be
syn-sedimentary or epigenetic and Variscan in age [36,37]. However, recent studies have reassessed
the age and origin of numerous metal deposits, assigning younger ages than previously admitted.
The arguments are essentially two-fold: (i) absolute dating and (ii) fluid chemistry by isotopic and
fluid inclusion study methods. Indeed geochronological methods (e.g., Re/Os, Ar/Ar) frequently give
younger ages than expected though possible resetting of dating materials and (or) ore remobilisation are
rarely discussed. For instance, the Imiter deposit would coincide with the Permo-Triassic boundary [38].
Similarly, a late Carboniferous age has been proposed for ore enrichment at BouAzzer with a possible
earlier pre-mineralising stage [39]. The strong association of mineral deposits with fluids of moderate
to high salinities, suggests interpretations favouring the influence of basinal brines in the ore-forming
processes [40,41]. Indeed, processes involving basin-related and(or) surface-related brines resulting
from evaporation of seawater in Triassic basins in the formation of ore deposits, up to now interpreted
as, deposits related to the late Neoproterozoic felsic magmatic event [42–48] have been defended by
several recent works [40,41,49] although the debate is still open [50]. This controversy mainly concerns
the Bou Azzer (Co–Ni–), Imiter (Ag–Hg) and Zgounder (Ag–Hg) mines that represent the main three
world-class ore deposits of the Moroccan Anti-Atlas after the Akka gold mine was closed few years
ago. The Imiter concentration has been proposed to be associated with the 550 ± 3 Ma rhyolitic
magmatism [51], a hypothesis recently controverted by Ar-Ar geochronology [38] and palaeo-fluid
geochemistry investigations [41]. The Zgounder deposit is supposed to be emplaced around 564 ± 5 Ma
at the same time as rhyolitic intrusions [46,52] although fluid geochemistry would suggest a Triassic
fluid contribution [41,49] sedimentary. The Bou Azzer Co–Ni–Fe–As(–Au–Ag) district hosts the only
mine in the world where Co is produced as a primary commodity directly from Co- and As-bearing
106
Minerals 2018, 8, 592
arsenide minerals [53]. It is interpreted as having experienced many ore remobilisations from Late
Neoproterozoic, Variscan to Triassic magmatic-hydrothermal stages [39,40,44,54,55].
ȱ
Figure 1. (a) Location of the Anti-Atlas belt at the northern limit of the West African Craton, after
Thiéblemont et al. [56]. (b) Main geological units and major mining districts of the Moroccan
Anti-Atlas [5,7,29,48,57–59]. Inliers–BD: Bas Drâa; If: Ifni; K: Kerdous; TA: Tagragra d’Akka; Im:
Igherm; TT: Tagragra de Tata; Ig: Iguerda; AM: Agadir-Melloul; Z: Zenaga.
The debate as to whether the Anti-Atlas ore deposits are mostly Neoproterozoic in age and
magmatic-related, or Phanerozoic and disconnected from any magmatic input, is similar to the one that
exists between the orogenic and intrusion-related gold deposit models [60–65]. For instance, as applied
to the Variscan gold ore deposits in the French Massif Central and beyond, this debate is focused on the
involvement of magmatic fluids in the formation of mineralised systems. Arguments supporting the
orogenic model are mostly based on fluid inclusion studies and isotopic data on quartz-bearing veins
and highlight the meteoric and/or metamorphic signatures of the fluids [66,67]. In such cases, heat
production from possible synchronous granitoids is supposed to generate only thermal convection
cells. Arguments in favour of an intrusion-related model highlight the systematic spatial association
between granite and hydrothermal systems, while fluid compositions and metal source are interpreted
as showing a magmatic signature [68–70].
107
Minerals 2018, 8, 592
Therefore, it may be relevant to clarify whether most of the ore deposits of the Moroccan Anti-Atlas
are linked to a Neoproterozoic magmatic input (the magmatic-related alternative), or to a more recent
stage related to the penetration and circulation of sedimentary brines and (or) metamorphic fluids
into the basement (the orogenic alternative). Indeed, as the Anti-Atlas domain may be considered
as an important Neoproterozoic magmatic province with respect to the abundance of plutonic and
volcanic rocks, evidences for a major Neoproterozoic metallogenetic province need to be deciphered.
In this work, we present a review of the global geology and metallogeny of the Jbel Saghro in the
Eastern Anti-Atlas with a specific emphasis on the formation, styles and tectonic controls of different
ore deposit occurrences in order to assess and discuss a tectono-magmatic evolution of the studied
area and a metallogenetic and geodynamic model. The world-class Bou Azzer deposit is not concerned
in this study because of its location far from the Jbel Saghro. Its characterisation will be the subject of
work currently in progress.
108
Minerals 2018, 8, 592
109
Minerals 2018, 8, 592
110
ȱ
Figure 2. Simplified geologic map of the Jbel Saghro, after Hindermeyer et al. [104], Tuduri [47] and Tuduri et al. [48]. U-Pb radiometric ages obtained on
zircon [29,48,51,78,105–108].
Minerals 2018, 8, 592
111
Minerals 2018, 8, 592
ȱ
Figure 3. Structural maps of the (a) Imiter inlier modified from SMI (Société Métallurgique d’Imiter)
data and Ighid et al. [121] and (b) Boumalne inlier (from Tuduri et al. [48]). Note that metabasaltic
rocks with pilloid structures have never been observed in the Imiter inlier. Sills have been described
and discussed as possibly syn-sedimentary in origin [100]. Stereoplots of structural orientation data:
(c) bedding, (d) foliation and refraction cleavage, and (e) lineation. Bedding data are from the Qal’at
112
Minerals 2018, 8, 592
MGouna, Boumalne and Imiter Lower Complex inliers whereas foliation, refraction cleavage and
lineation data are mainly from the Boumalne and Imiter Lower Complex inliers [47,48]. The dotted red
lines in the bedding stereoplot highlight zones where data are locally reoriented because of drag fault
zones and disturbance upon pluton emplacement.
113
Minerals 2018, 8, 592
been only observed in the eastern part of the Jbel Saghro, in the Boumalne and Imiter areas. Indeed,
these fabrics (Figures 3 and 4b) remain difficult to observe except in the hornfels zone surrounding the
diorite and granodiorite intrusions described above (Figure 4b). Foliations show a constant ENE strike
and dip toward the NW (N075◦ E 60◦ N, Figure 3d), except when reoriented in fault drag areas (i.e.,
NW–SE strikes) or when refracted (NE–SW strikes). Such foliation is axial planar and, thus, is generally
parallel to the bedding of the metaturbiditic rocks when localised along the limbs of the first-order tight
folds. It is called herein S0–1 . A S2 cleavage refraction (Figure 4c) may also occur in specific areas, that
define an obliquity with respect to the S0–1 . It is caused by localised shear developed in incompetent
layers usually associated with the drag fold structures. Stretching and mineral lineations (L1 ) are
always carried by the S0–1 foliation. They are discrete and their orientation (Figure 3e) also appear
fairly constant with a rather low dispersion from NW–SE to N–S trending directions (average: N170◦ E
55◦ N). According to mineralogic and micro-structural evidences, at least 2 distinct metamorphic
assemblages are observed but appear to be related to the same tectonic event defined as a hornfels
assemblage and a regional chlorite to amphibole assemblage. (i) The hornfels zone is observed in the
contact metamorphic aureoles caused by the diorite and granodiorite intrusions within the Cryogenian
greywackes [48,121,130]. It consists of a spotted phyllite zone in which foliation (phyllitic cleavage to
gneissic foliation) and lineation are easily discernable. Rocks are mostly characterised by K-feldspar,
muscovite, biotite and small spots of retrogressed andalusite and/or cordierite (Figure 4b). The latter
frequently highlights the lineation and, therefore, suggests a probable relationship between pluton
emplacement and the D1 tectonics. Garnet-amphibole assemblage and spessartine occurrences are
associated with skarns and skarnoids as reported by Benziane [130] and Tuduri [47]. (ii) When
observable far from intrusions, foliation (slaty to phyllitic cleavage) is defined by a planar-linear fabric
mainly formed by phyllosilicates such as chlorite, sericite to biotite [48]. In both assemblages, rolling
structures and S-C fabrics (Figure 4d) define unambiguous south- to southeast-verging noncoaxial
shear criteria, parallel to the average lineation trending direction. These structures indicate reverse
sense top-to-the-south deformation in the Boumalne and Imiter areas [47,48,77,82,121,123].
ȱ
Figure 4. (a) Drag fold from the Qal’at Mgouna metagreywacke inlier. (b) Regional S0–1 metamorphic
foliation and related lineation marked by elongate contact metamorphic minerals that affect the
metagreywacke sequence. (c) S2 cleavage refraction, from the Imiter metagreywacke inlier, consistent
with a dextral shearing. (d) Microphotograph showing syn- to late kinematic metamorphic mineral,
a probable cordierite (Crd) showing rolling structure consistent with a top-to-the-south shearing sense.
Note the asymmetric tails composed of biotite (Bt), plane polar light.
114
Minerals 2018, 8, 592
The D1 deformation is also well-expressed along the large-scale N070–090◦ E trending strike slip
faults, about 100 to 150 km long that occur in the northern and central part of the Jbel Saghro (Figures 2
and 3a,b). On both sides of these shear zones, deformation appears more intense especially in the
slaty to phyllitic cleavage domains where the S2 refraction cleavage oriented N040–050◦ E 80◦ N is
formed (Figure 4c) synchronously with drag-folds (Figure 4a). These structures are consistent with
a dextral sense of shearing [47]. Where plutons are emplaced within shear zones, they develop both
magmatic foliation and lineation (e.g., the Igoudrane pluton). There, the foliation is parallel to the
fault (N075–090◦ E 80◦ N) and the lineation roughly horizontal. All structures linked with the D1 events
and the geometry of the associated intrusions indicate that this first Pan-African event developed in
response to a NW–SE to WNW–ESE trending shortening.
3.2. The Upper Complex or the Inception of a Silicic Large Igneous Province
3.2.1. Generalities
Rocks of the Upper Complex cover nearly 80% of the surface of the Jbel Saghro (Figure 2).
This complex, only affected by very low-grade metamorphism and weak deformation, consists of
thick and regionally extensive felsic volcaniclastic sequences that are non-conformably above the more
deformed and metamorphosed rocks of the Lower Complex [29,30,47,112,131–133]. Related dykes
and plutons intrude both the rocks of the Lower and Upper Complexes. These rocks belong to the
Ouarzazate Supergroup that includes the lower Mançour Group and the upper Imlas Group [5,107].
Plutonic suites are attributed to the Tanghourt Suite. These plutonic and volcanic rocks were emplaced
between ca. 575 and 540 Ma [29,51,74,78,106–108,112,134,135]. Most of the volcanic rocks are ash-flow
tuffs, felsic lavas, resedimented volcaniclastic deposits with some andesitic lavas and rare mafic
intrusions [29,30,47,112,133] that cover an area of approximately 2000 km2 and reach a maximum
thickness of 1000–1500 m.
115
Minerals 2018, 8, 592
ȱ
Figure 5. Detailed geologic map of the Qal’at MGouna district showing the extra- and intra-caldera
rock units (from Benharref [132], Derré and Lécolle [113] and Tuduri [47]).
All the data presented above argue for the existence of a collapse caldera structure as shown on
Figure 5 [136–139]. The peripheral structure of the ash-flow caldera where ring dykes occur is herein
interpretated as having accommodated both subsidence (caldera collapse and intra-caldera sequences
deposition) as well as subsequent uplift and tilting (magmatic resurgence and ring dykes injections).
The geometry given by the NW–SE trending structural limit favours an elliptic shape rather than a
sub-circular one for the caldera [41,140,141]. The consequence of such a shape will be discussed further.
From bottom to top and according to Tuduri [47], the intra-caldera volcaniclastic sequence is detailed
below (Figures 5–7):
(i). A lowermost pyroclastic layer consists of a 400–500 m thick, unwelded to slightly welded,
moderately crystal-rich dacitic lapilli tuff (Figure 6c,d and Figure 7a,b) interpreted as an ash-flow
deposit [142,143]. Internal stratification of the ash-flow tuff is crude, oriented N060–080◦ E 80◦ NW,
as highlighted by discrete layers which are either pumice-richer, lithic-richer or entirely devitrified
with spherulites. In a specific layer 40 m thick, greenish fibrous pumices displaying silicified
tubular micro-vesicles and a silky/fibrous fabric can reach up to 5 cm long. Lithic clasts up to
20 cm in size are common throughout and consist mainly of basement greywackes, lavas and
quartz-rich ignimbrite fragments. Phenocrysts are mostly broken plagioclase and K-feldspar,
in various ratios with minor amounts of chloritised ferro-magnesian crystals (biotite and probable
amphibole) and very scarce quartz.
116
Minerals 2018, 8, 592
ȱ
Figure 6. (a) View of the structural limit of the caldera showing the contact between the intra-caldera
sequence, rhyolite ring dykes and basement, Taghia area (see location on Figure 5). (b) Detailed view
of the structural limit showing the ring dykes and the unconformity between the basement to the south
and the intra-caldera sequences to the north. (c) Microphotograph of the lower intra-caldera ash-flow
tuff. Slight compaction and welding of shards are characteristics of this tuff, Plane Polar Light. (d) View
towards the east showing the relationships between the lower and upper intra-caldera tuffs and
interbedded sedimentary ponded rock sequences Taghia area (see location on Figure 5). (e) Close-up
view of the upper intra-caldera tuff characterised by compacted fiammes (flattened pumices) and lithic
fragments of K-feldspar-rich granite. (f) In thin section, the upper ignimbrite shows strong compaction
and welding of shards, Plane Polar Light. Bt: biotite, Kfs: K-feldspar, Pl: plagioclase, Qtz: quartz.
117
Minerals 2018, 8, 592
(ii). Above the ash-flow unit lies a ca. 200 m thick volcano-sedimentary (epiclastic) unit with very thin
bedding (Figures 6d and 7a,b). The lower part (100 m thick) is made of layered tuffaceous breccias
containing ignimbrites fragments. The upper part consists of laminated reddish and greenish
mudstones and sandstone oriented N070◦ E 70◦ NW. When preserved from important silicification,
the identifiable components are microscopic broken crystals and lithic fragments. Beds are broadly
continuous laterally, being only sometimes disrupted by syn-sedimentary normal faults and
slump-like structures. Faults are roughly oriented NW–SE. Fluid escape textures are common
and allow assessment of the polarity of the intra-caldera sequence. All sedimentological features
argue for a subaqueous emplacement, at least for the upper part of the epiclastic unit. In our
model, and as no marine sediments have been hitherto recognised in the entire Jbel Saghro in the
Ediacaran formations, such subaqueous environment may be reasonably related to a caldera lake.
(iii). Above the volcano-sedimentary unit lies a ca. 200–300 m thick crystal-rich rhyolitic ash-
and-lapilli tuff (Figures 6d and 7a–c). Plastic deformation due to significant compaction is
evidenced by reddish flattened pumices (Figure 6e). Glass shards and broken phenocrysts are
visible under the microscope (Figure 6f). Phenocrysts are quartz, plagioclase, K-feldspar, and
scarce amounts of chloritised Fe–Mg minerals (biotite and amphibole). Up to 2 m-sized lithic
clasts are abundant, especially in the basal part, and consist of metagreywacke, ash-flow tuff
fragments, jasperoids and monzogranite (Figure 6e). This voluminous quartz- and pumice-rich
unit is readily interpreted here as a welded ignimbrite.
(iv). The top of the sequence is dominated by massive andesite lava flows and epiclastic polylithologic
breccias (Figures 5, 6d and 7a,b). The bottom of this whole sequence is intruded by a monzogranite
porphyry (Figures 5 and 7a,b). Plugs of similar porphyry facies also locally intrude the upper
parts of the sequence. Such porphyry is here interpreted as a resurgent pluton [137,144] that tilted
the intra-caldera sequence upon emplacement [47]. Because of the tilting, the thickness of the
intra-caldera sequence is exposed over 1500–2000 m, of which 800–1000 m consists of ash-flow
tuffs and epiclastic rocks. To the north, the intra-caldera sequence disappears beneath the young
sedimentary rocks of the Dadès valley (Figure 5).
Volcanic and pyroclastic rocks also occur outside the caldera structure to the south and west.
They form a broadly stratified pile up to 500 m thick with moderate dips toward the W–NW (Figure 5).
In the vicinity of the Awrir-n-Tamgalount (Figure 5), the extra-caldera sequences are made up of two
units overlain by the Tamgalount tuff (Figure 7d,e). The lower unit is dominated by hundreds of
meters of well bedded, normally graded crystal-rich sandstones and siltstones (Figure 7e). Under the
microscope some of the silt-sized layers are formed of formerly vitric, now devitrified, material and
might be primary ash fall deposit. The stratified lower unit dips towards the west at variable angles
(Figure 5), perhaps due to palaeo-topography effects and(or) syn-tectonic deposition. The upper unit
is dominated by rhyo-dacitic lavas that display distinctive spherulitic devitrification microtexture and
pilloïd texture in the field that suggests emplacement under water. The Tamgalount tuff (Figures 5
and 7d,e) is a porphyritic porphyritic rhyo-dacitic ash-flow tuff (ca. 25% phenocryst) with eutaxitic
texture. Broken phenocrysts consist of quartz, plagioclase, K-feldspar and chloritised Fe-Mg minerals.
Lithic fragments are abundant and composed of greywackes, ash-flow tuff fragments and jasperoids.
all extra-caldera units lie unconformably on the Lower Complex and are pervasively affected to various
degrees by hydrothermal alteration and/or silicification. The importance of this unconformity will be
discussed later.
Numerous intrusions are related to the Upper Complex (Figures 2 and 5). They have been
mapped according to their mineralogy, texture and geochemical features. Two types are distinguished:
(i) gabbros and biotite- and amphibole-rich, pink-coloured coarse-grained granites (monzo- to
syenogranites), which are calc-alkaline to highly potassic, abd coeval porphyries emplaced at shallower
levels (e.g., the intra-caldera resurgent granite and its apophyses); and (ii) Si-rich alkali (K-rich,
i.e., shoshonitic in composition) granites and related aplitic bodies (sills, dykes), which frequently
appear as late magmatic events in the Upper Complex history. The pink mozogranites contain
118
Minerals 2018, 8, 592
quartz, albite-oligoclase, Fe-edenite, annite, K-feldspar and accessory minerals such as thorite, zircon,
allanite, apatite, magnetite, sulphides and W and Mo-rich minerals. By analogy with the Isk-n-Alla
monzogranite in the central part of the Jbel Saghro (Figure 2), they may have been emplaced around
555 Ma [29,106]. The alkali granites are mainly composed of quartz, albite and K-feldspar displaying
granophyric intergrowths. Accessory minerals include tourmaline (fluor-schorl) and metamict
zircon [47]. They are associated with late N–S rhyolitic dykes that also display shoshonitic compositions
(Figure 5). We assume they may have emplaced later, between 550 and 530–520 Ma [74,107,108].
ȱ
Figure 7. Simplified cross section and graphic logs of the Qal’at Mgouna volcanic sequences according
to Tuduri [47]. (a) Schematic N–S cross section across the Qal’at MGouna ash-flow caldera. Summary
stratigraphic sections for the (b) Taghia and (c) Awjja-n-Wizargane intra-caldera sequences, and
extra-caldera sequences from (d) Awrir-n-Tamgalount and (e) Tawrirt-n-Cwalh (See location on
Figure 5).
119
Minerals 2018, 8, 592
According to the relative and absolute chronology of both volcanic and plutonic rocks in both
the Jbel Saghro and in the Qal’at MGouna area, three main ignimbrites flare-ups are herein evidenced.
The earliest flare-up corresponds to the lower intra-caldera ash-flow tuff emplacement. It is mostly
dacitic and may coeval with granodiorite plutons emplaced around 575 Ma [29,74]. Ignimbrites from
the Oued Dar’a caldera described by Walsh et al. [29] are herein interpreted as belonging to this earliest
flare-up. Then, the emplacement of the Tamgalount ash-flow tuff with rhyo-dacitic affinities, may
correspond to a second high-volume magmatic event emplaced around 565 Ma [78,107]. Such an
event would have produced similar tuffs that are coeval with the huge rhyo-dacitic dyke swarm [29]
observed in the western and southern parts of the Jbel Saghro (Figure 2). The later ignimbrite flare-up
corresponds to the upper intra-caldera rhyolitic ash-flow tuff and to the numerous rhyolitic lava
flows and domes reported in the literature [29,74,106], coeval with pink monzogranite plutons around
555 Ma.
120
Minerals 2018, 8, 592
121
Minerals 2018, 8, 592
ȱ
Figure 8. Main features of the Taghassa intrusion-related gold deposit (IRG) [47,48]. (a) General
map and stereoplots of structural orientation data of the Au–Ag Thaghassa intrusion-related gold
deposit. These reveal the high density of veins developed north of the 575–560 Ma Ikniwn granodiorite.
(b) Interpretative sketch illustrating the magmatic-hydrothermal model that involved a progressive
and continuous tectonic event including the aplo-pegmatitic dykes and sills emplacement, then
the intermediate veins and the hydrothermal and gold-rich striped quartz veins. (c) Pegmatite dyke
122
Minerals 2018, 8, 592
showing a dextral pull-apart geometry. (d) N120◦ E trending intermediate veins filled by quartz,
muscovite and feldspar assemblage, cross polar light. (e) Macrostructure illustrating the gold-bearing
quartz vein stage. The layering texture defined the striped aspect of veins. (f) Microtextural
characteristics of gold-bearing quartz veins. The internal texture shows elongate quartz grains with
obliquity with respect to vein walls suggesting a dextral shearing, cross polar light.
4.2. The Qal’at Mgouna Au–Ag (Cu, Mo, Bi, Te) District
The Qal’at Mgouna district is composed of three main exploration projects: the Isamlal,
Talat-n-Tabarought and Tawrirt-n-Cwalh districts, all of which located outside of the caldera structure
(Figures 5 and 9). Based on mineralogical, chemical, textural and structural constraints, two distinct
ore deposit types have been identified: an older porphyry ore deposit on which a younger epithermal
system is superimposed [47,113,153,154].
123
Minerals 2018, 8, 592
ȱ
Figure 9. Main features of the Qal’at Mgouna deposit types [47,158]. (a) Detailed map of the
Au(–Cu–Mo–Ag–Te–Bi) Qal’at Mgouna district (from Tuduri [47]). Note that the kinematics shown
by shear zones are consistent with a WNW–ESE direction of shortening. (b) Kriging interpolation
revealed that gold anomalies are correlated with both the quartz stockwork and the NW–SE faulted
corridor in the Isamal porphyry deposit. The red colour is indicative of the highest Au grades [158].
(c) Stereoplots of structural orientations data for: the quartz stockwork related to the porphyry stage
(mean orientation N120◦ E); the aplitic dykes and sills related to the alkali–syeno–granite stocks (mean
124
Minerals 2018, 8, 592
orientation N018◦ E); and the adularia-specularite-quartz veins from the epithermal stage (mean
orientation N008◦ E). (d) Quartz stockwork from the Isamlal porphyry Au(–Cu–Mo) system. (e) Typical
potassic alteration (pink coloured zones) of a porphyritic granodiorite (the Isamlal porphyry
Au(–Cu–Mo) system). (f) Typical alteration and vein development of the epithermal stage: pervasive
tourmalinites are cut by quartz and adularia-rich veins in the vicinity of the Tawrirt-n-Cwalh deposit,
plane polar light. (g) Economic paragenesis of the epithermal stage characterised by Au-Ag tellurides,
electrum and Bi-telluride veinlets within pyrite; Tawrirt-n-Cwalh deposit, SEM back scattered picture.
Chalcopyrite and gold were also observed in multiphase inclusions. Multiphase inclusions
have a high though variable salinity (30 to 45 wt. % NaCl equiv.) and are characterised
through homogenization by halite-disappearance. The large range of homogenization temperatures
(160–460 ◦ C) combined with a zoned potassic to propylitic alteration, stockwork structures and with
an Au(–Cu–Mo) paragenesis is interpretated as characteristic of Au(–Cu–Mo) porphyry environments.
This porphyry Au(–Cu–Mo) system is herein described as a vein-dominated deposit (stockwork)
that is consistent with the emplacement of a porphyry stock, then exsolution and cooling of a
magmatic-derived hydrothermal fluid. The overall system appears as mostly controlled by a ca.
WNW–ESE trending direction. Indeed, this pattern suggests that the stockwork structure both reflects
the magmatic stress associated with the porphyry emplacement and fluid exsolution, and also, a ca.
NNE–SSW-oriented minimum principal stress (i.e., extensional direction) associated with a regional
deformation that may be consistent with a WNW–ESE shortening direction although no clear tectonic
regime has been proposed for the hydrothermal stage.
125
Minerals 2018, 8, 592
4.3. The Zone des Dykes Intermediate Sulfidation Epithermal Au-Base Metal Deposit
The Zone des Dykes, also known as the Issarfane area, is located in the western part of the
Jbel Saghro inlier (Figure 2) in the vicinity of a huge N–S rhyolitic dyke swarm emplaced around
565 Ma [29]. The Zone des Dykes ore district consists of quartz veins systems hosted by two ash-flow
tuff units belonging to the Upper Complex (Figure 10). Therein, three mineralised systems, called
the F1, F5 and Bou Issarfane structures, respectively (Figure10a), are identified [47]. The F1 structure
consists of a 2 km long and 2 m width vein system that trends N180–160◦ E and dips about 50–60◦ to
the east. The vein system shows several step-over zones showing a left lateral pull-apart geometry
(Figure 10b) with a faint vertical component. The F5 structure also consists of a vein system, 1 km
long and 2 m width, that is roughly oriented N080◦ E 60–70◦ S and is characterised by right-lateral
shear structures (Figure 10c). The F1 and F5 structures are both interpreted as developed as conjugate
pairs (Figure 10d). Because cross-cutting relationships are observed, we interpret the N080◦ E direction
trend (i.e., F5) as the dominant direction. The F1 shear structures is herein interpreted as emplaced
along a pre-existing NNW–SSE fracture analogous to the ones occuring between the Bouskour and
Issarfane areas (Figure 2). The F5 structure is also emplaced along an important pre-existing fracture
set that corresponds to a main ENE–WSW regional fault. This may explain why the F1 and F5
shear fractures are almost orthogonal yet conjugated, but also why the F1 pull-apart structures are
always brecciated. The Bou Issarfane structure has a ca. N–S orientation like the F1 system, but
unlikely lies at dip angles of 20–40◦ to the east. It consists of a 1.5 km long and 5 to 10 m thick
silicified breccia system hosted by ignimbrites and rhyolitic tuff that is affected by E–W brittle faults
(Figure 10a). While an unsilicified rhyodacitic lava flow occurs at the hanging wall, the footwall
is made up of a 5–10 m thick anastomosed quartz stockwork. All veins are mostly filled in by
quartz (Figure 10e,g–i) with scarce amount of adularia, sericite, chlorite, calcite and rare fluorite [47].
Sulphides are also common and mainly consist of arsenian pyrite (Figure 10f). Chalcopyrite, sphalerite,
Pb–Cu–Bi assemblages (aikinite group) and electrum are accessory minerals and appear as inclusions
within the As-rich pyrites (Figure 10f). Chlorite has a pycnochlorite composition and a Fe/Fe + Mg
ratio close to 0.48 which is consistent with temperature of crystallisation bracketed between 210 and
280 ◦ C [47]. All veins are characterised by internal textures typical of epithermal deposits according to
Dong et al. [161]. The most representative quartz textures are those showing a partial replacement of a
silica gel precursor characterised by colloform and moss texture (Figure 10g). Such textures are specific
of siliceous sinters in active geothermal systems [161]. Ghost-bladed calcite textures [162] are also
observed (Figure 10h,i). The occurrence of platy calcite (ghost-bladed calcite) demonstrates that boiling
processes were active during vein formation [163–165]. Veins also show complex texture reflecting
several stages of crystallisation, replacement and re-crystallisation occur. Within this complex process
of vein formation, Tuduri [47] suggests that sulphides crystallised after the second stage of quartz
replacement while electrum is mostly located within fissures of the pyrite. While the structural control
of the F1 and F5 structures is clearly evidenced, the emplacement of the Bou Issarfane breccia remains
126
Minerals 2018, 8, 592
unclear and needs to be discussed. Fragments from the Bou Issarfane breccia are silicified, sub-angular
and of dimensions lower than 1 cm. Their probable volcanic origin suggest minor distance of transport.
The matrix is highly silicified and seems composed by detritus of rock fragments (host rocks), and (or)
by comminuted gangue minerals.
ȱ
Figure 10. Main features of the Zone des Dykes deposits [47]. (a) General map of the Zone des Dykes
Au-Ag deposit at the crossing between the F1 and F5 vein systems and Bou Issarfane area (after
Tuduri [47]). (b) Pull-apart texture indicative of a left-lateral shearing movement along the F1 structure.
127
Minerals 2018, 8, 592
The filling is composed by quartz. (c) Pull-apart geometry of the F5 structure mainly filled by
quartz and formed by dextral kinematics. (d) Global kinematic interpretation for the F1 and F5
structures integrating all the structural features observed in the field. (e) Microphotograph showing
internal microtexture of the F1 structure characterized by a saccharoidal layout of quartz grains, cross
polar light. (f) Typical gold-rich paragenesis of the F5 structure. Py: pyrite, Sp: sphalerite, Ccp:
chalcopyrite. (g) Microphotograph of silica spheroid aggregates displaying moss texture, cross polar
light. (h) Microphotograph of parallel ghost bladed calcites replaced by quartz, cross planar light and
(i) plane polarized light microscopy.
The origin of this breccia may be therefore compared with phreatic or phreatomagmatic breccia
pipes, although we cannot exclude a tectonic origin corresponding to a silicified cataclasite. According
to salinity and homogenization temperatures, two type of fluids are assumed to be at the origin of
the mineralised system [166]. The Bou Issarfane stockwork is characterised by primary multiphase
fluid inclusions composed of liquid, vapour and halite cubes. Values obtained using the FIA concept
indicate homogenization temperatures between 210 and 230 ◦ C and moderate salinities (14–17 wt. %
eq. NaCl). Secondary inclusions have lower homogenization temperatures between 130 and 180 ◦ C.
The F1–F5 veins consist of primary multiphase inclusions composed of liquid, vapour and halite cube
with CaCl2 assemblages.
Homogenization temperatures are bracketed between 160 and 180◦ C. Salinities are variable from
6% to 29% (wt. % NaCl + CaCl2 equiv.) and may reflect the effects of boiling processes. Such textural,
mineralogical and fluid characteristics suggest this hydrothermal Au–Ag(–Cu–Zn–Pb–Bi) system
is comparable with intermediate sulfidation epithermal deposits as the ones reported in the Sierra
Madre Occidental in Mexico [25,26,167]. The age of this ore deposit is still unknown. Considering
the N–S trending direction of the F1 and Bou Issarfane structures, and the N–S trending direction of
the rhyolitic dyke swarm, the mineralisation may have been formed coevally with this volcanic pulse
around 565 Ma, i.e., after the emplacement of the ash-flow tuffs, that host the mineralisation dated,
at 574 ± 7 and 571 ± 5, respectively [29]. However, we cannot exclude that the mineralisation is related
to a later magmatic-hydrothermal period (e.g., the 550 or 530–520 Ma event) whilst no study has so far
evidenced such later activity in the Bou Isserfane area.
128
Minerals 2018, 8, 592
ȱ
Figure 11. Main features of the world-class Ag–Hg Imiter mining district [47,102]. (a) General map
of the giant Ag–Hg Imiter mine (after Leistel and Qadrouci [127] and Tuduri [47]) and (b) stereoplots
of structural orientations and interpretative block diagram explaining the formation of the main
ore-bearing vein system. Thrusts are formed in core of transpressive push-up structures. Note
that the mineralised structures were everywhere controlled by a ESE–WNW direction of shortening.
(c) Pull-apart and tension-gashes structures of the economic stage filled by geodic quartz and formed
129
Minerals 2018, 8, 592
during dextral kinematics, F0 North vein systems, view realised towards the top of the mining gallery,
Imiter I. (d) Pull-apart geometry of the economic stage indicative of a reverse shearing towards the
north and showing void formations, R6 structure, Imiter II. (e) Pull-apart texture of the F0 South vein
systems, thrusting towards the NW–NNW. The filling is composed by quartz (economic stage) and
scarce pink dolomite in core of pull-apart texture. (f) Typical paragenesis of the economic stage 1
composed by quartz veins and huge concentration of Ag–Hg alloys, Imiter I, F0 south. (g) Tension
gashes and left lateral pull-apart structures filled with pink dolomite of the stage 2, F0 structure, Imiter I.
(h) Typical features of the stage 2 pink dolomite stage with large patches of Ag-rich galena, F0 structure,
Imiter I. Since photographs c and h were taken towards the top of exploration galleries, kinematics
interpretation must be inverted.
Tuduri et al. [102] further demonstrated that a stage 2 reactivated the transpression-related
structures in the opposite sense, and developed normal left-lateral motions associated with massive
pink dolomite crystallisation, as well as prismatic quartz and variable amounts of Ag-rich galena
and sphalerite, pyrite, chalcopyrite, arsenopyrite and freibergite (Figure 11g,h). Note that these
two economic stages were preceded by a barren quartz vein network stage associated with sericite,
illite-chlorite and base-metal sulphide minerals such as pyrite, galena, sphalerite with chalcopyrite
exsolutions [45,47,51,127,169–171].
The main driving mechanism for silver ore deposition is assumed to be the dilution of ore-bearing
fluids that were CaCl2 -dominated. Values obtained using the FIA concept [172,173] point to a general
temperature decrease from stage 1 (280–100 ◦ C) to stage 2 (110–60 ◦ C). Note that the deepest levels
of the mine workings (−220 m below the surface) record temperature in excess of 60 ◦ C (i.e., lowest
temperatures >160 ◦ C) with respect to the shallow levels (−100 m) where the lowest temperatures are
around 100◦ C. During stage 1, fluid salinities are moderate to high (8.4 to 26.1 wt. % NaCl + CaCl2
equiv.), whereas they are very high when stage 2 dolomite precipitates (24.6 to 30 wt. % NaCl + CaCl2
equiv.). Such value ranges are in agreement with data published by previous authors whether or not
they used the FIA concept [41,45,101,171]. At shallower levels, additional supergene enrichment has
been responsible for massive formation of native silver (1500 g/t Ag) associated with cerusite and
mimetite [164].
Work in progress shows that Ag–Hg sulfohalides could also be related to the supergene
processes [174]. Two opposing ore-forming models are strongly debated at Imiter. Some authors
taking into account halogen composition of fluid inclusions, stable (C, O, S) and radiogenic (Pb,
Re/O) isotope data together with noble gas (He) isotope compositions, suggest that the deposit is
consistent with an epithermal model related to the felsic volcanic event at the Precambrian–Cambrian
transition [45,47,51,101,127,170,175]. In that way, the huge Ag–Hg deposit would be comparable with
the ones from the Mexican Sierra Madre Occidental Ag–Pb–Zn–Au belt, and should be considered as
an intermediate sulfidation epithermal deposit [167,175]. On the other hand, a lithogene model [176]
has been alternatively proposed in which fluids, according to laser ablation inductively coupled
plasma-mass spectrometry (LA-ICP-MS) data on fluid inclusions, halogen signatures, and stable
isotopes (H, C, O), are the products of diagenetic brine–evaporite interactions within a sedimentary
basin [41]. The ore deposit might also be the result of basin inversion that expelled deep Ag-rich
brines during, or at the end, of the Palaeozoic orogeny [41,175]. Recent 40 Ar/39 Ar age measured at
255 ± 3 Ma on adularia from stage 1 quartz vein supports the late Palaeozoic brine model [38].
5. Discussion
130
Minerals 2018, 8, 592
Figure 12. Interpretative three-phase model (a–c) explaining the tectono-magmatic evolution as well as
the Lower and Upper Complexes definition of the Jbel Saghro and the formation of the ore-bearing
vein systems of the Zone des Dykes, Qal’at MGouna, Thaghassa and Imiter districts. The size of the
blue arrow relates with the inferred intensity of regional stress. See text for explanation. Note that the
Qal’at MGouna porphyry deposit may belong to either the late stage of event 1 or the earlier stage of
event 2 as the related porphyry stock emplaced at the transition between the two at 576 ± 5 Ma [29].
131
Minerals 2018, 8, 592
Previous interpretations tried to link each type of structures or pluton emplacement to one distinct
tectonic event, increasing the complexity. From this study, it appears more appropriate to interpret
all these features by the occurrence of a single and unique D1 tectono-magmatic event. According
to Gasquet et al. [83], as basins infilling with greywacke sequences was active until the onset of the
Ediacaran period between 630 and 610 Ma; we suggest that the D1 deformation probably occurred
coevally with the syn-tectonic calc-alkaline magmatic occurrences, those being dated from ca. 615 Ma
until ca. 575–565 Ma. This upper 575–565 Ma limit for the D1 age is an important issue and will be
further discussed below.
5.1.3. About the Transition between the Two Complexes and the D1 and D2 Tectonics
In the Eastern Anti-Atlas, the transition between the Lower and Upper Complexes is characterised
by a significant increase of the magmatic addition rate [178] mostly evidenced by the Ediacaran volcanic
flare-ups and by a change in the tectonic regime. However, such a transition appears as diachronous
whether we consider the start of the volcanic activity or the change in tectonic regime.
132
Minerals 2018, 8, 592
The key evidence highlighting such a transition is the existence of an angular unconformity
between the Lower and Upper Complex units [29,47,74,78]. However, the age of the transition remains
elusive and debated. If we consider the available ages earlier volcanics of the Upper Complex that
lie above the unconformity, the transition might occur between 580 and 570 Ma taking into account
the error bars of the radiometric ages. This has been well described in the Western Anti-Atlas on the
Tizgui geological map [29,179]. In the Central Anti-Atlas, Blein et al. [78] also suggest that most of
the regional deformation observed there was completed by ca. 580 Ma and followed by an important
erosional phase prior to the deposition of the Upper Complex volcaniclastic sequences. The 575 Ma
age, therefore, represents a mean value.
Dating the deformations and thus the change in the tectonic regime from D1 to D2 is matter
of more confusion. Indeed, the transition between the two complexes is estimated at ca. 575 Ma
with the incipient volcanic activity and the occurrence of a strong angular uncorformity. By contrast,
the transition between the two deformation events would be a little more recent and would have
occurred around 565 Ma. The age assigned to the D1 deformation appears thus mostly dependent on
the radiometric ages obtained on the syntectonic plutons, and the issue of what plutons are syntectonic
or not in the Eastern Anti-Atlas is clearly to be better assessed from field data. Currently, numerous
plutons emplaced between ca. 575 and 565 Ma have been interpreted as syntectonic as they develop
a coherent and homogenous ductile deformation in contact aureoles [48,105,106,121,123]. Given the
error bars on the ages, we can actually question on their belonging to the earlier Upper Complex
structuration, or to the later stage of the Lower Complex and D1 deformation. As a matter of fact,
discussions do exist about the possible intrusive character of some of these plutons in the lower part of
the Upper Complex (e.g., The Igoudrane, Taouzzakt and Ikniwn plutons, Figures 2 and 3), while they
may display an erosive roof on which lies Upper Complex volcaniclastics (at least the upper part of the
Upper Complex). In addition, we have shown above that magmatic mushes at the origin of plutons
emplaced around 575 and 565 Ma may be at the origin of the former ignimbritic flare-ups.
