Sie sind auf Seite 1von 31

Ore Geology Reviews 102 (2018) 474–504

Contents lists available at ScienceDirect

Ore Geology Reviews


journal homepage: www.elsevier.com/locate/oregeorev

Phanerozoic uranium mineralization in Variscan Europe – More than T


400 Ma of tectonic, supergene, and climate-controlled uranium
redistribution
Rolf L. Romera, , Michel Cuneyb

a
GFZ German Research Centre for Geosciences, Telegrafenberg, D-14473 Potsdam, Germany
b
Université de Lorraine, CNRS, CREGU, GeoRessources, F-54000 Nancy, France

ARTICLE INFO ABSTRACT

Keywords: Silurian organic-rich black shales are widespread and serve as basic source for most U mineralization throughout
Silurian black shales the entire Variscan Orogen of Central and Western Europe. In particular the lowermost unit of Silurian black
Hirnantian glaciation shales is regionally strongly enriched in U. The strong U-enrichment of these shales is the result of the Hirnantian
Variscan orogeny glaciation that removed old near-surface rocks that already had lost most leachable U during long-lasting al-
Permian European Extensional Province
teration and brought rocks with leachable U to the surface. Sea level rise due to the melting of the Hirnantian ice
Weathering
U-fertile granites
sheet inundated the glacially modeled continent margin and resulted in regionally restricted oceanic circulation
and anoxic conditions that are related to both low initial oxygen levels and sudden increase of organic pro-
duction. Organic-rich shales on the Gondwana shelf developed particularly high U content as they scavenged U
from the surface runoff that was leached after deglaciation from the newly exposed drainage area.
The present spatial distribution of Silurian U-rich black shales within the Varisan orogenic belt is largely
controlled by large-scale tectonic processes operating during the approach and eventual collision of Gondwana
and Laurussia, leading to the formation of western Pangea. During the Variscan orogeny, these U-anomalous
rocks became stacked and folded at the margins of some low-strain domains of the orogenic belt and exposed to
metamorphism and in part to crustal melting – contributing to the formation of U-fertile post-kinematic granites
– in the high-strain domains of the orogenic belt. There are no U deposits that are directly related to the em-
placement of these granites. Variscan granites, however, are the most important source of leachable U for
younger, major U mineralization. Late-Variscan reorganization of the plate movement between Gondwana and
Laurussia resulted in an extensional tectonic regime throughout Europe with important Permian sedimentary
basins. Heat input and fluid migration along Permian fault zones resulted in the formation of episyenite-bound
and vein-type U mineralization in the Variscan basement, whereby U was leached from granites and Silurian
sedimentary rocks and their metamorphic equivalents and was deposited in local episyenite and vein-type mi-
neralization (e.g., districts of Aue-Niederschlema, Jáchymov, Příbram, Forez, and Salamanca).
Late-Variscan erosion of the orogenic belt resulted in the redistribution of U from Variscan metamorphic rocks
and granites, but also the Silurian black shales, into late Carboniferous and Permian sedimentary basins. Permian
mobilization of fault zones controled the availability of U-rich source rocks for leaching, the transport of U into
the basins, and the distribution of organic-rich sediments that may trap U in organic-rich layers (e.g., Döhlen,
Lodève). The Permian U redistribution represents the major deposit-forming event throughout Variscan Europe,
whereby hydrothermal U mineralization in episyenites and veins in the Variscan basement and low-temperature
U mineralization in sandstones and organic-rich sediments are different expressions of the same tectonic process,
i.e., the development of the Central European Extensional Province. Later tectonic reactivations of these fault
zones during the opening of the Tethys and North Atlantic oceans and during the Alpine Orogeny mainly resulted
in the local redistribution of U mineralization along the same fault zones at c. 180 Ma, 150 Ma, and 120 Ma, and
more recently during the Alpine orogeny, in part with addition of other metals (e.g., Ag, Bi, Co, Ni, As) leached
from the local wall rocks.
Post-Permian intense chemical weathering, mostly in the Triassic and the Cretaceous, resulted in (i) U re-
distribution and additional U concentration within near-surface mineralization (e.g., Ronneburg, Aue-
Niederschlema), (ii) leaching from near-surface U-rich rocks followed by entrapment in sediments to form


Corresponding author.
E-mail address: romer@gfz-potsdam.de (R.L. Romer).

https://doi.org/10.1016/j.oregeorev.2018.09.013
Received 3 April 2018; Received in revised form 12 September 2018; Accepted 17 September 2018
Available online 18 September 2018
0169-1368/ © 2018 Elsevier B.V. All rights reserved.
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

lignite-type (e.g., black shales, peats) and roll-type mineralization (e.g., Königstein), and (iii) the dispersion of U
in the supergene environment.
The rocks of the Variscan belt host a history of more than 400 Ma of U redistribution, starting with the
deposition of early Silurian U-rich black shales. The formation of these black shales is climatically controlled.
Their spatial distribution eventually controls the spatial distribution of younger mineralization that is generally
closely related with the tectonic development and reactivation of the basement and the redistribution by
magmatic, hydrothermal, and erosional processes. Basement faults are pathways for fluids and their reactivation
controls the formation and inversion of basins and the uplift of basement blocks, eventually linking the en-
dogenic U cycle (magmatic and hydrothermal rocks) with the exogenic U cycle (making U-rich rocks available
for erosion and leaching and creating traps to precipitate mobilized U). The timing of post-Variscan U miner-
alization is controlled by these tectonic reactivations, whereas supergene enrichment or redistribution is par-
ticularly efficient in periods of intense chemical weathering.