In any case, we assume that between 575 and 565 Ma, syn-D1 plutons have developed in the Lower
Complex metagreywackes, both contact metamorphism and ductile deformation that we interpret
as the ongoing tectono-magmatic evolution between the Lower and Upper Complex structuration
as suggested by Tuduri et al. [48]. Possibly, the S2 cleavage refraction that developed after the main
first-order folding event may be related to this transitional stage. It also remains difficult to assess
whether some plutons were intruded within folds (e.g., the Bou Teglimt granodiorite in core of the
Imiter inlier anticline, Figure 3a) and thus after the folding event (i.e., plutons would belong to the
earlier stage of the Upper Complex) or if their emplacement played a part in the anticline structuring
(i.e., plutons would belong to the Lower Complex tectono-magmatic history). Consequently, taking
the important dioritic and granodioritic plutonism as systematically belonging to the Lower Complex
might be misleading. As well, it is not straightforward to math the change in the tectonic regime
(D1 -D2 transition) with the transition between the Lower and Upper Complexes.
Therefore, we herein suggest that the reported tectonic transition must have been an ongoing
process through the Lower-Upper Complex transition, given the deformation features (cleavage, weak
upright folding) observed in the volcaniclastic rocks of the lower part of the Upper Complex in the
western Saghro and Central Anti-Atlas [29,78]. Indeed, Blein et al. [78] recall that rapid variations in
thickness of the volcanic and volcaniclastic rocks of the Upper Complex suggest they were deposited
during active tectonics and on highly variable basement (i.e., Lower Complex units) topography.
Such variations in the topography may be due to the earlier transpressive stage but also to later
extensional-transtensional tectonics. In the rest of the Jbel Saghro no clear evidence of deformation
(except cleavage) has been described affecting the lower part of the Upper Complex rocks. Note that
the “Série molassique du Dadès” described by Walsh et al. [29] in the northern Qal’at Mgouna area,
as being bedded, deformed and weakly metamorphosed in lower greenschist facies corresponds
to the tilted intra-caldera volcanic sequence which underwent propylitic hydrothermal alteration.
This formation herein belongs to the Upper Complex.
133
Minerals 2018, 8, 592
The preservation of the same shortening direction and tectonic style (i.e., strike slip dominated)
further suggests that D2 is a continuation of D1 even though D1 was associated with transpressive
tectonics and D2 linked with transtension to extension. Eventually, the transition between the D1
and the D2 deformation regimes occurred, while the shortening direction (parallel to Z strain axis)
remained roughly constant (i.e., NWSE to WNW–ESE). However, one can note a shift in the two other
strain directions showing a decrease of the vertical extension that becomes horizontal. This explains the
numerous extensional features developed during the D2 stage (i.e., calderas formation and multiple
ore concentrations within open structures). We note that the transition from D1 to D2 must be achieved
prior the emplacement of the Thagassa IRGD around 565 Ma [48]. This transition is supposed to be
progressive and the earliest mineralisation stages, such as the ones emplaced in the Thaghassa area,
are the witnesses of this transition.
Therefore, we assume that the D1 deformation might affect the lower units of the Upper Complex.
Future works may focuse on testing this assumption. The contrasting structural levels that exist
between the Lower Complex and Upper Complex units may argue against such a continuum. However,
the combination of exhumation processes along strike-slip fault systems and of denudation history,
integrating erosion processes [180–182] provides the first elements of an answer. Moreover, ignimbrite
flare-ups and caldera formation are assumed to have been occurred rapidly as catastrophic events,
contributing unconformities between the volcaniclastic rocks of the Upper Complex and the Lower
Complex units.
134
Table 1. Synthetic table summarizing the magmatic-hydrothermal features of main ore deposits of the Saghro area.
135
albite, andalusite, paragonite and kaolinite [189] and microthermometric data results [41,45,48,155,159,160,166,170–173]. See Tuduri [47] for details.
Minerals 2018, 8, 592
• The first stage shows strong magmatic influence. It is characterised by emplacement of porphyry
stocks, aplite dykes and sills at high temperatures from 400 ◦ C up to 600 ◦ C. At Thaghassa, this
stage was responsible for the partial melting of the metagreywackes in response to the Ikniwn
granodiorite thermal effect and for the related genesis of leucocratic S-type haplogranitic sills.
In the Qal’at Mgouna district, this stage was responsible for the formation of stockscheider and
miarolitic cavities within sills and dykes.
• The second intermediate stage consists of magmatic-hydrothermal vein emplacement and
associated pervasive alteration. The persistence of the magmatic character is shown
by the occurrence of high-temperature alteration phases such as tourmaline, micas,
andalusite, apatite, K-feldspar with quartz. In the Qal’at Mgouna district, at Isamlal,
this stage can be compared with the classical potassic and magnetite alteration in some
porphyry type systems [190]. It is also marked by the wide pervasive development of
Al-silicate–Al-hydroxide–phosphate–muscovite–F-rich phlogopite and F-rich tourmaline
alteration, related to the late Si-rich alkali granites. K-feldspar, apatite, white mica along vein
rims are observed at the beginning of vein aperture at Thaghassa. Temperatures of formation are
bracketed between 250 and 500 ◦ C.
• The third stage is hydrothermal and formed at lower temperature (60 < T (◦ C) < 300) producing
gangue minerals except for the Thaghassa and Isamlal deposits where high-temperature minerals
were also formed (350–450 ◦ C). This stage end with the emplacement of economic ore.
The conditions of vein formation vary depending on their location in the Lower or Upper Complex
and are reflected by variations in mineralogy and internal texture. Such variations are basically due to
different structural levels of formation. Indeed, the geometry of the mineralised structures is controlled
by tectonics, hydrostatic pressure, effective vertical stress and volcano-related effects. The Thaghassa
prospect, hosted by the Lower Complex metagreywakes, exhibits texture and mineralogy indicative of
high-temperature conditions of formation. The Zone des Dykes vein system, entirely developed in
Upper Complex ash-flow tuffs, shows internal textures consistent with low-temperature epithermal
deposits. The Qal’at Mgouna deposits are developed at an intermediate structural level. Veins hosted
by the Lower Complex rocks are related to high-temperature formation (i.e., the Isamlal porphyry
Au(–Cu_Mo) deposit) whereas those formed at shallower levels reflect low-temperature conditions
(Talat-n-Tabarought and Tawrirt-n-Cwalh districts). Tuduri et al. [154] suggest that high temperature
systems emplaced at ca. 150–200 MPa, whereas the lower temperature systems were formed at a
lower depth (20–50 MPa). The regional variations clearly document the transition from magmatic
to hydrothermal conditions, i.e., from somewhat high temperature fluids (350 ◦ C and higher) at
Thaghassa and Isamlal, to lower temperature hydrothermal fluids (below 300 ◦ C) at the Zone des
Dykes and Imiter.
For all ore deposits described in this study, the Ag and Au economic concentrations (with Cu,
Zn, Pb) correlate with ore-forming fluids with moderate to high salinity. This is consistent with
transport by chloro-complexes and confirms the importance of brines in such ore formation. If brines
are frequent in the formation of porphyry copper deposit [183,191] and probably IRGDs [64,68], their
role in intermediate sulfidation epithermal systems remains a matter of debate. In large Mexican
epithermal silver deposits of intermediate sulfidation state, Wilkinson et al. [26] suggested that the
ore forming fluids, were injected into a geothermal circulation system in response to the ascent of
a magmatic intrusion. Such hydrothermal diapirs would be sourced from a stratified hyper-saline
brine reservoir, formed in response to incremental exsolution of magmatic fluid, and intense brine
condensation at depth, with halite precipitation, being stored above the source magma reservoirs [26,
192,193]. By contrast, Essarraj et al. [40,41,49] suggested that the ore-forming fluids were related to
deep-basinal sedimentary brines and that metals had no genetic relationship with Neoproterozoic
magmatism, on the basis of numerous deposits they investigated in the Eastern and Central Anti-Atlas.
They suggested that ore brines resulted from evaporatively concentrated seawater in Triassic basins
producing hot basinal brines, comparable with conditions for Mississippi Valley-Type (MVT) deposits.
136
Minerals 2018, 8, 592
We here suggest that all ore deposits described above are related to the Ediacaran
magmatic-hydrothermal complex emplaced from 575 to 530–520 Ma. The concept of stored hypersaline
brines is emphasised following Scott et al. [193]. Nevertheless, while a magmatic origin of such
stored brines is obvious, the origin for Ca-rich brines has yet to be defined. Though such Ca-rich
brines may derive from magmatic processes, they are also widespread in the deeper parts of many
sedimentary basins and involved in ore forming processes [194,195]. This is, pro parte, the reason why
Essarraj et al. [40,41,49] support a deep-basinal sedimentary brine model in the Anti-Atlas. Surprisingly,
they do not support an Ediacaran model for such a model but a more recent one that is Permo-Triasic
in age. Nevertheless, prior and during the early stage of development of the SLIP that characterise the
Upper Complex and the related ore deposits, periods that roughly belong to the Ediacaran Gaskiers
glaciation on the West African Craton and around [196,197], would have occurred between 595 and
565 Ma in the Anti-Atlas [198]. Such evidence may also point to a possible contribution of natural
cryogenic brines in the ore forming processes. Indeed, according to the Starinsky model of evolution
of a marine-cryogenic basin [199], these brines usually are seawater-derived and also enriched in Mg,
K, Na, Ca, Cl, SO4 and Br, even if they may have been heavily modified by fluid–rock interactions
and(or) dilution [200]. Future studies would have to address the possible role of Ediacaran glaciations
on the ore-forming processes in the Anti-Atlas.
The age of the mineralisations in the Jbel Saghro remains poorly constrained due to the lack of
absolute dating. Nevertheless, we can assume that the transtensional regime characteristic of the
Upper Complex provided extensional structures for melt and fluid circulations that are favourable for
magmatic-hydrothermal ore forming processes. This period of time (575–550 Ma and 550–520 Ma) was
suitable for ore emplacement as suggested above. We thus assume that most of the mineralised systems
may have been emplaced between ca. 575 and 520 Ma. Indeed, these ore deposits seem to have been
emplaced when tectonic regime changes at the transition between the Lower and Upper Complexes
and when the typical medium-K calc-alkaline arc magmatism became more abundant in between
580 and 570 Ma. Some porphyry Au(–Cu–Mo) deposit may have been emplaced earlier around
575 Ma. In the Sirwa mountains belonging to the Central Anti-Atlas, Cu–Mo ± Au mineralisations
associated with high-K calc-alkaline intrusions are also assumed to have been formed between
575 and 560 Ma [59,201]. Similar ages are also suggested in the vicinity of Bouskour where Re/Os
analyses on molybdenite related to a Cu-rich mineralised stage, yield a weighted average age of
574.9 ± 2.4 Ma [103]. Further, the Thaghassa deposit displays strong similarities with the earlier base
metal assemblage observed at Imiter and Tiouit [43,45,47,48,51]. Indeed, in the Tiouit gold deposit,
the ore body is closely associated with the Ikniwn granodiorite dated at 564 ± 6 [48]. At Imiter,
40 Ar/39 Ar dating on sericites related to the earlier base metal sulphide veins range from 577 ± 4 to
563 ± 5 Ma, in good agreement with the synchronous Taouzzakt granodiorite dated at 572 ± 5 Ma and
the Thaghassa model of emplacement (U-Pb radiometric ages on zircon, [48,51]). The intermediate
sulfidation epithermal Au(–Ag–Cu–Pb–Zn) deposit of the Zone des Dykes may have been emplaced
during lull periods soon after the emplacement of the second ignimbrite flare-up and related huge
dyke swarm in the western Saghro (around 560 Ma). Similarly, the huge intermediate sulfidation
epithermal Ag(–Hg–Pb–Zn) deposit at Imiter may have been emplaced following the third ignimbrite
flare-up around 550 Ma. This assumption adopts the age alreadyproposed for the Imiter mineralisation
on the basis of absolute dating of felsic volcanism at 550 Ma [45,51,101]. Lastly, we suggest that the
intermediate to low sulfidation epithermal Au–Ag–Te deposits were emplaced later when magmatism
became less abundant, more silicic and alkaline between 550 and 530–520 Ma [74,107,108]. This is
supported by the fact that some late rhyolitic dykes do not cross the Cambrian boundary while they
cut across mineralised structures, at Qal’at Mgouna for example (Figure 9a)
137
Minerals 2018, 8, 592
5.3. Implication of the Tectonic Regime Changes for the Late Neoproterozoic—Early Cambrian Geodynamic
Evolution and Ore Deposit Emplacement
In the light of our results, a geodynamic evolution model is proposed (Figure 13), showing the
spatial and temporal distribution of metal deposits in the Eastern Anti-Atlas that may be reasonably
extended to the whole Anti-Atlas. Considering the tectonic and metallogenetic framework at the
scale of the Anti-Atlas regional scale, we suggest a first-order influence of subduction dynamics
on the shift with time of metal concentrations and ore deposit types in the Anti-Atlas. In our
scenario (Figure 13a,b) and according to Walsh et al. [29], the 615–575 Ma period is characterised
by an Andean-type subduction of a wide continuous slab along the northern Gondwana margin (i.e.,
the currently northern side of South America and Africa). This long period of subduction (ca. 40 Myr)
probably occurs because the large slab width (>2000 km) increases the viscous resistance of the mantle
on the slab [202]. In that model, oceanic lithosphere subducts beneath the continental lithosphere that
progressively becomes thicker due to the presence of the West African Craton. Suction between ocean
and continent increases, favouring slab flattening and mantle wedge closure [203].
ȱ
Figure 13. Compiled lithospheric-scale reconstructions (a–e) of the Pan-African/Cadomian belt systems
from the West African Craton foreland to the Iapetus oceanic domain showing the possible progressive
slab retreat since ca. 575 Ma to 500 Ma as suggested in text. Zones of partial melting in both the
subcontinental lithospheric mantle (SCLM) and the lower crust, as well as the zones of storage and
transfer of melts are shown in red. Kinematic vectors related to the West African Craton drifting are
from Merdith et al. [204] using the GPlates software [205]. The palaeotectonic map reconstruction of
Gondwana (f) has been realised using the GPlates software and the global plate models with kinematic
continuity of Domeier [206].
Both slab flattening and probably relatively high convergence rate control the D1 deformation in
the overriding plate. A subduction vector (Figure 13b) oblique to the continental margin may explain
138
Minerals 2018, 8, 592
the overall transpression regime that characterises the Lower Complex. Indeed, the lack of nappes,
fold nappes, mylonitic fabrics and metamorphic gradient preclude a continental collision setting in
that period. Possibly, a volcanic arc accretion event may locally emphasize this deformation. Melt
generation remains limited given the restricted size of the mantle wedge while their ascent may occur
when regional or local extensional regime occurred (Figure 13a,b). Ore forming processes are limited in
this setting as they depend on the generation and ascent of fertile melts. However, melts produced and
stored at depth in the MASH (melting, assimilation, storage and homogenization) zone may become
more fertile in the following stages.
From 575 Ma, the ongoing evolution of the flattened subduction of oceanic beneath cratonic
lithospheres causes a dynamic push on the slab surface [203]. This occurs as the rate of wedge
closure increases, pushing the slab backward and initiating slab roll-back and high-volume magmatic
production (Figure 13c), ultimately leading to the first ignimbrite flare-up. Ongoing slab roll-back along
with tectonic regimes becoming more extensional initiate the second, and then the third ignimbrite
flare-ups (Figure 13d). Slab tearing and (or) breakoff provide important asthenospheric flow that
probably catalyse partial melting of both the asthenospheric mantle wedge and subcontinental
lithosphere mantle (SCLM). The latest alkali magmas are assumed to have been emplaced at an
extensional stage related to the Adoudounian rift in the Western Anti-Atlas [87,106] that corresponds
to a back-arc setting. Such ongoing extensional tectonics [206,207] will ultimately result in the opening
of the future Rheic Ocean (Figure 13e,f).
A collision then post-collision scenario cannot be applied to the Anti-Atlas region between 615 and
520 Ma, since subduction do not cease in our model but progressively migrate towards lower latitudes
(Figure 13). Eventually, such subduction dynamics determine the dominant stress regime in the
overriding plate, which influences the metals mobilization in the MASH zone and, their ascent through
the crust, and thus controls the distribution of resulting metal occurrences [208–212].
According to Tosdal and Richards [211], most of the mineralised structures are suggested to have
formed during the D2 -related transtensive regime caused by shortening in a WNW–ESE direction
and extension along the NNE–SSW direction (Figures 12c and 13c–e). In our interpretation, porphyry
Au(–Cu–Mo) and intrusion-related gold deposits are emplaced earlier than the first and(or) second
ignimbrite flare-ups (i.e., 575 and(or) 565 Ma). Intermediate sulfidation epithermal Au, Ag deposits
may be emplaced during lull periods after the second and (or) the third flare-ups (i.e., 560 and(or)
550 Ma). Compressive structures are indicated in the Zone des Dykes and Imiter districts, as the result
of likely a specific structures geometry with respect to the shortening direction. Intermediate to low
sulfidation epithermal Au–Ag–Te deposits are emplaced late and in relation with the felsic alkaline
magmatic stage (550–520 Ma).
We have stressed here the existence of a long period of magmatism, i.e., over a 95 Myr duration,
which is characterised by an increase in produced magmatic volumes, probably in response to
geodynamical controls, marked by a late magmatic paroxysm in the form of several ignimbrite
flare-ups over a shorter duration of ca. 25 Myr. Such long-lived magmatic activity is paralleled by a
tectonic progressive evolution from beginning within transpression conditions to transtension then
extension, allowing the mineralisations to take place.
At the regional scale, we may question a possible diachronism of the magmatic activity between
the western Bou Azzer and Siroua and the eastern Saghro inliers (Figure 2). Indeed, the main volcanic
event occurred between 580 and 560 Ma in Bou Azzer, Siroua and western Saghro inliers. In the
central and eastern Jbel Saghro area, volcanic rocks as reported are somewhat younger and mostly
dated between 570 and 550 Ma from west to east [51,74]. This suggests that the main volcanic stage
progresses toward the east along with the D2 tectonic regime. If we compare with the Sierra Madre
Occidental as a model of large silicic volcanic province [14,147,148,213], the widespread ash-flow tuff
deposits of the Anti-Atlas domain should be emplaced as ignimbrite flare-ups and are correlated with
progressive and diachronic formation toward the east of caldera collapse structures, broadly oriented
E–W to NW–SE. This volcano-structural framework developed coevally with a transtensive tectonic
139
Minerals 2018, 8, 592
regime characterised by both NNE–SSW extension and WNW–ESE shortening. In terms of melting
processes, the main controlling factor in the generation of such a SLIP is a crustal setting located along
a long-lived active subduction zone that evolves into a post-subduction domain via slab-roll back
processes. According to Bryan and Ferrari [14] and Ernst [12], a huge thermal pulse is at the origin of a
large-scale crustal anataxis of fertile and hydrous lower-crustal materials as well as metasomatised
subcontinental lithospheric mantle. Flare-up models are in part inherited from the late evolution of arc
settings that underwent slab-roll back, slab-breakoff and slab-window [14,147,148,214], as propoded
here and correlated with the D1 and D2 tectonic model (Figure 13). Because they represent transient
events of high magmatic fluxes from the mantle [215], volcanic flare ups are considered here as highly
potential for Ag(–Au) economic deposits emplacement.
6. Conclusions
We document in this paper a long-lived tectono-magmatic event that produced two main
litho-structural units we refer to as the Lower and Uper Complexes, respectively. The Lower Complex
is coeval with the main Pan-African D1 deformation consisting of a transpressive regime responsible
for folding, faulting and pluton emplacement under the effects of a NW–SE to WNW–ESE direction
of shortening from 615 to 575 Ma. The Upper Complex is characterised by the emplacement of large
volumes of ash-flow tuffs and volcanoclastic rocks and related intrusions. These were linked with the
formation of ash-flow caldera structures which are uncommom examples of preserved Precambrian
ash-flow calderas.
Ore deposits (porphyry, intermediate and low sulfidation epithermal deposit types, and IRGD)
show strong spatial and temporal relationships with the emplacement of the widespread magmatic
units belonging to the Upper Complex. For each ore deposit, fluid circulations associated with plutonic
and/or volcanic systems can be invoked to be at the origin of the genesis of economic paragenesis.
We suggest that magmatism of the Upper Complex and ore concentrations were both coeval with a D2
deformation stage (575–540 Ma) and were controlled by the same transtensive tectonic regime under
the effect of a nearly WNW–ESE shortening direction.
Despite an incomplete record by absolute datations, we infer that the Jbel Saghro was affected
over a long period of time (i.e., 95 Myr) by successively magmatic, magmato-hydrothermal and
hydrothermal events which formed a large mineralised province with substantial economic potential.
Previous authors have already envisioned such a long period of magmatism and hydrothermalism in
the area [29,47,74]. We further defend such a view and contend the Jbel Saghro province compares
in this respect to numerous examples of large magmatic-related mineralised systems documented
elsewere during the Archaean to Caenozoic times. It may be questioned whether the particular
longevity of the magmatic activity in areas dominated by transpressive and transtensive tectonics
could be related to the specific behaviour of ancient continental domains in which tectonics are
dominated by vertical forces and are linked with especially huge magmatism [216–218]. This debate
is currently open and our results illustrate one additional example of long-lived tectono-magmatic
event that characterises the late Precambrian in this part or the African continent. Considering the
large volume of ash-flow tuffs that crop out in the Western, Central and Eastern Anti-Atlas and their
long-lived magmatic activity (575–550 Ma), we infer that the whole Anti-Atlas area (i.e., 700 km
long) belongs to a continental silicic large igneous province as defined by Bryan and Ferrari [14],
and Ernst [12], and emplaced in a subduction to post-subduction setting and that may be linked to
a wide area including the Cadomian segments [28,206,207,219]. Anyway, our results offer further
evidence that the majority of metal deposits in the Moroccan Anti-Atlas could be formed during the
Neoproterozoic times coevally with the tectonic and magmatic evolutions in this period. They also
demonstrate that structural geology can provide relevant constraints for debating the age and mode of
formation of ore deposits, specifically in the context of a large silicic magmatic provinces.
Author Contributions: J.T., A.C., L.B. (Luc Barbanson), J.-L.B., M.L., A.E., L.B. (Lakhlifi Badra), C.E.-L., M.D.
and S.S. took part in the field investigation; M.L., A.E. and L.M. supported the field investigation; J.T., A.C., L.B.
140
Minerals 2018, 8, 592
(Luc Barbanson), J.-L.B., C.E.-L., S.S., and M.D. interpreted the data and took part in the discussion; J.T., A.C.,
J.-L.B. and M.D. wrote the original draft; J.T. and A.C. reviewed and edited the paper.
Funding: This work has been undertaken with the help of the French-Moroccan programs “Action Intégrée No
222/STU/00”.
Acknowledgments: The REMINEX exploration team and SMI mining company provided funds and logistics
for field and laboratory studies. We particularly acknowledge El Hajj Bouiroukouten and the intern geologists
of MANAGEM for their constant help and support. We are grateful to the Masters students from the BRGM
Campus, University of Orléans, University of Lille and LaSalle Beauvais who were involved in the field for
geological mapping exercises. Olivier Rouer and Gilles Drouet are warmly thanked for assistance and help
with electronic microprobe analyses. The manuscript benefitted considerably from constructive reviews by four
anonymous referees.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Dostal, J.; Caby, R.; Keppie, J.D.; Maza, M. Neoproterozoic magmatism in Southwestern Algeria (Sebkha
el Melah inlier): A northerly extension of the Trans-Saharan orogen. J. Afr. Earth Sci. 2002, 35, 213–225.
[CrossRef]
2. Ennih, N.; Liegeois, J.-P. The boundaries of the West African craton, with special reference to the basement of
the Moroccan metacratonic Anti-Atlas belt. Geol. Soc. Lond. Spec. Publ. 2008, 297, 1–17. [CrossRef]
3. Liégeois, J.P.; Latouche, L.; Boughrara, M.; Navez, J.; Guiraud, M. The LATEA metacraton (Central Hoggar,
Tuareg shield, Algeria): Behaviour of an old passive margin during the Pan-African orogeny. J. Afr. Earth Sci.
2003, 37, 161–190. [CrossRef]
4. Ouzegane, K.; Kienast, J.-R.; Bendaoud, A.; Drareni, A. A review of Archaean and Paleoproterozoic evolution
of the In Ouzzal granulitic terrane (Western Hoggar, Algeria). J. Afr. Earth Sci. 2003, 37, 207–227. [CrossRef]
5. Thomas, R.J.; Fekkak, A.; Ennih, N.; Errami, E.; Loughlin, S.C.; Gresse, P.G.; Chevallier, L.P.; Liegeois, J.-P. A
new lithostratigraphic framework for the Anti-Atlas Orogen, Morocco. J. Afr. Earth Sci. 2004, 39, 217–226.
[CrossRef]
6. El Bahat, A.; Ikenne, M.; Söderlund, U.; Cousens, B.; Youbi, N.; Ernst, R.; Soulaimani, A.; El Janati, M.H.;
Hafid, A. U–Pb baddeleyite ages and geochemistry of dolerite dykes in the Bas Drâa Inlier of the Anti-Atlas
of Morocco: Newly identified 1380Ma event in the West African Craton. Lithos 2013, 174, 85–98. [CrossRef]
7. Ikenne, M.; Söderlund, U.; Ernst, R.E.; Pin, C.; Youbi, N.; El Aouli, E.H.; Hafid, A. A c. 1710 Ma mafic
sill emplaced into a quartzite and calcareous series from Ighrem, Anti-Atlas—Morocco: Evidence that the
Taghdout passive margin sedimentary group is nearly 1 Ga older than previously thought. J. Afr. Earth Sci.
2017, 127, 62–76. [CrossRef]
8. Youbi, N.; Kouyaté, D.; Söderlund, U.; Ernst, R.E.; Soulaimani, A.; Hafid, A.; Ikenne, M.; El Bahat, A.;
Bertrand, H.; Rkha Chaham, K.; et al. The 1750Ma Magmatic Event of the West African Craton (Anti-Atlas,
Morocco). Precambrian Res. 2013, 236, 106–123. [CrossRef]
9. Davies, J.H.F.L.; Marzoli, A.; Bertrand, H.; Youbi, N.; Ernesto, M.; Schaltegger, U. End-Triassic mass extinction
started by intrusive CAMP activity. Nat. Commun. 2017, 8, 15596. [CrossRef]
10. Knight, K.B.; Nomade, S.; Renne, P.R.; Marzoli, A.; Bertrand, H.; Youbi, N. The Central Atlantic Magmatic
Province at the Triassic-Jurassic boundary: Paleomagnetic and 40 Ar/39 Ar evidence from Morocco for brief,
episodic volcanism. Earth Planet. Sci. Lett. 2004, 228, 143–160. [CrossRef]
11. Marzoli, A.; Callegaro, S.; Dal Corso, J.; Davies, J.H.F.L.; Chiaradia, M.; Youbi, N.; Bertrand, H.; Reisberg, L.;
Merle, R.; Jourdan, F. The Central Atlantic Magmatic Province (CAMP): A Review. In The Late Triassic World:
Earth in a Time of Transition; Tanner, L.H., Ed.; Springer International Publishing: Cham, Switzerland, 2018;
pp. 91–125.
12. Ernst, R.E. Large Igneous Provinces; Cambridge University Press: Cambridge, UK, 2014.
13. Ernst, R.; Bleeker, W. Large igneous provinces (LIPs), giant dyke swarms, and mantle plumes: Significance
for breakup events within Canada and adjacent regions from 2.5 Ga to the Present. Can. J. Earth Sci. 2010, 47,
695–739. [CrossRef]
14. Bryan, S.E.; Ferrari, L. Large igneous provinces and silicic large igneous provinces: Progress in our
understanding over the last 25 years. GSA Bull. 2013, 125, 1053–1078. [CrossRef]
141
Minerals 2018, 8, 592
15. Chappell, B.W.; White, A.J.R. Two contrasting granite types: 25 years later. Aust. J. Earth Sci. 2001, 48,
489–499. [CrossRef]
16. Bonin, B. A-type granites and related rocks: Evolution of a concept, problems and prospects. Lithos 2007, 97,
1–29. [CrossRef]
17. Ernst, R.E.; Jowitt, S.M. Large Igneous Provinces (LIPs) and Metallogeny. In Tectonics, Metallogeny, and
Discovery: The North American Cordillera and Similar Accretionary Settings; Colpron, M., Bissig, T., Rusk, B.G.,
Thompson, J.F.H., Eds.; Society of Economic Geologists: Littleton, CO, USA, 2013; Volume 17, pp. 17–51.
18. Arndt, N.T.; Lesher, C.M.; Czamanske, G.K. Mantle-derived magmas and magmatic Ni-Cu-(PGE) deposits.
In Economic Geology 100th Anniversary Volume; Hedenquist, J.W., Thompson, J.F.H., Goldfarb, R.J., Richard, J.P.,
Eds.; Society of Economic Geologists: Littleton, CO, USA, 2005; pp. 5–23.
19. Barnes, S.J.; Holwell, D.A.; Le Vaillant, M. Magmatic Sulfide Ore Deposits. Elements 2017, 13, 89–95.
[CrossRef]
20. Goodenough, K.M.; Schilling, J.; Jonsson, E.; Kalvig, P.; Charles, N.; Tuduri, J.; Deady, E.A.; Sadeghi, M.;
Schiellerup, H.; Müller, A.; et al. Europe’s rare earth element resource potential: An overview of REE
metallogenetic provinces and their geodynamic setting. Ore Geol. Rev. 2016, 72, 838–856. [CrossRef]
21. Ernst, R.E.; Bell, K. Large igneous provinces (LIPs) and carbonatites. Mineral. Petrol. 2010, 98, 55–76.
[CrossRef]
22. Rao, N.V.C.; Lehmann, B. Kimberlites, flood basalts and mantle plumes: New insights from the Deccan
Large Igneous Province. Earth-Sci. Rev. 2011, 107, 315–324. [CrossRef]
23. Barton, M.D. 13.20—Iron Oxide(–Cu–Au–REE–P–Ag–U–Co) Systems A2—Holland, Heinrich, D. In Treatise
on Geochemistry (Second Edition); Turekian, K.K., Ed.; Elsevier: Oxford, UK, 2014; pp. 515–541.
24. Camprubí, A. Tectonic and Metallogenetic History of Mexico. In Tectonics, Metallogeny, and Discovery:
The North American Cordillera and Similar Accretionary Settings; Colpron, M., Bissig, T., Rusk, B.G.,
Thompson, J.F.H., Eds.; Society of Economic Geologists: Littleton, CO, USA, 2013; Volume 17, pp. 201–243.
25. Camprubí, A.; Albinson, T. Epithermal deposits in México—Update of current knowledge, and an empirical
reclassification. Geol. Soc. Am. Spec. Pap. 2007, 422, 377–415.
26. Wilkinson, J.J.; Simmons, S.F.; Stoffell, B. How metalliferous brines line Mexican epithermal veins with silver.
Sci. Rep. 2013, 3, 2057. [CrossRef]
27. Ballèvre, M.; Le Goff, E.; Hébert, R. The tectonothermal evolution of the Cadomian belt of northern Brittany,
France: A Neoproterozoic volcanic arc. Tectonophysics 2001, 331, 19–43. [CrossRef]
28. Linnemann, U.; Pereira, F.; Jeffries, T.E.; Drost, K.; Gerdes, A. The Cadomian Orogeny and the opening of
the Rheic Ocean: The diacrony of geotectonic processes constrained by LA-ICP-MS U–Pb zircon dating
(Ossa-Morena and Saxo-Thuringian Zones, Iberian and Bohemian Massifs). Tectonophysics 2008, 461, 21–43.
[CrossRef]
29. Walsh, G.J.; Benziane, F.; Aleinikoff, J.N.; Harrison, R.W.; Yazidi, A.; Burton, W.C.; Quick, J.E.; Saadane, A.
Neoproterozoic tectonic evolution of the Jebel Saghro and Bou Azzer—El Graara inliers, eastern and central
Anti-Atlas, Morocco. Precambrian Res. 2012, 216–219, 23–62. [CrossRef]
30. Boyer, C.; Leblanc, M. Les appareils émissifs de la formation volcanique infracambriennes de Ouarzazate,
Anti-Atlas (Maroc). Comptes rendus hebdomadaires des séances de l’Académie des sciences 1977, 285, 641–644.
31. Moume, W.; Youbi, N.; Marzoli, A.; Bertrand, H.; Gärtner, A.; Linnemann, U.; Gerdes, A.; Ernst, R.;
Söderlund, U.; Hachimi Hind, E.; et al. The distribution of the Central Iapetus Magmatic Province (CIMP)
into West African craton: U-Pb dating, geochemistry and petrology of Douar Eç-çour and Imiter mafic
Dyke Swarms (High and Anti-Atlas, Morocco). In Proceedings of the 2nd Colloquium of the International
Geoscience Programme (IGCP638), Casablanca, Morocco, 7–12 November 2017.
32. Puffer, J.H. A late Neoproterozoic eastern Laurentian superplume: Location, size, chemical composition, and
environmental impact. Am. J. Sci. 2002, 302, 1–27. [CrossRef]
33. Barbey, P.; Oberli, F.; Burg, J.P.; Nachit, H.; Pons, J.; Meier, M. The Palaeoproterozoic in western Anti-Atlas
(Morocco): A clarification. J. Afr. Earth Sci. 2004, 39, 239–245. [CrossRef]
34. Burkhard, M.; Caritg, S.; Helg, U.; Robert-Charrue, C.; Soulaimani, A. Tectonics of the Anti-Atlas of Morocco.
Comptes Rendus Geosci. 2006, 338, 11–24. [CrossRef]
35. Missenard, Y.; Zeyen, H.; Frizon de Lamotte, D.; Leturmy, P.; Petit, C.; Sébrier, M.; Saddiqi, O. Crustal versus
asthenospheric origin of relief of the Atlas Mountains of Morocco. J. Geophys. Res. Solid Earth 2006, 111,
B03401. [CrossRef]
142
Minerals 2018, 8, 592
36. Bourque, H.; Barbanson, L.; Sizaret, S.; Branquet, Y.; Ramboz, C.; Ennaciri, A.; El Ghorfi, M.; Badra, L.
A contribution to the synsedimentary versus epigenetic origin of the Cu mineralizations hosted by terminal
Neoproterozoic to Cambrian formations of the Bou Azzer–El Graara inlier: New insights from the Jbel
Laassel deposit (Anti Atlas, Morocco). J. Afr. Earth Sci. 2015, 107, 108–118. [CrossRef]
37. Pouit, G. Paléogéographie et répartition des minéralisations stratiformes de cuivre dans l’Anti-Atlas
occidental (Maroc). Chronique de la Recherche Minière 1966, 34, 279–289.
38. Borisenko, A.S.; Lebedev, V.I.; Borovikov, A.A.; Pavlova, G.G.; Kalinin, Y.A.; Nevol’ko, P.A.; Maacha, L.;
Kostin, A.V. Forming conditions and age of native silver deposits in Anti-Atlas (Morocco). Dokl. Earth Sci.
2014, 456, 663–666. [CrossRef]
39. Oberthur, T.; Melcher, F.; Henjes-Kunst, F.; Gerdes, A.; Stein, H.; Zimmerman, A.; El Ghorfi, M. Hercynian
age of the cobalt-nickel-arsenide-(gold) ores, Bou Azzer, Anti-Atlas, Morocco: Re-Os, Sm-Nd, and U-Pb age
determinations. Econ. Geol. 2009, 104, 1065–1079. [CrossRef]
40. Essarraj, S.; Boiron, M.-C.; Cathelineau, M.; Banks, D.A.; Benharref, M. Penetration of surface-evaporated
brines into the Proterozoic basement and deposition of Co and Ag at Bou Azzer (Morocco): Evidence from
fluid inclusions. J. Afr. Earth Sci. 2005, 41, 25–39. [CrossRef]
41. Essarraj, S.; Boiron, M.-C.; Cathelineau, M.; Tarantola, A.; Leisen, M.; Boulvais, P.; Maacha, L. Basinal Brines
at the Origin of the Imiter Ag-Hg Deposit (Anti-Atlas, Morocco): Evidence from LA-ICP-MS Data on Fluid
Inclusions, Halogen Signatures, and Stable Isotopes (H, C, O). Econ. Geol. 2016, 111, 1753–1781. [CrossRef]
42. Abia, E.H.; Nachit, H.; Marignac, C.; Ibhi, A.; Saadi, S.A. The polymetallic Au-Ag-bearing veins of Bou
Madine (Jbel Ougnat, eastern Anti-Atlas, Morocco): Tectonic control and evolution of a Neoproterozoic
epithermal deposit. J. Afr. Earth Sci. 2003, 36, 251–271. [CrossRef]
43. Al Ansari, A.E.; Sagon, J.P. Le gisement d’or de Tiouit (Jbel Saghro, Anti-Atlas, maroc). Un système
mésothermal polyphasé à sulfures-or et hématite-or dans une granodiorite potassique d’âge Protérozoïque
supérieur. Chronique de la Recherche Minière 1997, 527, 3–25.
44. Leblanc, M.; Lbouabi, M. Native silver mineralization along a rodingite tectonic contact between serpentinite
and quartz diorite (Bou Azzer, Morocco). Econ. Geol. 1988, 83, 1379–1391. [CrossRef]
45. Levresse, G.; Cheilletz, A.; Gasquet, D.; Reisberg, L.; Deloule, E.; Marty, B.; Kyser, K. Osmium, sulphur, and
helium isotopic results from the giant Neoproterozoic epithermal Imiter silver deposit, Morocco: Evidence
for a mantle source. Chem. Geol. 2004, 207, 59–79. [CrossRef]
46. Marcoux, E.; Wadjinny, A. Le gisement Ag–Hg de Zgounder (Jebel Siroua, Anti-Atlas, Maroc): Un épithermal
néoprotérozoïque de type Imiter. Comptes Rendus Geosci. 2005, 337, 1439–1446. [CrossRef]
47. Tuduri, J. Processus de formation et relations spatio-temporelles des minéralisations à or et argent
en contexte volcanique Précambrien (Jbel Saghro, Anti-Atlas, Maroc). Implications sur les relations
déformation-magmatisme-volcanisme-hydrothermalisme. Ph.D. Thesis, University of Orléans, Orléans,
France, 2005.
48. Tuduri, J.; Chauvet, A.; Barbanson, L.; Labriki, M.; Dubois, M.; Trapy, P.-H.; Lahfid, A.; Poujol, M.; Melleton, J.;
Badra, L.; et al. Structural control, magmatic-hydrothermal evolution and formation of hornfels-hosted,
intrusion-related gold deposits: Insight from the Thaghassa deposit in Eastern Anti-Atlas, Morocco. Ore Geol.
Rev. 2018, 97, 171–198. [CrossRef]
49. Essarraj, S.; Boiron, M.-C.; Cathelineau, M.; Banks, D.A.; El Boukhari, A.; Chouhaidi, M.Y. Brines related to
Ag deposition in the Zgounder silver deposit (Anti-Atlas, Morocco). Eur. J. Mineral. 1998, 10, 1201–1214.
[CrossRef]
50. Levresse, G.; Bouabdellah, M.; Gasquet, D.; Cheilletz, A. Basinal Brines at the Origin of the Imiter Ag-Hg
Deposit (Anti-Atlas, Morocco): Evidence from LA-ICP-MS Data on Fluid Inclusions, Halogen Signatures,
and Stable Isotopes (H, C, O)—A Discussion. Econ. Geol. 2017, 112, 1269–1272. [CrossRef]
51. Cheilletz, A.; Levresse, G.; Gasquet, D.; Azizi-Samir, M.R.; Zyadi, R.; Archibald, A.D.; Farrar, E. The giant
Imiter silver deposit: Neoproterozoic epithermal mineralization in the Anti-Atlas, Morocco. Miner. Depos.
2002, 37, 772–781. [CrossRef]
52. Pelleter, E.; Cheilletz, A.; Gasquet, D.; Mouttaqi, A.; Annich, M.; Camus, Q.; Deloule, E.; Ouazzani, L.;
Bounajma, H.; Ouchtouban, L. U/Pb Ages of Magmatism in the Zgounder Epithermal Ag–Hg Deposit,
Sirwa Window, Anti-Atlas, Morocco. In Mineral Deposits of North Africa; Bouabdellah, M., Slack, J.F., Eds.;
Springer International Publishing: Cham, Switzerland, 2016; pp. 143–165.
143
Minerals 2018, 8, 592
53. Ahmed, A.H.; Arai, S.; Ikenne, M. Mineralogy and Paragenesis of the Co-Ni Arsenide Ores of Bou Azzer,
Anti-Atlas, Morocco. Econ. Geol. 2009, 104, 249–266. [CrossRef]
54. Ennaciri, A.; Barbanson, L.; Touray, J.C. Mineralized hydrothermal solution cavities in the Co-As Ait Ahmane
mine (Bou Azzer, Morocco). Miner. Depos. 1995, 30, 75–77. [CrossRef]
55. Leblanc, M. Co-Ni arsenide deposits, with accessory gold, in ultramafic rocks from:Morocco. Can. J. Earth
Sci. 1986, 23, 1592–1602. [CrossRef]
56. Thiéblemont, D.; Chêne, F.; Liégeois, J.-P.; Ouabadi, A.; Le Gall, B.; Maury, R.C.; Jalludin, M.; Ouattara
Gbélé, C.; Tchaméni, R.; Fernandez-Alonso, M. Geological Map of Africa at 1:10 Million Scale, 35th International
Geology Congress ed; CCGM-BRGM: Orléans, France, 2016.
57. Hollard, H.; Choubert, G.; Bronner, G.; Marchand, J.; Sougy, J. Carte géologique du Maroc, échelle:
1/1.000.000. Notes et Mémoires du Service Géologique du Maroc 1985, 260.
58. Mouttaqi, A.; Rjimati, E.; Maacha, A.; Michard, A.; Soulaimani, A.; Ibouh, H. Les principales mines du Maroc.
Notes et Mémoires du Service Géologique du Maroc 2011, 564, 375.
59. Thomas, R.J.; Chevallier, L.P.; Gresse, P.G.; Harmer, R.E.; Eglington, B.M.; Armstrong, R.A.; de Beer, C.H.;
Martini, J.E.J.; de Kock, G.S.; Macey, P.H.; et al. Precambrian evolution of the Sirwa Window, Anti-Atlas
Orogen, Morocco. Precambrian Res. 2002, 118, 1–57. [CrossRef]
60. Goldfarb, R.J.; Groves, D.I. Orogenic gold: Common or evolving fluid and metal sources through time. Lithos
2015, 233, 2–26. [CrossRef]
61. Groves, D.I.; Santosh, M.; Goldfarb, R.J.; Zhang, L. Structural geometry of orogenic gold deposits:
Implications for exploration of world-class and giant deposits. Geosci. Front. 2018, 9, 1163–1177. [CrossRef]
62. Hart, C.J. Reduced intrusion-related gold systems. In Mineral Deposits of Canada: A Synthesis of Major Deposit
Types, District Metallogeny, the Evolution of Geological Provinces, and Exploration Methods; Special Publication;
Geological Association of Canada, Mineral Deposits Division: St. John’s, NL, Canada, 2007; pp. 95–112.
63. Kontak, D.; O’Reilly, G.; MacDonald, M.; Horne, R.; Smith, P. Gold in the Meguma Terrane, Southern
Nova Scotia: Is There a Continuum between Mesothermal Lode Gold and Intrusion-related Gold Systems?
In Proceedings of the 49th Annual Meeting of the GAC-MAC, St. Catharines, ON, Canada, 12–14 May2004;
p. 128.
64. Lang, J.R.; Baker, T. Intrusion-related gold systems: The present level of understanding. Miner. Depos. 2001,
36, 477–489. [CrossRef]
65. Walshe, J.; Neumayr, P.; Cooke, D. Two boxes we don’t need: Orogenic and intrusion-related gold systems.
In Proceedings of the STOMP 2005: Structure, Tectonics and Ore Mineralisation Processes, Townsville,
Australia, 29 August–2 September 2005; EGRU: Townsville, Australia, 2005; p. 143.
66. Boiron, M.-C.; Cathelineau, M.; Banks, D.A.; Fourcade, S.; Vallance, J. Mixing of metamorphic and surficial
fluids during the uplift of the Hercynian upper crust: Consequences for gold deposition. Chem. Geol. 2003,
194, 119–141. [CrossRef]
67. Vallance, J.; Cathelineau, M.; Boiron, M.C.; Fourcade, S.; Shepherd, T.J.; Naden, J. Fluid-rock interactions and
the role of late Hercynian aplite intrusion in the genesis of the Castromil gold deposit, northern Portugal.
Chem. Geol. 2003, 194, 201–224. [CrossRef]
68. Baker, T.; Lang, J.R. Fluid inclusion characteristics of intrusion-related gold mineralization,
Tombstone–Tungsten magmatic belt, Yukon Territory, Canada. Mineral. Depos. 2001, 36, 563–582. [CrossRef]
69. Chauvet, A.; Volland-Tuduri, N.; Lerouge, C.; Bouchot, V.; Monié, P.; Charonnat, X.; Faure, M.
Geochronological and geochemical characterization of magmatic-hydrothermal events within the Southern
Variscan external domain (Cévennes area, France). Int. J. Earth Sci. 2012, 101, 69–86. [CrossRef]
70. Mustard, R.; Ulrich, T.; Kamenetsky, V.S.; Mernagh, T. Gold and metal enrichment in natural granitic melts
during fractional crystallization. Geology 2006, 34, 85–88. [CrossRef]
71. Gouiza, M.; Charton, R.; Bertotti, G.; Andriessen, P.; Storms, J.E.A. Post-Variscan evolution of the Anti-Atlas
belt of Morocco constrained from low-temperature geochronology. Int. J. Earth Sci. 2017, 106, 593–616.
[CrossRef]
72. Teixell, A.; Ayarza, P.; Zeyen, H.; Fernàndez, M.; Arboleya, M.-L. Effects of mantle upwelling in a
compressional setting: The Atlas Mountains of Morocco. Terra Nova 2005, 17, 456–461. [CrossRef]
73. Choubert, G. Histoire géologique du Précambrien de l’Anti-Atlas de l’Archéen à l’aurore des temps primaires.
Notes et Mémoires du Service Géologique du Maroc 1963, 162, 352.
144
Minerals 2018, 8, 592
74. Gasquet, D.; Levresse, G.; Cheilletz, A.; Azizi-Samir, M.R.; Mouttaqi, A. Contribution to a geodynamic
reconstruction of the Anti-Atlas (Morocco) during Pan-African times with the emphasis on inversion
tectonics and metallogenic activity at the Precambrian-Cambrian transition. Precambrian Res. 2005, 140,
157–182. [CrossRef]
75. Hefferan, K.; Soulaimani, A.; Samson, S.D.; Admou, H.; Inglis, J.; Saquaque, A.; Latifa, C.; Heywood, N.
A reconsideration of Pan African orogenic cycle in the Anti-Atlas Mountains, Morocco. J. Afr. Earth Sci. 2014,
98, 34–46. [CrossRef]
76. Leblanc, M.; Lancelot, J.R. Interprétation géodynamique du domaine panafricain (Précambrien terminal)
de l’Anti-Atlas (Maroc) à partir de données géologiques et géochronologiques. Can. J. Earth Sci. 1980, 17,
142–155. [CrossRef]
77. Saquaque, A.; Benharref, M.; Abia, H.; Mrini, Z.; Reuber, I.; Karson, J.A. Evidence for a Panafrican volcanic
arc and wrench fault tectonics in Jbel Saghro, Morocco. Geol. Rundsch. 1992, 81, 1–13. [CrossRef]
78. Blein, O.; Baudin, T.; Chèvremont, P.; Soulaimani, A.; Admou, H.; Gasquet, P.; Cocherie, A.; Egal, E.; Youbi, N.;
Razin, P.; et al. Geochronological constraints on the polycyclic magmatism in the Bou Azzer-El Graara inlier
(Central Anti-Atlas Morocco). J. Afr. Earth Sci. 2014, 99, 287–306. [CrossRef]
79. El Hadi, H.; Simancas, J.F.; Martínez-Poyatos, D.; Azor, A.; Tahiri, A.; Montero, P.; Fanning, C.M.; Bea, F.;
González-Lodeiro, F. Structural and geochronological constraints on the evolution of the Bou Azzer
Neoproterozoic ophiolite (Anti-Atlas, Morocco). Precambrian Res. 2010, 182, 1–14. [CrossRef]
80. Inglis, J.D.; D’Lemos, R.S.; Samson, S.D.; Admou, H. Geochronological constraints on late Precambrian
intrusions, metamorphism, and tectonism in the Anti-Atlas mountains. J. Geol. 2005, 113, 439–450. [CrossRef]
81. Inglis, J.D.; MacLean, J.S.; Samson, S.D.; D’Lemos, R.S.; Admou, H.; Hefferan, K. A precise U-Pb zircon age
for the BleIda granodiorite, Anti-Atlas, Morocco: Implications for the timing of deformation and terrane
assembly in the eastern Anti-Atlas. J. Afr. Earth Sci. 2004, 39, 277. [CrossRef]
82. Saquaque, A.; Admou, H.; Karson, J.; Hefferan, K.; Reuber, I. Precambrian accretionary tectonics in the Bou
Azzer-El Graara region, Anti-Atlas, Morocco. Geology 1989, 17, 1107–1110. [CrossRef]
83. Gasquet, D.; Ennih, N.; Liégeois, J.-P.; Soulaimani, A.; Michard, A. The Pan-African Belt. In Continental
Evolution: The Geology of Morocco; Michard, A., Saddiqi, O., Chalouan, A., Frizon de Lamotte, D., Eds.;
Springer-Verlag: Berlin/Heidelberg, Germany, 2008; pp. 33–64.
84. Álvaro, J.J.; Benziane, F.; Thomas, R.; Walsh, G.J.; Yazidi, A. Neoproterozoic–Cambrian stratigraphic
framework of the Anti-Atlas and Ouzellagh promontory (High Atlas), Morocco. J. Afr. Earth Sci. 2014, 98,
19–33. [CrossRef]
85. Choubert, G. In Essai d’application de la notion d’Infracambrien aux formations anciennes de l’Anti-Atlas
(Maroc). In Proceedings of the 19th International Geological Congress, Alger, Algeria, 8–15 September 1952;
pp. 33–71.
86. Soulaimani, A.; Michard, A.; Ouanaimi, H.; Baidder, L.; Raddi, Y.; Saddiqi, O.; Rjimati, E.C. Late
Ediacaran–Cambrian structures and their reactivation during the Variscan and Alpine cycles in the Anti-Atlas
(Morocco). J. Afr. Earth Sci. 2014, 98, 94–112. [CrossRef]
87. Soulaimani, A.; Bouabdelli, M.; Piqué, A. The Upper Neoproterozoic-Lower Cambrian continental extension
in the Anti-Atlas (Morocco). Bulletin de la Société Géologique de France 2003, 174, 83–92. [CrossRef]
88. Chèvremont, P.; Blein, O.; Razin, P.; Baudin, T.; Barbanson, L.; Gasquet, D.; Soulaimani, A.; Admou, H.;
Youbi, N.; Bouabdelli, M.; et al. Carte géologique du Maroc (1/50 000), feuille de Bou Azer. Notes et Mémoires
du Service Géologique du Maroc 2013, 535bis, 153.
89. Ducrot, J.; Lancelot, J.R. Problème de la limite Précambrien–Cambrien: Étude radiochronologique par
la méthode U–Pb sur zircons du volcan du Jbel Boho (Anti-Atlas marocain). Can. J. Earth Sci. 1977, 14,
2771–2777. [CrossRef]
90. Maloof, A.C.; Schrag, D.P.; Crowley, J.L.; Bowring, S.A. An expanded record of Early Cambrian carbon
cycling from the Anti-Atlas Margin, Morocco. Can. J. Earth Sci. 2005, 42, 2195–2216. [CrossRef]
91. Fekkak, A.; Pouclet, A.; Ouguir, H.; Ouazzani, H.; Badra, L.; Gasquet, D. Géochimie et signification
géotectonique des volcanites du Cryogénien inférieur du Saghro (Anti-Atlas oriental, Maroc). Geodin. Acta
2001, 13, 1–13.
92. Ouguir, H.; Macaudière, J.; Dagallier, G. Le Protérozoïque supérieur d’Imiter, Saghro oriental, Maroc:
Un contexte géodynamique d’arrière arc. J. Afr. Earth Sci. 1996, 22, 173–189. [CrossRef]
145
Minerals 2018, 8, 592
93. Baidder, L.; Raddi, Y.; Tahiri, M.; Michard, A. Devonian extension of the Pan-African crust north of the West
African craton, and its bearing on the Variscan foreland deformation: Evidence from eastern Anti-Atlas
(Morocco). Geol. Soc. Lond. Spec. Publ. 2008, 297, 453–465. [CrossRef]
94. Malusà, M.G.; Polino, R.; Feroni, A.C.; Ellero, A.; Ottria, G.; Baidder, L.; Musumeci, G. Post-Variscan tectonics
in eastern Anti-Atlas (Morocco). Terra Nova 2007, 19, 481–489. [CrossRef]
95. Michard, A.; Soulaimani, A.; Hoepffner, C.; Ouanaimi, H.; Baidder, L.; Rjimati, E.C.; Saddiqi, O.
The South-Western Branch of the Variscan Belt: Evidence from Morocco. Tectonophysics 2010, 492, 1–24.
[CrossRef]
96. Frizon de Lamotte, D.; Tavakoli-Shirazi, S.; Leturmy, P.; Averbuch, O.; Mouchot, N.; Raulin, C.;
Leparmentier, F.; Blanpied, C.; Ringenbach, J.-C. Evidence for Late Devonian vertical movements and
extensional deformation in northern Africa and Arabia: Integration in the geodynamics of the Devonian
world. Tectonics 2013, 32, 107–122. [CrossRef]
97. Alvaro, J.J.; Macouin, M.; Ezzouhairi, H.; Charif, A.; Ayad, N.A.; Ribeiro, M.L.; Ader, M. Late Neoproterozoic
carbonate productivity in a rifting context: The Adoudou Formation and its associated bimodal volcanism
onlapping the western Saghro inlier, Morocco. Geol. Soc. Lond. Spec. Publ. 2008, 297, 285–302. [CrossRef]
98. Álvaro, J.J. Late Ediacaran syn-rift/post-rift transition and related fault-driven hydrothermal systems in the
Anti-Atlas Mountains, Morocco. Basin Res. 2013, 25, 348–360. [CrossRef]
99. Sebti, S.; Saddiqi, O.; El Haimer, F.Z.; Michard, A.; Ruiz, G.; Bousquet, R.; Baidder, L.; Frizon de Lamotte, D.
Vertical movements at the fringe of the West African Craton: First zircon fission track datings from the
Anti-Atlas Precambrian basement, Morocco. Comptes Rendus Geosci. 2009, 341, 71–77. [CrossRef]
100. Caritg, S.; Burkhard, M.; Ducommun, R.; Helg, U.; Kopp, L.; Sue, C. Fold interference patterns in the Late
Palaeozoic Anti-Atlas belt of Morocco. Terra Nova 2004, 16, 27–37.
101. Levresse, G.; Bouabdellah, M.; Cheilletz, A.; Gasquet, D.; Maacha, L.; Tritlla, J.; Banks, D.; Moulay Rachid, A.S.
Degassing as the Main Ore-Forming Process at the Giant Imiter Ag–Hg Vein Deposit in the Anti-Atlas
Mountains, Morocco. In Mineral Deposits of North Africa; Bouabdellah, M., Slack, J.F., Eds.; Springer
International Publishing: Cham, Switzerland, 2016; pp. 85–106.
102. Tuduri, J.; Chauvet, A.; Ennaciri, A.; Barbanson, L. Modèle de formation du gisement d’argent d’Imiter
(Anti-Atlas oriental, Maroc). Nouveaux apports de l’analyse structurale et minéralogique. Comptes Rendus
Geosci. 2006, 338, 253–261. [CrossRef]
103. Bouabdellah, M.; Maacha, L.; Jébrak, M.; Zouhair, M. Re/Os Age Determination, Lead and Sulphur Isotope
Constraints on the Origin of the Bouskour Cu–Pb–Zn Vein-Type Deposit (Eastern Anti-Atlas, Morocco) and
Its Relationship to Neoproterozoic Granitic Magmatism. In Mineral Deposits of North Africa; Bouabdellah, M.,
Slack, F.J., Eds.; Springer International Publishing: Cham, Switzerland, 2016; pp. 277–290.
104. Hindermeyer, J.; Choubert, G.; Destombes, J.; Gauthier, H. Carte géologique de l’Anti-Atlas oriental: Feuille
Dadès et Jbel Saghro 1/200 000. Notes et Mémoires du Service Géologique du Maroc 1977, 161.
105. Baidada, B.; Ikenne, M.; Barbey, P.; Soulaimani, A.; Cousens, B.; Haissen, F.; Ilmen, S.; Alansari, A. SHRIMP
U–Pb zircon geochronology of the granitoids of the Imiter Inlier: Constraints on the Pan-African events in
the Saghro massif, Anti-Atlas (Morocco). J. Afr. Earth Sci. 2018. [CrossRef]
106. De Wall, H.; Kober, B.; Errami, E.; Ennih, N.; Greiling, R.O. Age de mise en place et contexte géologique
des granitoïdes de la boutonnière d’Imiter (Saghro oriental, Anti-Atlas, Maroc). In Proceedings of the
2ème Colloque International 3MA (Magmatisme, Métamorphisme & Minéralisations Associées), Marrakech,
Maroc, 10–12 May 2001; p. 19.
107. O’Connor, E.; Barnes, R.; Beddoe-Stephens, B.; Fletcher, T.; Gillespie, M.; Hawkins, M.; Loughlin, S.; Smith, M.;
Smith, R.; Waters, C. Geology of the Drâa, Kerdous, and Boumalne districts, Anti-Atlas, Morocco; British Geological
Survey: Nottingham, UK, 2010; p. 310.
108. Schiavo, A.; Taj Eddine, K.; Algouti, A.; Benvenuti, M.; Dal Piaz, G.V.; Eddebi, A.; El Boukhari, A.;
Laftouhi, N.; Massironi, M.; Ounaimi, H.; et al. Carte géologique du Maroc au 1/50000, feuille Imtir.
Notes et Mémoires du Service Géologique du Maroc 2007, 518.
109. Charlot, R.; Choubert, G.; Faure-Muret, A.; Tisserant, D. Etude géochronologique du Précambrien de
l’Anti-Atlas (Maroc). Notes et Mémoires du Service Géologique du Maroc 1970, 30, 99–134.
110. Choubert, G. Sur le Précambrien marocain. Comptes rendus hebdomadaires des séances de l’Académie des sciences
1945, 221, 249–251.
146
Minerals 2018, 8, 592
111. Hindermeyer, J. Le Précambrien I et le Précambrien II du Saghro. Comptes rendus hebdomadaires des séances de
l’Académie des sciences 1953, 237, 921–923.
112. Hindermeyer, J. Le Précambrien III du Saghro. Comptes rendus hebdomadaires des séances de l’Académie des
sciences 1953, 237, 1024–1026.
113. Derré, C.; Lécolle, M. Altérations hydrothermales dans le Protérozoïque supérieur du Saghro (Anti-Atlas
oriental). Relations avec les minéralisations. Chronique de la Recherche Minière 1999, 536–537, 39–61.
114. Fekkak, A.; Boualoul, M.; Badra, L.; Amenzou, M.; Saquaque, A.; El-Amrani, I.E. Origine et contexte
géotectonique des dépôts détritiques du Groupe Néoprotérozoïque inférieur de Kelaat Mgouna (Anti-Atlas
Oriental, Maroc). J. Afr. Earth Sci. 2000, 30, 295–311. [CrossRef]
115. Fekkak, A.; Pouclet, A.; Badra, L. The Pre-Panafrican rifting of Saghro (Anti-Atlas, Morocco): Exemple of
the middle Neoproterozoic Basin of Boumalne. Bulletin de la Société Géologique de France 2002, 173, 25–35.
[CrossRef]
116. Fekkak, A.; Pouclet, A.; Benharref, M. The Middle Neoproterozoic Sidi Flah Group (Anti-Atlas, Morocco):
Synrift deposition in a Pan-African continent/ocean transition zone. J. Afr. Earth Sci. 2003, 37, 73–87.
[CrossRef]
117. Fekkak, A.; Pouclet, A.; Ouguir, H.; Badra, L.; Gasquet, D. The Kelaat Mgouna early Neoproterozoic Group
(Saghro, Anti-Atlas, Morocco): Witness of an initial stage of the pre-Pan-African extension. Bulletin de la
Société Géologique de France 1999, 170, 789–797.
118. Marini, F.; Ouguir, H. Un nouveau jalon dans l’histoire de la distension pré-panafricaine au Maroc:
Le Précambrien II des boutonnières du Jbel Saghro nord-oriental (Anti-Atlas, Maroc). Comptes Rendus
de l’Académie des Sciences Série II Mécanique-physique Chimie, Sciences de l’univers, Sciences de la Terre 1990, 310,
577–582.
119. Errami, E.; Bonin, B.; Laduron, D.; Lasri, L. Petrology and geodynamic significance of the post-collisional
Pan-African magmatism in the Eastern Saghro area (Anti-Atlas, Morocco). J. Afr. Earth Sci. 2009, 55, 105–124.
[CrossRef]
120. Liégeois, J.-P.; Fekkak, A.; Bruguier, O.; Errami, E.; Ennih, N. The Lower Ediacaran (630–610 Ma) Saghro
group: An orogenic transpressive basin development during the early metacratonic evolution of the
Anti-Atlas (Morocco). In Proceedings of the IGCP485 4th Meeting, Algiers, Algeria, 2 September 2006;
p. 57.
121. Ighid, L.; Saquaque, A.; Reuber, I. Plutons syn-cinématiques et la déformation panafricaine majeure dans le
Saghro oriental (boutonnière d’Imiter, Anti-Atlas, Maroc). Comptes Rendus de l’Académie des Sciences Série II
Mécanique-physique Chimie, Sciences de l’univers Sciences de la Terre 1989, 309, 615–620.
122. El Baghdadi, M.; El Boukhari, A.; Jouider, A.; Benyoucef, A.; Nadem, S. Calc-alkaline arc I-type granitoid
associated with S-type granite in the Pan-African belt of eastern Anti-Atlas (Saghro and Ougnat, South
Morocco). Gondwana Res. 2003, 6, 557–572. [CrossRef]
123. Errami, E.; Olivier, P. The Iknioun granodiorite, tectonic marker of Ediacaran SE-directed tangential
movements in the Eastern Anti-Atlas, Morocco. J. Afr. Earth Sci. 2012, 69, 1–12. [CrossRef]
124. Karl, A.; de Wall, H.; Rieger, M.; Schmitt, T.; Errami, E.; Kober, B.; Greiling, R.O. Petrography and
geochemistry of the Bou Teglimt, Taouzzakt and Igoudrane intrusions in the Eastern Saghro (Anti Atlas,
Morocco). In Magmatic evolution of a Neoproterozoic island-arc: Syn- to post-orogenic igneous activity in the
Anti-Atlas (Morocco); de Wall, H., Greiling, R.O., Eds.; Forschungszentrum Jülich, International Cooperation,
Scientific Series: Jülich, Germany, 2001; Volume 45, pp. 243–253.
125. Ouguir, H.; Macaudière, J.; Dagallier, G.; Qadrouci, A.; Leistel, J.-M. Cadre structural du gîte Ag-Hg d’Imiter
(Anti-Atlas, Maroc); implication métallogénique. Bulletin de la Société Géologique de France 1994, 165, 233–248.
126. Massironi, M.; Moratti, G.; Algouti, A.; Benvenuti, M.; Dal Piaz, G.V.; Eddebi, A.; El Boukhari, A.; Laftouhi, N.;
Ounaimi, H.; Schiavo, A.; et al. Carte géologique du Maroc au 1/50000, feuille Boumalne. Notes et Mémoires
du Service Géologique du Maroc 2007, 521.
127. Leistel, J.-M.; Qadrouci, A. Le gisement argentifère d’Imiter (Protérozoïque supérieur de l’Anti-Atlas, Maroc).
Contrôles des mineralisations, hypothèses génétiques et perspectives pour l’exploration. Chronique de la
Recherche Minière 1991, 502, 5–22.
128. Benkirane, Y. Les minéralisations à W (Sn, Mo, Au, Bi, Ag, Cu, Pb, Zn) du granite de Taourirt-Tamellalt
dans leur cadre géologique, la boutonnière protérozoïque du SE de Boumalne du Dadès (Saghro oriental,
Anti-Atlas, Maroc). In 3ème Cycle; Université de Paris VI: Paris, France, 1987.
147
Minerals 2018, 8, 592
129. Lécolle, M.; Derré, C.; Nerci, K. The Proterozoic sulphide alteration pipe of Sidi Flah and its host series. New
data for the geotectonic evolution of the Pan-African Belt in the eastern Anti-Atlas (Morocco). Ore Geol. Rev.
1991, 6, 501–536. [CrossRef]
130. Benziane, F. Lithostratigraphie et évolution géodynamique de l’anti-Atlas (Maroc) du paléoprotérozoïque au
néoprotérozoïque: Exemples de la boutonnière de Tagragra Tata et du Jebel Saghro. In 3ème Cycle; Université
de Chambéry: Chambéry, France, 2007.
131. Bajja, A. Volcanisme syn à post orogénique du Néoprotérozoïque de l’Anti-Atlas: Implications
pétrogénétiques et géodynamiques. Ph.D. Thesis, Université Chouaib Doukkali, El Jadida, Maroc, 1998.
132. Benharref, M. Le Précambrien de la boutonnière d’El Kelaa des M’Gouna (Saghro, Anti-Atlas, Maroc).
Pétrographie et structures de l’ensemble. Implications lithostratigraphiques et géodynamiques. In 3ème Cycle;
Université Cadi Ayyad: Marrakech, Maroc, 1991.
133. Bouladon, J.; Jouravsky, G. Les ignimbrites du Précambrien III de Tiouine et du sud marocain. Notes et
Mémoires du Service Géologique du Maroc 1954, 120, 37–59.
134. Fauvelet, E.; Hindermeyer, J. Note préliminaire sur les granites associés à des coulées rhyolitiques au Sud de
Ouarzazate (Anti-Atlas central) et dans le Sarho. C. R. Hebd. Seances Acad. Sci. 1952, 234, 2626–2628.
135. Mifdal, A.; Peucat, J. Datation U-Pb et Rb-Sr du volcanisme acide de l’Anti-Atlas marocain et du socle
sous-jacent dans la région de Ouarzazate. Apport au problème de la limite Précambrien-Cambrien. Sci. Géol.
Bull. 1985, 38, 185–200.
136. Acocella, V. Understanding caldera structure and development: An overview of analogue models compared
to natural calderas. Earth-Sci. Rev. 2007, 85, 125–160. [CrossRef]
137. Lipman, P.W. The roots of ash flow calderas in western north america: Windows into the tops of granitic
batholiths. J. Geophys. Res. Solid Earth 1984, 89, 8801–8841. [CrossRef]
138. Lipman, P.W. Subsidence of ash-flow calderas: Relation to caldera size and magma-chamber geometry.
Bull. Volcanol. 1997, 59, 198–218. [CrossRef]
139. Williams, H. Calderas and their origin. University of California publications. Bull. Dep. Geol. Sci. 1941, 25,
239–346.
140. Acocella, V.; Korme, T.; Salvini, F.; Funiciello, R. Elliptic calderas in the Ethiopian Rift: Control of pre-existing
structures. J. Volcanol. Geotherm. Res. 2003, 119, 189–203. [CrossRef]
141. Holohan, E.P.; Troll, V.R.; Walter, T.R.; Münn, S.; McDonnell, S.; Shipton, Z.K. Elliptical calderas in active
tectonic settings: An experimental approach. J. Volcanol. Geotherm. Res. 2005, 144, 119–136. [CrossRef]
142. Ross, C.S.; Smith, R.L. Ash-flow tuffs: Their origin, geologic relations and identification. Geol. Surv. Prof. Pap.
1961, 366, 81.
143. Smith, R.L. Ash flows. Geol. Soc. Am. Bull. 1960, 71, 795–842. [CrossRef]
144. Smith, R.L.; Bailey, R.A. Resurgent cauldrons. Geol. Soc. Am. Mem. 1968, 116, 613–662.
145. Bellier, O.; Sébrier, M. Relationship between tectonism and volcanism along the Great Sumatran Fault Zone
deduced by image analyses. Tectonophysics 1994, 233, 215–231. [CrossRef]
146. Chesner, C.A.; Rose, W.I. Stratigraphy of the Toba Tuffs and the evolution of the Toba Caldera Complex,
Sumatra, Indonesia. Bull. Volcanol. 1991, 53, 343–356. [CrossRef]
147. Ferrari, L.; Valencia-Moreno, M.; Bryan, S. Magmatism and tectonics of the Sierra Madre Occidental and its
relation with the evolution of the western margin of North America. Geol. Soc. Am. Spec. Pap. 2007, 422,
1–39.
148. Ferrari, L.; Lopez-Martinez, M.; Rosas-Elguera, J. Ignimbrite flare-up and deformation in the southern
Sierra Madre Occidental, western Mexico: Implications for the late subduction history of the Farallon plate.
Tectonics 2002, 21. [CrossRef]
149. Lécuyer, F.; Bellier, O.; Gourgaud, A.; Vincent, P.M. Tectonique active du Nord-Est de Sulawesi(Indonésie) et
contrôle structural de la caldeira de Tondano. Comptes Rendus de l’Academie des Sciences Ser. IIA Earth Planet.
Sci. 1997, 325, 607–613. [CrossRef]
150. Van Wyk de Vries, B.; Merle, O. Extension induced by volcanic loading in regional strike-slip zones. Geology
1998, 26, 983–986. [CrossRef]
148
Minerals 2018, 8, 592
151. Tuduri, J.; Chauvet, A.; Barbanson, L.; Labriki, M.; Badra, L. In Atypical gold mineralization within
the Neoproterozoic of Morocco. Structural and mineralogical constraints from the Thaghassa prospect
(Boumalne inlier, Jbel Saghro, Eastern Anti-Atlas). In Proceedings of the Mineral Exploration and Sustainable
Development, Athens, Greece, 24–28 August 2003; Eliopoulos, D.G., Ed.; Millpress: Athens, Greece;
pp. 537–540.
152. Goldstein, R.H.; Reynolds, T.J. Systematics of Fluid Inclusions in Diagenetic Minerals; Society for Sedimentary
Geology: Broken Arrow, OK, USA, 1994; Volume 31, p. 199.
153. Lécolle, M.; Derré, C.; Rjimati, E.C.; Fonteilles, M.; Azza, A.; Benanni, A. Une altération hydrothermale
peralumineuse à silicates, phosphates et rutile dans le Protérozoïque supérieur du Saghro (Anti-Atlas, Maroc).
Genèse et implications métallogéniques. Comptes Rendus de l’Académie des Sciences Série II Mécanique-physique
Chimie Sciences de l’univers Sciences de la Terre 1993, 316, 123–130.
154. Tuduri, J.; Dubois, M.; Try, E.; Chauvet, A.; Barbanson, L.; Ennaciri, A. The porphyry to epithermal transition
in atypical late Neoproterozoic REE-Au-Ag-Te occurrences. Acta Mineral.-Petrogr. Abstr. Ser. 2010, 6, 288.
155. Delapierre, A. Etude de la minéralisation aurifère d’Isamlal (Jbel Saghro, Anti-Atlas, Maroc). In Mem.
Diplôme; Université de Lausanne: Lausanne, Switzerland, 2000; p. 128.
156. Leloix, C. Etude des minéralisations aurifères épithermales d’Isamlal. District de Kelaat M’Gouna (Anti-Atlas,
Maroc). In Rapport Reminex; Université d’Orléans: Orléans, France, 1999; p. 44.
157. Sizaret, S. Etude des minéralisations aurifères d’Isamlal (district de Kelâa M’Gouna, Anti-Atlas, Maroc).
Master’s Thesis, Université d’Orléans, Orléans, France, 1999.
158. Gaspard, E. Etude du prospect d’Isamlal (Anti-Atlas Oriental- Maroc): Caractérisation d’un porphyre à
Au-Cu-Mo. Master’s Thesis, Université d’Orléans—ENAG, Orléans, France, 2014.
159. Try, E.; Dubois, M.; Tuduri, J.; Ventalon, S.; Potdevin, J.-L.; Chauvet, A.; Barbanson, L. The transition
between porphyric and epithermal styles: Insights from F.I. of the Kelâa M’Gouna prospect, Morocco.
In Proceedings of the ECROFI-XX 20th Biennial Conferences, Granada, Spain, 21–27 September 2009; Volume
20, pp. 261–262.
160. Tomczyk, C. Âge de mise en place et modèle génétique du stockwerk du prospect à Au-Ag-Te de Kelâa
M’Gouna (Maroc). Master’s Thesis, University of Lille, Lille, France, 2010.
161. Dong, G.; Morrison, G.; Jaireth, S. Quartz textures in epithermal veins, Queensland; classification, origin and
implication. Econ. Geol. 1995, 90, 1841–1856. [CrossRef]
162. Etoh, J.; Izawa, E.; Watanabe, K.; Taguchi, S.; Sekine, R. Bladed quartz and its relationship to gold
mineralization in the Hishikari low-sulfidation epithermal gold deposit, Japan. Econ. Geol. 2002, 97,
1841–1851. [CrossRef]
163. André-Mayer, A.-S.; Leroy, J.L.; Bailly, L.; Chauvet, A.; Marcoux, E.; Grancea, L.; Llosa, F.; Rosas, J. Boiling
and vertical mineralization zoning: A case study from the Apacheta low-sulfidation epithermal gold-silver
deposit, southern Peru. Mineral. Depos. 2002, 37, 452–464. [CrossRef]
164. Chauvet, A.; Bailly, L.; André, A.-S.; Monié, P.; Cassard, D.; Tajada, F.; Vargas, J.; Tuduri, J. Internal vein
texture and vein evolution of the epithermal Shila-Paula district, southern Peru. Mineral. Depos. 2006, 41,
387–410. [CrossRef]
165. Simmons, S.F.; Christenson, B.W. Origins of calcite in a boiling geothermal system. Am. J. Sci. 1994, 294,
361–400. [CrossRef]
166. Saule, A. La Zones des Dykes, Anti-Atlas Marocain: Caractérisation des fluides minéralisateurs et du
gisement. Master’s Thesis, Institut National Polytechnique de Lorraine, Nancy, France, 2012.
167. Albinson, T.; Norman, D.I.; Cole, D.; Chomiak, B. Controls on Formation of Low-Sulfidation Epithermal
Deposits in Mexico: Constraints from Fluid Inclusion and Stable Isotope Data. In New Mines and Discoveries
in Mexico and Central America; Society of Economic Geologists: Littleton, CO, USA, 2001; Volume 8, pp. 1–32.
168. Guillou, J.-J.; Monthel, J.; Picot, P.; Pillard, F.; Protas, J.; Samana, J.-C. L’imitérite, Ag2 HgS2 , nouvelle espèce
minérale; propriétés et structure cristalline. Bull. Mineral. 1985, 108, 457–464.
169. Guillou, J.-J.; Monthel, J.; Samama, J.-C.; Tijani, A. Morphologie et chronologie relative des associations
minérales du gisement mercuro-argentifère d’Imiter (Anti-Atlas—Maroc). Notes et Mémoires du Service
Géologique du Maroc 1988, 44, 215–228.
170. Levresse, G. Contribution à l’établissement d’un modèle génétique des gisements d’Imiter (Ag-Hg), Bou
Madine (Pb-Zn-Cu-Ag-Au), Bou Azzer (Co, Ni, As, Au, Ag) dans l’Anti-Atlas marocain. In 3ème Cycle;
Institut National Polytechnique de Lorraine: Nancy, France, 2001.
149
Minerals 2018, 8, 592
171. Baroudi, Z.; Beraaouz, E.H.; Rahimi, A.; Chouhaidi, M.Y. Minéralisations polymétalliques argentifères
d’Imiter (Jbel Saghro, Maroc): Minéralogie, évolution des fluides minéralisateurs et mécanismes de dépôt.
Chronique de la Recherche Minière 1999, 536–537, 91–111.
172. Hulin, C.; Dubois, M.; Tuduri, J.; Chauvet, A.; Boulvais, P.; Gaouzi, A.; Mouhajir, M.; Essalhi, M.;
Outhounjite, S. New fluid inclusions and oxygen isotope data to constrain a formation model for the
Imiter Ag world class deposit (Anti-Atlas, Morocco). In Proceedings of the ECROFI XXII 22nd Biennial
Conferences, Antalya, Turkey, 4–9 June 2013; pp. 78–79.
173. Hulin, C.; Dubois, M.; Tuduri, J.; Chauvet, A.; Boulvais, P.; Gaouzi, A.; Mouhajir, M.; Essalhi, M.;
Outhounjite, S. A fluid inclusion and stable isotope study of the world class Imiter silver deposit (Morocco).
In Proceedings of the 24ème Réunion des Sciences de la Terre, Pau, France, 27–31 October 2014; p. 373.
174. Tuduri, J.; Pourret, O.; Chauvet, A.; Barbanson, L.; Gaouzi, A.; Ennaciri, A. Rare earth elements as proxies
of supergene alteration processes from the giant Imiter silver deposit (Morocco). In Let’s Talk Ore Deposits,
Proceeding of the Eleventh Biennial SGA Meeting; Barra, F., Reich, M., Campos, E., Tornos, F., Eds.; Ediciones
Universidad Católica del Norte: Antofagasta, Chile, 2011; Volume 2, pp. 826–828.
175. Tuduri, J.; Pourret, O.; Boulvais, P.; Chauvet, A.; Barbanson, L.; Gaouzzi, A.; Hulin, C.; Dubois, M.
A reassessment of fluid-mineral relations in the world-class Imiter silver deposit (Anti-Atlas, Morocco).
In Proceedings of the SEG 2012 Conference, Lima, Peru, 23–26 September 2012; Society of Economic
Geologists: Lima, Peru, 2012.
176. Graybeal, F.T.; Vikre, P. A review of silver-rich mineral deposits and their metallogeny. In SEG Special
Publication: The Challenge of Finding New Mineral Resources: Global Metallogeny, Innovative Exploration, and New
Discoveries; Goldfarb, R.J., Marsh, E.E., Monecke, T., Eds.; Society of Economic Geologists: Littleton, CO,
USA, 2010; Volume 15, pp. 85–117.
177. Azizi Samir, M.R.; Ferrandini, J.; Tane, J.L. Tectonique et volcanisme tardi-Pan Africains (580-560 M.a.) dans
l’Anti-Atlas Central (Maroc): Interpretation geodynamique a l’echelle du NW de l’Afrique. J. Afr. Earth Sci.