1. Introduction During later tectonic reactivation, both the U-rich metamorphosed Si-
lurian sedimentary rocks and the U-fertile Variscan granites served as
The area of the Variscan Orogen in Central and Western Europe source rocks for low-temperature hydrothermal-type, sandstone-type,
hosts a major U province, which in its (historic) importance only is and coal/lignite-type mineralization (e.g., Tonndorf, 2000; Cuney and
surpassed by the unconformity related deposits of Canada (Athabasca Kyser, 2015; Ballouard et al., 2018). We demonstrate that the dis-
and Thelon basins) and Australia (Ranger, Jabiluka, and Nabarlek), the tribution of U-rich Silurian shales of former peri-Gondwana essentially
quartz-pebble conglomate deposits of South Africa and Canada, the controls the spatial distribution of all U mineralization in Variscan
sandstone-hosted deposits of the Central Asian Mesozoic to Tertiary Europe, whereas type and timing of mineralization are related to large-
basins of Kazhakstan and China and of the Colorado Plateau, the me- scale tectonic processes and climate. We present a generic model
tasomatic, metamorphic, and magmatic deposits of the Ukraine, Brazil, linking the distribution and type of U mineralization to the tectonic
and Namibia, and some IOCG and breccia deposits, in particular evolution in the large-scale context of the assembly of supercontinent
Olympic Dam (Dahlkamp, 2016; Cuney and Kyser, 2015; Bruneton and Pangea and its subsequent fragmentation during the opening of the
Cuney, 2016). Whereas U mineralization of most major U provinces Tethys and Atlantic oceans.
occurs in Archean and Paleoproterozoic domains, mineralization in
Central and Western Europe occurs in Phanerozoic rocks and formed
predominatly after the Variscan orogeny (e.g., Marignac and Cuney, 2. Different types of U mineralization
1999; Hiller and Schuppan, 2008; Cuney and Kyser, 2009, 2015). Ur-
anium mineralization in Central and Western Europe encompasses a Uranium is a redox-sensitive element that is highly mobile under
wide range of different mineralization types, e.g., hydrothermal-type oxidized conditions and forms mineralization through the entire geo-
related to granites (veins and episyenites), sandstone-type, hydro- logical cycle in magmatic, metamorphic, and sedimentary rocks (e.g.,
thermally remobilized black shales, and coal/lignite-type, that formed Dahlkamp, 1993, 2016; Cuney and Kyser, 2015). Most U mineralization
at different times and show evidence for repeated redistribution of both is the result of multiple processes and may involve enriched source
U and other metals (e.g., Bi, Co, Ni, Ag, As; Oelsner, 1958; Förster and rocks and multiple stages of U redistribution with U enrichment and
Haack, 1995; Herrmann et al., 1995; Kuschka, 2002; Hiller and dispersion within a given deposit (e.g., Robertson et al., 1978; Nash,
Schuppan, 2008). Spatially, U mineralization is mostly controlled by 1981, 2010; Nash et al., 1981; Dahlkamp, 1993; Cuney, 2009, 2010,
the distribution of Variscan basement massifs. Uranium mineralization 2014). The redox-sensitive character of U, however, does not only ac-
and redistribution occurred over a period of more than 400 Ma. The count for the wide range of different types of U deposits in both en-
oldest U mineralization in Variscan Europe is hosted in early Silurian dogenic and exogenic environments, but also for changes in type of U
shales (Unterer Graptolitnenschiefer, e.g., Ronneburg, Lange and mineralization through time that are not related to preservation po-
Freyhoff, 1991), whereas the youngest U mineralization presently forms tential alone (Table 1). The most important parameter affecting the
in soils, marshes and/or calcretes in the drainage system of leached U- distribution of certain U mineralization is the availability of oxygen,
bearing granites (cf., Wilson, 1984; Regenspurg et al., 2010). most importantly the level of oxygen in the atmosphere, and the
In this paper we present the temporal and spatial distribution of U abundance of organic matter in sedimentary environments and rocks.
mineralization in Central and Western Europe. We demonstrate that For instance, detrital uraninite, typically along with sulfide minerals,
early Silurian black shales, similar to those of Ronneburg (e.g., Lange only occurs in Archean and early Paleoproterozoic conglomerates that
and Freyhoff, 1991; Bolonin and Gradovsky, 2012), represent one of the were deposited before the increase in atmospheric oxygen (e.g., Cuney,
main source rocks of most U mineralization in Variscan Europe. In the 2010). After this increase in atmospheric oxygen, redox-related mobi-
low-strain domains of the Variscan Belt, U mineralization formed when lization and aqueous transport of U become the dominant processes for
the U content in the Silurian shales was locally concentrated to form the formation of near-surface U mineralization (e.g., Cuney, 2010).
mineralization during later tectonic and weathering events. In the high- Similarly, the presence of organic matter and authigenic sulfides in
strain domains of the Variscan Belt, the U-anomalous Silurian rocks marine sediments and – after the Silurian – of vascular plants in ter-
were metamorphosed or melted during the Variscan orogeny, giving restrial sediments provides efficient traps for U, giving rise to black
rise to graphite-bearing schists and U-fertile1 granites, respectively. shale related mineralization in marine sediments and sandstone-type
mineralization in terrestrial sediments (e.g., Lovley et al., 1991;
Dahlkamp, 1993, 2016; Spirakis, 1996; Jaireth et al., 2008; Cuney and
1
We use the term U-fertile rather than U-specific, U-rich, or U-anomalous to
highlight that U content is not the only factor that controls whether a granite (footnote continued)
may serve as U source or not. If U is hosted in silicates, niobates, and other whose U content therefore becomes accessible for redistribution. It should be
chemically stable minerals, it is generally not accessible to mobilization by noted that U- and-Th rich silicate minerals such as zircon, allanite, and U-rich
oxidized fluids. In the way used here, U-rich refers to granites that have high U- thorite cummulate radiation damage and become with time increasingly me-
contents, whereas U-fertile refers to granites that host a significant protion of tamict. Once the crystal lattice of these minerals is strongly damaged, they also
their U budget in uraninite or minerals that readily become metamict, i.e., are sources of leachable U.

475
Table 1
Types of U mineralization.
Typea Uranium source and Uranium trap Basic requirements Age rangeb Examples Referencesc
mobilization
R.L. Romer, M. Cuney