1990, 10, 549–563. [CrossRef]
178. Ducea, M.N.; Paterson, S.R.; DeCelles, P.G. High-Volume Magmatic Events in Subduction Systems. Elements
2015, 11, 99–104. [CrossRef]
179. Harrison, R.W.; Yazidi, A.; Benziane, F.; Quick, J.E.; El Fahssi, A.; Stone, B.D.; Yazidi, M.; Saadane, A.;
Walsh, G.J.; Aleinikoff, J.N.; et al. Carte géologique au 1/50 000, Feuille Tizgui. Notes et Mémoires du Service
Géologique du Maroc 2008, 470, 131.
180. McQuarrie, N.; Barnes, J.B.; Ehlers, T.A. Geometric, kinematic, and erosional history of the central Andean
Plateau, Bolivia (15–17◦ S). Tectonics 2008, 27, TC3007. [CrossRef]
181. Till, A.B.; Roeske, S.; Sample, J.C.; Foster, D.A. Exhumation Associated with Continental Strike-Slip Fault Systems;
The Geological Society of America: Boulder, CO, USA, 2007; Volume 434, p. 264.
182. Willett, S.D.; Brandon, M.T. On steady states in mountain belts. Geology 2002, 30, 175–178. [CrossRef]
183. Hedenquist, J.W.; Lowenstern, J.B. The role of magmas in the formation of hydrothermal ore deposits. Nature
1994, 370, 519–527. [CrossRef]
184. Monier, G.; Robert, J.L. Muscovite solid solutions in the system K2 O, MgO, FeO, AL2 O3 , SiO2 , H2 O: An
experimental study at 2 kbar PH2O and comparison with natural Li-free white micas. Mineral. Mag. 1986, 50,
257–266. [CrossRef]
185. Cathelineau, M.; Nieva, D. A chlorite solid solution geothermometer: The Los Azufres (Mexico) geothermal
system. Contrib. Mineral. Petrol. 1985, 91, 235–244. [CrossRef]
186. Kranidiotis, P.; MacLean, W.H. Systematics of chlorite alteration at the Phelps Dodge massive sulfide deposit,
Matagami, Quebec. Econ. Geol. 1987, 82, 1898–1911. [CrossRef]
187. Kretschmar, U.; Scott, S.D. Phase relations involving arsenopyrite in the system Fe-As-S and their application.
Can. Mineral. 1976, 14, 364–386.
188. Sundblad, K.; Zachrisson, E.; Smeds, S.A.; Berglund, S.; Aalinder, C. Sphalerite geobarometry and
arsenopyrite geothermometry applied to metamorphosed sulfide ores in the Swedish Caledonides. Econ. Geol.
1984, 79, 1660–1668. [CrossRef]
189. Lynch, G.; Ortega, J. Hydrothermal alteration and tourmaline-albite equilibria at the Coxheat porphyry
Cu-Mo-Au deposit, Nova Scotia. Can. Mineral. 1997, 35, 79–94.
190. Sillitoe, R.H. Porphyry Copper Systems. Econ. Geol. 2010, 105, 3–41. [CrossRef]
150
Minerals 2018, 8, 592
191. Kouzmanov, K.; Pokrovski, G.S. Hydrothermal controls on metal distribution in porphyry Cu (-Mo-Au)
systems. In Geology and Genesis of Major Copper Deposits and Districts of the World: A Tribute to Richard H. Sillitoe;
Hedenquist, J.W., Harris, M., Camus, F., Eds.; Special Publications of the Society of Economic Geologists:
Littleton, CO, USA, 2012; Volume 16, pp. 573–618.
192. Rottier, B.; Kouzmanov, K.; Casanova, V.; Wälle, M.; Fontboté, L. Cyclic Dilution of Magmatic Metal-Rich
Hypersaline Fluids by Magmatic Low-Salinity Fluid: A Major Process Generating the Giant Epithermal
Polymetallic Deposit of Cerro de Pasco, Peru. Econ. Geol. 2018, 113, 825–856. [CrossRef]
193. Scott, S.; Driesner, T.; Weis, P. Boiling and condensation of saline geothermal fluids above magmatic intrusions.
Geophys. Res. Lett. 2017, 44, 1696–1705. [CrossRef]
194. Letsch, D.; Large, S.J.E.; Buechi, M.W.; Winkler, W.; von Quadt, A. Ediacaran glaciations of the west African
Craton—Evidence from Morocco. Precambrian Res. 2018, 310, 17–38. [CrossRef]
195. Pinneker, Y.V.; Lomonosov, I.S. Concentrated brines of Siberian Platform and their counterparts in Asia,
Europe, Africa and America. Int. Geol. Rev. 1968, 10, 431–442. [CrossRef]
196. Richard, A.; Pettke, T.; Cathelineau, M.; Boiron, M.-C.; Mercadier, J.; Cuney, M.; Derome, D. Brine–rock
interaction in the Athabasca basement (McArthur River U deposit, Canada): Consequences for fluid
chemistry and uranium uptake. Terra Nova 2010, 22, 303–308. [CrossRef]
197. Linnemann, U.; Pidal, A.P.; Hofmann, M.; Drost, K.; Quesada, C.; Gerdes, A.; Marko, L.; Gärtner, A.; Zieger, J.;
Ulrich, J.; et al. A ~565 Ma old glaciation in the Ediacaran of peri-Gondwanan West Africa. Int. J. Earth Sci.
2018, 107, 885–911. [CrossRef]
198. Vernhet, E.; Youbi, N.; Chellai, E.H.; Villeneuve, M.; El Archi, A. The Bou-Azzer glaciation: Evidence for
an Ediacaran glaciation on the West African Craton (Anti-Atlas, Morocco). Precambrian Res. 2012, 196–197.
[CrossRef]
199. Starinsky, A.; Katz, A. The formation of natural cryogenic brines. Geochim. Cosmochim. Acta 2003, 67,
1475–1484. [CrossRef]
200. Toner, J.D.; Catling, D.C.; Sletten, R.S. The geochemistry of Don Juan Pond: Evidence for a deep groundwater
flow system in Wright Valley, Antarctica. Earth Planet. Sci. Lett. 2017, 474, 190–197. [CrossRef]
201. Belkacim, S.; Ikenne, M.; Souhassou, M.; Elbasbas, A.; Toummite, A. The Cu-Mo±Au mineralizations
associated to the High-K calc-alkaline granitoids from Tifnoute valley (Siroua massif, anti-atlas, Morocco):
An arc-Type porphyry in the late neoproterozoic series. J. Environ. Earth Sci. 2014, 4, 90–106.
202. Loiselet, C.; Husson, L.; Braun, J. From longitudinal slab curvature to slab rheology. Geology 2009, 37, 747–750.
[CrossRef]
203. Manea, V.C.; Pérez-Gussinyé, M.; Manea, M. Chilean flat slab subduction controlled by overriding plate
thickness and trench rollback. Geology 2012, 40, 35–38. [CrossRef]
204. Merdith, A.S.; Collins, A.S.; Williams, S.E.; Pisarevsky, S.; Foden, J.D.; Archibald, D.B.; Blades, M.L.;
Alessio, B.L.; Armistead, S.; Plavsa, D.; et al. A full-plate global reconstruction of the Neoproterozoic.
Gondwana Res. 2017, 50, 84–134. [CrossRef]
205. Boyden, J.A.; Müller, R.D.; Gurnis, M.; Torsvik, T.H.; Clark, J.A.; Turner, M.; Ivey-Law, H.;
Watson, R.J.; Cannon, J.S. Next-generation plate-tectonic reconstructions using GPlates. In Geoinformatics:
Cyberinfrastructure for the Solid Earth Sciences; Keller, G.R., Baru, C., Eds.; Cambridge University Press:
Cambridge, UK, 2011; pp. 95–113.
206. Domeier, M. A plate tectonic scenario for the Iapetus and Rheic oceans. Gondwana Res. 2016, 36, 275–295.
[CrossRef]
207. Torsvik, T.H.; Cocks, L.R.M. Gondwana from top to base in space and time. Gondwana Res. 2013, 24, 999–1030.
[CrossRef]
208. Richards, J.P. Postsubduction porphyry Cu-Au and epithermal Au deposits: Products of remelting of
subduction-modified lithosphere. Geology 2009, 37, 247–250. [CrossRef]
209. Richards, J.P. Magmatic to hydrothermal metal fluxes in convergent and collided margins. Ore Geol. Rev.
2011, 40, 1–26. [CrossRef]
210. Sillitoe, R.H.; Hedenquist, J.W. Linkages between volcanotectonic settings, ore-fluid compositions and
epithermal precious metal deposits. In Volcanic, Geothermal and Ore-Forming Fluids; Rulers and Witnesses of
Processes within the Earth; Simmons, S.F., Graham, I., Eds.; Society of Economic Geologist Special Publication:
Littleton, CO, USA, 2003; Volume 10, pp. 315–343.
151
Minerals 2018, 8, 592
211. Tosdal, R.; Richards, J. Magmatic and structural controls on the development of porphyry Cu±Mo±Au
deposits. Rev. Econ. Geol. 2001, 14, 157–181.
212. Menant, A.; Jolivet, L.; Tuduri, J.; Loiselet, C.; Bertrand, G.; Guillou-Frottier, L. 3D subduction dynamics:
A first-order parameter of the transition from copper- to gold-rich deposits in the eastern Mediterranean
region. Ore Geol. Rev. 2018, 94, 118–135. [CrossRef]
213. Bryan, S.E.; Orozco-Esquivel, T.; Ferrari, L.; López-Martínez, M. Pulling apart the Mid to Late Cenozoic
magmatic record of the Gulf of California: Is there a Comondú Arc? Geol. Soc. Lond. Spec. Publ. 2013, 385.
[CrossRef]
214. Thorkelson, D.J.; Breitsprecher, K. Partial melting of slab window margins: Genesis of adakitic and
non-adakitic magmas. Lithos 2005, 79, 25–41. [CrossRef]
215. de Silva, S. Arc magmatism, calderas, and supervolcanoes. Geology 2008, 36, 671–672. [CrossRef]
216. Chauvet, A.; Alves Da Silva, F.C.; Faure, M.; Guerrot, C. Structural evolution of the Paleoproterozoic Rio
Itapicuru granite-greenstone belt (Bahia, Brazil): The role of synkinematic plutons in the regional tectonics.
Precambrian Res. 1997, 84, 139–162. [CrossRef]
217. Hickman, A.H. Two contrasting granite-greenstone terranes in the Pilbara Craton, Australia: Evidence for
vertical and horizontal tectonic regimes prior to 2900 Ma. Precambrian Res. 2004, 131, 153–172. [CrossRef]
218. Van Kranendonk, M.J.; Collins, W.J.; Hickman, A.; Pawley, M.J. Critical tests of vertical vs. horizontal
tectonic models for the Archaean East Pilbara Granite-Greenstone Terrane, Pilbara Craton, Western Australia.
Precambrian Res. 2004, 131, 173–211. [CrossRef]
219. Nance, R.D.; Murphy, J.B.; Strachan, R.A.; Keppie, J.D.; Gutiérrez-Alonso, G.; Fernández-Suárez, J.;
Quesada, C.; Linnemann, U.; D’lemos, R.; Pisarevsky, S.A. Neoproterozoic-early Palaeozoic tectonostratigraphy
and palaeogeography of the peri-Gondwanan terranes: Amazonian v. West African connections. Geol. Soc.
Lond. Spec. Publ. 2008, 297, 345–383. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
152
minerals
Article
Fault Zone Evolution and Development of
a Structural and Hydrological Barrier: The Quartz
Breccia in the Kiggavik Area (Nunavut, Canada)
and Its Control on Uranium Mineralization
Alexis Grare 1, *, Olivier Lacombe 1 , Julien Mercadier 2 , Antonio Benedicto 3 , Marie Guilcher 2 ,
Anna Trave 4 ID , Patrick Ledru 5 and John Robbins 5
1 Sorbonne Université, CNRS-INSU, Institut des Sciences de la Terre de Paris, ISTeP UMR 7193,
F-75005 Paris, France; [email protected]
2 Université de Lorraine, CNRS, CREGU, GeoRessources lab, 54506 Vandoeuvre-lès-Nancy, France;
[email protected] (J.M.); [email protected] (M.G.)
3 UMR Geops, Université Paris Sud, 91405 Orsay, France; [email protected]
4 Departament de Mineralogia, Universitat de Barcelona (UB), Petrologia i Geologia Aplicada,
Facultat de Ciències de la Terra, 08028 Barcelona, Spain; [email protected]
5 Orano Canada Inc., 817 45th Street, West Saskatoon, SK S7L 5X2, Canada; [email protected] (P.L.);
[email protected] (J.R.)
* Correspondence: [email protected]
Abstract: In the Kiggavik area (Nunavut, Canada), major fault zones along, or close to, where uranium
deposits are found are often associated with occurrence of thick quartz breccia (QB) bodies.
These bodies formed in an early stage (~1750 Ma) of the long-lasting tectonic history of the Archean
basement, and of the Proterozoic Thelon basin. The main characteristics of the QB are addressed
in this study; through field work, macro and microscopic observations, cathodoluminescence
microscopy, trace elements, and oxygen isotopic signatures of the quartz forming the QB. Faults
formed earlier during syn- to post-orogenic rifting (1850–1750 Ma) were subsequently reactivated,
and underwent cycles of cataclasis, pervasive silicification, hydraulic brecciation, and quartz
recrystallization. This was synchronous with the circulation of meteoric fluids mixing with
Si-rich magmatic-derived fluids at depth, and were coeval with the emplacement of the Kivalliq
igneous suite at 1750 Ma. These processes led to the emplacement of up to 30 m thick QB,
which behaved as a mechanically strong, transverse hydraulic barrier that localized later fracturing,
and compartmentalized/channelized vertical flow of uranium-bearing fluids after the deposition
of the Thelon Basin (post 1750 Ma). The development and locations of QB control the location of
uranium mineralization in the Kiggavik area.
1. Introduction
Fault zones are often associated with enhanced, focused, repeated fluid circulations in the earth’s
crust [1–7]. These fluids may have different origins: Meteoric, magmatic, metamorphic or basinal,
and possibly transport metals to a favorable area of deposition [8,9]; that will ultimately allow for
the formation of potential economic ore deposits. In many conceptual models of the formation
of ore deposits, fault zones are important structural features acting as pathways [2,10] and/or
as traps for fluids, and related metals [11]. In the uppermost crust, deformation is dominantly
brittle and breccias are commonly observed in fault zones [12–15]. Among the different families of
breccias, hydrothermal breccias are one sub-class that would develop early, in response to fracture
propagation processes [13], through interaction between brecciated rocks and hydrothermal solutions.
Hydrothermal breccias can be of various types depending on several parameters, such as pressure,
temperature, depth of emplacement, and elements in the fluids [14]. Among them, quartz-cemented
breccias can have an economic interest, being possibly associated with ore deposits such as epithermal
(Au-Ag-Cu-Pb-Zn-Sb, [16,17]), orogenic gold (Au, [18]), and porphyric (Cu-Mo-Au-Ag, [19,20]).
They display thickness from meter to several meters, thicker hydrothermal breccias being relatively
rarely described. Quartz breccias in fault zones form progressively during several cycles of fluid
pressure growth, seismogenic fault slip and quartz precipitation [21,22]. Unaltered, quartz-rich bodies
have a lowered porosity and thus have an impact on later fluid circulation within the fault zone.
Such silicification would be comparable to fluid-flow being constrained by horizontal barriers, such as
sedimentary layers indurated through diagenesis (aquitards, [23,24]), or impermeable (clay-rich) layers
in roll-front uranium deposits [25]. In addition, the likely hardening of the fault rocks in response to
multiple cycles of quartz brecciation and healing may cause a significant rheological contrast between
the “strong” fault zone and the expectedly “weaker” hosting terranes, possibly controlling localization
of subsequent deformation.
In this contribution, we focus on one structural feature encountered in many fault zones within
the Uranium (U)-rich district of the Kiggavik area (Nunavut, Canada): The so-called hydrothermal
Quartz Breccia (QB). The importance of this breccia, only briefly described by previous authors [26–30]
was recently highlighted by Grare et al. [31] who documented the control exerted by this breccia on
later fracturing events, hydrothermal alterations and uranium mineralization at the Contact uranium
prospect. However, despite observations in several locations of the Kiggavik area and its seemingly
strong control on the current distribution of the uranium mineralization, the genetic model of the QB
remains poorly characterized and explained to date. Grare et al. [31] showed that the QB emplaced
along faults of inferred Archean age, and that this emplacement was a key event within a long-lasting
(~1000 Ma) complex brittle tectonic history that led to uranium mineralization within or in the vicinity
of the quartz breccia (Figure 1C). In order to better constrain the nature, emplacement, significance
and role of the QB, we carried out a structural analysis combined with vein cement petrography using
optical and cathodoluminescence observations, trace elements, and oxygen stable isotope analysis of
quartz. Our study addresses the structural, mineralogical and geochemical characteristics of the QB.
Combined with the reconstructed geochemical signature of the fluids, a model of formation of the QB
is proposed and its role in controlling uranium mineralization in the Kiggavik area is highlighted.
2. Geological Setting
154
Minerals 2018, 8, 319
Figure 1. (A) Outline of Canada and location of the Thelon basin in yellow; (B) geological map of
the Churchill-Wyoming craton showing the location of the Thelon basins and the Kiggavik area on
its Eastern border; (C) simplified geological map of the Kiggavik area (Orano internal document)
highlighting the occurrence of the QB (yellow) along the major faults; and (D) cross-section from the
Thelon fault to the Judge Sisson fault. Deposits and prospects are indicated with red circles.
155
Minerals 2018, 8, 319
followed by uplift, extensive erosional peneplanation and regolith formation, over which deposited
the eolian sandstones and conglomeratic red-beds of the Thelon formation (ca. 1670–1540 Ma [32,33]),
linked to thermal subsidence in the sag, fault-controlled intracratonic Thelon basin [36,38,39].
This volcano-sedimentary pile unconformably overlies a metamorphosed basement consisting of
Archean rocks that include Mesoarchean (ca. 2870 Ma) granitic gneisses, 2730–2680 Ma, supracrustal
rocks of the Woodburn Lake Group [40], and a distinctive package of 2620–2580 Ma felsic volcanic and
related hypabyssal rocks known as the Snow Island Suite [41–47].
Before emplacement of the Thelon formation, the Archean to Paleoproterozoic rocks of the
Churchill province where intruded by three magmatic suites: (i) The late syn-orogenic (ca. 1830 Ma)
Hudson Suite [48], (ii) the Dubawnt Minette Suite (contemporaneous of the Hudson Suite),
with ultrapotassic intrusions, minette dikes and lamprophyres, and (iii) the anorogenic (ca. 1750 Ma)
Kivalliq Igneous Suite (KIS) [46,49–51].
156
Minerals 2018, 8, 319
Figure 2. (A) Outcrop view looking east on the N80-trending steeply dipping to the north Judge
Sisson fault (JSF) underlain by at least 10 m of white quartz veins; (B) heterogeneous size, pervasively
hematized clasts cemented by a white quartz matrix; (C) right lateral relay step, N80 trending main
veins (outcrop on the JSF); (D) optical microphotograph picture (OM): Clasts bearing quartz veins in the
Thelon sandstones; (E) oriented data of thick quartz veins for deposits and prospects; and (F) histogram
of all measured quartz vein dips in the Kiggavik area.
157
Minerals 2018, 8, 319
deposits. All samples were studied from the macro- to the micro-scale in order to characterize the
macroscopic texture of the quartz breccia and its relationships with predating and postdating fracturing
and faulting events. Thirty-five thin sections were prepared for petrographic and microstructural
studies. Thin sections were observed through optical microscopy (plane polarized transmitted and
reflected light microscope Motic BA310 POL Trinocular, equipped with a 5 M pixel Moticam camera)
(Motic Instruments Inc., Richmond, BC, Canada), and cathodoluminescence microscopy (CITL Cold
Cathodoluminescence device Model MK5-1, made at University of Barcelona (Barcelona, Spain),
for deciphering quartz generations.
set-up the instrument and correct for drifts and fractionations using a standard bracketing approach.
The internal precision for δ18 O was between 0.06 and 0.1‰ (measurements on the standards Brésil
and Brésil-2 and on the different quartz generations of Kiggavik). δ18 O values are reported relative to
the V-SMOW standard.
4. Results
158
Minerals 2018, 8, 319
(Figure 2D), indicating that QB predates formation of the Thelon Basin, as already suggested by several
authors [29,31] and crosscuts, thus postdates, Hudsonian intrusions (ca. 1.83 Ga). Fault zones outlined
by the QB are presumably better preserved in the field due to the silicification process that increases
their resistance to erosion.
The outcrop shown in Figure 2B illustrates the complexity of the identification of the main
structural trends on limited exposures. We considered that the most regionally significant structural
trend of the breccia bodies is given by the thicker (>10 cm) veins and breccias, because where they
are visible, minor quartz veins are more randomly oriented or give a mean statistical value that is
different between two (2) nearby drill holes. By plotting the orientations of thick veins we infer the
true orientation of the quartz breccia (Figure 2E), which was revealed to be consistent with the major
fault trends in map view (Figure 1D). The QB usually displays a consistent high angle dip, reflecting
the orientation of the main fault trend: N30, dip to the NW at Contact, N175, dip to the W at Bong,
N50 and N90, dip to the NW and to the S, respectively, at End (Figure 2E). Even though the majority of
minor quartz veins display throughout the Kiggavik area a steep dip (60–90◦ ), a significant amount of
veins (Figure 2F) shows relatively shallow dip angles (<30◦ ).
Figure 3 summarizes the data collected on drill holes at the Contact prospect (Figure 3A). The QB
bodies usually display two main distinct zones, an outer zone and an inner (core) zone. The outer
zone (blue in Figure 3B) is represented by a dense to scarce network of millimeter to centimeter-thick
quartz veins, while the inner (core) zone (red in Figure 3B) is represented by thick (>10 cm thick) quartz
veins and a dense quartz vein network, where angular clasts of the fragmented host rock are barely
observable. Several QB core zones were crosscut by drill holes (Cont-24, Cont-16, Cont-06). These core
zones are discontinuous from the SW to the NE. They are tapering toward their ends (Figure 3C) both
laterally (for example, between Cont-26 and Cont-25, Figure 3B), and vertically (for example, between
Cont-10 and Cont-11, Figure 3B). This supports that they have elliptical shapes, connected by quartz
vein networks. This observation explains the important changes in thickness of the QB between two
nearby drill holes (e.g., Cont-06 and Cont-13).
Figure 3. Organisation of inner (core) and outer zones of the quartz breccia (QB) crosscut in drillholes
at Contact. (A) Plan view of the drill holes; (B) lateral variation in thickness of QB inner (core) and
outer zones; and (C) simplified interpretative drawing of the QB intersected in drill holes (grey plane).
159
Minerals 2018, 8, 319
One observation not highlighted by previous studies in the Kiggavik area is the presence of
a large (20–100 m) brittle fault zone predating emplacement of the QB but systematically spatially
associated with it. Macroscopically, the QB consists of thin to massive quartz veins as described in
Figure 2; however, our detailed observations document numerous quartz healing events crosscutting
clay-altered cataclastic to ultra-cataclastic fault rocks that are now silicified and “preserved”. Clasts
are monomictic, sub-rounded, millimetric to centimetric in size and clay altered, embeded in a light
red to brown matrix (Figure 4A,B).
Figure 4. (A) Pervasively silicified cataclastic fault rock; (B) same as (A), crosscut by a white quartz
vein of the QB; (C) pervasively silicified fault zone crosscut by late fracturing and clay alteration event
(End deposit); and (D) typical intersection of the QB displaying deep purple hematized rock, massive
and minor white quartz veins. Jigsaw textures are locally observable (e.g., at 189 m, yellow arrow;
Contact prospect).
The quartz veins of the QB were observed in several locations as cutting across the cataclasites
(Figure 4B). These early cataclastic fault rocks therefore predate the QB; they could be related to
extensional to trans-tensional faulting during formation of the Baker Lake Basin [31]. This early,
160
Minerals 2018, 8, 319
now silicified fault zones and the QB are spatially associated, indicating that the pervasive silicification
likely occurred at the onset of emplacement of the QB. However, even though the pervasive silicification
of the fault zone is spatially and likely roughly temporally associated with the QB, we differentiate
hereafter these two features: The silicified fault zone on one hand and the QB that results from
brecciation sealed by quartz on the other hand.
Both features display different thicknesses: In Figure 4C, the pervasively silicified fault zone with
its light reddish color is observable along 40 m of drill core and is cut by numerous small quartz veins
and a 4 m thick core zone of the QB. A late faulting and white clay alteration pattern is observed at
depth 389–395 m (Figure 4C, post ore faulting f7). In Figure 4D, the silicified fault zone is observed
along 5 m of drill core and is cut by 23 m of QB.
The pre-QB silicified fault zone displays evidence of multiple events of tectonic brecciation
and comminution. In the sample observed at micro-scale under transmitted light (Figure 5A),
three generations of cataclastic fault rocks are observed, with each generation of cataclasis consuming
the previous one. They are crosscut by at least three generations of quartz veins, building a complex
pattern (Figure 5B,C). Minerals from the original host rock (psammo-pelitic gneiss with quartz, apatite,
illite, muscovite, pyrite) are preserved in the first generation of clasts (pink, Figure 5B). A closer look at
the cataclastic fault rocks reveals that the different cements are made of micro-crystalline quartz and
white micas (Figure 5D,E). The superimposition of multiple generations of cataclasites indicates that
the localized zone of deformation was repeatedly reactivated during progressive deformation.
Figure 5. (A,B) Thin section of a polyphase cataclastic fault rock crosscut by several generations of
quartz veins of the QB. White arrow indicates a late microcrystalline quartz veinlet; (C) simplified
chronology of the events; (D) zoom on the different generations of clasts; and (E) matrix of the latest
cataclastic event displaying white micas and micro-crystalline quartz.
161
Minerals 2018, 8, 319
In order to better understand and characterize the influence of silicification on fluid circulation,
we selected porosity data measured in the field for four types of rocks: Fresh host rock (granitic
gneiss, before fracturing and alteration), silicified type 1 (pervasively silicified fault zone), silicified
type 2 (typical white QB), and clay-altered/fractured samples. Results are presented in Figure 6.
Fresh granitic gneiss yields the lowest porosity values, <2%. Fault rock and samples displaying quartz
brecciation and pervasive silicification yield values slightly higher but <5%. Fractured and clay altered
fault rock display much higher values, up to 40%. Cataclastic fault rocks formed before the QB should
have displayed a high porosity, but after pervasive silicification they have a porosity comparable to
fresh rock (Figure 6), unlike strongly clay altered and fractured samples (Figure 6).
Figure 6. Porosity measured for fresh samples (granitic gneiss from the Contact prospect), pervasively
silicified cataclastic fault rock, thick quartz veins within granitic gneiss, and clay-altered, fractured host
rock (examples for each category are displayed on the right of the chart).
162
Minerals 2018, 8, 319
Figure 7. Optical microscope microphotograph (OM): (A) Disseminated hematite (Hem) and specular
hematite (Spec Hem). Qtz: Quartz; (B) banded microcrystalline quartz (Qtz) with synchronous anhedral
hematite and magnetite; (C) euhedral quartz crystals and arrays of dense monophase fluid inclusions
(vapor rich); (D) euhedral clear quartz cement a fracture that crosscuts previous quartz generations;
(E) trends of microcrystalline quartz (yellow); (F) comb quartz grains (example in yellow) engulfed
in a fine-grained quartz matrix; (G) moss quartz texture; and (H) bladed lattice calcite (white arrow)
replaced by quartz.
163
Minerals 2018, 8, 319
Microscopic observations also document a variety of quartz textures (Figure 7C–F for example),
mutually crosscutting each other, and defining different conditions of quartz precipitation. The two
most common types of quartz are: Euhedral quartz (comb quartz, ~200 μm in size) and microcrystalline
quartz (~<50 μm). Two generations of euhedral white quartz can be distinguished: One (millimetric
quartz) being characterized by dense arrays of monophase fluid inclusions (vapor rich), usually at
the tip of the quartz crystal (Figure 7C); and the second, clearer, nearly fluid inclusions-free, usually
observed as a late quartz generation (~100 μm, Figure 7D). In addition to the banded microcrystalline
quartz-hematite texture, microcrystalline quartz is also observed filling vugs, and as conjugate “trends”
(Figure 7E) in subhedral quartz veins. In other samples, subhedral quartz can be found as clasts in
a microcrystalline quartz mass (Figure 7F). In term of quartz texture, comb quart, microcrystalline
quartz and “moss” textures were observed (Figure 7G). Additionally, rare recrystallized bladed calcite
were found (Figure 7H).
Quartz observed under cathodoluminescence display weak luminescence intensity, with a 20 s
exposure time required in order to get enough signal for imaging. The most recurrent color observed
under cathodoluminescence is a deep blue observed for microcrystalline quartz veins, sometimes
synchronous with hematite (Figure 8A), and quartz cementing microbreccias. The fluid inclusion
(FI)-rich euhedral quartz crystals exhibit alternating growth zones of brown and blue luminescence
(oscillatory growth-zoning, Figure 8B). Clasts of euhedral quartz crystals are found within a blue
luminescent quartz matrix (Figure 8C). The brown luminescence is also observed in breccias where
the quartz has likely completely recrystallized, leaving the breccia texture only observable under
cathodoluminescence; the “cement” of the breccia displays a brown luminescence (Figure 8D).
These colors characterize the main generations of quartz in the QB.
The luminescence of the latest generation of quartz (i.e., euhedral quartz filling vugs and open
fractures), is dark blue with rare concentric zoning. It also displays greenish luminescence associated
with primary to pseudo-secondary fluid inclusions (Figure 8E). In terms of spatial occurrence of this
quartz generation, it is more frequently observed in the vicinity of the QB than in its inner zone.
Quartz which was formerly in contact with uranium minerals displays a characteristic
luminescence: Red/pink close to uranium-bearing minerals and yellow/greenish further from the
uranium-bearing mineral (Figure 8F). This is especially well observed in quartz veins that were later
microfractured as described by Grare et al. [31] and in quartz of the host rock (when not dissolved
by circulation of the uranium-bearing fluid). This peculiar luminescence is brighter than the original
luminescence of the quartz and displays a nearly uniform circular shape of 35–45 μm width (Figure 8F).
The latest quartz generation, which fills vugs and open fractures, is characterized by dark blue
luminescence with rare concentric zoning. This generation is observed more in the vicinity of the QB.
164
Minerals 2018, 8, 319
165
Minerals 2018, 8, 319
Figure 9. Concentration of trace elements measured in quartz through LA-ICP-MS, for main quartz
generations: (A) Li, K, and Ba; (B) Zr, Ti, and B; (C) Mg, Na, Al, Fe, and Ca; and (D) Al vs. Ti
concentrations of main quartz generation. Zones correspond to values of hydrothermal quartz from
low T ◦ C, orogenic Au, and porphyry-type deposits [60].
In some analyses, trace elements display extreme values above 10,000 ppm (e.g., Al or Mg,
Figure 9) which likely represent analysis of undetected solid inclusions, hence are not displayed in
Figure 9. The high Fe content in both the deep blue microcrystalline and brown-blue euhedral quartz
could reflect the analysis of micro-inclusions of iron oxides, related to the pervasive hematization
synchronous with the QB event. Microcrystalline quartz in banded veins associated with iron oxides
shows a small range of values for all elements except for K and Fe. Brown-blue euhedral quartz
displays bimodal concentrations for most of the elements consistent with observed concentric zoning.
For all quartz generations, Li, K, and Na are positively correlated with Al. Dark blue vuggy quartz
usually display a small range of values for most elements compared to other quartz generation, except
for Li (23–248 ppm) and Al (255–2593 ppm).
Li contents are homogeneous between the three quartz generations and are below 250 ppm.
Such values correlate positively with Al concentrations, Li balancing the replacement of Si by
Al [61]. K is enriched in deep blue microcystalline quartz with values up to 1700 ppm, compared to
166
Minerals 2018, 8, 319
concentrations below 100 ppm in the case of the two other quartz generations. Fe yields high values
(up to 7000 ppm) in the case of the two quartz generations of the QB (deep blue microcrystalline
and brown-blue euhedral quartz). B displays concentrations below 20 ppm except for 2 analysis.
Concentrations in B are lower in the case of post-QB dark blue vuggy quartz.
Deep blue microcrystalline and dark blue vuggy quartz yield low values of Ti (<20 ppm for most
measurements). Bi-modal concentrations of Ti were measured for brown-blue euhedral quartz, with
one group of values below 20 ppm and the other above 40 ppm.
Table 1. δ18 Oquartz measured in main quartz generations and calculated values of δ18 Ofluid . To calculate
δ18 Ofluid , an average temperature of 250 ◦ C was used for quartz generations of the QB (lines 1–3 of the
table), while an average temperature of 150 ◦ C was used for late druzy quartz (lines 4–5).
Quartz Type CL Luminescence Color δ18 Oquartz Avg. n δ18 Ofluid Avg.
Banded microcrystalline quartz
Deep blue 12.0–14.0 12.9 14 2.4–5.3 3.9
(alternated with iron oxides)
Alternating blue and brown
Euhedral quartz with concentric zoning 7.5–9.3 8.4 18 −1.6–0.3 −0.6
luminescence
“Late” microcrystalline quartz Deep blue 18.8–23.9 22.1 10 11.5–14.8 12.4
Vuggy quartz Dark blue 14.4–15.5 14.9 29 −5.3–2.5 −3.3
Quartz alteration associated with fluid
Green 16.2–22.0 17.8 5 −2.5–3.8 −0.3
inclusions
δ18 Oquartz and temperatures measured by fluid inclusion microthermometry in quartz veins in
the area [26,29,62] were used to calculate the δ18 Ofluid following the equation of Clayton et al. [63],
set for measuring oxygen isotope exchange between quartz and water (assuming that the fluid was
in equilibrium with the quartz at the temperature of mineralisation). We considered homogenisation
temperatures in the range of 200–300 ◦ C, avg. 250 ◦ C (i.e., the range of temperatures revealed
by low salinity fluid inclusions), to be representative for the quartz generations of the QB and of
100–200 ◦ C, avg. 150 ◦ C (i.e., the range of temperature revealed by high salinity fluid inclusions),
to be representative for late druzy quartz probably precipitating from basinal brines [29]. However,
a microthermometric study on primary fluid inclusions for each quartz generation is missing actually
and would give a more accurate calculation of fluid isotopic values. Results are displayed in
Table 1. The δ18 Ofluid value range from 2.4‰ to 5.3‰ (+3.9‰ on average) for microcrystalline
quartz associated with hematite. In contrast, late veinlets of micro-crystalline quartz display a much
higher δ18 Ofluid value: Between +11.5‰ and +14.8‰ (+12.4‰ on average). Brown-blue quartz
precipitated from a fluid with a lighter δ18 Ofluid value comprised between −1.6‰ and +0.3‰ (−0.6‰
167
Minerals 2018, 8, 319
on average). Late vug-filling euhedral quartz yield lighter isotopic values from −5.3‰ to −3.3‰
(−3.3‰ on average).
168
Minerals 2018, 8, 319
be explained by a local effect of rock buffering within breccia cavity. As a result, veinlets would have
formed from an isotopically isolated fluid reservoir, thus yielding higher δ18 Oquartz values.
Cathodoluminescence observations support the above interpretations, as blue-purple quartz
luminescence is commonly found in quartz precipitated in magmatic/hydrothermal environments [75]
while brown luminescence is rather observed in sedimentary-diagenetic (i.e., lower temperature)
environments [76]).
The very low B content of the post-QB dark blue vuggy quartz could be explained by
co-crystallization of other minerals enriched in B in the uranium deposits of the Kiggavik area; such as
magnesiofoitite (dravite), which is commonly observed in environments seeing brine circulations [77].
Accordingly, the isotopic values for vug-filling euhedral quartz (post-QB) are consistent with those
obtained for quartz precipitated from relatively low temperature (~150 ◦ C) brines in the Proterozoic
Athabasca [78] and Kombolgie basins [79,80] and linked to the formation of U deposits. Considering
that the QB predates the formation of the Thelon Basin, from which brines are likely derived, this
supports that brines circulated after the emplacement of the QB, in agreement with the findings of
Grare et al. [31]. In the Kiggavik area, the fluid inclusion studies by Pagel [26] on the hydrothermal
quartz, at Andrew Lake, and by Chi et al. [29] on the hydrothermal quartz, at End, are in agreement
with our observations. Indeed, these authors documented low temperature (100–200 ◦ C)-high salinity
(25–38 wt % NaCl) fluids, low temperature (150–200 ◦ C)-low salinity (<9 wt % NaCl) fluids, and high
temperature (200–300 ◦ C)-low salinity (<9 wt % NaCl) fluids within the QB. Our study indicates that
low-T ◦ C/high salinity fluids (brines) circulated after formation of the QB, while the high-T ◦ C/low
salinity fluids are more characteristic of the QB that formed earlier in the history of the Kiggavik area.
Using data from several deposits type, it has been shown that deposits linked to low (Mississippy
Valley type, Carlin, Epithermal) and high (porphyry Cu-Au) fluid temperatures can be distinguished
one from another based on Al and Ti concentrations in quartz associated with orebodies [60].
Figure 9D plots Al-Ti concentrations for the main quartz generations in our study. The epithermal
domain of Figure 9D was built after data from low temperature (~100–350 ◦ C) hydrothermal fluids.
The distribution is scattered even within one quartz generation (e.g., deep blue microcrystalline).
Primary (comb quartz) and secondary (“moss” quartz) textures [81,82] indicate primary quartz
deposition and recrystallization. Bobis [83] also attributed the rounded shapes of the moss texture
to recrystallisation of silica gel, which preserved the original structure and impurities of the silicate
phase. These quartz textures, along with recrystallized bladed calcite, also characterize phases of silica
precipitation by boiling and non-boiling hydrothermal fluids in a geothermal/epithermal system [84].
Some textures observed within the QB are typical of epithermal deposits, but they are rare. Commonly
encountered precious metals (e.g., Au, Ag) are lacking within the breccia even though they were
observed at the nearby Mallery Lake deposit [62]. However, such environment of formation is
consistent with the geochemical signature of quartz in the QB and is much more plausible for the
formation of the QB than orogenic Au and porphyric deposits.
To sum up, even though it is difficult to be truly conclusive with the measured trace elements
concentrations only, the combination of these data with oxygen isotope values and quartz textures
points toward a scenario in which high (magmatic-derived) and low (meteoric-derived) temperature
fluids interacted and mixed during silicification of the fault zone that led to the formation of the QB.
The important volumes of Si would have been provided by intrusive bodies of the KIS emplaced at
depth and related to the rift-related extensional tectonics that occurred at ca. 1750 Ma (Figure 10).
169
Minerals 2018, 8, 319
Figure 10. Cross-section after Peterson et al. [46] and zoom in the zone of formation of the QB.
5.2. Fault Zone Processes Leading to the Formation of the QB: Cataclasis, Silicification and
Hydraulic Brecciation
Macroscopic observations of the QB and petrographic and textural observations on quartz,
although lacking a simple and clear chronology of events, provide additional constraints on
the processes behind its formation. Before emplacement of quartz cemented veins and breccias,
the superimposition of multiple generations of cataclasites indicates that a localized zone
of deformation was repeatedly reactivated during progressive deformation. The presence of
microcrystalline quartz in clasts generated before emplacement of quartz veins show that the pervasive
silicification of the fault zone was a syn-tectonic process. Regarding quartz-cemented fractures, the
common macroscopic textural observation of quartz cemented breccia with angular fragments and
jigsaw pattern indicates hydraulic brecciation [14] of the host rock. The pervasive silicification of the
fault zone was a first step (Figure 11A,B) before emplacement of the quartz veins and breccias of the
QB: It likely triggered fluid pressure build up in the fault zone leading to hydraulic brecciation of the
host rock, hence to the “building” of the so-called QB.
The complex patchwork of quartz textures observed under optical microscope shows a still
more complex pattern under cathodoluminescence, but highlights several events of quartz fracturing
(reworked quartz clasts) and recrystallization. The white quartz mass which displays fine-grained
subhedral quartz crystals also shows in some locations numerous fragments of earlier aggregates.
The conjugate trends of microcrystalline quartz likely reflect shearing in the quartz mass and
synchronous quartz recrystallization. A better evidence for such fracturing and synchronous quartz
crystallization is provided by white quartz veins in which euhedral quartz grains are surrounded
by microcrystalline quartz and other quartz of heterogeneous sizes (Figure 7F). We interpret this as
cataclasis and tectonic comminution (i.e., fracture propagation and wear abrasion) of previously formed
quartz mass and recrystallization of quartz (i.e., syn-tectonic). This process differs from the formation
of sub-horizontal quartz veins and hydraulic breccias related to transient fluid overpressurization
followed by fluid pressure drop and quartz precipitation [21].
To summarize, textures and crosscutting relationships of quartz cements reveal the following
sequence of events: (1) Episodes of brittle faulting and cataclasis, before silicification and
quartz-brecciation; (2) pervasive silicification of the fault zones (beginning of the QB event);
and (3) episodes of brittle fracturing synchronous with the circulation of silica-rich fluids (QB event).
During this last event, there were stages of hydrothermal hydraulic brecciation with slow and rapid
silica precipitation in relation to boiling of magmatic and/or meteoric fluids (trace elements and
δ18 O data inconclusive). This boiling process is supported by the monophase fluid inclusions within
the quartz generations of the QB. The presence of 100% of monophase vapor inclusions can be only
explained by a boiling process, affecting either magmatic fluids (with a spatial separation between
170
Minerals 2018, 8, 319
vapor and brines) or meteoric fluids heated due to emplacement of a magmatic intrusion at low depth
at ca. 1750 Ma. Boiling process can be marked in other geological environments by the presence of
monophase vapor fluid inclusions spatially associated with multiphase and of relatively high-salinity
brines due to demixion of the magmatic fluids. The absence of two-phase fluid inclusion in the
observed samples of the QB could indicate that the vapor migrated farther than the magmatic brines.
The hydraulic brecciation alternated with stages of fluid-assisted cataclasis and quartz recrystallization
(Figure 11A). Arrays of monophase fluid inclusions (vapor-rich) also indicate abrupt pressure drops
following rupture of the “seal” of the system [78]. The so-called QB therefore appears to be a composite
structural feature much more complex than previously thought, which consists of a mass of quartz
emplaced by alternating quartz healed hydraulic brecciation and tectonic-induced cataclasis with
synchronous quartz recrystallization during fault zone reactivation.
Figure 11. (A) Evolution of inferred fault strength (frictional shear resistance) and fluid pressure in the
fault zone as a response of fluid pulse, fracturing and silica-precipitation. (B) Scheme depicting the
succession of events that produced the QB.
Intense multi-episodic hydraulic brecciation of the early fault zone at the time of QB formation
would have occurred during the interaction of two isotopically distinct fluids: Meteoric water,
mixed with a magmatic-derived fluid. The processes of faulting/fracturing discussed in this section,
that led to the formation of the QB, likely occurred at shallow depth (~2 km, [29]) and, looking at
relative chronology and geochemical constraints, were likely initiated by the emplacement of the
KIS (Figure 10). To a first glance, the fluid temperature of ~350 ◦ C is not easy to reconcile with this
shallow depth even if considering an abnormal thermal gradient related to the emplacement of the
Kivalliq intrusions. We infer that hydrothermal fluids originated at a greater depth (about 5 km,
171
Minerals 2018, 8, 319
which may indicate a 70◦ /km geothermal gradient), and flowed upward sufficiently fast to prevent
any significant cooling before they mixed with downward-moving meteoric fluids and precipitated the
quartz generations of the QB in thermal disequilibrium with the hosting basement rocks. Interestingly,
such a quartz cemented breccia and its complex spatial organization are similar to the meter-thick
hydrothermal quartz breccia related to the emplacement of an igneous intrusion described by Tanner
et al. [85] in Scotland.
We therefore propose a conceptual tectono-hydrological model for the QB formation involving
mixing of deep silica-rich fluids of igneous origin with downward-moving meteoric fluids. The possible
mechanisms allowing for such meteoric fluid downward flow in fault zones are either active seismic
pumping or passive meteoric infiltration throughout a permeable fault zone. We favor a mechanism of
syn-tectonic seismic pumping because beside the formation of quartz-filled fractures, the intrinsic low
permeability of the unaltered basement rock surrounding the fault zone and the impermeabilization of
the fault zone—including its damage zone—through multiple silicification events presumably make
a simple, gravity-driven downflow of meteoric fluids difficult, hence unlikely. Fluid pressure built
up at depth through the input of meteoric fluids and magmatic-derived fluids together with likely
pore cementation of basement rocks (that decreased porosity). Fluid mixed and flowed upward to
higher crustal levels along the fault zones which served as conduits (Figures 10 and 11). This upward
flow likely occurred cyclically as the fluid pressure evolved between hydrostatic and supralithostatic
(Figure 11A), depending, among other factors, on the sealing effectivity of the reactivated fault zone
by quartz precipitation [86]. In turn, fluid pressure increased during the QB event also likely favored
multiple reactivations of the high angle fault zone under the regional stress field.
5.3. Evolution of the Fault Zone Properties though Time and Structural Control on Later
Uranium Mineralization
The QB is found along many segments of the main fault (Figure 1C) trends, and uranium orebodies
are systematically spatially associated with more or less thick bodies of QB along these fault zones.
Even though a systematical study of QB thickness could not be undertaken throughout the area, the
QB was observed as being usually thinner where it is not associated with uranium mineralization,
which implies a possible control on later uranium mineralization by the thickness of the QB in
fault zones.
Cataclastic fault rocks formed before the QB should have displayed an initial high porosity,
but after pervasive silicification they likely ended with a low porosity comparable to that of the fresh
basement rocks, unlike strongly clay altered and fractured samples (Figure 6). Since the evolution of
the porosity can be to some extent directly linked to the evolution of permeability since it is controlled
by fracturing and mineralogical destabilization/dissolution, we can safely infer that these multiple
events of pervasive silicification, faulting/fracturing and quartz cementation caused the destruction
of the porosity (hence of the permeability) and thus directly impacted the fluid circulation within
the conduit.
At all deposits and prospects in Kiggavik, three main fracturing events postdate emplacement of
the QB (two stages of faulting/fracturing and uranium mineralization and one stage of faulting
and strong clay alteration [31]). The distribution of fractures and mineralization in some drill
holes intersecting uranium orebodies in the vicinity of the QB is shown in Figure 12A for Contact,
End, Bong and Andrew Lake. Post-QB fracturing and uranium mineralization are clearly restricted
to the hanging wall of the QB in Contact, where the thickness of the breccia is far greater compared
to the earlier silicified fault zone. At End, Andrew Lake and Bong, post-QB fracturing and uranium
mineralization are observed in both the hanging wall and the footwall, but still not within the QB
(inner zone). In the case of End, the QB displays lateral variations in thickness comparable—even less
important—to what is observed at Contact. This distribution indicates that post-QB fracturing was
preferentially localized in the hanging-wall and/or in the footwall of the QB, along its contact with the
host rocks, while most of the QB (core zone) remained poorly fractured.
172
Minerals 2018, 8, 319
Figure 12. (A) Distribution of fracture density and uranium mineralization as a function of depth, for
selected drillholes from Contact (Cont), End, Andrew Lake (And) and Bong. Fracture density as black
lines; 0: Non-fractured drill-core, 10: Intensely fractured drill-core. Uranium mineralization in red:
U in ppm measured by assays, logarithmic scale). (B) Number of quartz veins as a function of depth
for Contact and End. (C) Simplified cross sections.
173
Minerals 2018, 8, 319
A significant amount of quartz veins (Figure 12B) were observed up to the top of drill holes,
i.e., in the transition from the outer zone of the QB to the host rock. Such quartz veins are typically
re-opened and were also used as pathways for uranium bearing fluids at the first stage of uranium
mineralization [31]. Ore minerals are observed along the vein boundaries (Figure 13A) and cementing
orthogonal microfractures (Figure 13B,C; see also Chi et al. [29]). Quartz with uranium-oxides in their
vicinity display specific luminescence which has been described in many places worldwide [87–90];
it has been explained by the destabilization of the crystal lattice by radiation damages (due to liberation
of alpha particles through U238 decay series).
Figure 13. (A) Plane polarized light picture and interpretation drawing of a quartz vein network
guiding the mineralizing fluid along its boundaries. Quartz and iron oxides display evidence for
dissolution; (B,C) macroscopic drill core sample scan and interpretation drawing: Examples of a QB
related quartz vein bearing pitchblende (Pch) along edges or in orthogonal microfractures.
This change of the luminescence, together with the fact that the micro-fractures cemented with
uranium minerals crosscut several generations of quartz, show that formation of the QB and deposition
of uranium bearing minerals are two distinctive events, supporting isotope data on dark blue vuggy
quartz. Uranium minerals and associated specific luminescence are only observed in the vicinity of the
QB. Vuggy quartz precipitated from basinal brines (potentially U-bearing; [29]) are observed mainly in
the outer zone and in the vicinity of the QB and further demonstrate the barrier role played by the QB
in fluid flow partitioning. These observations suggest that the QB behaved as a rigid and hard body
compared to the weaker host rocks, so that later deformation preferentially localized in the host rocks
along the contact with the strong QB body.
Quartz veins in the outer zone of the QB form a network that, when microfractured, helped
focusing mineralizing fluid flow—thus creating local traps for uranium deposition. Post QB fractures
located in its vicinity acted as preferential pathways parallel to the QB for later, uranium-rich brines,
leading to deposition of uranium ore bodies at ca. 1500–1300 Ma for main stages [28–30]. To conclude,
174
Minerals 2018, 8, 319
the silicification processes ultimately led to the building of a complex quartz-cemented breccia body,
up to tens of meters thick, acting as a transverse hydraulic barrier depending on the vertical and
lateral variations in thickness and the degree of quartz cementation. As a result, the distribution of
mineralization in the Kiggavik area was heavily controlled—at different scales—by the mechanical and
hydraulic properties of the reactivated pre-existing fault zones where the QB was emplaced. The QB
behaved as a mechanically strong, transverse hydraulic barrier, that localized later fracturing and
compartmentalized/channelized vertical flow of uranium-bearing fluids, hence orebodies (Figure 12C)
in its hanging-wall and/or footwall during fault zone reactivation.
Even if the 3D architecture is not perfectly constrained, and would deserve proper 3D geometrical
and kinematic modelling, we could expect that relays within the QB, vertical and horizontal
variations in thickness, and overlap between QB bodies would likely influence fluid flow properties
(e.g., fluid velocity)—hence would impact uranium deposition rate [91]. The quartz breccia in
the Kiggavik area seems to be a good example of the “physical seal” developed by McCuaig and
Hronsky [92] in their mineral system concept, in conjunction with other factors to generate ore in
a considered area.
6. Conclusions
Based on a structural analysis combined with vein cement petrography, trace elements, and
oxygen stable isotope analysis of quartz, we constrain the nature, emplacement and significance of the
QB which strongly controlled the location of uranium deposits in the Kiggavik area (Nunavut, Canada).
The formation of the breccia bodies appears to be linked to fluctuations in pressure, temperature and
compositions of fluids during tectonic reactivation of the fault zones along which the QB was emplaced.
Faults formed during syn- to post-orogenic rifting processes and formation of the Baker Lake basin
(ca. 1850–1750 Ma) were subsequently reactivated, and cycles of underwent pervasive silicification,
hydraulic brecciation, and quartz recrystallization linked to cataclasis. This was associated with the
circulation of meteoric-derived fluids mixing with Si-rich magmatic-derived fluids at depth. This is
interpreted to be linked to the emplacement of the KIS at ca. 1750 Ma.
Post-QB fracturing at 1500–1300 Ma was constrained in the hanging wall and footwall of the
QB, with flow of basin-derived brines being channeled along the fault zones where QB emplaced.
The network of quartz veins in the vicinity of the QB was a favorable pathway for circulation of these
uranium-rich fluids and related uranium precipitation, as they were re-opened and micro-fractured.
Thus, the QB bodies likely exerted a major structural and hydrological control on the formation
of significant uranium orebodies in the Kiggavik area. Beyond regional implications, this study
demonstrates how an unconventional trap was built in impermeable Archean basement rocks. It also
emphasizes the importance of the spatial organization, and long-term evolution of fault zones in the
location of uranium orebodies of economic interest.
Author Contributions: A.G. and O.L. conceptualized both the study and the final model and wrote the original
draft. J.M., A.B., A.T., P.L., J.R. reviewed and edited the draft. A.G. performed macro to micro-scale petrographic
and microstructural characterization; A.G. and A.T. performed cathodoluminescence microscopy; A.G. and J.M.
performed trace elements and isotopes analyses and their interpretation. P.L. and J.R. gave their validation.
Funding acquisition and project administration were performed by A.B. and P.L., and geochemical analyses were
funded by O.L. and J.M.
Funding: This research was funded by Orano Canada and the laboratory ISTeP.
Acknowledgments: The authors thank ORANO and ORANO Canada for the full financial support and access
to the Kiggavik camp and exploration data. Special thanks to geologists R. Zerff, R. Hutchinson, K. Martin,
and D. Hrabok for their help and enriching discussions during field work. The authors also want to acknowledge
the first exploration geologists (Cogema and/or Orano) that worked on, and developed preliminary concepts on
the quartz breccia in the Kiggavik area: D. Baudemont, N. Flotte, J.-L. Feybesse, J.-L. Lescuyer. The authors thank
The SIMS team of the CRPG (Vandoeuvre-lès-Nancy, France) for their assistance in measuring the O isotopic
composition of the quartz by SIMS. Special thanks to A. Pêtre for his thoughtful comments.
175
Minerals 2018, 8, 319
Conflicts of Interest: The authors declare no conflict of interest. The founding sponsors had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, and in the
decision to publish the results”.
References
1. Sibson, R.H.; Robert, F.; Poulsen, K.H. High-angle reverse faults, fluid-pressure cycling, and mesothermal
gold-quartz deposits. Geology 1988, 16, 551–555. [CrossRef]
2. Blundell, D.J.; Karnkowski, P.H.; Alderton, D.H.M.; Oszczepalski, S.; Kucha, H. Copper mineralization of
the polish Kupferschierfer: A proposed basement fault-fracture system of fluid flow. Econ. Geol. 2003, 98,
1487–1495. [CrossRef]
3. Micklethwaite, S.; Cox, S.F. Fault-segment rupture, aftershock-zone fluid flow, and mineralization. Geology
2004, 32, 813–816. [CrossRef]
4. Cox, S.F. Coupling between deformation, fluid pressures, and fluid flow in ore-producing hydrothermal
systems at depth in the crust. Econ. Geol. 2005, 100th Anniv. Vol, 39–75. [CrossRef]
5. Muchez, P.; Heijlen, W.; Banks, D.; Blundell, D.; Boni, M.; Grandia, F. 7: Extensional tectonics and the timing
and formation of basin-hosted deposits in Europe. Ore Geol. Rev. 2005, 27, 241–267. [CrossRef]
6. Micklethwaite, S.; Sheldon, H.A.; Baker, T. Active fault and shear processes and their implications for mineral
deposit formation and discovery. J. Struct. Geol. 2010, 32, 151–165. [CrossRef]
7. Caine, J.S.; Evans, J.P.; Forster, C.B. Fault zone architecture and permeability structure. Geology 1996, 24,
1025–1028. [CrossRef]
8. McCuaig, T.; Kerrich, R. P-T-t-deformation-fluid characteristics of lode gold deposits: Evidence from
alteration systematics. Ore Geol. Rev. 1998, 12, 381–453. [CrossRef]
9. Ridley, J.R.; Diamond, L. Fluid chemistry of orogenic lode gold deposits and implications for genetic models.
Rev. Econ. Geol. 2000, 13, 141–162.
10. Kolb, J.; Rogers, A.; Meyer, F.M.; Vennemann, T.W. Development of fluid conduits in the auriferous shear
zones of the Hutti Gold Mine, India: Evidence for spatially and temporally heterogeneous fluid flow.
Tectonophysics 2004, 378, 65–84. [CrossRef]
11. Sibson, R. Earthquake rupturing as a mineralizing agent in hydrothermal systems. Geology 1987, 15, 701–704.
[CrossRef]
12. Sibson, R.H. Fault rocks and fault mechanisms. J. Geol. Soc. Lond. 1977, 133, 191–213. [CrossRef]
13. Phillips, W.J. Hydraulic fracturing and mineralization. J. Geol. Soc. Lond. 1972, 128, 337–359. [CrossRef]
14. Jébrak, M. Hydrothermal breccias in vein-type ore deposits: A review of mechanisms, morphology and size
distribution. Ore Geol. Rev. 1997, 12, 111–134. [CrossRef]
15. Laznicka, P. Breccias and Coarse Fragmentites: Petrology, Environments, Associations, Ores; Developments
in Economic Geology Series; Elsevier Science & Technology Books: New Yrok, NY, USA, 1988;
ISBN 9780444412508.
16. Simmons, S.F.; White, N.C.; John, D. Geological characteristics of epithermal precious and base metal
deposits. Econ. Geol. 2005, 100, 485–522.
17. Zhong, J.; Pirajno, F.; Chen, Y.-J. Epithermal deposits in South China: Geology, geochemistry, geochronology
and tectonic setting. Gondwana Res. 2017, 42, 193–219. [CrossRef]
18. Goldfarb, R.; Christie, A.; Bierlein, F. The orogenic gold deposit model and New Zealand: Consistencies and
anomalies. In Proceedings of the 2005 New Zealand Minerals Conference: Realising New Zealand’s Mineral
Potential, Auckland, New Zealand, 13–16 November 2005; pp. 105–114.
19. Cannell, J.; Cooke, D.R.; Walshe, J.L.; Stein, H. Geology, mineralization, alteration, and structural evolution
of the El Teniente Porphyry Cu-Mo Deposit. Econ. Geol. 2005, 100, 979–1003. [CrossRef]
20. Landtwing, M.R.; Furrer, C.; Redmond, P.B.; Pettke, T.; Guillong, M.; Heinrich, C.A. The Bingham Canyon
Porphyry Cu-Mo-Au deposit. III. Zoned copper-gold ore deposition by magmatic vapor expansion.
Econ. Geol. 2010, 105, 91–118. [CrossRef]
21. Henderson, I.H.C.; McCaig, A.M. Fluid pressure and salinity variations in shear zone-related veins, central
Pyrenees, France: Implications for the fault-valve model. Tectonophysics 1996, 262, 321–348. [CrossRef]
176
Minerals 2018, 8, 319
22. Rusk, B.; Reed, M. Scanning electron microscope-cathodoluminescence analysis of quartz reveals complex
growth histories in veins from the Butte porphyry copper deposit, Montana. Geology 2002, 30, 727–730.
[CrossRef]
23. Tóth, J.; Corbet, T. Post-Palaeocene evolution of regional groundwater flow systems and their relation to
petroleum accumulations, Taber Area, southern Alberta, Canada. Geol. Soc. Lond. Spec. Publ. 1987, 34, 45–77.
[CrossRef]
24. Hiatt, E.E.; Palmer, S.E.; Kyser, K.; O’Connor, E.; Terrence, K.H. Basin evolution, diagenesis and uranium
mineralization in the Paleoproterozoic Thelon Basin, Nunavut, Canada. Basin Res. 2010, 22, 302–323.
[CrossRef]
25. Reynolds, R.L.; Goldhaber, M.B. Origin of a South Texas roll-type uranium deposit; I, Alteration of
iron-titanium oxide minerals. Econ. Geol. 1978, 73, 1677–1689. [CrossRef]
26. Pagel, M.; Ahamdach, N. Etude des Inclusions Fluides dans les Quartz des Gisements U de l’Athabasca et du Thelon;
Internal Report of Centre de Recherches sur la Géologie des Matières Premières Minérales et Energétiques
(CREGU): Nancy, France, 1995.
27. Riegler, T.; Lescuyer, J.-L.; Wollenberg, P.; Quirt, D.; Beaufort, D. Alteration related to uranium deposits in
the kiggavik-andrew lake structural trend, Nunavut, Canada: New insights from petrography and clay
mineralogy. Can. Mineral. 2014, 52, 27–45. [CrossRef]
28. Sharpe, R.; Fayek, M.; Quirt, D.; Jefferson, C.W. Geochronology and genesis of the Bong Uranium deposit,
Thelon Basin, Nunavut, Canada. Econ. Geol. 2015, 110, 1759–1777. [CrossRef]
29. Chi, G.; Haid, T.; Quirt, D.; Fayek, M.; Blamey, N.; Chu, H. Petrography, fluid inclusion analysis, and
geochronology of the End uranium deposit, Kiggavik, Nunavut, Canada. Miner. Depos. 2017, 52, 211–232.
[CrossRef]
30. Shabaga, B.M.; Fayek, M.; Quirt, D.; Jefferson, C.W.; Camacho, A. Mineralogy, geochronology, and genesis
of the Andrew Lake uranium deposit, Thelon Basin, Nunavut, Canada. Can. J. Earth Sci. 2017, 54, 850–868.
[CrossRef]
31. Grare, A.; Benedicto, A.; Lacombe, O.; Trave, A.; Ledru, P.; Blain, M.; Robbins, J. The Contact uranium
prospect, Kiggavik project, Nunavut (Canada): Tectonic history, structural constraints and timing of
mineralization. Ore Geol. Rev. 2018, 93, 141–167. [CrossRef]
32. Hiatt, E.E.; Kyser, K.; Dalrymple, R.W. Relationships among sedimentology, stratigraphy, and diagenesis in
the Proterozoic Thelon Basin, Nunavut, Canada: Implications for paleoaquifers and sedimentary-hosted
mineral deposits. J. Geochem. Explor. 2003, 80, 221–240. [CrossRef]
33. Davis, W.J.; Gall, Q.; Jefferson, C.W.; Rainbird, R.H. Fluorapatite in the Paleoproterozoic Thelon Basin:
Structural-stratigraphic context, in situ ion microprobe U-Pb ages, and fluid-flow history. GSA Bull. 2011,
123, 1056–1073. [CrossRef]
34. Jefferson, C.; Thomas, D.J.; Gandhi, S.; Ramaekers, P.; Delauney, G.; Brisbin, D.; Cutts, C.; Portella, P.;
Olson, R. Unconformity-associated uranium deposits of the Athabasca Basin, Saskatchewan and Alberta.
Bull. Geol. Surv. Can. 2007, 588, 23–67.
35. AREVA (Areva, Paris, France). Internal Reference Document 2015. Available online: http://www.sa.areva.
com/EN/finance-1176/regulated-financial-information.html (accessed on 24 July 2018).
36. Hadlari, T.; Rainbird, R.H. Retro-arc extension and continental rifting: A model for the Paleoproterozoic
Baker Lake Basin, Nunavut1Geological Survey of Canada Contribution 2010 04 36. Can. J. Earth Sci. 2011, 48,
1232–1258. [CrossRef]
37. Rainbird, R.H.; Davis, W.J.; Stern, R.A.; Peterson, T.D.; Smith, S.R.; Parrish, R.R.; Hadlari, T. Ar-Ar and U-Pb
geochronology of a Late Paleoproterozoic Rift Basin: Support for a Genetic Link with Hudsonian Orogenesis,
Western Churchill Province, Nunavut, Canada. J. Geol. 2006, 114, 1–17. [CrossRef]
38. Rainbird, R.H.; Davis, W.J. U-Pb detrital zircon geochronology and provenance of the late Paleoproterozoic
Dubawnt Supergroup: Linking sedimentation with tectonic reworking of the western Churchill Province,
Canada. GSA Bull. 2007, 119, 314. [CrossRef]
39. Rainbird, R.H.; Hadlari, T.; Aspler, L.B.; Donaldson, J.A.; LeCheminant, A.N.; Peterson, T.D. Sequence
stratigraphy and evolution of the paleoproterozoic intracontinental Baker Lake and Thelon basins, western
Churchill Province, Nunavut, Canada. Precambrian Res. 2003, 125, 21–53. [CrossRef]
40. Pehrsson, S.J.; Berman, R.G.; Eglington, B.; Rainbird, R. Two Neoarchean supercontinents revisited: The case
for a Rae family of cratons. Precambrian Res. 2013, 232, 27–43. [CrossRef]
177
Minerals 2018, 8, 319
41. Peterson, T.D. Geological setting and geochemistry of the ca. 2.6 Ga Snow island Suite in the central Rae
Domain of the Western Churchill Province, Nunavut. Geol. Surv. Can. Open File 2015, 7841. [CrossRef]
42. Jefferson, C.; Pehrsson, S.; Peterson, T.; Chorlton, L.; Davis, B.; Keating, P.; Gandhi, S.; Fortin, R.; Buckle, J.;
Miles, W.; et al. Northeast Thelon region geoscience framework–new maps and data for uranium in Nunavut.
Geol. Surv. Can. 2011, 288791. [CrossRef]
43. McEwan, B. Structural style and regional comparison of the Paleoproterozoic Ketyet River group in the
region North-Northwest of Baker Lake, Nunavut. Master’s Thesis, University of Regina, Regina, SK, Canada,
2012; p. 155.
44. Tschirhart, V.; Morris, W.A.; Jefferson, C.W. Framework geophysical modelling of granitoid vs. supracrustal
basement to the northeast Thelon Basin around the Kiggavik uranium camp, Nunavut. Can. J. Earth 2013, 50,
667–677. [CrossRef]
45. Tschirhart, V.; Jefferson, C.W.; Morris, W.A. Basement geology beneath the northeast Thelon Basin, Nunavut:
Insights from integrating new gravity, magnetic and geological data. Geophys. Prospect. 2017, 65, 617–636.
[CrossRef]
46. Peterson, T.D.; Scott, J.M.J.; LeCheminant, A.N.; Jefferson, C.W.; Pehrsson, S.J. The Kivalliq Igneous Suite:
Anorogenic bimodal magmatism at 1.75 Ga in the western Churchill Province, Canada. Precambrian Res.
2015, 262, 101–119. [CrossRef]
47. Johnstone, D.; Bethune, K.M.; Quirt, D. Lithostratigraphic and structural controls of uranium mineralization
in the Kiggavik East, Centre and Main Zone deposits, Nunavut. In Proceedings of the Geological
Association of Canada-Mineralogival Association of Canada, Joint Annual Meeting, WhiteHorse, YT, Canada,
1–3 June 2016.
48. Peterson, T.D.; Van Breemen, O.; Sandeman, H.; Cousens, B. Proterozoic (1.85–1.75 Ga) igneous suites of
the Western Churchill Province: Granitoid and ultrapotassic magmatism in a reworked Archean hinterland.
Precambrian Res. 2002, 119, 73–100. [CrossRef]
49. Hoffman, P.F. United Plates of America, The Birth of a Craton: Early Proterozoic Assembly and Growth of
Laurentia. Annu. Rev. Earth Planet. Sci. 1988, 16, 543–603. [CrossRef]
50. Van Breemen, O.; Peterson, T.D.; Sandeman, H.A. U-Pb zircon geochronology and Nd isotope geochemistry
of Proterozoic granitoids in the western Churchill Province: Intrusive age pattern and Archean source
domains. Can. J. Earth Sci. 2005, 42, 339–377. [CrossRef]
51. Scott, J.M.J.; Peterson, T.D.; Davis, W.J.; Jefferson, C.W.; Cousens, B.L. Petrology and geochronology of
Paleoproterozoic intrusive rocks, Kiggavik uranium camp, Nunavut. Can. J. Earth Sci. 2015, 52, 495–518.
[CrossRef]
52. Zaleski, E.; Pehrsson, N.D.; Davis, W.J.; L’Heureux, R.; Greiner, E.; Kerswill, J.A. Quartzite Sequences and Their
Relationships, Woodburn Lake Group, Western Churchill Province, Nunavut; West. Churchill NATMAP Project;
Geological Survey of Canada: Ottawa, ON, Canada, 2000; pp. 1–10.
53. Rainbird, R.H.; Davis, W.J.; Pehrsson, S.J.; Wodicka, N.; Rayner, N.; Skulski, T. Early Paleoproterozoic
supracrustal assemblages of the Rae domain, Nunavut, Canada: Intracratonic basin development during
supercontinent break-up and assembly. Precambrian Res. 2010, 181, 167–186. [CrossRef]
54. LeCheminant, A.N.; Heaman, L.M. Mackenzie igneous events, Canada: Middle Proterozoic hotspot
magmatism associated with ocean opening. Earth Planet. Sci. Lett. 1989, 96, 38–48. [CrossRef]
55. Heaman, L.M.; LeCheminant, A.N. Paragenesis and U-Pb systematics of baddeleyite (ZrO2 ). Chem. Geol.
1993, 110, 95–126. [CrossRef]
56. Anand, A.; Jefferson, C.W. Reactivated fault systems and their effects on outcrop patterns of thin-skinned
early thrust imbrications in the Kiggavik uranium camp, Nunavut. Geol. Surv. Can. 2017. [CrossRef]
57. Peter, J.K.; Ulrike, W.; Brigitte, S.; Dmitry, K.; Qichao, Y.; Ingrid, R.; Andreas, S.; Karin, B.; Detlef, G.;
Jacinta, E. Determination of Reference Values for NIST SRM 610–617 Glasses Following ISO Guidelines.
Geostand. Geoanal. Res. 2010, 35, 397–429. [CrossRef]
58. Paton, C.; Hellstrom, J.; Paul, B.; Woodhead, J.; Hergt, J. Iolite: Freeware for the visualisation and processing
of mass spectrometric data. J. Anal. At. Spectrom. 2011, 26, 2508–2518. [CrossRef]
59. Hervig, R.L.; Williams, L.B.; Kirkland, I.K.; Longstaffe, F.J. Oxygen isotope microanalyses of diagenetic quartz:
Possible low temperature occlusion of pores. Geochim. Cosmochim. Acta 1995, 59, 2537–2543. [CrossRef]
60. Kempe, U.; Götze, J.; Dombon, E.; Monecke, T.; Poutivtsev, M. Quartz: Deposits, Mineralogy and Analytics;
Springer: Berlin, Germany, 2012; pp. 331–355.
178
Minerals 2018, 8, 319
61. Dennen, W.H. Stoichiometric substitution in natural quartz. Geochim. Cosmochim. Acta 1966, 30, 1235–1241.
[CrossRef]
62. Turner, W.; Richards, J.; Nesbitt, B.; Muehlenbachs, K.; Biczok, J. Proterozoic low-sulfidation epithermal
Au-Ag mineralization in the Mallery Lake area, Nunavut, Canada. Miner. Depos. 2001, 36, 442–457.
[CrossRef]
63. Clayton, R. Oxygen isotope exchange between quartz and water. J. Geophys. Res. 1972, 77, 3057–3067.
[CrossRef]
64. Lowenstern, J.B.; Sinclair, W.D. Exsolved magmatic fluid and its role in the formation of comb-layered quartz
at the Cretaceous Logtung W-Mo deposit, Yukon Territory, Canada. Trans. R. Soc. Edinb. Earth Sci. 1996, 87,
291–303. [CrossRef]
65. Thomas, J.B.; Bruce Watson, E.; Spear, F.S.; Shemella, P.T.; Nayak, S.K.; Lanzirotti, A. TitaniQ under pressure:
The effect of pressure and temperature on the solubility of Ti in quartz. Contrib. Mineral. Petrol. 2010, 160,
743–759. [CrossRef]
66. Perny, B.; Eberhardt, P.; Ramseyer, K.; Pankrath, R. Microdistribution of Al, Li, and Na in α quartz: Possible
causes and correlation with short-lived cathodoluminescence. Am. Mineral. 1992, 77, 534–544.
67. Rusk, B.G.; Lowers, H.A.; Reed, M.H. Trace elements in hydrothermal quartz: Relationships to
cathodoluminescent textures and insights into vein formation. Geology 2008, 36, 547–550. [CrossRef]
68. Spear, F.; Wark, D. Cathodoluminescence imaging and titanium thermometry in metamorphic quartz.
J. Metamorph. Geol. 2009, 27, 187–205. [CrossRef]
69. Leeman, W.P.; MacRae, C.M.; Wilson, N.C.; Torpy, A.; Lee, C.-T.; Student, J.J.; Thomas, J.B.; Vicenzi, E.P. A
Study of cathodoluminescence and trace element compositional zoning in natural quartz from volcanic
rocks: Mapping titanium content in quartz. Microsc. Microanal. 2012, 18, 1322–1341. [CrossRef] [PubMed]
70. Rusk, B.; Koenig, A.; Lowers, H. Visualizing trace element distribution in quartz using cathodoluminescence,
electron microprobe, and laser ablation-inductively coupled plasma-mass spectrometry. Am. Mineral. 2011,
96, 703–708. [CrossRef]
71. Huang, R.; Audétat, A. The titanium-in-quartz (TitaniQ) thermobarometer: A critical examination and
re-calibration. Geochim. Cosmochim. Acta 2012, 84, 75–89. [CrossRef]
72. Taylor, H.P. The Application of oxygen and hydrogen isotope studies to problems of hydrothermal alteration
and ore deposition. Econ. Geol. 1974, 69, 843–883. [CrossRef]
73. Bettencourt, J.S.; Leite, W.B.; Goraieb, C.L.; Sparrenberger, I.; Bello, R.M.S.; Payolla, B.L. Sn-polymetallic
greisen-type deposits associated with late-stage rapakivi granites, Brazil: Fluid inclusion and stable isotope
characteristics. Lithos 2005, 80, 363–386. [CrossRef]
74. Bowen, G.J. Statistical and Geostatistical Mapping of Precipitation Water Isotope Ratios. In Isoscapes:
Understanding Movement, Pattern, and Process on Earth through Isotope Mapping; West, J.B., Bowen, G.J.,
Dawson, T.E., Tu, K.P., Eds.; Springer: Dordrecht, The Netherlands, 2010; pp. 139–160, ISBN 978-90-481-3354-3.
75. Vollbrecht, A.; Oberthür, T.; Ruedrich, J.; Weber, K. Microfabric analyses applied to the Witwatersrand gold-
and uranium-bearing conglomerates: Constraints on the provenance and post-depositional modification of
rock and ore components. Miner. Depos. 2002, 37, 433–451. [CrossRef]
76. Kraishan, G.; Rezaee, R.; Worden, R. Significance of trace element composition of quartz cement as a key to
reveal the origin of silica in sandstones: An example from the cretaceous of the Barrow Sub-Basin, Western
Australia. Quartz Cementation Sandstones 2009, 29, 317–331.
77. Rosenberg, P.E.; Foit, F.F. Magnesiofoitite from the uranium deposits of the Athabasca Basin, Saskatchewan,
Canada. Can. Mineral. 2006, 44, 959–965. [CrossRef]
78. Richard, A.; Boulvais, P.; Mercadier, J.; Boiron, M.-C.; Cathelineau, M.; Cuney, M.; France-Lanord, C.
From evaporated seawater to uranium-mineralizing brines: Isotopic and trace element study of
quartz–dolomite veins in the Athabasca system. Geochim. Cosmochim. Acta 2013, 113, 38–59. [CrossRef]
79. Polito, P.A.; Kyser, T.K.; Thomas, D.; Marlatt, J.; Drever, G. Re-evaluation of the petrogenesis of the Proterozoic
Jabiluka unconformity-related uranium deposit, Northern Territory, Australia. Miner. Depos. 2005, 40,
257–288. [CrossRef]
80. Derome, D.; Cathelineau, M.; Fabre, C.; Boiron, M.-C.; Banks, D.; Lhomme, T.; Cuney, M. Reconstitution
of paleo-fluid composition by Raman LIBS and crush-leach techniques: Application to mid-Proterozoic
evaporitic brines (Kombolgie Formation basin, Northern Territory, Australia). Chem. Geol. 2007, 237, 240–254.
179
Minerals 2018, 8, 319
81. Bodnar, R.J.; Reynolds, T.J.; Kuehn, C.A. Fluid-Inclusion Systematics in Epithermal Systems. In Geology and
Geochemistry of Epithermal Systems; Berger, B.R., Bethke, P.M., Eds.; Society of Economic Geologists: Littleton,
CO, USA, 1985.
82. Dong, G.; Morrison, G.; Jaireth, S. Quartz textures in epithermal veins, Queensland; classification, origin and
implication. Econ. Geol. 1995, 90, 1841–1856. [CrossRef]
83. Bobis, E.R. A review of the description, classification, and origin of quartz textures in low-sulphidation
epithermal veins. J. Geol. Soc. Philipp. 1994, 99, 15–39.
84. Moncada, D.; Mutchler, S.; Nieto, A.; Reynolds, T.J.; Rimstidt, J.D.; Bodnar, R.J. Mineral textures and fluid
inclusion petrography of the epithermal Ag–Au deposits at Guanajuato, Mexico: Application to exploration.
J. Geochem. Explor. 2012, 114, 20–35. [CrossRef]
85. Tanner, P.W. The giant quartz-breccia veins of the Tyndrum-Dalmally area, Grampian Highlands, Scotland:
Their geometry, origin and relationship to the Cononish gold-silver deposit. Earth Environ. Sci. Trans. R.
Soc. Edinb. 2012, 103, 51–76. [CrossRef]
86. Sibson, R.H. Implications of fault-valve behaviour for rupture nucleation and recurrence. Tectonophysics
1992, 211, 283–293. [CrossRef]
87. Meunier, J.D.; Sellier, E.; Pagel, M. Radiation-damage rims in quartz from uranium-bearing sandstones.
J. Sediment. Res. 1990, 60, 53–58. [CrossRef]
88. Hu, B.; Pan, Y.; Botis, S.; Rogers, B.; Kotzer, T.; Yeo, G. Radiation-induced defects in drusy quartz, Athabasca
basin, Canada: A new aid to exploration of uranium deposits. Econ. Geol. 2008, 103, 1571–1580. [CrossRef]
89. MacRae, C.M.; Wilson, N.C.; Torpy, A. Hyperspectral cathodoluminescence. Mineral. Petrol. 2013, 107,
429–440. [CrossRef]
90. Cerin, D.; Götze, J.; Pan, Y. Radiation-Induced Damage In Quartz At the Arrow Uranium Deposit,
Southwestern Athabasca Basin, Saskatchewan. Can. Mineral. 2017, 55, 457–472. [CrossRef]
91. Zhang, Y.; Robinson, J.; Schaubs, P.M. Numerical modelling of structural controls on fluid flow and
mineralization. Geosci. Front. 2011, 2, 449–461. [CrossRef]
92. McCuaig, T.C.; Hronsky, J.M.A. The Mineral System Concept: The Key to Exploration Targeting. In Building
Exploration Capability for the 21st Century; Kelley, K.D., Golden, H.C., Eds.; Society of Economic Geologists:
Littleton, CO, USA, 2014; ISBN 9781629491424.