Endogenic U mineralization
Magmatic deposits Enrichment of U by magmatic No trap; U-enrichment during Late-stage magmatic fluids – Kvanefjeld, Poços de Caldas, 1, 2, 3, 4
differentiation of peralkaline fractionation of peralkaline magmas; that redistribute U in Lovozero, Bokan Mountain
magmas in particular in highly evolved late- episyenites, albitites, and
stage magmas; U mostly stored in pegmatites
silicates and REE and/or Nb, Ta
minerals
Migmatite-related deposits Leucocratic melt derived from Reducing conditions during U anomalous protoliths that – Rössing, Wollaston and Mudjatik 5, 6, 7
anatexis of U-rich crystallization of melts prevent partition U preferentially into domains, Bancroft, Mont Laurier
metasedimentary rocks significant loss of U to fluid phases that partial melts
leave the system
High-grade metamorphism Leaching of granitic and Na- (less commonly K-) metasomatism High-temperature Paleo- and Meso- Central Ukraine U Province, 8, 9, 10, 11
and Na-metasomatism metamorphic rocks by high- resulting in regional scale quartz hydrothermal alteration by Proterozoic Lagoa Real, Michelin, Valhalla,
temperature alkaline fluids dissolution and albitization, generally highly alkaline fluid Lianshanguan
followed by Ca-metasomatism is
induced by high-temperature fluids
that also introduce U into the altered
rocks
Vein-type and episyenite Magmatic and metamorphic Mineralization is not related to granite Precipitation by fluid-mixing – Marnac, Fanay, Les Brugeauds, 12, 13, 14, 15,
low-temperature rocks with leachable U as well emplacement, even if there is a close or fluid-rock interaction Bois Noirs-Limouzat, Pontivy- 16, 17, 18, 19,
hydrothermal as black shales (and other U- spatial relation. Instead, the granites leading to reduction, ligand Rostrenen complex, Schlema- 20, 21, 22, 23
mineralization bearing sedimentary rocks) act as U source and control the loss, or change in fluid Alberoda, Schneeberg,
position of medium to low- composition Tellerhäuser, and Jáchymov
temperature hydrothermal alteration districts, Příbram, Mina Fé

476
and faults. Episyenites (rock affected
by medium-temperature quartz
dissolution) develop locally high
porosity that focuses later fluid flow.
Uranium precipitation in episyenites
and veins from medium to low
temperature fluids
Unconformity related Basinal brines leaching and Spatially related to reduction fronts U precipitation within the Proterozoic Cigar Lake, McArthur River, 45, 46
deposits transporting U near the contact between basement or the sandstones by Millenium (all Athabasca Basin),
Paleoproterozoic basement and interaction of the fluid with Kiggavik, End, Andrew (all
overlying sandstone reducing sections of the host- Thelon Basin), Jabiluka, Ranger,
rocks or by fluid mixing Koongara (all Kombolgie Basin)

Shallow and exogenic U mineralization


Quartz-pebble Primary U mineralization with Reduction of gradient in fluvial Reducing conditions in Older than c. 2.4 Ga Elliot Lake, Witwatersrand 24, 25, 26, 27
conglomerate uraninite at erosion level transport system with deposition of atmosphere (before the increase
coarse sediments and heavy minerals, of oxygen content
including uraninite, pyrite, and gold in the atmosphere)
Sandstone-hosted Leaching of U from near- Organic material (remains of land U migration before or during Commonly Silurian Powder River Basin, Colorado 25, 28, 29, 30,
surface rocks by circulating plants) in predominantly fluvial and diagenesis and younger Plateau, Karoo Basin, Mesozoic 31, 32, 33
groundwater lacustrine sandstones between basins in Kazakhstan, Mongolia,
relatively impermeable layers of shale China, and Transbaikal
Calcrete-hosted Leaching of U at or near the Sorption on silicates, clay minerals, Semi-arid to arid climate; Tertiary to recent Dry areas in Western Australia 34, 35, 36
surface by surface runoff and oxyhydroxides, incorporation in flood plain, delta, or playa (Yeelirrie, Lake Way), Namibia,
circulating groundwater carbonates, supersaturation of deposits South Africa, Somalia
transport agent by evaporation or
ligand loss (e.g., CO2 loss)
Lignite/peat/bog deposits Leaching of U at or near the Organic material in surface sediments Oxidizing fluids meet reducing Tertiary to recent Flodelle Creek 37, 38, 39
surface by surface runoff and and peat surface sediments
circulating groundwater
Ore Geology Reviews 102 (2018) 474–504

(continued on next page)


R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

1 = Sørensen et al., 2011; 2 = Cuney, 2014; 3 = Dostal et al., 2014; 4 = Waber et al., 1992; 5 = Corvino and Pretorius, 2013; 6 = Basson and Greenway, 2004; 7 = McKeough et al., 2013; 8 = Cuney et al., 2012;
9 = Polito et al., 2009; 10 = Alexandre, 2010; 11 = Veríssimo et al., 2016; 12 = Hiller and Schuppan, 2008; 13 = Hiller and Schuppan, 2016; 14 = Ondruš et al., 2003a; 15 = Ballouard et al., in press; 16 = Leroy and

2005; 33 = Hall et al., 2017; 34 = Jaireth et al., 2016; 35 = Hou et al., 2017; 36 = Bowell and Davies, 2017; 37 = Zielinski and Burruss, 1991; 38 = IAEA, 1984; 39 = Regenspurg et al., 2010; 40 = Lecomte et al., 2014;
Mineralization styles largely based on grouping and compilation of Cuney and Kyser (2009) and Cuney (2009). Note, there are transitions between different styles and many deposits are the result of the superposition of

Typical formation ages or only indicated for deposit types that are restricted to particular periods. For U mineralization in the exogenic environment, the most critical parameters controlling the occurrence of
several processes. Above grouping does not represent a classification. Instead, it allows for combining a wide range of deposits that are classified differently, but that are different expressions of the same large-scale

Holliger, 1984; 17 = Naumov et al., 2017; 18 = Cuney, 1978; 19 = Cathelineau et al., 1990; 20 = Kříbek et al., 1999; 21 = Kříbek et al., 2009; 22 = Škácha et al., 2009; 23 = Martin-Izard et al., 2002; 24 = Anderson
et al., 1987; 25 = Dahlkamp, 1993; 26 = Roscoe and Minter, 1993; 27 = Pretorius, 2012; 28 = Turner-Peterson and Fishman, 1986; 29 = Boberg, 2010; 30 = Cai et al., 2007; 31 = Bonnetti et al., 2015; 32 = Min et al.,
Kyser, 2015; Cuney, 2010). A short summary of the most important
types of U mineralization is given in Table 1.2 Several of these deposit
40, 41, 42, 43,
types are restricted to Archean and early Proterozoic rocks. These types
Referencesc

will not be discussed in detail, as U mineralization in the Variscides and


their later reworking are Phanerozoic exclusively.
44

During the evolution of the Earth, U is concentrated in the felsic


crust through progressive extraction from the mantle. Although there is
some reinjection into the mantle by subduction of altered oceanic crust
Talvivaara, Ranstad, Ronneburg,

and sedimentary material (e.g., Fyfe and Brown, 1979; Albarède and
Michard, 1986), eventually accounting for the high-238U/204Pb (HIMU)
mantle component (Zindler and Hart, 1986), most U is recycled within
the continental crust or is returned by subduction zone magmatism
from the lithospheric mantle to the continental crust (e.g., Cuney,
Chattanooga