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
180
minerals
Article
Structural Control on Clay Mineral Authigenesis in
Faulted Arkosic Sandstone of the Rio do Peixe
Basin, Brazil
Ingrid B. Maciel 1 , Angela Dettori 2 , Fabrizio Balsamo 2, * , Francisco H.R. Bezerra 1,3 ,
Marcela M. Vieira 3 , Francisco C.C. Nogueira 4 , Emma Salvioli-Mariani 2 and
Jorge André B. Sousa 5
1 Post-Graduation Program on Geodynamics and Geophysics, Universidade Federal do Rio Grande do Norte,
Natal, RN 59078-970, Brazil; [email protected] (I.B.M.); [email protected] (F.H.R.B.)
2 Department of Chemistry, Life Sciences and Environmental Sustainability, University of Parma,
I-43124 Parma, Italy; [email protected] (A.D.); [email protected] (E.S.-M.)
3 Department of Geology, Federal University of Rio Grande do Norte, Natal, RN 59078-970, Brazil;
[email protected]
4 Department of Petroleum Engineering, Federal University of Campina Grande,
Campina Grande, PB 58100-000, Brazil; [email protected]
5 Petrobras Research Center—CENPES, Rio de Janeiro, RJ 21941-915, Brazil; [email protected]
* Correspondence: [email protected]; Tel.: +39-0521-905365
Abstract: Clay minerals in structurally complex settings influence fault zone behavior and
characteristics such as permeability and frictional properties. This work aims to understand the role
of fault zones on clay authigenesis in arkosic, high-porosity sandstones of the Cretaceous Rio do
Peixe basin, northeast Brazil. We integrated field, petrographic and scanning electron microscopy
(SEM) observations with X-ray diffraction data (bulk and clay-size fractions). Fault zones in the field
are characterized by low-porosity deformation bands, typical secondary structures developed in
high-porosity sandstones. Laboratory results indicate that in the host rock far from faults, smectite,
illite and subordinately kaolinite, are present within the pores of the Rio do Peixe sandstones.
Such clay minerals formed after sediment deposition, most likely during shallow diagenetic processes
(feldspar dissolution) associated with meteoric water circulation. Surprisingly, within fault zones
the same clay minerals are absent or are present in amounts which are significantly lower than those
in the undeformed sandstone. This occurs because fault activity obliterates porosity and reduces
permeability by cataclasis, thus: (1) destroying the space in which clay minerals can form; and (2)
providing a generally impermeable tight fabric in which external meteoric fluid flow is inhibited.
We conclude that the development of fault zones in high-porosity arkosic sandstones, contrary to
other low-porosity lithologies, inhibits clay mineral authigenesis.
1. Introduction
Clay minerals have important economic applications in industry—e.g., [1]. Additionally, in most
geological settings clay minerals can occur in faults, thus influencing their permeability, frictional
properties [2–8] and subsurface fluid flow [4,9–14]. Therefore, the understanding of the feedback
between faulting and clay mineral authigenesis has important implications for seismicity, the migration
and accumulation of oil and gas in the subsurface, and contaminant transport in aquifers.
Faults in high-porosity sandstones are generally considered as barriers to fluid flow, due to the
combined effect of grain size and porosity reduction within fault cores and associated deformation
bands in damage zones [15–22]. In this context, clay minerals are commonly described as mechanically
weak minerals, and because of this weakness their presence in faults commonly contributes to stable
sliding failures [23,24]. Furthermore, the origin and distribution of clays in sandstone are also important
in oil industry, because these minerals contribute to increases in the sealing potential of faults and can
determine reservoir compartmentalization [3,6,8].
Several studies have described the clay mineralogy of fault zones [24–26], however little attention
has been paid to the role of faults in determining the type and amount of clay mineral transformation in
faulted, arkosic sandstones. The goal of this study is to investigate how fault zones in arkosic sandstones
(composed of a fault core surrounded by deformation bands) modify grain-scale fabric and control
clay mineral authigenesis at shallow burial depths. We selected the Cretaceous Rio do Peixe basin in
northeast Brazil (Figure 1) as a case study, due to its excellent exposures of undeformed sandstones
and well-preserved fault zones. By integrating field analysis with laboratory data, we conclude that
deformation-band faulting in arkosic, high-porosity sandstones inhibits clay mineral authigenesis,
rather than promoting alteration and clay mineral formation.
182
Minerals 2018, 8, 408
Figure 1. Simplified geological map of the Rio do Peixe Basin, showing major faults and the
lithostratigraphics units. The location of the four selected outcrops is indicated. Modified from [27,29,35].
The sedimentary rocks of the Rio do Peixe basin were affected by two main tectonic phases:
an Early to Late Cretaceous NW–SE oriented extension [29] followed by a basin inversion in a strike–slip
regime from the Late Cretaceous to Cenozoic [35]. The extensional faults that developed during the
first phase are dominated by deformation bands, often associated with slickensided surfaces [34].
The deformation bands occur as cm-thick tabular structures developed in the fault damage zone,
and are arranged as single elements or in clusters. Within deformation bands, a cataclastic foliation
was formed by preferential grain alignment and the selective fragmentation of feldspar grains [34].
183
Minerals 2018, 8, 408
using the light petrographic microscope, we focused on grain size and roundness, sorting, packing,
porosity, mineral composition and amount of clay. We also described depositional and diagenetic
features in undeformed and faulted samples. Small representative samples were analyzed using
a scanning electron microscope (SUPERSCAN SSX-550, Shimadzu Corporation, Kyoto, Japan) to
improve clay mineral identification and textural analysis. EDS was used to identify the main chemical
elements and mineral composition of the samples. X-ray diffraction (XRD) analyses were performed
using a Bruker (Billerica, MA, USA) D2 Phaser powder diffractometer (CuKα radiation, voltage of
30 kV, current of 10 mA, step size of 0.018, interval of 1 s per step) on powdered bulk samples (n = 10)
and fraction samples <2 μm in size (n = 6) for clay mineral identification in undeformed and faulted
rocks. The oriented samples of the clay fractions were analyzed under three different conditions: air
dried; ethylene glycol saturated; and heated to a temperature of 550 ◦ C. The powdered bulk samples
were measured in the range 2–80◦ 2θ, and the clay fraction samples were measured in the range
2–20◦ 2θ. Mineral phase identification and semi-quantitative estimations were performed using the
DIFFRAC.EVA suite software provided by Bruker Corporation (Billerica, MA, USA). The results of the
XRD analyses, together with sample description and location, are listed in Table 1.
4. Results
Figure 2. Conceptual sketch showing the typical architecture of fault zones in the Rio do Peixe basin.
The host rock represents the sandstones and conglomerates with pristine textures and sedimentary
structures not affected by faults. The fault core is the most deformed part of the fault zone, where
slip surfaces were frequently developed and where several movements (different slicken lines) are
observed. The damage zone is the deformed rock volume next to the fault core that has single or
clusters of deformation bands (Sites 2, 3 and 4 in this study). Colors are indicative of the amount of
weathering observed in the field.
184
Minerals 2018, 8, 408
Concerning the host rock, the original undeformed fluvial facies of the Rio do Peixe basin exhibit
a massive laminated structure with trough–festoon crossbedding stratification (Figure 3). These units
vary from silty sandstones to fine conglomerates. In a few cases, thin silt lenses are also observed
(Figure 3). The sandstones are generally clast-supported with a granular texture, and grains are locally
fractured. The grain sizes vary between silt and gravel.
Figure 3. Schematic profile and compositional classification of the host rock (Antenor Navarro
Formation) in the Rio do Peixe basin, Site 1. (A) Outcrop photograph showing fluvial sedimentary
structures with tabular and lenticular shapes of fine sandstones and trough–festoon conglomerates.
Note the intense red-orange coloration of undeformed rocks. (B) Vertical sedimentary log showing
sampling position (Samples 1 to 9) in the fluvial succession. (C) Compositional classification of analyzed
Samples 1 to 9, based on [37]. Key: c—clay; s—sand; fs—fine sand; ms—medium sand; cs—coarse
sand; g—gravel; Qz—quartz; Fd—feldspar; Rf—rock fragments.
The fault cores range from 0.1 m to 0.3 m in thickness, whereas the width of the damage zones
broadly range from ~5 m to 10 m (in small faults of Sites 2 and 3) up to ~200 m (in the hanging wall
damage zone of Site 4). The fault cores and the inner damage zone generally form topographic relief
up to 1 m in height with respect to the surrounding undeformed rock (Figure 4A). The sandstones in
the fault core show a strong decrease in grain size and a preferential grain alignment, which forms a
tectonic foliation visible at the hand scale (Figure 4B). Most offsets are extensional or slightly oblique.
The footwall and hanging wall damage zones consist of clusters of anastomosing deformation bands
(Figure 4C) and isolated single deformation bands (Figure 4D). Deformation bands also form a small
positive relief. The fault cores and deformation bands exhibit lighter colors than surrounding host
rocks (Figure 4B–D) and in some cases an orange to red coloration is also observed.
185
Minerals 2018, 8, 408
Figure 4. Field photographs showing the main structural features of studied fault zones in the Rio
do Peixe basin. (A) Example of an extensional fault zone with m-scale offset showing positive relief
with respect to the host sandstone, Site 2. The dotted line indicates the approximate position of the
fault core. (B) Foliated fault core rock (sample UT13 in Table 1) showing light grey to red colors, Site 2.
The diameter of the coin is ~2.5 cm. (C) Example of a 12.0 cm-thick cluster of deformation bands,
in positive relief, developed in the fault damage zone, Site 4. The length of the white scale is ~8.0 cm.
(D) Whitish single deformation band in positive relief developed in the damage zone, Site 2. The length
of the pen is ~14.0 cm. Key: FWDZ—footwall damage zone; HWDZ—hanging wall damage zone;
FC—fault core; DB—deformation band; CDB—cluster of deformation bands.
4.2. Petrography
186
Minerals 2018, 8, 408
exhibit primary intergranular porosity and, subordinately, secondary moldic porosity associated with
the selective dissolution of feldspar grains (Figure 5E,F). Fracture porosity is also observed (Figure 5C),
although fractures are mostly filled by clay minerals. In thin sections, visual porosity was observed
to be 17% and 28% in fine and coarse sandstones, respectively. Generally, grains are coated by thin
layers of clay minerals; these are even more abundant within intergranular pores and microfractures.
In some samples, pores show shrinkage (Figure 5B,C). In rare cases, the porosity is almost entirely
filled by clay minerals (Figure 5A).
Figure 5. Optical microscopic images of the undeformed Rio do Peixe basin sandstones at Site 1.
(A) Host rock, with very fine grains and abundant small pores which are frequently filled by clay
minerals (pore filling, PF). The porosity is mainly secondary porosity (SP). (B) Coarse-grained sandstone
showing sub-rounded grains and primary porosity, partially filled by smectite and illite. (C) Example of
a feldspar clast dissolved and replaced by abundant clays. (D–F) show intergranular primary porosity
(PP) between quartz grains (Qtz) and SP resulting from the dissolution of feldspar grains (Fsp).
187
Minerals 2018, 8, 408
grains within the deformation bands (Figure 6C). In places, fractures are open and filled by fine-grained
angular cataclastic material (Figure 6D).
Samples from fault cores exhibit a strong reduction in grain size with abundant fractions ranging
from fine sand to silt (Figure 6E,F). In strong contrast to the undeformed host rock, fault core samples
are clearly matrix-supported and very poorly sorted (Figure 6E,F). The brown colored, fine-grained
matrix consists of crushed feldspar grains, in agreement with recent observations [34]. In the fault cores,
the visual porosity determined using the optical microscope is practically zero due to the presence of
the cataclastic matrix (Figure 6F).
Figure 6. Fault rocks viewed under optical microscope. (A) Example of tight fabric in a deformation
band within the damage zone, Site 2. (B) Detail of primary porosity filled by small angular clasts
generated by the cataclasis process in a deformation band, Site 2. Note that feldspar grain dissolution
is limited. (C) Secondary porosity (SP) developed by intragranular microfractures in a deformation
band. (D) Detail of intragranular fractures filled by fine-grained angular cataclastic material in a small
fault core. (E) Fine-grained matrix in the fault core resulting from high grain comminution, in which
the pore space was completely destroyed, Site 3. (F) A high degree of cataclasis within the fault core,
showing a dramatic reduction in grain size and porosity, Site 4. The brown crushed material in (E) and
(F) mostly consist of very small feldspar grains developed during a cataclastic process cf. [34].
188
Minerals 2018, 8, 408
Figure 7. SEM images of clay minerals in undeformed sandstones. (A) Feldspar grain surrounded by
fractured smectite. (B) Well developed smectite flakes inside a pore. (C) Pore-line illite coatings capping
quartz grains. (D) Mixed layers of smectite and illite. Key: Sme—smectite; Fsp—feldspar; Ill—illite.
189
Minerals 2018, 8, 408
Table 1. Results of XRD analyses (bulk and clay fractions) performed in undeformed and faulted
rocks from Sites 1, 2 and 4. (I-S: illite-smectite; DB: deformation bands). Sample labels are the same as
Figures 8 and 9.
190
Minerals 2018, 8, 408
Figure 8. Examples of XRD analyses on bulk samples representative of undeformed sandstone (A),
foliated cataclasite in fault core rock (B) and deformation band (C).
191
Minerals 2018, 8, 408
XRD analyses of clay fractions, performed on the six samples which had sufficient clay minerals in
bulk analyses, indicate that the main types of clay minerals in both undeformed and faulted sandstones
are smectite and subordinately illite (Figure 9), as indicated by the comparison between the XRD
spectra of air-dried, glycolated and heated samples. Based on spectral peak intensities, in all the
spectra the amount of illite was found to be less than that of smectite. Both illite and smectite occur as
both distinct phases and mixed layers (Figure 9). In undeformed samples (Figure 9A–D) the spectral
peaks of illite and smectite have greater intensities than in faulted samples (Figure 9E,F). Smectite
is absent in the foliated fault-core rock (Figure 9E), which also shows the lowest amount of illite of
all the analyzed samples; this is consistent with the non-weathered, whitish foliated cataclasites that
are often observed in the field (e.g., Figure 4B) and the observed lack of clay minerals in thin sections
(Figure 7E,F). A very small amount of chlorite is also observed in the fault core rock sample (Figure 9E).
Figure 9. XRD diffractograms of aggregates of clay-size fractions of undeformed (A–D) and fault core
(E) rocks, and a deformation band in a fault damage zone (F). Air-dried (in grey), heated at 550 ◦ C
(in red), and treated with ethylene glycol (in blue) conditions are shown.
5. Discussion
It is well known that the development of fault zones in sandstones can significantly modify fluid
circulation pathways e.g., [12], thus influencing a variety of shallow diagenetic processes [21,24,38].
In this study, fault zones that developed in high-porosity arkosic sandstones are found to have
the typical architecture described in other settings [9,24,38], being organized in a foliated fault-core
surrounded by a damage zone hosting cataclastic deformation bands [34].
192
Minerals 2018, 8, 408
193
Minerals 2018, 8, 408
weathering and dissolution of feldspar grains in a semi-arid environment (as shown by moldic
porosity in thin sections) and clay mineral authigenesis in the high-porosity undeformed sandstones
and conglomerates; (5) the exhumation of faults during regional basin inversion and the formation of
positive reliefs of fault zones (due to differential surface erosion) caused by a tight cataclastic fabric.
Figure 10. Summary diagrams showing a comparison between diffractograms of undeformed host
rocks (green lines) and faulted rocks (red lines). (A) Bulk XRD analysis indicating that that the amount
of clay minerals in the undeformed samples is higher than that in faulted samples. (B) The relative
abundance of smectite in undeformed and faulted samples. (C) The relative abundance of illite in
undeformed and faulted samples.
6. Conclusions
We studied clay mineral assemblages in faulted, high-porosity arkosic sandstone of the Rio do
Peixe basin (northeast Brazil) to understand the role of faults in clay mineral authigenesis. We integrated
field observations with analysis of microstructures, optical and scanning electron microscopy and XRD
(bulk and clay-fraction) mineralogy. The results obtained in this study indicate the following conclusions:
194
Minerals 2018, 8, 408
(1) The bulk mineralogy of the Rio do Peixe sandstone does not change significantly between the
undeformed and faulted domains, consisting of lithic arkose with feldspar grains generally
comprising >50%.
(2) In both undeformed and faulted domains, clay minerals are <1–2% and consist of smectite and
illite, and subordinately illite–smectite mixed layers. Despite the similar mineralogy, the amount
of clay is systematically less in the faulted domain than in pristine rocks and in some cases is not
observed at all.
(3) Clay minerals in the studied arkosic sandstones most likely developed during feldspar weathering
processes in a shallow meteoric environment. A detrital origin of clay is excluded in the analyzed
sandstones and conglomerates.
(4) Contrary to the results of other fault rock studies in similar lithologies, clay is found to be less
abundant in the faulted domains (fault core and damage zone) than in the host rocks. We conclude
that this is due to the tight fabric that developed in the faulted porous sandstone, which inhibited
meteoric fluid circulation and clay mineral authigenesis.
We conclude that, contrary to several other faulted settings which have a high abundance of
authigenic clays, the development of fault zones in high-porosity arkosic sandstone in semi-arid
regions prevents the authigenesis of clay minerals. Consequently, clay authigenesis is more efficient in
undeformed sandstones than faulted domains, which has important implications for oil and water
reservoir quality in siliciclastic rocks and fault behavior in structurally complex settings.
Author Contributions: I.B.M. participated to the fieldwork and sampling, performed petrographic and SEM
analysis, contributed to manuscript writing. A.D. participated to the fieldwork, performed XRD analyses and
thin section observations. F.B. conceived the research, participated to the fieldwork and sampling, contributed to
manuscript writing; F.H.R.B. participated to fieldwork, contributed to manuscript writing; M.M.V. supervised
SEM and petrographic analyses; F.C.C.N. organized the overall fieldwork and participated to sampling; E.S.-M.
supervised and interpreted the XRD analysis; J.A.B.S. participated to the fieldwork and supported the project.
Funding: This research was funded by Petrobras/Federal University of Campina Grande project
(TC 0050.0096065.15.9 grant to Francisco C. C. Nogueira); Fieldwork of Angela Dettori and Fabrizio Balsamo was
funded by University of Parma, Italy (Overworld progam 2016–2017 grant to Fabrizio Balsamo); Ingrid Maciel
was supported by a Brazilian CAPES grant.
Acknowledgments: We kindly thank two anonymous reviewers which significantly improve the original early
version of this manuscript. We thank Luca Aldega and Luciana Mantovani for helful discussion on XRD data.
We also thank the Brazilian Agency of Oil, Gas, and Biofuels (Agência Nacional do Petróleo, ANP) for sharing data
on the Rio do Peixe basin. Fabrizio Balsamo wishes to dedicate this work in faulted sandstones to his mother
Isabella Bellina, died in Albano Laziale (Rome, Italy) the 14 June 2018.
Conflicts of Interest: The authors declare no conflicts of interest.
References
1. Xi, K.; Cao, Y.; Liu, K.; Jahren, J.; Zhu, R.; Yuan, G.; Hellevang, H. Authigenic minerals related to wettability
and their impacts on oil accumulation in tight sandstone reservoirs: An example from the Lower Cretaceous
Quantou Formation in the southern Songliao Basin, China. J. Asian Earth Sci. 2018. [CrossRef]
2. Rice, J.R. Fault stress states, pore pressure distributions, and the weakness of the San Andreas Fault.
In International Geophysics; Academic Press: Cambridge, MA, USA, 1992; pp. 475–503.
3. Fisher, Q.J.; Knipe, R.J. Fault sealing processes in siliciclastic sediments. In Faulting, Fault Sealing and Fluid
Flow in Hydrocarbon Reservoirs. Geol. Soc. Spec. Publ. 1998, 147, 117–134. [CrossRef]
4. Haines, S.H.; Van der Pluijm, B.A.; Ikari, M.J.; Saffer, D.M.; Marone, C. Clay fabric intensity in natural
and artificial fault gouges: Implications for brittle fault zone processes and sedimentary basin clay fabric
evolution. J. Geophys. Res. 2009, 114, B05406. [CrossRef]
5. Lander, R.H.; Bonnell, L.M. A model for fibrous illite nucleation and growth in sandstones. AAPG Bull. 2010,
94, 1161–1187. [CrossRef]
6. Faulkner, D.R.; Jackson, C.A.L.; Lunn, R.J.; Schlische, R.W.; Shipton, Z.K.; Wibberley, C.A.J.; Withjack, M.O.
A review of recent developments concerning the structure, mechanics and fluid flow properties of fault
zones. J. Struct. Geol. 2010, 32, 1557–1575. [CrossRef]
195
Minerals 2018, 8, 408
7. Balsamo, F.; Aldega, L.; De Paola, N.; Faoro, I.; Storti, F. The signature and mechanics of earthquake ruptures
along shallow creeping faults in sediments. Geology 2014, 42, 435–438. [CrossRef]
8. Buatier, M.D.; Cavailhes, T.; Charpentier, D.; Lerat, J.; Sizun, J.P.; Labaume, P.; Gout, C. Evidence of
multi-stage faulting by clay mineral analysis: Example in a normal fault zone affecting arkosic sandstones
(Annot sandstones). J. Struct. Geol. 2015, 75, 101–117. [CrossRef]
9. Antonellini, M.; Aydin, A. Effect of Faulting on Fluid Flow in Porous Sandstones: Petrophysical Properties.
AAPG Bull. 1994, 78, 355–377.
10. Rawling, G.C.; Goodwin, L.B.; Wilson, J.L. Internal architecture, permeability structure, and hydrologic
significance of contrasting fault-zone types. Geology 2001, 29, 43–46. [CrossRef]
11. Eichhbl, P.; Taylor, W.L.; Pollard, D.D.; Aydin, A. Paleo-fluid flow and deformation in the Aztec Sandstone at
the Valley of Fire, Nevada—Evidence for the coupling of hidrogeologic, diagenetic, and tectonic process.
GSA Bull. 2004, 116, 1120–1136. [CrossRef]
12. Fossen, H.; Schultz, R.A.; Shipton, Z.K.; Mair, K. Deformation bands in sandstone: A review. J. Geol. Soc.
2007, 164, 755–769. [CrossRef]
13. Caine, J.S.; Minor, S.A. Structural and geochemical characteristics of faulted sediments and inferences on the
role of water in deformation, Rio Grande Rift, New Mexico. GSA Bull. 2009, 121, 1325–1340. [CrossRef]
14. Vrolijk, P.; Van der Pluijm, B. Clay gouge. J. Struct. Geol. 1999, 21, 1039–1048. [CrossRef]
15. Gibson, R. Physical character and fluid-flow properties of sandstonederived fault gouge, in Structural
Geology in Reservoir Characterization. Geol. Soc. Spec. Publ. 1998, 127, 87–93. [CrossRef]
16. Fisher, Q.; Knipe, R.J. The permeability of faults within siliclastic petroleum reservoirs of the North Sea and
Norwegian Continental Shelf. Mar. Pet. Geol. 2001, 18, 1063–1081. [CrossRef]
17. Fossen, H.; Bale, A. Deformation bands and their influence on fluid flow. AAPG Bull. 2007, 91, 1685–1700.
[CrossRef]
18. Rotevatn, A.; Torabi, A.; Fossen, H.; Braathen, A. Slipped deformation bands: A new type of cataclastic
deformation bands in Western Sinai, Suez rift, Egypt. J. Struct. Geol. 2008, 30, 1317–1331. [CrossRef]
19. Eichhubl, P.; Davatzes, N.C.; Becker, S.P. Structural and diagenetic control of fluid migration and cementation
along the Moab Fault, Utah. AAPG Bull. 2009, 93, 653–681. [CrossRef]
20. Balsamo, F.; Storti, F.; Salvini, F.; Lima, C.C. Structural and petrophysical evolution of extensional fault zones
in low-porosity, poorly lithified sandstones of the Barreiras Formation NE Brazil. J. Struct. Geol. 2010, 32,
1806–1826. [CrossRef]
21. Balsamo, F.; Bezerra, F.H.; Vieira, M.; Storti, F. Structural control on the formation of iron oxide concretions
and Liesegang bands in faulted, poorly lithified Cenozoic sandstones of the Paraiba basin, Brazil. Bulletin
2013, 125, 913–931. [CrossRef]
22. Williams, J.N.; Toy, V.G.; Massiot, C.; McNamara, D.D.; Wang, T. Damaged beyond repair? Characterising
the damage zone of a fault late in its interseismic cycle, the Alpine Fault, New Zealand. J. Struct. Geol. 2016,
90, 76–94. [CrossRef]
23. Hoffman, U.; Endell, K.; Wilm, M.D. Kristallstruktur und Quellung von Montmorillonit. Z. Kristallogr.
Cryst. Mater. 1933, 86, 340–348. [CrossRef]
24. Solum, J.G.; Davatzes, N.C.; Lockner, D.A. Fault-related clay authigenesis along the Moab Fault: Implications
for calculations of fault rock composition and mechanical and hydrologic fault zone properties. J. Struct. Geol.
2010, 32, 1899–1911. [CrossRef]
25. Van der Pluijm, R. Out-of-Plane Bending of Masonry: Behaviour and Strength Eindhoven. Ph.D. Thesis,
Technische Universiteit Eindhoven, Eindhoven, The Neitherlands, 1999.
26. Solum, J.G.; Van der Pluijm, B.A.; Peacor, D.R. Neocrystallization, fabrics and age of clay minerals from an
exposure of the Moab Fault, Utah. J. Struct. Geol. 2005, 27, 1563–1576. [CrossRef]
27. Sénant, J.; Popoff, M. Early Cretaceous extension in northeast Brazil related to the South Atlantic opening.
Tectonophysics 1991, 198, 35–46. [CrossRef]
28. Matos, R.M.D. The Northeast Brazilian Rift System. Tectonics 1992, 11, 766–791. [CrossRef]
29. Françolin, J.B.L.; Cobbold, P.R.; Szatmari, P. Faulting in the early Cretaceous Rio do Peixe basin (NE Brazil)
and its significance for the opening of the Atlantic. J. Struct. Geol. 1994, 16, 647–661. [CrossRef]
30. De Castro, D.L.; De Oliveira, D.C.; Gomes Castelo Branco, R.M. On the tectonics of the Neocomian Rio do
Peixe Rift Basin, NE Brazil: Lessons from gravity, magnetics, and radiometric data. J. South. Am. Earth Sci.
2007, 24, 184–202. [CrossRef]
196
Minerals 2018, 8, 408
31. Albuquerque, J.P.T. Inventário Hidrogeológico do Nordeste; Folha 15; Sudene, Divisão de Documentação:
Recife, Brazil, 1970; p. 187.
32. Lima, M.R.; Coelho, M.P.C.A. Estudo palinológico da sondagem de Lagoa do Forno Bacia do Rio do Peixe
Cretáceo do Nordeste do Brasil. São Paulo. Bol. IG-USP Sci. 1987, 18, 67–83.
33. Córdoba, V.C.; Antunes, A.F.; Jardim de Sá, E.F.; Nunes da Silva, A.; Sousa, D.C.; Lins, F.A.P.L. Análise
estratigráfica e estrutural da Bacia do Rio do Peixe Nordeste do Brasil: Integração de dados a partir do
levantamento sísmico pioneiro 0295_rio_do_peixe_2d. Bol. Geoci. Petrobras 2008, 16, 53–68.
34. Nicchio, M.A.; Nogueira, F.C.C.; Balsamo, F.; Souza, J.A.B.; Carvalho, B.R.B.; Bezerra, F.H.R. Development of
cataclastic foliation in deformation bands in feldspar-rich conglomerates of the Rio do Peixe Basin, NE Brazil.
J. Struct. Geol. 2018, 107, 132–141. [CrossRef]
35. Nogueira, F.C.C.; Marques, F.O.; Bezerra, F.H.R.; de Castro, D.L.; Fuck, R.A. Cretaceous intracontinental
rifting and post-rift inversion in NE Brazil: Insights from the Rio do Peixe Basin. Tectonophysics 2015, 644,
92–107. [CrossRef]
36. Araujo, R.E.B.; Bezerra, F.H.R.; Nogueira, F.C.C.; Balsamo, F.; Carvalho, B.R.B.M.; Souza, J.A.B.;
Sanglard, J.C.D.; de Castro, D.L.; Melo, A.C.C. Basement control on fault formation and deformation
band damage zone evolution in the Rio do Peixe Basin, Brazil. Tectonophysics 2018, 745, 117–131. [CrossRef]
37. Folk, R.L. Petrology of Sedimentary Rocks; Hemphill Publishing Company: Austin, TX, USA, 1968; p. 182.
38. Balsamo, F.; Storti, F.; Grocke, D. Fault-related fluid flow history in shallow marine sediments from carbonate
concretions, Crotone Basin, south Italy. J. Geol. Soc. 2012, 169, 613–626. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
197
minerals
Article
Structural Controls on Copper Mineralization in the
Tongling Ore District, Eastern China: Evidence from
Spatial Analysis
Tao Sun 1,2, * ID
, Ying Xu 3 , Xuhui Yu 4 , Weiming Liu 1 , Ruixue Li 1 , Zijuan Hu 1 and Yun Wang 5
1 School of Resources and Environmental Engineering, Jiangxi University of Science and Technology,
Ganzhou 341000, China; [email protected] (W.L.); [email protected] (R.L.);
[email protected] (Z.H.)
2 Jiangxi Key Laboratory of Mining Engineering, Jiangxi University of Science and Technology,
Ganzhou 341000, China
3 Institute of Multipurpose Utilization of Mineral Resources, CAGS, Chengdu 610041, China;
[email protected]
4 College of Earth Sciences, Chengdu University of Technology, Chengdu 61005, China; [email protected]
5 School of Water Resource and Environment, China University of Geosciences, Beijing 341515, China;
[email protected]
* Correspondence: [email protected]; Tel.: +86-0797-831-2751
Abstract: Structures exert significant controls on hydrothermal mineralization, although such controls
commonly have cryptic expression in geological datasets dominated by 2D maps. Analysis of spatial
patterns of mineral deposits and quantification of their correlation with detailed structural features
are beneficial to understand the plausible structural controls on mineralization. In this paper, a series
of GIS-based spatial methods, including fractal, Fry, distance distribution and weights-of-evidence
analyses, were employed to reveal structural controls on copper mineralization in the Tongling ore
district, eastern China. The results indicate that Yanshanian intrusions exert the most significant
control on copper mineralization, followed by EW-trending faults, intersections of basement faults
and folds. The scale-variable distribution patterns of copper occurrences are attributed to the
different structural controls operating in the basement and sedimentary cover. In the basement,
EW-trending faults serve as pathways for channeling Yanshanian magma from a deep magma
chamber to structurally controlled trap zones in the caprocks, imposing an important regional
control on the spatial distribution of Cretaceous magmatic-hydrothermal system genetically related
to copper mineralization. In the sedimentary cover, bedding-parallel shear zones, formed during
the progressive folding and shearing in Indosinian and overprinted by tensional deformation in
Yanshanian, act as favorable sites for hosting, focusing and depositing the ore-bearing fluids, playing
a vital role in the localization of stratabound deposits at fine scale.
1. Introduction
Structural controls on hydrothermal mineralization at various scales have been widely
recognized [1–5]. At a global scale, hydrothermal systems usually form in specific tectonic settings,
e.g., porphyry systems mostly occur in magmatic arc settings [1–3]. At a regional scale, hydrothermal
deposits show close proximity to regional faults system or shear zones, which sever as pathways
for transporting ore-forming fluids from deep-seated sources to shallow depositing spots [6,7]. At a
deposit scale, hydrothermal replacement disseminations, breccias and veins, which are related to
subsidiary fracture zones of regional structures, serve as favorable sites for focusing and depositing
the ore-bearing fluids and are interpreted to be responsible for localization of orebodies [7]. However,
such controls may usually have cryptic expression in various sources of geological records, because
(i) structures, especially large-scale structures, may have variable expressions from depth to surface
(e.g., mylonite zone at depth and fault zone near the surface) [7]; (ii) spatial associations between
map-generalized structures and surface-projected deposits in 2D maps may lead to an inaccurate
view or even misunderstanding with respect to controls of mineralization; and (iii) structural features
together with structurally controlled mineralization may be formed through successive deformation
and polyphase tectonics [8]. Thus, it is a challenge to identify ore-related structural features and
elucidate structural controls, as well as measure their contributions to the formation of mineral deposits.
GIS-based spatial analysis has been well-established and developed in the last three decades,
assisting in identification of inherent patterns of ore-related geological features and delineation of
interplay of the processes that constrain the formation of mineral deposits [9–12]. More specifically,
with the help of quantitative methods and easy-to-use GIS software, delineating the spatial patterns
of known occurrences of mineral deposits and their associations with geological features (e.g.,
structural, lithological and geochemical features) can, in addition to field observations, geochemical
and mineralogical laboratory methods, provide insights into the controlling mechanisms operating at
different scales [10,13–15], especially in the brownfield areas where a relatively large number of mineral
deposits have been well-explored [15]. Furthermore, recognition of geological features controlling the
mineralization is critical for defining exploration criteria in future prospecting [16].
Since mineral occurrences are simplified to be represented as points on large-scale maps in
various applications of spatial analyses, methods of spatial analysis for point patterns have been
increasingly employed in studying spatial distribution and geological controls of mineral deposits,
mostly involving fractal geometry [17,18] and Fry analysis [19,20]. Through statistical calculation,
fractal and Fry analyses are able to highlight the distribution pattern of mineral deposits that may
be difficult to be recognized by exclusively relying on visual interpretation [5]. Moreover, distance
distribution [14,21] and weights-of-evidence (WofE) analyses [22,23] can further quantify the strength
of the spatial association between mineral deposits and geological features believed to be favorable
in predicting the location of the mineralization. A joint application of these methods is necessary
as individual methods only characterizes a particular aspect, such as non-random clustering of
deposits or preferential direction of deposits distribution, of complex spatial features of mineralization
systems [10,24].
The Tongling ore district (TOD) is one of the most important Cu producers in China, with totally
estimated reserves of over 5 Mt copper [25]. Large stratabound copper deposits constitute the
majority of the copper reserves in this area, e.g., the Donggushan deposit with 1 Mt Cu @ 1.01% [26]
and the Xinqiao deposit with 0.5 Mt Cu @ 0.71% [27], which have attracted many studies
focusing on their genesis [28–36]. These stratabound deposits were firstly considered to be of
SEDEX origin by many researchers because of their stratiform orebodies and massive sulfide
ores, and the major orebodies occurring in the Carboniferous strata were thought to be products
of Late Paleozoic (Hercynian) sedimentary exhalative system [28–30]. Some researchers further
proposed an exhalative origin overprinted by Yanshanian magmatic-hydrothermal processes, based on
the restricted occurrences of the stratabound orebodies in areas where Yanshanian intrusions are
particularly extensive [31–33]. In contrast, some authors advocated that the stratabound mineralization
is of epigenetic origin and genetically associated with the Jurassic-Cretaceous tectono-thermal
events [34–36]. The precise geochronological data derived from recent studies confirmed that the
massive sulfide and skarn orebodies were coeval with the Yanshanian intrusions [25,27,37], supporting
the magmatic-hydrothermal origin of these stratabound deposits.
Although the genesis of stratabound deposits is still disputable, most of the researchers
tend to agree with the epigenetic origin or at least the dominant contribution of Yanshanian
magmatic-hydrothermal activities in the superimposed ore-forming processes [38,39]. In the
199
Minerals 2018, 8, 254
magmatic-related genetic model, the stratiform orebodies were formed as a result of progressive
fluid-rock interaction along the bedding-parallel structurally controlled conduits and were integral but
distal parts of a large hydrothermal system that produced the proximal skarn orebodies at the contact
zones and porphyry orebodies in the Yanshanian intrusions [35]. Such hydrothermal system and
stratabound deposits are similar to their counterparts elsewhere [35], among which the porphyry-skarn
polymetallic deposits in the Ertsberg district of Indonesia and manto-type copper deposits in Chile are
two representative examples. In the Ertsberg district, the Ertsberg East skarn orebody, one of the largest
orebodies, is hosted by a bedding-parallel fault-bounded zone between the limestone of the Faumai
Formation and dolomitic carbonate of the Waripi Formation [40]. In the Punta del Cobre district in
Chile, the stratabound tabular orebodies occur in the andesite breccia horizons between underlying
massive andesite and overlying shale, while the sites of economic copper concentration appear to be
controlled by faults [41]. Since structure is an important controlling factor of these stratabound deposits,
some relevant studies have been conducted in the TOD, including the spatial patterns [34], deformation
model [42] and formative process [43] of ore-controlling structures. However, these studies mostly
focus on theoretical deduction and qualitative analysis, and lack quantitative analysis concerning
detailed structural features. Hence, this paper attempts to delineate the structural controls by both
qualitative and quantitative analytical methods, focusing on the structural controlling mechanisms
operating at different scales, which can facilitate the understanding of the formation of copper deposits
and provide criteria for future exploration in the TOD.
200
Minerals 2018, 8, 254
discovered in the TOD, mainly clustering in four ore fields designated as Tongguanshan, Shizishan,
Fenghuangshan and Shatanjiao ore field from west to east (Figure 1b).
The copper occurrences (including known deposits and prospects) and structural features
employed in this study were derived from Geological Database of Bureau of Geological and Mineral
Resources of Anhui Province based on 1:50,000 geological survey and complemented by the literature
available for the study area concerning regional geological settings [26,42,45,49,51]. The raw data were
examined before being inputted into a spatial database. Only those copper and copper-dominated
polymetallic deposits were included in the analysis, since the other types of copper-related polymetallic
deposits may be products of different structurally controlled processes when compared with copper
mineralization. The structural features were reclassified into three categories including the basement
faults, cover faults and folds. All the examined data were compiled to vector formats and imported
into the ArcGIS 10 platform (Environmental Systems Research Institute, Redlands, CA, USA) for the
subsequent spatial analyses.
201
Minerals 2018, 8, 254
Figure 1. Geological map of the study area: (a) simplified tectonic map showing the location of the
TOD; and (b) geological map of the TOD showing the locations of copper occurrences, modified from a
1:50,000 scale geological map [52] and [37,42,43,53].
where DB is the box-counting fractal dimension, and A is a constant. Practically, a graph of log(N(δ))
versus log(δ) is plotted and then a best-fit regression line is drawn by the least square method, while
the slope of the regression line represents the box-counting fractal dimension (Figure 2d).
202
Minerals 2018, 8, 254
Figure 2. Schematic diagram of box-counting analysis: (a) 7 boxes containing target points with box size
δ = 8; (b) 9 boxes counted with box size δ = 6; (c) 14 boxes counted with box size δ = 3; and (d) log-log
plot revealing the power-law relationship of counted box number N(δ) and box size δ, obtaining
box-counting fractal dimension DB = 0.7071.
In the radial-density method, fractal points, also called fractal dusts, have been demonstrated to
satisfy a radial-density relationship, which can be described as [55]:
d ∝ Br DR −2 (2)
where, d is the average point density of the circles with a radius r that center in every point, and B
is a constant, while DR is the radial-density fractal dimension. Likewise, DR is usually obtained by
calculating the slope of a regression line that presents the linear relationship of d and r in a log-log plot.
203
Minerals 2018, 8, 254
Figure 3. Schematic diagram for constructing a Fry plot: (a) the original sheet records raw points;
(b) the origin O is placed on one of the raw points; (c) distribution pattern of raw points according to
the origin O is transferred to the tracing sheet; (d,e) the origin O is re-placed on every raw point; (f) the
tracing sheet records all distribution patterns of raw points with respect to different origins; and (g) Fry
plot is constructed.
where M is the number of mineral occurrences that used to estimate DM , while N is the number
of non-occurrence locations using for calculating DN , and 9.21 is a constant for significance level
α = 0.01 [24].