2014). The crustal U cycle describing the recycling and redistribution of

41 = Leventhal, 1991; 42 = Bolonin and Gradovsky, 2012; 43 = Lippmaa et al., 2011; 44 = Huertas et al., 2013; 45 = Cuney & Kyser, 2009; 46 = Cuney and Kyser, 2015.
Examples

U in the continental crust includes two branches (Fig. 1), an endogenic


one, involving metamorphic, magmatic, and hydrothermal processes,
and an exogenic one, involving weathering and scavenging processes at
and near the interfaces of the geosphere to the hydrosphere and at-
mosphere. Endogenic processes redistribute U within the crust, even-
2.4 Ga (increase of
oxygen content in
Younger than c.

tually leading to overall depletion of the lower crust and transport of U


atmosphere)

from the lower crust to the upper crust, where erosion makes U avail-
Age rangeb

able for redistribution in the exogenic cycle. The exogenic cycle as used
here describes –after the increase of atmospheric oxygen – the mobili-
zation of U by oxidation, aqueous transport on the surface and within
sedimentary basins, and the fixation of U by reduction on mainly sul-
fides and organic matter, as well as phosphorites, and adsorbtion on
source, reducing conditions at

clay minerals and Ti-Fe-oxyhydroxides. Orogenic processes connect the


exogenic and endogenic branch of the U cycle, by burying sedimentary
Oxidizing conditions at

and magmatic rocks and eventually making U contained in these rocks


particular deposit types include the oxygen level of the atmosphere, the rise of land plants, and the preservation potential.
Basic requirements

available for redistribution by metamorphic, magmatic, and hydro-


thermal processes (e.g., Cuney, 2009, 2010; Cuney and Barbey, 2014;
Dahlkamp, 2016).
Endogenic redistribution of U within the continental crust results on
sink

the large scale, as demonstrated by the retarded 206Pb and 207Pb growth
in Archean granulites (e.g., Taylor et al., 1980), in the transfer of U
from the lower continental crust into the middle and upper continental
shallow marine sediments receive U

crust via metamorphic fluids and lower crustal melts. Partial melting of
Organic-rich fluvial, lacustrine, and

via surface runoff and groundwater

lower crust, driven by the emplacement of mantle-derived melts or


destabilization of hydrous minerals by the influx of CO2 (e.g., Cuney
and Barbey, 2014), results in the partitioning of U into the melt. De-
pending on protolith mineralogy, water availability, and melt volume,
partial melting in the crust may result in (i) U- and Th-rich leucocratic
melts that do not travel far in their high-grade metamorphic hosts rocks
Uranium trap

(e.g., anatectic pegmatoids as in Central Nambia at Rössing, as well as


Husab and many others in the Litsk district (Kola Peninsula), Wollaston
and Mudjatik domains (Saskatchewan), and the Bancroft (Ontario)

2
Because most U mineralization represents the superposition of a wide range
Alteration of near-surface

groundwater and surface

of different processes at different temperatures and in contrasting settings, most


conditions followed by
rocks under oxidizing
Uranium source and

classifications have limitations. There are two major types of deposit classifi-
transport of U via

cation, i.e., the IAEA (2018) classification is largely descriptive and bases on the
host-rocks of U mineralization independent whether these host-rocks are cri-
mobilization

tical for the location or formation of the deposits. This classification, first
published by the IAEA (OECD-NEA-IAEA, 1992), is largely based on the work of
runoff

Dahlkamp “Uranium Ore Deposits” (1993), including 16 principal types of U


deposits and occurrences distinguished according to host environment and/or
geometry and more than 40 subtypes and classes. In contrast, the classification
of Cuney (2012) is a genetic classification that focuses on the processes relevant
for deposit formation. Both classifications include a large number of subtypes
that broadly reflect that (i) a variety of processes may result in U mineralization
Black shale deposits
Table 1 (continued)

in a particular host-rock and (ii) the same process of U mineralization may


tectonic process.

operate in a wide range of different host rocks. For the purpose of this paper, it
is advantageous to combine different types of mineralization that are controlled
by similar processes and to highlight their relation to endogenic and exogenic
Typea

processes. If particular deposits are discussed, the classification and nomen-


b

clature of Cuney (2012) is used.


a

477
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Fig. 1. General geologic cycle for the formation of U


mineralization in the endogenic and exogenic en-
vironment (simplified from Cuney, 2014). There is
U mineralization related to mantle-derived mag-
matic rocks (not shown), in particular to peralkaline
rocks and alaskites. In most cases, the role of mag-
matic rocks, however, is to transport U from the
lower crust to the middle and upper crust and, thus,
to make U available for leaching and further redis-
tribution. In such a scenario, U-fertile granites re-
present the primary source lithology for U miner-
alization that occurs on structures defined by the
granites and obtains its metal content from leaching
of granites. Such structurally controlled granite-
bound U mineralization, however, may be distinctly
younger than the emplacement of the granites.
Furthermore, weathering and alteration of U-fertile
granites may mobilize U for exogenic U miner-
alization. Note, our distinction between endogenic
and exogenic U mineralization is not strictly bound
to a particular depth or fluid source, but broadly
reflects the relation of U mineralization to “hard
rocks” and “pre-diagenetic/diagenetic sediments
and sedimentary rocks” (Table 1).

domain) or (ii) U-rich melts that are emplaced in the upper crust (e.g., and paleovalley types (Renac et al., 2002; Cuney and Kyser, 2009,
Cuney and Kyser, 2015; Cuney, 2014; Cuney and Barbey, 2014). 2015; Chi et al., 2017).
Peralkaline granitic melts may have significant concentrations of U Uranium precipitates by a wide range of processes that operate in
(Peiffert et al., 1994, 1996), but commonly do not form economic U different environments. For instance, surficial calcrete-hosted U mi-
mineralization as most U is hosted in zircon, U-Th-Zr-silicophosphates, neralization may form in arid environments, involving leaching of
or niobates rather than in uraninite (Cuney and Kyser, 2009, 2015; source rocks followed by transport into a closed basin. Uranium may
Cuney, 2014). Peraluminous granites and volcanic rocks, in contrast, do precipitate, commonly along with V, in channel, flood plain, delta, or
not show extreme U contents (typically up to 20–35 ppm U, slightly playa facies by evaporation of the transporting fluid or by mixing with
higher in F-rich systems; Peiffert et al., 1996) and, therefore, U does not another fluid (e.g., Butt et al., 1984; Dahlkamp, 1993). In terrestrial or
develop mineralization at the magmatic stage. Instead, these U anom- marginal marine environments, redox-controlled U precipitation in-
alous granites and volcanic rocks may be excellent sources for other volving plant remains, sulfides, bacterial sulphate reduction, and mi-
types of U mineralization, largely depending on the chemistry of the grated hydrocarbons, as well as adsorption on Ti-Fe-oxyhydroxides and
rocks that controls whether U forms a mineral of its own or is in- clays, may form impregnations and replacements in fluvial, lacustrine,
corporated into silicates, niobates, or phosphates (e.g., Cuney, 2014; and deltaic sandstones (e.g., Dahlkamp, 1993, 2016; Cai et al., 2007;
Ballouard et al., 2018). The U-bearing minerals control whether U is Jaireth et al., 2008; Skirrow et al., 2009: Penney, 2012; Ingham et al.,
readily mobilized during later low-temperature or hydrothermal al- 2014; Bonnetti et al., 2015). Depending on the shape and the position
teration. Hydrothermal leaching of granites and U-bearing sedimentary within the sedimentary unit, different types of sandstone-hosted U
rocks, as well as prograde metamorphic devolatization, may mobilize U mineralization are distinguished, e.g., basal, tabular, or roll-front type.
that subsequently may concentrate in hydrothermal vein-type miner- The distribution of mineralization is controlled by permeability and the
alization or in exogenic mineralization. distribution of reduced and oxidized sections within the sandstones, but
Exogenic redistribution and formation of U mineralization involves also faults and breccia zones (e.g., Cuney and Kyser, 2015). Uranium
oxidation of U during mobilization, transport in solution, and pre- dissolved in the surface runoff and groundwater may be precipitated in
cipitation by reduction, scavenging, or supersaturation of the solution wetlands (peat or lignite) and lakes or it may reach the oceans, where a
(e.g., Cuney and Kyser, 2009, 2015). Before the oxygenization of the significant portion of dissolved U will be precipitated in the coastal
atmosphere, exogenic U mineralization involved the transport of det- environment in organic matter bearing sediments and phosphorites
rital uraninite with its concentration in placers (pebble-quartz con- (e.g., Church et al., 1996; Morford et al., 2005). Black shales that
glomerates; see Table 1). Major sources of U for the formation of exo- formed coast-near or on the proximal shelf – in analogy with Mo in
genic U mineralization include (i) granites and felsic volcanic rocks that black shales (Gilleaudeau and Kah, 2013) – may have higher U contents
have enhanced U contents and host U in readily leacheable form (e.g., than black shales that formed on the outer shelf or in open marine
uraninite, volcanic glass or metamic minerals such as allanite, U- basins (e.g., Schovsbo, 2002).
thorite, and zircon) (Förster et al., 2008; Cuney and Kyser, 2015; The endogenic and exogenic environments of U mobilization,
Trumbull et al., 2010; Cuney, 2014; Bonnetti et al., 2017a,b), (ii) older transport, and precipitation are connected by tectonic processes that
U mineralization that are exposed to the oxidation zone because of return U from the surface to the endogenic cycle or that make U
tectonic uplift or hydrological and climatic changes (Schmitt and Thiry, available for redistribution in the exogenic environment. (i) Subduction
1987; Sejkora et al., 2007; Bolonin and Gradovsky, 2012; Hiller and makes U available for redistribution by metamorphic fluids and partial
Schuppan, 2008, 2016), and (iii) oxidation of lithologies that earlier melts. (ii) Tectonic uplift exposes surface rocks to erosion, which re-
had scavenged dissolved U as for instance black shales (e.g., Fischer moves altered surface rocks that are U-depleted as they have lost U
et al., 2009; Bolonin and Gradovsky, 2012). Transport of U may be during alteration, and makes fresh rocks that have not lost their mo-
related to a wide range of different fluid types, including hot highly bilizable U available for alteration and leaching. Faults play a parti-
saline oxidized diagenetic brines, very acidic for unconformity related cularly important role as they control and localize uplift and act as
deposits and alkaline for tabular or tectonolithologic types of deposits pathways for deep and meteoric fluids to mobilize and transport U.
and collapse breccia pipes, as well as cold meteoric fluids for roll front Precipitation of U in basement faults may be related to the nature of the