204
Minerals 2018, 8, 254
D
Pprior = P( D ) = (4)
T
and the relative importance of spatial association between the geological feature Bi and mineralization
is estimated by a pair of weights, namely positive weight W + and negative weight W − , which can be
given by:
+ P( B| D ) − P( B D )
W = ln , W = ln (5)
P( B D ) P( B D )
where P denotes the corresponding probability; B and B are the presence and absence of geological
features; D and D are the presence and absence of mineral occurrences. P(B|D), for example, represents
the probability of B occurring given the presence of D. The contrast C is defined as an overall
measurement of spatial correlation, which is given by:
C = W+ − W− (6)
In order to evaluate the significance of the contrast C, the confidence of the contrast (denoted
as CS ), obtained from a Student t-test, is employed here and defined as:
C C
CS = = (7)
S(C ) S 2 (W + ) + S 2 (W − )
205
Minerals 2018, 8, 254
(a) (b)
Figure 4. Log-log plot defining the fractal dimensions of spatial pattern of copper occurrences in the
TOD: (a) box-counting linear relationship; and (b) radial-density linear relationship.
The results of fractal analyses in this study, including the multi-fractal dimension model and
fractal structures occurring within identical ranges, are consistent with those of some previous
studies [10,15,16,24]. It is considered that discrepancies in fractal dimensions are plausibly linked to
different geological controls operating at diverse scales, e.g., regional-, local- and prospect-scale [10].
Nevertheless, such scale-variable geological controls are still cryptic and need to be delineated by
further analysis.
Fry analysis has been performed to investigate the orientations of plausible controls on copper
mineralization. 3906 Fry points were delivered from 63 copper occurrences in the TOD (Figure 5a),
based on which rose diagrams were constructed. The rose diagram for all Fry points illustrates a
simply dominant EW trend (Figure 5b), suggesting a fundamental EW-trending control on copper
mineralization at regional scale. Since fractal analyses indicate variations in fractal dimensions around
1.5 and 4.5 km, we also analyzed the characteristics of Fry points within these ranges. The rose diagram
for Fry points within 4.5 km of each other indicates a preferential NNE trend, with subordinate NE
and EW trends (Figure 5c). The rose diagram for Fry points within 1.5 km of each other exhibits a main
NE-NEE trend, with subsidiary trends in EW and NS directions (Figure 5d).
The results of Fry analysis infer different directional controls at regional- (>4.5 km) and fine-
(<4.5 km) scales, which could be correlated to detailed structural features in the TOD. However, such
correlation is not specific. For example, the NE-trending control at fine scale may be related to the
NE-trending faults or be linked to the folds with NE-striking axes. Further analysis is necessary
so as to delineate the one-to-one correspondence between the scale-variable controls and detailed
structural features.
206
Minerals 2018, 8, 254
Figure 5. (a) Fry plot showing spatial distribution of Fry points derived from 63 copper occurrences;
and rose diagram for (b) all Fry points; (c) Fry points within 4.5 km; (d) Fry points within 1.5 km.
207
Minerals 2018, 8, 254
Figure 6. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
EW-trending faults.
Figure 7. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
NS-trending faults.
The folds exhibit a statistically significantly positive correlation with copper occurrences in the
buffers ranging from 1.5 to 3 km (Figure 9). There is 22% higher frequency of copper occurrences than
what would be expected at a 2.5 km buffer (Figure 9b).
208
Minerals 2018, 8, 254
Figure 8. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
intersections of basement faults.
Figure 9. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance
to folds.
209
Minerals 2018, 8, 254
The cover faults consisting of NE- and NW-trending faults as well as the intersections of these
faults show positive spatial association with copper occurrences. There are at most 11%, 10% and 9%
higher frequencies of occurrences than what would be expected within the optimal buffers of NE-,
NW-trending faults and their intersections, respectively (Figures 10–12). Nevertheless, none of these
structural features have a statistically significantly associated with copper occurrence at any buffer
distance (Figures 10b, 11b and 12b).
WofE analysis was also implemented to investigate the association of structural features with
copper occurrences. At corresponding optimal buffer distances, the contrast values and confidences of
contrast were calculated. As depicted in Figure 13 and Table 2, the EW-trending faults, intersections of
basement faults and folds have top three highest values of both contrast and confidence of contrast,
which are remarkably greater than those of the other structural features. The contrasts and confidences
of contrast, which can assist in evaluating the intensity of spatial association, show exactly the same
variations as the results derived from distance distribution analysis, implying that the EW-trending
faults, intersections of basement faults and folds are plausibly major structural controls on copper
mineralization in the TOD.
Figure 10. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
NE-trending faults.
210
Minerals 2018, 8, 254
Figure 11. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
NW-trending faults.
Figure 12. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
intersections of cover faults.
211
Minerals 2018, 8, 254
Figure 13. Graph showing variations of higher frequencies than expected, contrasts and confidences of
contrast of detailed structural features in the TOD.
212
Minerals 2018, 8, 254
of basement faults and EW-trending faults, respectively (Figure 15a,c), suggesting strong associations of
these structural features with intrusions. The NS-trending basement faults have a moderate correlation
with intrusion, delineated by 11% higher frequency of intrusion regions than what would be expected
(Figure 15b). In contrast, cover faults and their intersections show negative correlations with intrusion
regions within a 1.5 km buffer (Figure 15d–f). Beyond the buffer distance of 1.5 km, they show positive
but weak associations with intrusions. There are 6%, 9% and 5% higher frequencies of intrusion regions
than what would be expected at the optimal buffers of NE-, NW-trending faults and their intersections,
respectively (Figure 15d–f).
It is noteworthy that the EW-trending faults and intersections of basement faults, which show the
strongest correlations with intrusions, also exhibit significant associations with copper occurrences
in the previous distance distribution analysis. It is necessary to evaluate what extent of these
structural controls on intrusion determine their strong correlations with copper mineralization.
The EW-trending faults and Yanshanian intrusions were buffered with their optimal distances, and the
copper occurrences located within the corresponding buffered zones were counted. It appears that
98% (47 out of 48) of the copper occurrences distributed within the buffers of EW-trending faults are
located in the overlapping zones of the buffered EW-trending faults and intrusions which account for
33.58% of total area. Only one occurrence is included in the buffered zones where intrusions are absent
(occupying 66.42% of total area) (Figure 16). Likewise, 96% (49 out of 51) of copper occurrences located
within the buffers of intersections of basement faults are included in the overlapping zones of buffered
intersections of faults and intrusions that occupy 37.11% of total area (Figure 17). It is inferred that
the significantly strong associations of EW-trending faults and intersections of basement faults with
copper mineralization are attributed to the controls of these structural features on intrusions.
Figure 14. (a) Buffer analysis and (b) graph of cumulative relative frequency concerning distance to
boundaries of intrusions.
213
Minerals 2018, 8, 254
Figure 15. Graph of cumulative relative frequency concerning distance to (a) EW-trending faults;
(b) NS-trending faults; (c) intersections of basement faults; (d) NE-trending faults; (e) NW-trending
faults and (f) intersections of cover faults.
214
Minerals 2018, 8, 254
Figure 16. Buffer analysis showing the distribution of copper occurrences in buffered EW-trending
faults and intrusions.
Figure 17. Buffer analysis showing the distribution of copper occurrences in buffered intersections of
basement faults and intrusions.
215
Minerals 2018, 8, 254
movement (ca. 135 Ma) characterized by transformation from contraction to extension since early
Cretaceous [35], which induced the formation of widespread intermediate-felsic intrusions and
associated mineralization. The multi-stage tectonic evolution is responsible for the structural features
in both basement and sedimentary cover that are related to epigenetic copper mineralization.
The basement structures are dominated by EW- and NS-trending faults. These faults, totally
overlain by Mesozoic strata, are considered to be formed before Indosinian period and reactivated in
the Mesozoic [45], although the detailed geometrical and kinematical characteristics of these faults are
still not clear. In previous studies, faults have been proven to act as favorable pathways for transporting
ore-related magma and ore-forming fluids from deep sources to shallow trap zones, resulting in a
strong association of these faults with hydrothermal mineral deposits [10,16,19]. In the study area,
the petrological data and geophysical profiles evidence that a magma chamber was developed in the
Mesozoic at about −10 km from the surface [42]. The EW-trending basement faults are interpreted to
play a vital role of channeling the magma from the magma chamber to the shallow trap zones during
the Yanshanian period. This significant control of the EW-trending faults on Yanshanian intrusions
is supported by distance distribution and WofE analyses, which is fully responsible for the strong
correlation between the EW-trending faults and copper mineralization. This interpretation can explain
the result of Fry analysis that exhibits a predominant EW trend at regional scale.
The known copper deposits are situated in the sedimentary cover where the folds with
sigmoidal-shaped axes are the dominant structures, thus delineating the formative process of the
folds is crucial for understanding the structural framework and copper mineralization in the cover
level. Since the youngest stratum involved in the folds is Middle Triassic, it is deduced that the folds
were formed during the Indosinian movement which resulted in the angular unconformity between
Middle Triassic and Lower Jurassic (Table 1). A classic model of dextral simple-shear deformation
in a strike-slip fault zone is introduced to illustrate the formation of folds and faults under the
deformation regime of Indosinian movement dominated by NW-SE compression and dextral shear
(Figures 18 and 19). As the fault zone initiates, a structural system forms consisting of (i) conjugate
strike-slip faults, (ii) folds, (iii) reverse faults, and (iv) normal faults (Figure 19a) [61,62]. The initially
formed folds and reverse faults trend perpendicular to the direction of the greatest shortening,
while the normal faults trend parallel to the direction of the greatest shortening. Subsequently,
the continued strike-slip shearing can lead to a rotation of the elements in this system [62]. The axes
of previously formed folds turn to sigmoidal shape. The earlier formed normal faults accommodate
sinistral strike-slip motion, and the reverse faults accommodate dextral strike-slip motion (Figure 19b).
The NE-trending thrust faults observed in the field [63] and sinistral strike-slip motion of NW-trending
faults identified in the geological map (Figure 1) support the rationality of this model.
In the Mesozoic strata, there existed several interfaces between two adjacent strata which have
distinct mechanical properties, some of which also represented the interfaces of disconformity, e.g.,
the interface between the quartz sandstone of Upper Devonian and limestone of Upper Carboniferous.
During the formative process of the folds in Indosinian period, the abovementioned interfaces were
subjected to the progressive deformation of folding and shearing, leading to extensive bedding-parallel
shear zones [43] (Figures 20 and 21). In particular, the bedding detachments occur in the cores of
the folds due to the layer-parallel slippage in the formative process of folds. These shear zones
were overprinted by tensional deformation in the Cretaceous when the tectonic regime in this
region changed from compression to extension, thus being favorable for trapping and localizing
mineralized fluids. This inference is supported by (i) the clearly discordant boundaries between
stratiform orebodies and wall rocks which suggest that the ores were deposited in mechanical dilation
spaces (Figure 22) [27,35,64], and (ii) the result of a numerical modeling on the Dongguashan deposit
which demonstrates that the stratiform high dilation zones induced by extensional stress are favorable
for fluids focusing and consistent with those positions where orebodies actually occur [64]. In addition,
the bedding-parallel trap zones are located near the contacts of intrusions where sufficient sources of
heat and fluid are available, and hosted in a set of carbonate strata suitable for forming skarn (Figure 20).
216
Minerals 2018, 8, 254
Therefore, the bedding-parallel structures in the folded strata are favorable for hosting, focusing and
depositing ore-bearing fluids, assisting in the formation of the stratabound orebodies in this area.
The thickening of orebodies in the cores of folds is attributed to the detachments occurring there (e.g.,
major orebody within C2 in Figure 20). This interpretation is supported by distance distribution and
WofE analyses, which both exhibit strong spatial association of the folds with copper mineralization.
It is also inferred that the dominance of NE, NNE and NEE trends in the rose diagrams of Fry points at
fine scales (<4.5 km) is attributed to the control of the folds with NE-striking axes, rather than those
of the NE-trending faults which show poor correlation with copper mineralization through spatial
analyses. Moreover, neither cover faults of various orientations nor the intersections of these faults
show statistically significant correlation with copper mineralization, suggesting that they may only
play a role in migrating the ore-bearing fluids towards the favorable host structures (i.e., multi-layered
bedding-parallel shear zones) where fluid concentration and mineral deposition actually occurred,
therefore leading to a lesser significant association of these cover faults with copper occurrences.
Figure 18. Stress regime during the formative process of the folds with sigmoidal axes, modified from [42].
Figure 19. Deformation model of dextral shearing in a strike-slip fault zone, modified from [61].
(a) a structural system formed in initial stage of deformation; and (b) a rotation of structural elements
during continued strike-slip shearing.
217
Minerals 2018, 8, 254
Figure 20. Typical cross-section of Shizishan ore field showing the characteristic stratabound skarn
orebodies hosted in the folds, modified from [43].
Figure 21. Field photograph of outcropped bedding-parallel shear zone between limestone of Upper
Carboniferous and quartz sandstone of Upper Devonian in the Xinqiao deposit.
218
Minerals 2018, 8, 254
Figure 22. Photographs showing the discordant boundaries between stratiform orebodies and wall
rocks in the Xinqiao deposit. (a) the boundary between orebody and underlying Upper Devonian quartz
sandstone; and (b) the boundary between orebody and overlying Upper Carboniferous limestone.
4. Conclusions
(i) Fractal dimensions obtained from box-counting and radial-density analyses suggest that different
structural controls operate at diverse scales of <1.5 km, 1.5–4.5 km and >4.5 km. This scale-variable
controlling behavior is supported and explored by the results of Fry analysis, which illustrates
a dominant EW trend at regional scale (>4.5 km) and preferential NE-NNE-NEE trends at fine
scale (<4.5 km).
(ii) The spatial associations of detailed structural features with copper mineralization are
further investigated by quantitative spatial analyses. The Yanshanian intrusions, EW-trending
faults, intersections of basement faults, and folds have significant associations with copper
mineralization, indicated by their high values of quantitative parameters in both distance
distribution and WofE analyses.
(iii) The interpretation of structural controls on copper mineralization is made in combination of
foregoing analytical results. The scale-variable patterns of mineral occurrences are attributed
to the different structural controls operating in the basement and sedimentary cover. In the
basement, the EW-trending faults serve as pathways for channeling magma from a magma
chamber into trap zones in the caprocks during Yanshanian period. The significant control of the
EW-trending faults on Yanshanian intrusion is fully responsible for the strong correlation between
the EW-trending faults and copper mineralization. This inference is supported by the result of Fry
analysis which shows a dominant EW trend at regional scale (>4.5 km). In the sedimentary cover,
the bedding-parallel shear zones formed during Indosinian folding and shearing and overprinted
by tensional deformation in Yanshanian period act as favorable sites for hosting, focusing and
depositing the ore-bearing fluids, which is responsible for the dominance of NE-NNE-NEE trends
at fine scale (<4.5 m) in the results of Fry analysis. Such bedding-parallel structures, together
with the contact zones of intrusion, exert an important control on the formation of characteristic
stratabound skarn deposits in the TOD.
Author Contributions: T.S. conducted the GIS-based computational experiments, analyzed the results and wrote
the draft paper; Y.X. and X.Y. participated in the analysis of experimental results; W.L. and R.L. revised the
calculation scheme; Z.H. and Y.W. collected the original data.
Acknowledgments: The research leading to this paper was jointly supported by National Natural Science
Foundation of China (Grant No. 41602335), Natural Science Foundation of Jiangxi Province (Grant
No. 20161BAB213084), Science and Technology Project of Jiangxi Provincial Department of Education (Grants
No. GJJ150625 and No. GJJ170537), Program of Qingjiang Excellent Young Talents (Grant No. JXUSTQJYX2017001)
and Doctoral Scientific Research Foundation of Jiangxi University of Science and Technology (Grant No. jxxjbs15002).
219
Minerals 2018, 8, 254
We would like to express our gratitude to two anonymous Minerals reviewers for their constructive comments
and suggestions that greatly improved the manuscript. Thanks are also given to Zhongfa Liu from Central South
University for assistance with field evidences.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Sillitoe, R.H. A plate tectonic model for the origin of porphyry copper deposits. Econ. Geol. 1972, 67, 184–197.
[CrossRef]
2. Tosdal, R.M.; Richards, J.P. Magmatic and structural controls on the developments of porphyry
Cu ± Mo ± Au deposits. Rev. Econ. Geol. 2001, 14, 157–181.
3. Kwelwa, S.D.; Dirks, P.H.G.M.; Sanislav, I.V.; Blenkinsop, T.; Kolling, S.L. Archaean gold mineralization in
an extensional setting: The structural history of the Kukuluma and Matandani Deposits, Geita Greenstone
Belt, Tanzania. Minerals 2018, 8, 171. [CrossRef]
4. Cox, S.F.; Knackstedt, M.A.; Braun, J. Principles of structural control on permeability and fluid flow in
hydrothermal systems. Rev. Econ. Geol. 2001, 14, 1–24.
5. 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. [CrossRef]
6. Sillito, R.H. Iron oxide-copper-gold deposits: An Andean view. Miner. Deposita 2003, 38, 787–812. [CrossRef]
7. Zeng, M.; Zhang, D.; Zhang, Z.; Liu, 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. [CrossRef]
8. Chauvet, A.; Piantone, P.; Barbanson, L.; Nehlig, P.; Pedroletti, I. Gold deposit formation during collapse
tectonics: Structural, mineralogical, geochronological, and fluid inclusion constraints in the Ouro Preto Gold
Mines, Quadrilátero Ferrífero, Brazil. Econ. Geol. 2001, 96, 25–48. [CrossRef]
9. Bonham-Carter, G.F. Geographic Information System for Geoscientists, Modeling with GIS; Pergamon: Elmsford,
NY, USA, 1994; pp. 238–333.
10. Haddad-Martim, P.M.; Filho, C.R.D.S.; Carranza, E.J.M. Spatial analysis of mineral deposit distribution:
A review of methods and implications for structural controls on iron oxide-copper-gold mineralization in
Carajás, Brazil. Ore Geol. Rev. 2017, 81, 230–244. [CrossRef]
11. Schetselaar, E.; Ames, D.; Grunsky, E. Integrated 3D geological modeling to gain insight in the effects of
hydrothermal alteration on post-ore deformation style and strain localization in the Flin Flon Volcanogenic
Massive Sulfide Ore System. Minerals 2018, 8, 3. [CrossRef]
12. Sun, T.; Wu, K.X.; Chen, L.K.; Liu, W.M.; Wang, Y.; Zhang, C.S. Joint application of fractal analysis and
weights-of-evidence method for revealing the geological controls on regional-scale tungsten mineralization
in Southern Jiangxi Province, China. Minerals 2017, 7, 243. [CrossRef]
13. Li, X.H.; Yuan, F.; Zhang, M.M.; Jia, C.; Jowitt, S.M.; Ord, A.; Zheng, T.K.; Hu, X.Y.; Li, Y. Three-dimensional
mineral prospectivity modeling for targeting of concealed mineralization within the Zhonggu iron orefield,
Ningwu Basin, China. Ore Geol. Rev. 2015, 71, 633–654. [CrossRef]
14. Xie, J.Y.; Wang, G.W.; Sha, Y.Z.; Liu, J.J.; Wen, B.T.; Nie, M.; Zhang, S. GIS prospectivity mapping and 3D
modeling validation for potential uranium deposit targets in Shangnan district, China. J. Afr. Earth Sci. 2017,
128, 161–175. [CrossRef]
15. Carranza, E.J.M. Developments in GIS-based mineral prospectivity mapping: An overview. In Proceedings of
the Mineral Prospectivity, Current Approaches and Future Innovations, Orléans, France, 24–26 October 2017.
16. Parsa, M.; Maghsoudi, A.; Yousefi, M. Spatial analyses of exploration evidence data to model skarn-type
copper prospectivity in the Varzaghan district, NW Iran. Ore Geol. Rev. 2018, 92, 97–112. [CrossRef]
17. Mandelbrot, B.B. Fractals: Form, Chances and Dimension; W.H. Freeman: New York, NY, USA, 1977; pp. 1–23.
18. Roberts, S.; Sanderson, D.J.; Gumiel, P. Fractal analysis of Sn-W mineralization from central Iberia; insights
into the role of fracture connectivity in the formation of an ore deposit. Econ. Geol. 1998, 93, 360–365.
[CrossRef]
19. Carranza, E.J.M.; Owusu, E.A.; Hale, M. Mapping of prospectivity and estimation of number of undiscovered
prospects for lode gold, Southwestern Ashanti Belt, Ghana. Miner. Deposita. 2009, 44, 915–938. [CrossRef]
220
Minerals 2018, 8, 254
20. Mehrabi, B.; Ghasemi, S.M.; Tale, F.E. Structural control on epithermal mineralization in the Troud-Chah
Shirin belt using point pattern and Fry analyses, North of Iran. Geotectonics 2015, 49, 320–331. [CrossRef]
21. Agterberg, F.P.; Bonham-Carter, G.F.; Wrigh, D.F. Statistical pattern integration for mineral exploration.
In Computer Application in Resource Estimation Prediction and Assessment for Metals and Petroleum; Gaal, G.,
Merriam, D.F., Eds.; Pergamon: Elmsford, NY, USA, 1990; pp. 1–21.
22. Cheng, Q.M.; Agterberg, F.P. Fuzzy weights of evidence method and its application in mineral potential
mapping. Nat. Resour. Res. 1999, 8, 27–35. [CrossRef]
23. Yuan, F.; Li, X.H.; Zhang, M.M.; Jowitt, S.M.; Jia, C.; Zheng, T.K.; Zhou, T.F. Three-dimensional weights
of evidence-based prospectivity modeling: A case study of the Baixiangshan mining area, Ningwu Basin,
Middle and Lower Yangtze Metallogenic Belt, China. J. Geochem. Explor. 2014, 145, 82–97. [CrossRef]
24. Carranza, E.J.M. Controls on mineral deposit occurrence inferred from analysis of their spatial pattern and
spatial association with geological features. Ore Geol. Rev. 2009, 35, 383–400. [CrossRef]
25. Cao, Y.; Zheng, Z.; Du, Y.; Gao, F.; Qin, X.; Yang, H.; Lu, Y.; Du, Y. Ore geology and fluid inclusions of the
Hucunnan deposit, Tongling, Eastern China: Implications for the separation of copper and molybdenum in
skarn deposits. Ore Geol. Rev. 2017, 81, 925–939. [CrossRef]
26. Liu, L.M.; Zhao, Y.L.; Zhao, C.B. Coupled geodynamics in the formation of Cu skarn deposits in the
Tongling–Anqing district, China: Computational modeling and implications for exploration. J. Geochem. Explor.
2010, 106, 146–155. [CrossRef]
27. Zhang, Y.; Shao, Y.J.; Li, H.B.; Liu, Z.F. Genesis of the Xinqiao Cu–S–Fe–Au deposit in the Middle-Lower
Yangtze River Valley metallogenic belt, Eastern China: Constraints from U–Pb–Hf, Rb–Sr, S, and Pb isotopes.
Ore Geol. Rev. 2017, 86, 100–116. [CrossRef]
28. Fu, S.G.; Yan, X.Y.; Yuan, C.X. Geologic feature of submarine volcanic eruption-sedimentary pyrite type
deposit in Carboniferous in the Middle-Lower Yangtze River Valley metallogenic belt, Eastern China.
J. Nanjing Univ. Nat. Sci. Ed. 1977, 4, 43–67. (In Chinese)
29. Gu, L.X.; Xu, K.Q. On the carboniferous submarine massive sulfide deposit in the lower reaches of the
Yangtze River. Acta Geol. Sin. 1986, 60, 176–188. (In Chinese)
30. Gu, L.X.; Hu, W.X.; He, J.X. Regional variations in ore composition and fluid features of massive sulfide
deposits in South China: Implications for genetic modeling. Episodes 2000, 23, 110–118.
31. Yang, D.F.; Fu, D.X.; Wu, N.X. Genesis of pyrite type copper in Xinqiao and its neighboring region according
to ore composition and structure. Issue Nanjing Inst. Geol. Miner. Resour. Chin. Acad. Geol. Sci. 1982, 3, 59–68.
(In Chinese)
32. Xie, H.G.; Wang, W.B.; Li, W.D. The genesis and metallogenetic of Xinqiao Cu–S–Fe deposit, Anhui Province.
Volcanol. Miner. Resour. 1995, 16, 101–107. (In Chinese)
33. Zhou, T.F.; Zhang, L.J.; Yuan, F.; Fang, Y.; Cooke, D.R. LA-ICP-MS in situ trace element analysis of pyrite
from the Xinqiao Cu–Au–S Deposit in Tongling, Anhui, and its constrains on the ore genesis. Earth Sci. Front.
2010, 17, 306–319. (In Chinese)
34. Chang, Y.F.; Liu, X.G. Layer control type skarn type deposit—Some deposits in the Middle-Lower Yangtze
Depression in Anhui Province as an example. Miner. Depos. 1983, 2, 11–20. (In Chinese)
35. Pan, Y.; Dong, P. The lower Changjiang (Yangtzi/Yangtze River) metallogenic belt, East-center China:
Intrusion and wall rock hosted Cu–Fe–Au, Mo, Zn, Pb, Ag deposits. Ore. Geol. Rev. 1999, 15, 177–242.
[CrossRef]
36. Mao, J.W.; Shao, Y.J.; Xie, G.Q.; Zhang, J.D.; Chen, Y.C. Mineral deposit model for porphyry-skarn
polymetallic copper deposits in Tongling ore dense district of Middle-Lower Yangtze Valley metallogenic
belt. Miner. Depos. 2009, 28, 109–119. (In Chinese)
37. Zhang, Y.; Shao, Y.; Zhang, R.; Li, D.; Liu, Z.; Chen, H. Dating ore deposit using garnet U–Pb geochronology:
Example from the Xinqiao Cu–S–Fe–Au deposit, Eastern China. Minerals 2018, 8, 31. [CrossRef]
38. Zhou, T.; Wang, S.; Fan, Y.; Yuan, F.; Zhang, D.; White, N.C. A review of the intracontinental porphyry
deposits in the Middle-Lower Yangtze River Valley metallogenic belt, Eastern China. Ore Geol. Rev. 2015, 65,
433–456. [CrossRef]
39. Hu, R.Z.; Chen, W.T.; Xu, D.R.; Zhou, M.F. Reviews and new metallogenic models of mineral deposits in
South China: An introduction. J. Asian Earth Sci. 2017, 137, 1–8. [CrossRef]
40. Mertig, H.J.; Rubin, J.N.; Kyle, J.R. Skarn Cu–Au orebodies of the Gunung Bijih (Ertsberg) district, Irian Jaya,
Indonesia. J. Geochem. Explor. 1994, 50, 179–202. [CrossRef]
221
Minerals 2018, 8, 254
41. Sato, T. Manto type copper deposit in Chile—A review. Bull. Geo. Surv. Japan 1984, 35, 565–582.
42. Wang, Q.F.; Deng, J.; Huang, D.H.; Xiao, C.H.; Yang, L.Q.; Wang, Y.R. Deformation model for the Tongling
ore cluster region, East-Central China. Int. Geol. Rev. 2011, 53, 562–579. [CrossRef]
43. Wu, G.G.; Zhang, D.; Zang, W.S. Study of tectonic layering motion and layering mineralization in the
Tongling metallogenic cluster. Sci. China Ser. D Earth Sci. 2003, 46, 852–863. [CrossRef]
44. Li, Y.; Li, J.W.; Li, X.H.; Selby, D.; Huang, G.H.; Chen, L.J.; Zheng, K. A carbonate replacement origin for the
Xinqiao stratabound massive sulfide deposit, middle-lower Yangtze Metallogenic Belt, China. Ore Geol. Rev.
2017, 80, 985–1003. [CrossRef]
45. Chang, Y.F.; Liu, X.P.; Wu, Y.C. The Copper–Iron Belt of the Low and Middle Reaches of the Changjiang River;
Geological Publish House: Beijing, China, 1991; pp. 1–359. (In Chinese)
46. Liu, L.M.; Yang, G.Y.; Peng, S.L.; Zhao, C.B. Numerical modeling of coupled geodynamical processes and its
role in facilitating predictive ore discovery: An example from Tongling, China. Resour. Geol. 2005, 55, 21–31.
[CrossRef]
47. Liu, W.C.; Li, D.X.; Gao, D.Z. Analysis on the time sequence of compounding of structural deformation
systems and resulting effects in Tongling area. J. Geomech. 1996, 2, 42–48. (In Chinese)
48. Lü, Q.T.; Hou, Z.Q.; Zhao, J.H.; Shi, D.N.; Wu, X.Z.; Chang, Y.F.; Pei, R.F.; Huang, D.D.; Kuang, C.Y. Complex
crustal structure of Tongling ore district: Insights from deep seismic reflection profiling. Sci. China Ser. D
2003, 33, 442–449. (In Chinese)
49. Liu, Z.F.; Shao, Y.J.; Wei, H.T.; Wang, C. Rock-forming mechanism of Qingshanjiao intrusion in Dongguashan
copper (gold) deposit, Tongling area, Anhui province, China. Trans. Nonferr. Met. Soc. China 2016, 26,
2449–2461. [CrossRef]
50. Xie, J.C.; Yang, X.Y.; Sun, W.D.; Du, J.G. Early Cretaceous dioritic rocks in the Tongling region, Eastern China:
Implications for the tectonic settings. Lithos 2012, 150, 49–61. [CrossRef]
51. Liu, L.M.; Peng, S.L. Prediction of hidden ore bodies by synthesis of geological, geophysical and geochemical
information based on dynamic model in Fenghuangshan ore field, Tongling district, China. J. Geochem. Explor.
2004, 81, 81–98. [CrossRef]
52. 321 Geological Team. Structural Maps of Tongling Area; Bureau of Geological and Mineral Resources of Anhui
Province: Hefei, China, 1989; pp. 1–33.
53. Du, Y.L. Ore-Controlling Factors and Metallogenic Model of Stratabound Skarn Deposits in Tongling Area,
Anhui Province. Ph.D. Thesis, China University of Geosciences, Beijing, China, 2013. (In Chinese)
54. Zuo, R.G.; Wang, J. Fractal/multifractal modeling of geochemical data: A review. J. Geochem. Explor. 2016,
164, 33–41. [CrossRef]
55. Mandelbrot, B.B. The Fractal Geometry of Nature: Updated and Augmented; W.H. Freeman: New York, NY, USA,
1983; pp. 1–31.
56. Berman, M. Distance distributions associated with poisson processes of geometric figures. J. Appl. Probab.
1977, 14, 195–199. [CrossRef]
57. Berman, M. Testing for spatial association between a point process and another stochastic process. J. R. Stat.
Soc. C Appl. 1986, 35, 54–62. [CrossRef]
58. Allek, K.; Boubaya, D.; Bouguern, A.; Hamoudi, M. Spatial association analysis between hydrocarbon fields
and sedimentary residual magnetic anomalies using weights of evidence: An example from the Triassic
Province of Algeria. J. Appl. Geophys. 2016, 135, 100–110. [CrossRef]
59. Sang, X.J.; Xue, L.F.; Liu, J.W.; Zhan, L. A novel Workflow for geothermal prospectively mapping
weights-of-evidence in Liaoning Province, Northeast China. Energies 2017, 10, 1069. [CrossRef]
60. Deng, J.; Huang, D.H.; Wang, Q.F.; Hou, Z.Q.; Lü, Q.T.; Yao, L.Q.; Xin, H.B.; Zhang, Q.; Wei, Y.G. Formation
mechanism of “drag depressions” and irregular boundaries in intraplate deformation. Acta Geol. Sin. 2004,
78, 267–272.
61. Waldron, J.W.F. Extensional fault arrays in strike-slip and transtension. J. Struct. Geol. 2005, 27, 23–34.
[CrossRef]
62. David, G.H.; Reynolds, S.J.; Kluth, C.F. Structural Geology of Rocks and Regions, 3rd ed.; JohnWiley & Sons,
Inc.: Westwood, MA, USA, 2011; pp. 336–338.
222
Minerals 2018, 8, 254
63. Wang, Q.F. Model study of the tectonic-magmatic-metallogenical system in Tongling ore cluster area.
Ph.D. Thesis, China University of Geosciences, Beijing, China, 2005. (In Chinese)
64. Liu, L.M.; Sun, T.; Zhou, R.C. Epigenetic genesis and magmatic intrusion’s control on the Dongguashan
stratabound Cu-Au deposit, Tongling, China: Evidence from field geology and numerical modeling.
J. Geochem. Explor. 2014, 144, 97–114. [CrossRef]
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
223
minerals
Article
The Hajjar Regional Transpressive Shear Zone
(Guemassa Massif, Morocco): Consequences on the
Deformation of the Base-Metal Massive Sulfide Ore
Safouane Admou 1,2, *, Yannick Branquet 2,3 , Lakhlifi Badra 1 , Luc Barbanson 2 ,
Mohamed Outhounjite 4 , Abdelali Khalifa 4 , Mohamed Zouhair 4 and Lhou Maacha 4
1 Département des Sciences de la Terre, Faculté des Sciences, Université Moulay Ismaïl de Meknès,
B.P. 11201 Zitoune Meknès, Morocco; badra_lakhlifi@yahoo.fr (L.B.)
2 Institut des Sciences de la Terre d’Orléans (ISTO), Université Orléans, CNRS BRGM UMR7327,
Campus Géosciences 1A, rue de la Férollerie, 45071 Orléans, CEDEX 2, France;
[email protected] (Y.B.); [email protected] (L.B.)
3 Géosciences Rennes (GR), Université de Rennes 1, CNRS UMR6118, Campus de Beaulieu, CS 74205,
35042 Rennes CEDEX, France
4 Groupe MANAGEM, Twin center, Tour A, BP 5199, Casablanca, Morocco;
[email protected] (M.O.); [email protected] (A.K.);
[email protected] (M.Z.); [email protected] (L.M.)
* Correspondence: [email protected]
Abstract: The genesis of the base-metal massive sulfide deposits hosted within the Moroccan
Hercynian Jebilet and Guemassa Massifs is still under debate. No consensus currently exists
between the two models that have been proposed to explain the deposits, i.e., (1) syngenetic
volcanogenic massive sulfide mineralization, and (2) synmetamorphic tectonic fluid-assisted
epigenetic mineralization. Conversely, researchers agree that all Hercynian massive sulfide deposits
in Morocco are deformed, even though 3D structural mapping at the deposit scale is still lacking.
Therefore, while avoiding the use of a model-driven approach, the main aim of this contribution
is to establish a first-order structural pattern and the controls of the Hajjar base metal deposit. We
used a classical structural geology toolbox in surface and subsurface mining work to image finite
strain at different levels. Our data demonstrate that: i) the Hajjar area is affected by a single foliation
plane (not two) which developed during a single tectonic event encompassing a HT metamorphism.
This syn-metamorphic deformation is not restricted to the Hajjar area, as it is widespread at the
western Meseta scale, and it occurred during Late Carboniferous times; ii) the Hajjar ore deposit
is hosted within a regional transpressive right-lateral NE-trending shear zone in which syn- to
post-metamorphic ductile to brittle shear planes are responsible for significant inflexion (or virgation)
of the foliation yielding an anastomosing pattern within the Hajjar shear zone. Again, this feature
is not an exception, as various Late Carboniferous-Permian regional scale wrenching shear zones
are recognized throughout the Hercynian Meseta orogenic segment. Finally, we present several
lines of evidence emphasizing the role of deformation in terms of mechanical and fluid-assisted
ore concentrations.
Keywords: Hajjar; shear zone; base metal massive sulfide deposits; structural control; remobilization
1. Introduction
Most Volcanogenic Massive Sulfide Deposits (VMSDs) are assumed to form within extensional
and subsiding basins during both divergent and convergent plate tectonic settings (e.g., [1]). As a result,
in convergent settings leading to continental collision for instance, many VMSDs underwent
deformation, burial, and metamorphism. During these transformations, syngenetic massive sulfide
bodies (e.g., stratoid lenses, chimneys and stockwerks) were reworked, and primary metallic bearing
mineral assemblages may have been remobilized (e.g., either depleted or enriched). For this reason,
the deformation and (re)mobilization of the primary sulfide concentration is a fundamental and
economic matter which has been recognized and studied for a long time (e.g., [2–6]).
However, in spite of recent advances in modern textural (e.g., electron backscatter diffraction
coupled to chemistry) and opaque mineral strain characterization (e.g., [7–10]), it still remains difficult
for economic geologists dealing with deformed VMSD to decipher the respective parts of primary
syngenetic vs. epigenetic mineralizing processes. As a result, metallogenic models of very large base
metal concentrations all over the world are still ambiguous and under debate.
Currently, the genesis of polymetallic base-metal massive sulfide deposits (MSD) from the western
Meseta domain in Morocco are currently under debate. This debate is particularly relevant for MSD
from the Central Jebilet unit (Figure 1), e.g., the Kettara, Draa Sfar, Koudiat Aïcha, and Lachach
deposits. Many authors consider these MSD as metamorphosed and deformed primary VMS and/or
sedimentary exhalative (SEDEX) deposits [11–17]; however, other authors argue for a fluid-assisted
syn-metamorphic origin during the major Hercynian deformation event [18–22]. In contrast, the Hajjar
MSD located in the Hercynian Guemassa Massif (Figure 1) is considered as a metamorphosed and
deformed syngenetic VMS/SEDEX deposit [12,23–25]. Although Hajjar shares many similar geological
and mineralogical features (e.g., predominance of pyrrhotite) with the Central Jebilet MSD to the north,
the hypothesis of either an epigenetic or a syn-metamorphic origin has not yet been put forward.
Since the pioneering works of Hibti (1993) [23] on the Hajjar MSD, very few studies dealing with
the structural controls of this ore deposit have been carried out and published in the international
literature. However, on a larger scale, much thermal and geochronological data dealing with
the tectono-magmatic evolution of the western segment of the Hercynian Meseta have been
published [26–28]. Therefore, using data collected from new outcrops, the aim of this work is to
complete the Hajjar MSD structural dataset and to re-evaluate the structural context and controls; this
is a prerequisite to being able to have a potential syngenetic vs. syn-metamorphic debate, if required.
Our approach is to perform structural mapping at each subsurface exploitation level, yielding a 3D
view of the deformation pattern. This pattern is then compared to the structural map of the surface
outcrops in the Guemassa Massif.
225
Minerals 2018, 8, 435
These syn-sedimentary structures are encountered both within host rocks and sulfide mineralized
bodies; ii) a D1 event corresponding to the incipient Hercynian deformation and responsible for a steep
NW-SE foliation (S1) in the Oriental Guemassa associated with folding under regional greenschist facies
metamorphic conditions. It should be noted that S1 cannot be observed clearly within the Hajjar MSD;
iii) a D2 Hercynian tectono-thermal event with P2 folds and associated S2 planar cleavage oriented
NE-SW under low-grade metamorphism with sericite. S2 is the predominant foliation observable in
the Hajjar mine; and iv) finally, a post-kinematic thermal event, likely related to “hidden plutons”,
responsible for the crystallization of static biotite porphyroblasts with cordierite and andalousite
locally described at Hajjar. In this ore deposit, this thermal event has been dated using “hydrothermal”
biotite at ca. 301 Ma [33]. Moreover, for Carboniferous times, the Guemassa Massif is affected by
intense multiscale ductile to brittle faulting [34,35], with probable components of Atlasic reactivation
during the Tertiary High Atlas orogen (the Guemassa Massif is 15 km to the north of the Atlasic
thrusting front, Figure 1B). On a structural map (Figure 1B), these faults and shear zones cross-cut
and delineated several blocks within the Guemassa Massif. In the Oriental Guemassa, in which the
Hajjar mine is located, the N’Fis block appears to present a peculiar “anarchic” foliation orientation
with respect to the bulk NNE-trend of the main Hercynian foliation in the western Meseta domain.