478
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

wall rocks, i.e., organic-rich and/or sulfide-rich rocks (Kříbek et al., started at c. 400 Ma with the collision of the leading parts of the Ar-
1999; Runge, 2001; Hiller and Schuppan, 2008; Cuney and Kyser, morican Spur (Armorica and Teplá-Barrandian Unit) with Laurussia and
2015), reaction of hydrothermal fluids with carbonate-bearing wall continued in Europe until c. 340 Ma, involving a series of intra-con-
rocks resulting in phase separation (saline aqueous fluid and CO2-rich tinental subduction zones. The collision at c. 400 Ma resulted in major
fluid) eventually resulting in ligand retraction at the carbonate-silicate dextral stike-slip movement within the Acadian orogen and the opening
contact (e.g., Romer et al., 2005) or by secondary boiling (e.g., of the Paleotethys in future East-Pangea (Kroner et al., 2016). The final
Romberger, 1984; Evert et al., 1997). Fluid mixing is a process of U closure of the Rheic Ocean involved displacement of Gondwana from
precipitation that is particularly important for unconformity-related U Laurussia along the Ancestral Gibraltar Fault, leading to the Allegha-
mineralization (e.g., Derome et al., 2005; Richard et al., 2012). This nian-Mauritanide orogeny, the development of the Central European
type of U mineralization typically is related to basement faults that cut Extensional Province, and the opening of the Neo-Tethys (e.g., Kroner
through highly permeable sedimentary rocks and may form in the et al., 2016).
basement at some distance from the unconformity and in the sedi- The complex shape of the Rheic Suture within the Variscan orogen
mentary rocks at the unconformity or within particular stratigraphic reflects the distribution of blocks of thick continental crust and domains
levels. The origin of the U in this type of deposit is debated: U was of stretched continental crust within the Armorican Spur, which was an
suggested to be derived from the basement or to have been introduced integral part of Gondwana, and late escape movements of different
by basinal brines (for discussion see Kyser et al., 2000; Kyser, 2007; continental blocks, in particular Iberia (e.g., Romer and Kroner, 2018;
Cuney and Kyser, 2009; Alexandre et al., 2009). Element associations Stephan et al., 2018). The Armorican Spur formed during the opening
and mineral assemblages of U mineralization are variable. For instance of the Rheic Ocean in the early Ordovician, when the northwestern part
in the Athabasca basin, there are simple and complex types of U mi- of Gondwana, i.e., the former Cadomian magmatic arc, became ex-
neralization. Deposits of the simple type tend to be hosted in the tended and fragmented (Linnemann et al., 2000; Kroner and Romer,
basement beneath the unconformity and the main ore element is U, as 2013). Areas of thick continental crust, which typically represent those
for instance at McArthur River, Rabbit Lake, and Eagle Point. Deposits parts of the Cadomian magmatic arc that have been extensively in-
of the complex type tend to occur in the basin sediments and include truded by 540–520 Ma granodiorites (Linnemann et al., 2008), were
apart from uraninite also sulfides and/or arsenides of Ni, Co, Cu, Pb, Zn, uplifted and their sedimentary cover in part was removed as indicated
and Mo, as for instance at Cigar Lake, Key Lake, Collins Bay, and by the gap in the Cambrian sedimentary record (Linnemann et al.,
Midwest (e.g. Fayek and Kyser, 1997; Alexandre et al., 2005, 2009; 2004, 2010a). The erosional debris was deposited in the evolving basins
Polito et al., 2005, 2006; Richard et al., 2012). This contrasting geo- between these blocks of thick continental crust and may reach several
chemical fingerprint may indicate that U and the sulfide/arsenide- thousand meters thickness (e.g., Noblet and Lefort, 1990; Linnemann
bound elements were transported by different fluids or that the sulfide/ et al., 2010b; Romer and Hahne, 2010). The various blocks of thick
arsenide-bound elements were already present in the sedimentary host- continental crust later developed into the low-strain domains of the
rocks of the ore or were introduced after U was intrioduced. Variscan orogen, as these blocks did not subduct. In contrast, the
thinned continental crust with its thick sedimentary cover was sub-
ducted during the Variscan orogeny – locally to mantle depth – during
3. The Variscan orogen
continental collision, was exhumed after 340 Ma, and was emplaced
onto the margins of the thick blocks, where they now form high-strain
The closure of the Rheic Ocean, eventually leading to the assembly
domains (Fig. 2; Kroner and Romer, 2010, 2013).
of western Pangea, is closely related with the Acadian, Variscan, and
Paleozoic sedimentary rocks recording the opening and the closure
Alleghanian orogenies (Kroner and Romer, 2013; Kroner et al., 2016).
of the Rheic Ocean in Variscan Europe are most completely preserved in
The closure of the Rheic Ocean between Gondwana and Laurussia

Fig. 2. Simplified map of Central and Western Europe


showing the position of Variscan Massifs and major
sutures (modified from Stussi, 1989; for details see
Kroner et al., 2007, 2010, 2016; Kroner & Romer, 2010,
2013). The spatial extent of Early Paleozoic rocks (in-
cluding Silurian black shales) is shown schematically as
grey shading (Thickpenny and Leggett, 1987). The
grouping of mineralization is for practical porposes
only and does not represent a classification sensu
stricto, whereby hydrothermal-type includes vein and
episyenite U mineralization, sandstone-type as used
here includes roll-front type, basal type, and evapo-
transpiration U mineralization, and lignite-type in-
cludes black shale and peat related U mineralization.
Abbreviations: (bold font) AM = Armorican Massif;
BM = Bohemian Massif; C = Cornwall; FMC = French
Massif Central; IM = Iberian Massif; O-S = Odenwald-
Spessart; P = Pyrenees; RM = Rhenic Massif;
S = Schwarzwald; V = Vosges; (regular font)
CAD = Central Armorican Domain; CIZ = Central
Iberian Zone; NAD = North Armorican Domain;
RHZ = Rheno-Hercynian Zone; STZ = Saxo-Thur-
ingian Zone; SAD = South Armorican Domain;
TBU = Teplá-Barrandian Unit. Numbers refer to U mi-
neralization mentioned in the text, but not shown on
small-scale maps: (1) South Terras Mine; (2) Kruth; (3)
Menzenschwand; (4) Wittichen; (5) Müllenbach; (6)
Stockheim; (7) Coutras.

479
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

the Schwarzburg area of the Saxo-Thuringian zone (Fig. 2). As all


Variscan areas to the south of the Rheic Ocean formed at the northern
margin of Gondwana, i.e., were part of the same shelf, the sedimentary
record of the Schwarzburg area is representative for these areas, al-
though there are regional variations in thickness of the various sedi-
mentary units and facies changes are not strictly coeval (e.g., Noblet
and Lefort, 1990). After the regionally diachronous initiation of rifting
in peri-Gondwana, these chemically intensely altered sediments
(CIA > 95%; CIA = Chemical Index of Alteration; Nesbitt and Young,
1982) redistributed from the interior of the stable craton to its ex-
tending margins, where they form voluminous sedimentary packages.
The deposition age of these sediments ranges from the late Cambrian in
the Middle East, Arabia, the Sahara, and Nova Scotia to the early Or-
dovician in Saxo-Thuringia (e.g., Noblet and Lefort, 1990). In contrast,
the diamectites of the Hirnantian glaciation and the post-glacial early
Silurian black shales are broady correlative, although they have re-
gionally different thickness. Silurian black shales of the peri-Gondwana
shelf, furthermore, also show regionally contrasting contents of organic
material and generally have high U contents (up to 70 ppm), but also up
to 130 ppm Mo and up to 5000 ppm V (e.g., Dabard and Paris, 1986;
Maletz and Katzung, 2003; Fischer et al., 2009; Romer et al., 2014a).
The drastic increase in sedimentation rate and volume with the onset of
the Variscan orogeny is regionally diachronous on the Armorican Spur.
The chemically intensely altered, mostly Cambrian to Ordovician
sedimentary rocks are the most voluminous Paleozoic sedimentary unit
of the northern Gondwana shelf (e.g., Noblet and Lefort, 1990; Avigad
et al., 2005; Meinhold et al., 2013) and of the Variscan metamorphic
units. Thus, their geochemical signature (in particular high Rb/Sr and
anomalously high 87Sr/86Sr) is also present in the sedimentary debris of
Variscan orogen (e.g., Romer and Hahne, 2010).
The Variscan Plate Boundary Zone is complex because of the seg-
mented lower plate, i.e., the Armorican Spur sensu Kroner and Romer
(2013). The Armorican Spur is part of the Gondwana plate. During the
early stages of the Variscan orogeny, the collision of blocks of thick
continental continental of the Armorican Spur with Laurussia, results in
the initiation of new subduction zones behind the various blocks and
culminated during the final stages of continent-to-continent collision
between Gondwana and Laurussia in the subduction of thinned con-
tinental crust (Fig. 3). The Rheic Ocean closed in the area of the
Amorican Spur with the collision of Armorica with Laurussia at c.
400 Ma and shortly thereafter of the Teplá-Barrandian Unit with
Laurussia. These collisions resulted in establishing new, intra-con-
tinental subduction zones behind these two blocks and the formation of
extensional zones along the former suture, most prominently preserved
in the Carpathian, Sleza, Lizard, and Careón ophiolitic units (see Kroner
and Romer, 2013). The initiation of new subduction zones also changed
the position of the collided blocks that originally were in a lower plate
position and after the collision were in an upper plate position, i.e., that
before the collision were part of Gondwana and after the collision and
initiation of a new subduction zone were part of Laurussia. The con-
tinued convergence of Laurentia and Gondwana resulted in the accre-
tion of magmatic arcs, as recorded by local c. 360 Ma blueschist facies
metamorphism (e.g., South Armorican Domain; Ballèvre et al. 2009)
and major strike slip movements.
The final continental collision between Gondwana and Laurussia
resulted in intra-continental subduction. Thinned continental crust and
its cover of sedimentary rocks were subducted, reaching high pressure
(HP) to ultra-high pressure (UHP) or high temperature (HT) to ultra-
high temperature (UHT) metamorphic conditions around 340 Ma (e.g.,
Schaltegger et al., 1996; O'Brien and Rötzler, 2003; Puelles et al., 2005;
Gébelin et al., 2009), and thereaftere were rapidly exhumed to the
middle crust. Remains of these rocks form the high-strain domains of (caption on next page)
the Variscan belt. The continental collision eventually resulted in a
reorganization of plate movements and a relocation of the Euler pole
that describes the relative movement of Gondwana and Laurussia
(Kroner et al., 2016). The final closure of the Rheic Ocean during the

480
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Fig. 3. Schematic reconstruction of the convergence between Laurussia and Province. These three groups fall into similar respective age ranges
Gondwana eventually leading to the assembly of Pangea (Kroner and Romer, along the entire orogen, as they are controlled by the same processes,
2013; Romer and Kroner, 2018). After the initial collision at 400 Ma, in- but have in different tectonic zones contrasting importance (Fig. 3). (i)
tracontinental subduction of thinned continental crust in the Armorican spur The tectonic emplacement of high-strain and high-grade metamorphic
led to continued closure of the Rheic Ocean in the west and opening of the
units occurred along deep-reaching structures that also acted as path-
Paleotethys in the east, with separation of continental blocks from the southern
ways for mantle-derived melts, which sampled the metasomatized
continent. After initial collision, plate movement between mainland Gondwana
and Laurussia is described by the Paleotethys Pole (PT in panels A and B). After mantle. Examples for such rocks include c. 340 Ma durbachites (e.g.,
the end of intracontinental subduction within the Armorican Spur, continued Janoušek and Holub, 2007) and the 340–330 Ma high-K Meissen Pluton
plate-convergence resulted in reorganization of plate movement (described by (e.g., Wenzel et al., 1997) in the Bohemian Massif, numerous c. 350 Ma
Euler Pole NT in panel C; Kroner et al., 2016) that led to in widespread ex- high-K calc-alkaline granites in the Western Massif Central, and c.
tension within the Variscan area of Central and Western Europe (Central Eur- 340–335 Ma high-K calc-alkaline granites and vaugnerites in the
opean Extensional Province, CEEP) and final closure of the Rheic Ocean along Eastern Massif Central and Vosges (Cuney et al., 2001; Laurent et al.,
the Mauritanite-Appalachian belt. Mainland Gondwana is separated from the 2017). In addition, there are c. 350 Ma peraluminous cordierite-biotite
Armorican Spur by the Ancient Gibraltar Fault (AGF) and the South Atlas Fault. granites (Guéret type) in the Eastern Massif Central (Laurent et al.,
NVDF and SVDF correspond to the northern and southern Variscan deformation
2017). Exhumation of dry UHT units, such as the Saxon Granulites, may
fronts, respectively. (A) Distribution of Silurian hot black shales on Gondwana
have induced dehydration melting of mica-bearing wall rocks (e.g.,
according to Le Heron et al. (2009). Silurian hot shales on the Armorian Spur
are shown schematically. Note, although there are Silurian black shales on
Förster and Romer, 2010), facilitating the emplacement of the UHT
Avalonia, Laurentia, and the East European Craton, which became assembled to rocks in the middle crust. (ii) Heating of the crust after emplacement of
Laurussia before the onset of the Variscan orogeny, these shales are not hot. the high-stain units largely was accomplished by advective heat trans-
Their distribution is shown schematically, largely based on Thickpenny and port into the orogenically thickened crust rather than by internal
Leggett (1987). (B) Spatial distribution of post-kinematic Variscan granites heating. Advective heat transport include the emplacement of UHT
based on compilations of Stussi (1989) and Kroner and Romer (2013). Note, the units in the middle crust and the extension of the thickened crust,
Cornish granites are not shown as they are related to the development of the eventually resulting in the upwelling and partial melting of the litho-
Central European Extensional Province. (C) Spatial distribution of post-Variscan spheric mantle, as for instance for the late-orogenic granites of the
U mineralization. Bohemian Massif and the high-K–high-Mg granites of the French Massif
Central (e.g., Ledru et al., 2001; Romer and Kroner, 2015; Moyen et al.,
2017). Similar granites are also widespread in the Iberian Massif Cen-
Alleghanian orogeny, forming the Appalachians and the Mauritanides
tral (e.g., Montero et al., 2004; Castiñeiras et al., 2008; Teixeira et al.,
and later the Ouachita, Marathon, and Sonora belts, was coeval with
2012). By volume, the most important Variscan granitic magmatism
extension in Central and Western Europe and the opening of the Neo-
falls in the age range 327–315 Ma and includes a wide range of per-
Tethys farther to the east (see Kroner and Romer, 2013; Kroner et al.,
aluminous granites (e.g., Cuney et al., 1990; Förster et al., 1999; Castro
2016).
et al., 1999; Couzinié et al., 2014; Moyen et al., 2017). Among these
Closure of the Rheic Ocean during the Variscan orogeny is not
granites, leucocratic two-mica granites represent the most important
strictly coeval as the different low-strain blocks did not collide at the
source for later U mineralization, largely due to the fact that magmatic
same time. Post-collision Variscan magmatism falls into three groups,
uraninite is the major U host in these rocks (e.g., Förster, 1999; Cuney,
i.e., (i) magmatism related to the emplacement of the high-strain do-
2014; Ballouard et al., 2017). (iii) The final closure of the Rheic Ocean
mains into the middle crust, (ii) magmatism related to crustal melting
resulted in large-scale extension within the Variscan belt of Europe with
in orogenically thickened high-strain domains, and (iii) magmatism
major rift basins and related mantle-derived magmatism (Oslo Rift,
related to the development of the Central European Extensional
North German Basin, northern part of the Armorican Massif, Western