These “anarchic” foliation orientations have been explained by deflection or virgations (here defined
as a bulk inflexion of foliation plane trajectories) induced by conjugate shear zones during or shortly
after a broad E-W-oriented D1 shortening [34,35]: the dominant and earlier shear zones are dextral
and trend ENE–WSW (e.g., the Imi-In-Tanout Fault, the eastern branch of the Amizmiz Fault, and the
Guemassa Fault, Figure 1B), whereas WNW-ESE-trending shear zones are sinistral, such as the Lalla
Takerkoust Fault (Figure 1B). This “virgation model” is compatible with a W–E horizontal shortening,
in contrast to Hibti’s hypothesis (1993) [23], which argued for a NE–SW horizontal shortening during
the D1 event (cf. supra).
Figure 1. (A) Structural map of Morocco showing the major bounding-fault domains. The arrows
indicate the sense of shear for the late Variscan structures (modified from Hoepffner et al., 2005 [36]);
(B) Geological and structural map of the central domain of the Hercynian belt (from [35,37]. The main
foliation trajectories in the Jebilet are reported from Essaifi, 1995 [18]). Within the Guemassa Massif,
the Hajjar base metal deposit is located in the N’Fis block which presents an “anarchic” foliation
orientation with respect to the bulk N to NNE trend reported in the Jebilet, Occidental Guemassa and
western High Atlas Variscan Massifs.
226
Minerals 2018, 8, 435
The geology of the Guemassa Massif is similar to the Central Jebilet domain (Figure 1B).
Both Massifs host the major MSD of the Occidental Meseta. Thus, recent advances in the
tectono-metamorphic and magmatic history of the Jebilet [26,28] may help better constrain the
Guemassa Massif evolution. Based on petro-structural data, new absolute dating and thermal
investigations, these authors improve the time constraints and the succession of the deformational
events as follows: i) from 370 to 325 Ma (D0 of Delchini, 2018 [26]), the Jebilet area was a basin
filled with syn- to post-rift sediments (the Sarhlef and Teksmin formations, respectively) intruded by
shallow sills and dykes and deeper plutonic laccoliths originating from a tholeitic bimodal magmatism
(e.g., the mafic/ultramafic Kettara and Sarhlef intrusions) and from a calc-alkaline magmatic suite
(e.g., the Oulad-Ouaslam granodiorite) respectively; ii) from 325 to around 310 Ma, a first Hercynian
event (D1) is marked by the emplacement of shallow thin skinned nappes with syn-sedimentary
breccias. The internal strain is very low and no regional foliation/cleavage (S1) is reported; iii) from
ca. 310 Ma to 280 Ma, the main Hercynian deformation (D2), which is polyphased and characterized
by a first regional metamorphism (M2a), locally reaches the amphibolite facies (Grt-St) and a second
HT/BP “contact” metamorphism in the syn-to post tectonic hornfels facies (M2b, biot + Crd + And)
is associated with the leucogranite emplacement around 295 Ma. The successive foliations (S2a and
S2b), sub-vertical and oriented N0/30, marked a homoaxial progressive and continual strain regime
from a coaxial to a non-coaxial transpression with a broad horizontal NW-SE-trending shortening axis.
Last, the D2 increments correspond to a right-lateral transpression accommodated and located along
the vertical and conjugate ductile shear zones as the sinistral MSZ (Figure 1B). Therefore, the tectonic
scenario proposed by Hibti (1993) [23] for the Guemassa which implies strain axis rotation between
D1 and D2 and post-tectonic HT/LP metamorphism diverges from the one proposed by Delchini
(2018) [26] for the Jebilet domain.
227
Minerals 2018, 8, 435
Figure 2. Geological map and surface structural data of the N’Fis block and the Souktana Massif. All
of the sedimentary formations are Carboniferous in age and are affected by both metamorphism and
deformation. IF: Imarine Fault; TF: Lala Takerkoust Fault; AKF: Ait Khaled Fault. The interpolation of
the foliation/shear planes is also supported by sub-surface structural data from underground mine
works (cf. infra). Note that the S1 trajectories depict a dextral drag fold against the ENE-trending
Tiferouine mineralized body.
Locally, the NW-trending S1 is marked by elongated and aligned biotite porphyroblasts, parallel
to the stretching of pyrrhotite grains (Figure 3D), suggesting a syn-tectonic growth of biotite.
No stretching lineation has been observed in the N’Fis block. Decimeter-scale sinistral WNW to
NW-trending vertical ductile shear planes, occurring sparsely and slightly oblique to S1, are responsible
for the local deflection of the S1 planes in the Imarine outcrops (six observations plotted on the
stereogram, Figure 2). Brittle faults and joints show a predominant NE-trending orientation with
a sub-vertical dip (Figure 2). Due to unfavorable rock materials, the precise kinematics of brittle faults
are difficult to establish, which enable the reconstruction of the paleo-stress using the right dihedral
method, for instance. However, when it can be observed, the apparent map offsets of the NE-trending
decimeter-scale faults indicate a dominant dextral sense of shear.
228
Minerals 2018, 8, 435
Finally, the Tiferouine outcrop (Figures 2 and 3E) shows a N70-trending gossan which corresponds
to the weathered part of a magnetite-bearing body recognized at depth [12]. The supergene alteration
appears to overprint an early cataclasite. Along and within the cataclased mineralized body, the S1
foliation orientation is strongly disturbed (Figure 3E), suggesting drag folding along a right-lateral
N70-trending wrench fault (also, see Figure 2 for a map view of the drag folding in the Tiferouine area).
Figure 3. Structures observed in the outcrops. (A) syn-sedimentary and soft sediment deformation
occurring as slumps and convolutes are widespread in the sandy limestones of the N’Fis block;
(B) obliquity between the S1 foliation plane and recumbent fold axial plane suggests that some isoclinal
folds are former slumps rather than P1 folds; (C) NW-trending S1 foliation plane developed within the
P1 hinge zone; (D) thin sections (cross polars normal to foliation) of metapelite with sulfide ribbons (Po:
pyrrhotite) from the N’Fis block. The bedding plane is transposed by the S1 foliation plane, the sulfide
ribbon and patches disseminated in the matrix are flattened. Biotite porphyroblasts are elongated
broadly parallel to the foliation plane; (E) mineralized Tiferouine body (see location in Figure 2) with
an associated gossan inside an ENE-trending dextral shear zone evidenced by cataclasites and the
re-orientation of S1.
229
Minerals 2018, 8, 435
230
Minerals 2018, 8, 435
Figure 4. Structural maps of the five main exploitation levels (decreasing altitude from A to E) in the
Hajjar mine. High strain corridors are marked by the development of dense foliation and shear planes.
The light blue traces are galleries. The coordinates are taken from the mine’s own system. Note that the
scale is slightly different for each level. The following acronyms are used for the ore bodies (translated
from French): CP = main body; CNE = north-eastern body; CWD = western body; CEWD = extreme
western body. The CP ore body has been intensively exploited and some zones are no longer accessible,
structural data from Hibti (1993) [23] were then added and carefully projected in these areas (see text
for explanation).
231
Minerals 2018, 8, 435
stretching lineation without the occurrence of reverse shear planes (called a “flattening corridor” below,
Figure 6).
Within reverse corridors, shear planes present dominant reverse rather than strike-slip kinematics.
In the map view (Figure 4), the obliquity between S1 and the shear planes, which seems to indicate
a sinistral sense of shear, is an artifact as the strike-slip component which is low and dextral when it
is observed. The noticeable meso-scale structures are: i) eastward verging thrusts and decollements,
most of time using a weak pyrrhotite-rich layer/body as the sole, which is near-parallel to the bedding
in the foot-wall (Figure 5A–C). The associated folds in the hanging-walls developed an axial-planar
cleavage S1. Typical meter-scale detachment folds, with thickening of the sulfide-rich decollement
level, are frequent (Figure 5C), which might explain the “corrugation” observed along the decollement
plane (Figure 5B); ii) the high strain corridors are characterized by the development of an intense
foliation associated with similar upright NS-oriented folds (Figure 5A and D) which are frequently
in association with reverse shear bands responsible for “pop ups” (Figure 5E). Local evidence of the
oblique-slip component is provided by oblique stria, the “pop ups” then corresponding to dextral
positive flower structures (Figure 5E).
Within the flattening corridors where thrusting is not observed, bedding marked by sulfide-rich
ribbons is fully overprinted by the S1 foliation which bears a horizontal NS stretching lineation
(Figure 6A). With increasing strain, the rock color changes to a very dark and black tint. To the west,
a massive sulfide body is exploited (CEWD). This body is not continuous as it is instead composed
of several distinct massive sulfide lenses aligned parallel to S1. The termination of the sulfide lenses
is wavy due to the occurrence of small-scale folds of sulfide ribbons or host rocks. This sulfide lens
morphology is frequently observed throughout the mine (e.g., Figure 8D). Near the termination, these
lenses integrate clasts of host rocks (Figure 6). Cm- to dm-thick veins are abundant along the high
strain corridor (Figure 6A). Locally, tips of massive sulfide lenses present triangular veins (or “saddle
reef”) at a “triple junction” position with respect to the foliation (Figure 6).
232
Minerals 2018, 8, 435
Figure 5. Structures and deformation of the Hajjar ore deposit. A to E are from the N to NNE-trending
high strain “reverse”corridors; F and G are from the NE to ENE-trending high strain corridors.
(A) Cross-section along the gallery from level 600-580 (see location in Figure 4A). S0 is shown in
green, S1 is in black, the brittle to ductile shear planes are given in blue, the main massive sulfide bodies
are shown in orange. The section is located within the footwall of the CP and is mainly composed of
stratified greso-pelites and tuffs with mm- to cm-thick sulfide ribbons (containing mostly pyrrhotite
and pyrite with a small amount of chalcopyrite) with no economic interest. The intensity/spacing of
the foliation and high frequency of the shear planes can be used to depict the high strain corridors.
Most of the brittle to ductile shear planes have an apparent reverse component: (B) an east-verging
thrust developed within a pyrrhotite-rich massive sulfide deposit acting as a decollement layer. In the
hanging-wall, the bedding is not observed whereas the S1 cleavage is curved by top-to-the-east drag
folding. Both massive sulfide wallrocks are corrugated (c. sp: corrugated shear plane); (C) Detachment
fold above a pyrrhotite-rich sulfide layer thickened within the core of a disharmonic fold hinge The S1
axial planar foliation is well-developed in the hanging-wall; (D) Upright similar fold with associated
axial planar S1 cleavage. The pyrrhotite-rich red ribbons are extremely thinned in the limbs and
thickened within the hinge zone; (E) Positive flower structure associated with similar drag folds and S1
cleavage (line drawing from level 600-580, Figure 4A). Along the N15E-trending faults, high dipping
stria show that the reverse component is dominant relatively to the dextral strike-slip one; (F) Ductile
dextral NE-trending near the vertical shear planes (C) and associated S1 foliation within a NE-trending
right-lateral high strain corridor in meta-siltstones (location CNE area, Figure 4D); (G) ENE-trending
steep dextral shear zones marked by foliated gouges and various branches (sense of shear is determined
in the gallery roof, location in Figure 4D).
233
Minerals 2018, 8, 435
Figure 6. Outcrops of the extreme western body (CEWD) gallery, a typical N-trending “flattening”
corridor. (A) 3D man-made sketch of the CEWD cross-section located in Figure 6A. The exploited
massive ore bodies correspond to meter-scale lenses aligned within a high strain zone marked by
an intense foliation in dark host rocks with high biotite and sulfide content. X and Z are the long
and short axis of the strain ellipsoid respectively; (B,C) Thin-section photographs of the triangular
veins developed at the massive ore lenses termination (RL). The vein is mainly filled with quartz
associated with a polymetallic assemblage. Pyrrhotite is replaced by pyrite along cracks (B,C) and
sphalerite/galena (± chalcopyrite) veinlets crosscut the former pyrrhotite and arsenopyrite grains
(not shown).
234
Minerals 2018, 8, 435
porphyroblasts replaced by white mica aggregates indicate the presence of this mineral (e.g., Figure 8E).
The second assemblage is made of quartz + chlorite + muscovite (± carbonate). This last assemblage
can also be observed in sandy-pelites with sulfide-rich ribbons, where it post-dates and locally replaces
the biotites (Figure 7B). Foliation-parallel veinlets are filled with quartz and large biotite crystals
associated with calcite in the geodic cavities (Figure 7A).
Figure 7. Thin section microphotographs of the S1 foliation and associated porphyroblasts within the
Hajjar host rocks. A, C, D, E and F are from the flattening corridors; B is from the reverse corridors.
(A) Quartz (qz), calcite (cal) and biotite (biot) vein parallel to the incipient S1 foliation, vertical section,
see location in Figure 6, NAPL. The host rock presents a fine-grained granoblastic texture composed
of biotite and andalousite grains with a local preferred orientation defining an incipient foliation
plane; (B) Footwall of the thrust (see location in Figure 5) with the So plane marked by sulfide-rich
ribbons (in blue) and discrete S1 planes (in red) characterized by muscovite (white laths) crystallization
(NAPL). Please note that the non-oriented biotite (i.e., “static”) porphyroblasts are replaced by chlorite
(pale green); (C–F) Horizontal thin sections parallel to the stretching lineation showing the main
foliation plane S1 marked by elongated sulfides (sulf) and particularly pyrrhotite (po) and sphalerite
(sph); see location in Figure 6. Like the fine-grained foliation, the pressure shadows and caps around
the andalousite (and) and biotite (biot) grains are composed of quartz, white micas, chlorite and
local carbonates. In the high strain area, asymmetric pressure shadows around the biotite indicate
a non-coaxial regime with a dextral sense of shear (E,F).
235
Minerals 2018, 8, 435
In areas where the foliation is weakly developed, biotite and andalousite porphyroblasts show
a granoblastic “static” texture with a very subtle preferred orientation locally (Figure 7A). With
increasing strain, biotite porphyroblasts are generally coarser and present a preferred orientation
parallel to the fine-grained S1 foliation, a planar axial surface with micro-folds (Figure 8C). In high
strain zones, pressure shadows and strain caps are found around some biotite and andalousite
crystals (Figure 7C to F), whereas other biotite crystals remain nearly free of foliation deflection
(e.g., a biotite crystal growing around a sphalerite core in Figure 7C). The pressure shadows are
generally composed of quartz, muscovite, and chlorite, i.e., the same assemblage constituting the
fined-grained foliation (Figure 7D,E). Asymmetric pressure shadows around biotite are common
in flattening corridors attesting to a non-coaxial regime, at least locally (e.g., dextral in the CEWD
outcrop, Figure 6, Figure 7E,F). Therefore, in high strain and non-coaxial zones, biotite crystals appear
as pre-tectonic prophyroblasts, suggesting severe non-coaxial strain increments after the HT/LP
metamorphism peak.
236
Minerals 2018, 8, 435
Figure 8. Deformation and textures of the sulfides in the Hajjar deposit. (A) The folding and associated
S1 foliation of fine-grained sediments containing early sulfide-rich ribbons parallel to the bedding
(S0). Note the cleavage refraction and thickening of the hinge zone due to the plastic behavior of
pyrrhotite; (B) Details of A with pyrrhotite flowing along the stretching direction whereas the behavior
of chalcopyrite and sphalerite is less plastic. A metamorphic assemblage mainly composed of muscovite
and chlorite (± biotite) grows parallel to S1; (C) Micro-fold affecting a sphalerite and pyrrhotite-rich
thin ribbon (CEWD, location in Figure 6A). The axial planar cleavage S1 is marked by the stretching of
sulfides and elongated biotite blasts; (D) Massive sulfide lenses separated by strongly foliated host rock
slices (south of CP, altitude 500 m). The ore bodies are internally banded parallel to the S1 foliation;
(E) Texture of deformed pyrrhotite-rich massive sulfide (RL) parallel to the S1 foliation. The dark
grey areas correspond to a muscovite/chlorite (replacing biotite locally) assemblage. Andalousite or
cordierite porphyroblast ghosts are replaced by white micas (arrow); (F) textural and mineralogical
banding within a massive sulfide body in the sole thrust (see location Figure 5A,B). Note the elongation
of the quartz grains; (G) Massive sulfide sample affected by ductile shearing and mylonitization (SE
part of the CP, level 400, the local name is “la bande Sud-Est”); (H) Details of G, thin section, RL.
The sulfide mylonites present typical C/S structures. It should be noted that sphalerite appears to be
“localized” in the C planes. The sample view from the bottom shows a dextral sense of shear.
237
Minerals 2018, 8, 435
5. Interpretation
5.1. Hajjar Mine and N’Fis Block: One Single Foliation (Not Two)
The rocks of the Hajjar mine are affected by one single flattening XY plane which is near vertical
and trends from N0 to N45. The maps of the S1 trajectories (Figure 4) show that the deformation
is not homogenous at the mine scale. In the high strain corridors, this XY plane corresponds to
a S1 penetrative foliation overprinting the entire rock, whereas in less deformed areas, S1 is a slaty
cleavage that is axial planar in similar folds. Host rocks and sulfide bodies present the same silicate
metamorphic assemblages (Figures 7 and 8). With respect to this foliation, the qtz + biot and assemblage
presents either a “static” granoblastic texture when the strain is low (i.e., weakly developed foliation,
Figure 7A) or pre- to syn-tectonic features when the foliation is strongly expressed (Figures 7C to E, 8C).
The texture, shapes, and aggregates of the biotite and andalousite (± suspected cordierite) are typical
of HT/LP “contact” metamorphism in the hornfels facies. The syn-tectonic assemblage is composed of
quartz + chlorite + white micas (± calcite) and partially replaced the former biotite and andalousite
blasts (Figures 7B and 8E).
Similarly, surface data from the N’Fis block (Figures 2 and 3) show the occurrence of a single
sub-vertical XY plane oriented N130. This flattening plane is a penetrative foliation secant to slumps
(Figure 3B) and axial-planar to P1 folds (Figure 3C). Contact metamorphic biotite blasts are elongated
parallel to the foliation and appear as flattened sulfide grains (Figure 3D).
Therefore, these data imply that the Hajjar MSD and the N’Fis block are affected by a single
foliation which encompasses a HT/LP contact metamorphism. Although a single Variscan foliation
was similarly recognized by Dias et al. (2011) [35] at the regional scale, our results disagree with the
previously published works on the Hajjar mine/N’Fis area: i) first, two foliations were identified and
consequently two successive tectonic events with sub-normal horizontal shortening directions were
invoked [12,23]. In particular, the N20-30 dry joints affecting the N’Fis block at the surface (Figure 2)
cannot be related to the N0-30 penetrative and ductile foliation observed in the Hajjar mine. Moreover,
there has been no direct observation of an early foliation/cleavage in the Hajjar galleries during our
study; ii) second, the biotite blasts were interpreted as post-tectonic with respect to the last deformation
event [25].
238
Minerals 2018, 8, 435
in the location near the Hajjar mine, a regional scale shear zone of this type has not been previously
recognized and constitutes a key structural feature of the Guemassa Hercynian orogenic segment.
Figure 9. Simplified and conceptual map view model of the internal strain pattern within the Hajjar
transpressive right-lateral shear zone (see text for explanation).
The shear planes of the Hajjar MSD present both ductile and brittle features (Figures 5 and 8G, H).
The last brittle increments cross-cut and offset the former S1 foliation along the gouge zones (Figure 5G).
Asymmetric biotite blasts with pressure shadows filled with the chlorite and white mica assemblage
(Figure 7E,F) argue for simple shearing after the thermal peak of the HT/LP contact metamorphism.
Lower or retrograde metamorphic conditions during simple shearing are also indicated via the
cataclasis of sphalerite and chalcopyrite along the shear planes within mylonitic zones affecting
massive sulfides (Figure 8H). Therefore, the Hajjar shear zone records simple shearing increments
during and after the development of the widespread S1 foliation.
Last, the Atlasic brittle reactivation of this Hercynian shear zone cannot be ruled out, however it
is still difficult to precisely depict this.
239
Minerals 2018, 8, 435
we present clear evidence of tectonic thickening within the fold hinge zone. The wavy termination of
the metric-scale massive sulfide lenses parallel to S1 suggests that these lenses were likely thickened
by folding before they were flattened within the XY plane of S1 (Figures 6 and 8D). This mechanism
is enhanced by the high “plasticity” of pyrrhotite, which is by far the dominant sulfide at Hajjar.
The pre-to syntectonic HT/LP metamorphism greatly favor the ductile behavior and recrystallization
of sulfides including chalcopyrite and sphalerite. This is observable at the thin section scale, where the
tectonic thickening induced the stress-oriented recrystallization of sphalerite, leading to an incipient
“banding” of sphalerite-rich/sphalerite-poor slices parallel to S1 (Figure 8C). We suggest that, in Hajjar
MSD, this solid-state thickening and remobilization are effective at a larger scale, but further modern
textural and mineralogical studies are required in order to be able to investigate this point.
Remobilization of the primary metal stock by fluids (e.g., the fluid state processes and chemical
remobilization described by Gilligan and Marshal (1987) [3] is also expressed in the Hajjar MSD. Even
though the metal mass balance quantification is outside the scope of this study, the polymetallic veins
argue for hydrothermal fluid-assisted remobilization during deformation. In particular, the polymetallic
triangular veins at the tips of the massive sulfide lenses indicate such remobilization. This type of vein
with a polymetallic assemblage associated with quartz, newly formed sphalerite and galena veinlets,
and pyrrhotite replacement by vermicular pyrite (Figure 6B,C), is similar to the so-called “piercement
veins” described by authors working on deformed MSD (e.g., [3,38–40]). It has been hypothesized
that the metamorphic fluids liberated during the prograde HT/LP contact metamorphism (e.g., quartz
veins with biotite in Figure 7A), combined with potential advective hot magmatic fluids exsolved from
deeper granitic bodies, are able to chemically rework the primary sulfides and concentrate metals into
dilatant sites as triple junction veins during the last increments of deformation [3]. Due to high reactive
chemistry, the fluid-assisted chemical reworking of primary VMSD is common in many metamorphic
contexts other than HT/LP metamorphic conditions (e.g., [41] and references therein).
240
Minerals 2018, 8, 435
Massif are structurally and temporally similar to the D2b tectono-metamorphic event described in the
Jebilet Massif to the north (see the section on geological settings above and [26]). It is noteworthy that
this thermal event is not restricted to the Hajjar mine, as it has been traced by Raman Spectroscopy
of Carbonaceous Materials geothermometry method (RSCM) throughout the whole N’Fis block [27].
In the Jebilet as in the Guemassa Massifs, it has been reported that this thermal event is the consequence
of hidden plutonic intrusions. Our data suggests the presence of fluid-assisted HT/LP “contact”
metamorphism (Figure 7A). Therefore, the vigorous advection of hot fluids exsolved from melts
and/or which come from metamorphic devolatilization may also partly explain the large extent of this
HP/LP metamorphism observed in the Guemassa and Jebilet Massifs close to 300 Ma. This regional
thermal anomaly is represented in Figure 10B. No former foliation/cleavage has been observed in
either the N’Fis block or in the Hajjar mine, suggesting that the D2a/M2a event identified by Delchini
(2018) [26] in the Jebilet Massif is not expressed in the Guemassa Massif. This is in agreement with
the fact that the D2a/M2a event, which reaches the garnet-staurolite amphibolite facies, is poorly
represented in the Jebilet Massif, and better expressed northward in the Rehamna Massif. Thus,
the S1 foliation/cleavage characterized in this study matches the S2b foliation identified in the Jebilet
Massif to the north by Delchini et al. (2016) [28]. Biotites related to the HT/LP metamorphism are not
post-kinematic, as proposed by Hibti (1993) [23]. They are pre- to syn-kinematic, which implies that
deformation occurred during the peak of the HT/LP “contact” metamorphism (Figure 9B). Based on
the biotite blasts vs. strain relationship, we suggest that during the HT peak, the deformation was
predominantly coaxial before shifting to a bulk non-coaxial regime.
Figure 10. Tectono-metamorphic model of the Hajjar shear zone and associated MSD. The name of the
tectonic events (D0, D2) corresponds to the tectonic events that have recently been established for the
Jebilet Massif by Delchini (2018) [26]. D1 has not been identified in this study. See text for explanations.
Third, during the Early Permian, the D2 event identified in the Jebilet Massif ended with
transpressive conjugate regional shear zones, oriented NE/ENE and SE/SSE with a dextral and
sinistral (e.g., the MSZ, Figure 1) sense of shear respectively (i.e., the D2c event described by Delchini,
241
Minerals 2018, 8, 435
2018) [26]. This led to the development of a regional scale “flower structuration” of the Jebilet Massif.
This strain localization along the shear zones appears to post-date the HT/LP contact metamorphism.
Our data from the Guemassa Massif are fully compatible with this scenario (Figure 10C): the Hajjar
regional shear zone we recognized in this study appears to be one of the dextral shear zones responsible
for the large virgation of the main foliation planes. As observed in the Jebilet Massif, this shear zone
corresponds to a progressive strain localization during the retrograde metamorphism when the D2
event ended. Last, as proposed by Dias et al. (2011) [35], conjugate WNW-ESE trending sinistral shear
zones activated as the Lalla Takerkoust fault (Figure 10 C). This sinistral wrench zone accentuated and
is responsible for the virgation of the S1 foliation, resulting in the “anarchic” WNW-orientation of the
foliation observed through the N’Fis block.
7. Conclusion
The Guemassa Massif and the Hajjar base-metal massive sulfide deposit have been affected
by a single foliation during a major Late Carboniferous-Early Permian Hercynian tectonic event.
This foliation is strongly affected and deflected by regional scale shear zones such as the Hajjar
N70-trending and right-lateral shear zone. Structural mapping in the Hajjar mine demonstrates that
the Hajjar shear zone is complex with anastomosing shear plane patterns combined with thrusting
and folding. This deformation is partially coeval, with a large thermal anomaly responsible for the
HT/LP metamorphism. The tectono-metamorphic evolution of the Oriental Guemassa Hercynian
segment is highly compatible with the evolution depicted for the Jebilet Massif. Strain under a high
heat flux favored the deformation of the massive sulfides bodies which partly underwent fluid-assisted
remobilization in the Hajjar mine. The tectonic thickening of the mineralization is observed at the
meter scale, and must be re-examined at a larger scale.
Author Contributions: S.A. and Y.B. conceptualized both the study and the final model and wrote the original
draft. L.B (Lakhlifi Badra) and L.B. (Luc Barbanson) reviewed and edited the draft. M.O., A.K., M.Z., L.M. gave
their validation, Funding acquisition and project administration.
Funding: The PhD thesis of S. Admou has been partly funded by the “Office Mediterannéen de la Jeunesse”
through a partnership between Orleans University (France) and Moulay Ismael University (Meknès, Morocco).
Acknowledgments: We are grateful to S. Janiec from ISTO and X. Le Coz from Geosciences Rennes who performed
high quality thin sections. Our discussion with S. Delchini was greatly appreciated. We thank the reviewers
and specially R. Dias for very fruitful and constructive review. The Guest Editor A. Chauvet is also thanked for
inviting us to submit our work.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Cawood, P.A.; Hawkesworth, C.J. Temporal relations between mineral deposits and global tectonic cycles.
In Ore Deposits in an Evolving Earth; Jenkin, G.R.T., Lusty, P.A.J., Mcdonald, I., Smith, M.P., Boyce, A.J.,
Wilkinson, J.J., Eds.; Geological Society of London: London, UK, 2013; pp. 9–21.
2. Graf, J.; Skinner, B. Strength and deformation of pyrite and pyrrhotite. Econ. Geol. 1970, 65, 206–215.
[CrossRef]
3. Marshall, B.; Gilligan, L.B. An introduction to remobilisation: information from ore-body geometry and
experimental considerations. Ore Geol. Rev. 1987, 2, 87–131.
4. Marshall, B.; Spry, P.G. Discriminating between regional metamorphic remobilization and syntectonic
emplacement in the genesis of massive sulfide ores. Rev. Econ. Geol. 1998, 11, 39–80.
5. Marignac, C.; Diagana, B.; Cathelineau, M.; Boiron, M.-C.; Banks, D.; Fourcade, S.; Vallance, J. Remobilisation
of base metals and gold by Variscan metamorphic fluids in the south Iberian pyrite belt: evidence from the
Tharsis VMS deposit. Chem. Geol. 2003, 194, 143–165. [CrossRef]
6. Chauvet, A.; Onézime, J.; Charvet, J.; Barbanson, L.; Faure, M. Syn- to late-tectonic stockwork emplacement
within the spanish section of the iberian pyrite belt: Structural, textural, and mineralogical constraints in the
tharsis and la zarza areas. Econ. Geol. 2004, 99, 1781–1792. [CrossRef]
242
Minerals 2018, 8, 435
7. Barrie, C.D.; Boyle, A.P.; Prior, D.J. An analysis of the microstructures developed in experimentally deformed
polycrystalline pyrite and minor sulphide phases using electron backscatter diffraction. J. Struct. Geol. 2007,
29, 1494–1511. [CrossRef]
8. Barrie, C.D.; Boyle, A.P.; Cook, N.J.; Prior, D.J. Pyrite deformation textures in the massive sulfide ore deposits
of the Norwegian Caledonides. Tectonophysics 2010, 483, 269–286. [CrossRef]
9. Barrie, C.D.; Peare, M.A.; Boyle, A.P. Reconstructing the pyrite deformation mechanism map. Ore Geol. Rev.
2011, 39, 265–276. [CrossRef]
10. Reddy, S.M.; Hough, R.M. Microstructural evolution and trace element mobility in Witwatersrand pyrite.
Contrib. Mineral. Petrol. 2013, 166, 1269–1284. [CrossRef]
11. Bernard, A.J.; Maier, O.W. Aperçus sur les amas sulfurés Massifs des hercynides Marocaines. Miner. Depos.
1988, 23, 104–114. [CrossRef]
12. Hibti, M. Les amas Sulfurés des Guemassa et des Jebilet (Meseta Sud-Occidentale, Maroc): Temoins de
L’hydrothermalisme Précoce dans le Bassin Mesetien. Ph.D Thesis, University Cadi Ayyad, Marrakech,
Morocco, 2001.
13. Belkabir, A.; Gibson, H.L.; Marcoux, E.; Lentz, D.; Rziki, S. Geology and wall rock alteration at the Hercynian
Draa Sfar Zn–Pb–Cu massive sulphide deposit, Morocco. Ore Geol. Rev. 2008, 33, 280–306. [CrossRef]
14. Marcoux, E.; Belkabir, A.; Gibson, H.L.; Lentz, D.; Ruffet, G. Draa Sfar, Morocco: A Visean (331 Ma)
pyrrhotite-rich, polymetallic volcanogenic massive sulphide deposit in a Hercynian sedimentdominant
terrane. Ore Geol. Rev. 2008, 33, 307–328. [CrossRef]
15. Moreno, C.; Sáez, R.; González, F.; Almodóvar, G.; Toscano, M.; Playford, G.; Alansari, A.; Rziki, S.; Bajddi, A.
Age and depositional environment of the Draa Sfar massive sulfide deposit, Morocco. Miner. Depos. 2008, 43,
891–911. [CrossRef]
16. Ben aissi, l. Contribution à L’étude Gîtologique des Amas Sulfurés Polymétalliques de Draa Sfar et de
Koudiat Aïcha: Comparaison avec les Gisements de Ben Slimane et de Kettara (Jebilet Centrales, Maroc
Hercynien). Ph.D Thesis, University Cadi Ayyad, Marrakech, Morocco, 2008.
17. Lotfi, F.; Belkabir, A.; Brown, A.C.; Marcoux, E.; Brunet, S.; Maacha, L. Geology and Mineralogy of the
Hercynian Koudiat Aïcha Polymetallic (Zn-Pb-Cu) Massive Sulfide Deposit, Central Jebilet, Morocco.
Explor. Min. Geol. 2008, 17, 145–162. [CrossRef]
18. Essaifi, A. Relations entre Magmatisme-Déformation et al.tération Hydrothermale: L’exemple des Jebilet
Centrales (Hercynien, Maroc). Ph.D Thesis, Unversity of RennesI, Rennes, France, 1995.
19. Essaifi, A.; Hibti, M. The hydrothermal system of Central Jebilet (Variscan Belt, Morocco): A genetic
association between bimodal plutonism and massive sulphide deposits? J. Afr. Earth Sci. 2008, 50, 188–203.
[CrossRef]
20. Essaifi, A.; Goodenough, K.M.; Lusty, P.A.J.; Outigua, A. Microstructural and Textural Evidence for
Protracted Polymetallic Sulphide Mineralization in the Jebilet Massif (Variscan Belt of Morocco). Min.
Resour. Sustain. World 2015, 1–5, 1603–1606.
21. Lusty, P.A.J.; Goodenough, K.M.; Essaifi, A.; Maacha, L. Developing the lithotectonic framework and model
for sulfide mineralization in the Jebilet Massif, Morocco: implications for regional exploration. In Mineral
Resources in a Sustainable World, Proceedings of the 13th Biennial SGA Meeting, Nancy, France, 24–27 August 2015;
André-Mayer, A.S., Cathelineau, M., Muchez, P.h., Pirard, E., Sindern, S., Eds.; Society for Geology Applied
to Mineral Deposits (SGA): Genéve, Switzerland, 2015; pp. 1635–1638.
22. N’Diaye, I.; Essaifi, A.; Dubois, M.; Lacroix, B.; Goodenough, K.M.; Maacha, L. Fluid flow and polymetallic
sulfide mineralization in the Kettara shear zone (Jebilet Massif, Variscan Belt, Morocco). J. Afr. Earth Sci.
2016, 119, 17–37. [CrossRef]
23. Hibti, M. L’amas Sulfuré de Hajjar, Contexte Géologique de mie en Place et Déformations Superposées
(Haouz de Marrakech, Méseta Sudoccidentale, Maroc). Ph.D Thesis, University Cadi Ayyad, Marrakech,
Morocco, 1993.
24. Zouhry, S. Étude Métallogénique D’un amas Sulfuré Viséen à Zn Pb Cu: cas de Hajar, Guemassa, Maroc.
Ph.D Thesis, Ecole polytechnique de Montréal, Montréal, Canada, 1999.
25. Hibti, M.; Marignac, C. The Hajjar deposit of Guemassa (SW Meseta, Morocco): A metamorphosed
syn-sedimentary massive sulfide ore body of the Iberian type of volcano-sedimentary massive sulfide
deposits. In Mineral Deposits at the Beginning of the 21st Century, Proceedings of the Joint Sixth Biennial SGA-SEG
Meeting, Krakow, Poland, 26–29 August 2001; A.A. Balkema: Lisse, The Netherlands, 2001; pp. 281–284.
243
Minerals 2018, 8, 435
26. Delchini, S. Etude Tectono-Thermique D’un Segment Orogénique Varisque à Histoire Géologique Complexe:
Analyse Structurale, Géochronologique et Thermique du Massif des Jebilet, de L’extension à la Compression.
Ph.D Thesis, University of Orléans, Orléans, France, 2018.
27. Delchini, S.; Lahfid, A.; Ramboz, C.; Branquet, Y.; Maacha, L. New Peak Temperature Constraints using
RSCM Geothermometry on the Hajjar Zn-Pb-Cu Mine and its Surroundings (Guemassa Massif, Morocco).
In Proceedings of the 13th SGA Biennial Meeting, Nancy, France, 24–27 August 2015.
28. Delchini, S.; Lahfid, A.; Plunder, A.; Michard, A. Applicability of the RSCM geothermometry approach in
a complex tectono-metamorphic context: The Jebilet Massif case study (Variscan Belt, Morocco). Lithos 2016,
256, 1–12. [CrossRef]
29. Haimmeur, J. Contribution à L’étude de L’environnement Volcano-Sédimentaire et du Minerai de Douar
Lahjar (Guemassa, Maroc), Lithologie, Paléo-Volcanisme, Géochimie et Métallogénie. Ph.D Thesis, École
Nationale Supérieure de Géologie, Nancy, France, 1988.
30. Raqiq, H. Le bassin Carbonifère des Guemassa (Meseta Sud occidentale, Maroc): Lithostratigraphie,
sédimentologie et évolution structurale. Ph.D Thesis, University Cadi Ayyad, Marrakech, Morocco, 1997.
31. Ouadjou, A. Pétrographie, Géochimie et Structure des Roches Magmatiques Antéschisteuses des Massifs
Hercyniens des Guemassa et Souktana. Ph.D Thesis, University Cadi Ayyad, Marrakech, Morocco, 1997.
32. Ed Debi, A.; Saquaque, A.; Kersit, M.; Chbiti, A. L’amas sulfuré de Hajar (Guemassa, Maroc). Chronique de la
Recherche Minière 1998, 531–532, 45–54.
33. Watanabe, Y. 40Ar/39Ar geochronologic constraints on the timing of massive sulfide and vein-Type Pb-Zn
mineralization in the Western Meseta of Morocco. Econ. Geol. 2002, 97, 147–157. [CrossRef]
34. Soulaimani, A. L’évolution structurale des Massifs hercyniens du Haouz de Marrakech: Guemassa- N’fis
(Maroc). Ph.D Thesis, University Cadi Ayyad, Marrakech, Morocco, 1991.
35. Dias, R.; Hadani, M.; Leal Machado, I.; Adnane, N.; Hendaq, Y.; Madih, K.; Matos, C. Variscan structural
evolution of the western High Atlas and the Haouz plain (Morocco). J. Afr. Earth Sci. 2011, 61, 331–342.
[CrossRef]
36. Hoepffner, C.; Soulaimani, A.; Piqué, A. The Moroccan Hercynides. J. Afr. Earth Sci. 2005, 43, 144–165.
[CrossRef]
37. Saadi, M.; Hilali, E.A.; Bensaîd, M.; Boudda, A.; Dahmani, M. Carte géologique du Maroc, échelle 1:1 000 000.
Notes Mém. Serv. Géol. Maroc. 1985. Available online: https://geodata.mit.edu/catalog/mit-gfcc2renabn5c
(accessed on 6 October 2018).
38. Pedersen, F.D. Remobilization of the massive sulfide ore of the Black Angel Mine, central West Greenland.
Econ. Geol. 1980, 75, 1022–1041. [CrossRef]
39. Maiden, K.J.; Chimimba, L.R.; Smalley, T.J. Cuspate ore-wall rock interfaces, piercement structures and the
localization of some sulfide ores in deformed sulfide deposits. Econ. Geol. 1986, 81, 1464–1472. [CrossRef]
40. Plimer, I.R. Remobilization in high-grade metamorphic environments. Ore Geol. Rev. 1987, 2, 231–245.
[CrossRef]
41. Gu, L.; Zheng, Y.; Tang, X.; Zaw, K.; Della-Pasque, F.; Wu, C.; Tian, Z.; Lu, J.; Li, X.; Yang, F.; et al. Copper,
gold and silver enrichment in ore mylonites within massive sulphide orebodies at Hongtoushan VHMS
deposit, NE China. Ore Geol. Rev. 2007, 30, 1–29. [CrossRef]
42. Aarab, E.M.; Beauchamp, J. Le magmatisme carbonifère pré-orogénique des Jebilet centrales (Maroc).
Précisions pétrographiques et sédimentaires. Implications géodynamiques. CR Acad. Sci. Paris 1987, 304,
169–174.
© 2018 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
244
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com