Fig. 4. Reconstruction of Gondwana at 250 Ma


(inset) highlighting the spatial distribution of U
mineralization and early Silurian hot shales (black
shales with elevated U contents) in the Variscan and
Appalachian orogenic belts. To avoid crowding, the
distribution of U mineralization is shown schemati-
cally independent of the actual size of the various
deposits. Although this over-emphasizes areas with
only economically insignificant U mineralization, it
highlights the regional pattern of U mineralization.
Early Silurian black shales in low-strain domains are
typically covered by younger sedimentary rocks and
crop out only at basin margins. The shown gen-
eralized distribution of Early Silurian (hot) black
shales (inset) largely bases on data from oil and gas
exploration. Distribution of U mineralization ac-
cording to Dahlkamp (2016) and references quoted
in the text. Distribution of Early Silurian black shales
in northern Africa and Arabia according to Lüning
et al. (2000a,b), Verniers et al. (2008), Al-Juboury
and Al-Hadidy (2009), Le Heron et al. (2009), and
Saberi et al. (2016). Extent of the Hirnantian ice
sheet according to Le Heron et al. (2009). Distribu-
tion of Hirnantian tillites and dropstones is taken
from the compilation of Stephan et al. (in press).
Tectonic reconstruction according to Kroner and
Romer (2013) and Kroner et al. (2016). Location of deposits mentioned in the text (but not shown on smaller-scale maps): TB = Timahdit Basin; OAB = Oulad
Abdoun Basin; JW = Jebbel Wafagga; MB = Millet Brook.

481
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Approaches Trough; e.g., Ziegler, 1987; Neumann et al., 2004; stacked during the Variscan orogeny. Fig. 4 (inset) shows the dis-
Timmerman, 2004; Ballouard et al., 2017). Within the orogen and its tribution of Silurian black shales on Gondwana. Note, black shales with
foreland, numerous fault-bound basins developed with voluminous particularly high U contents are also called hot shales. The early Si-
deposits of the rapidly eroding Variscan belt and predominantly felsic lurian black shales on northern Gondwana have higher U contents than
volcanic rocks (e.g., Schneider et al., 2006; Breitkreuz and Kennedy, middle Silurian and younger black shales (e.g. Lüning et al., 2000a,b;
1999), but also minor intrusions and hydrothermal fluid circulation Seidel, 2003).
that locally resulted in mineralization and the formation of episyenites During the Variscan orogeny, sedimentary rocks from the former
(e.g., Leroy and Turpin, 1988; Romer et al., 2010; Neiva et al., 2011; Gondwana orogen were intensely reworked, in part preserved in vari-
López-Moro et al., 2013). ably metamorphosed rocks of the high-strain domains, in part recycled
The continued closure of the Rheic Ocean had a major effect on the by crustal melting, transferring their U-content to U-rich granites.
Permian climate, enhancing and enforcing the expansion of arid/semi- Although many of these granites contain U minerals, in particular ur-
arid environments due to the drift of northern Pangea in the arid cli- aninite, monazite, xenotime, allanite, and brabantite (Cuney and
matic belt, interrupted by five wetter periods that correspond to periods Friedrich, 1987; Friedrich et al., 1987; Förster, 1998a,b, 1999, 2010;
with an extended icecap in southern Gondwana (Roscher and Förster et al., 2008; Ballouard et al., 2018), they do not show magmatic
Schneider, 2006; Schneider et al., 2006). The same structures that were or late-magmatic hydrothermal U mineralization, except for minor U
active during the closure of the Rheic Ocean were repeatedly re- enrichment in late-magmatic shear zones (Friedrich et al., 1987). In-
activated during the opening of the Tethys and Atlantic oceans, as in- stead, some of these granites show spatially closely related vein-type U
dicated by their role in controlling the age and distribution of sedi- mineralization that is typically 40–50 Ma younger (e.g., Scaillet et al.,
mentary basins. The age of basin formation and inversion and the age of 1996; Förster, 1996; Romer et al., 2010; Ballouard et al., 2018).
structurally controlled mineralization correspond to events of changes Post-Variscan U mineralization formed mainly around 280 Ma in
in the oceanic spreading pattern (Vogt and Tucholke, 1989) and, thus, two different settings, vein and episyenite mineralization in meta-
to periods of changing stress fields on the continent and the reactivation morphic rocks close to late-Variscan granites (e.g., the Gera-Aue-
of old zones of structural weakness (e.g., Romer et al., 2010). Jáchymov Zone, the South Armorican Shear Zone; Hiller and Schuppan,
2008; Ballouard et al., 2018) and sandstone-type mineralization in
4. Distribution of uranium occurrences in Central and Western Permian sedimentary basins (e.g., Barthel, 1974; Dill, 1987; Mathis
Europe et al., 1990; Pagel, 1990). The timing and location of individual U oc-
currences in Central and Western Europe is controlled by large-scale
Pre-Variscan U mineralization is uncommon in Central and Western crustal extension that is coeval with the final stages of the closure of the
Europe, whereas U-rich lithologies are widespread. The Ronneburg area Rheic Ocean along the Alleghanian-Mauritanide belt. Later reactiva-
is the exception. It hosts the most important U deposits in Central and tions of these extensional zones as well as intense weathering resulted
Western Europe. These deposits are closely related to Silurian and in younger redistribution of U, which may have resulted in the de-
Devonian black shales and the underlying glaciomarine sedimentary struction or enrichment of older U mineralization and in rare cases even
rocks of the Lederschiefer Formation that have been tectonically in the formation of new U mineralization such as the U deposits hosted

Fig. 5. Simplified tectonic map of the Bohemian


Massif showing the distribution of major tectonic
elements and the distribution of U mineralization
grouped according to type. Arrows indicate direc-
tion of U-transport during the formation of the
Cretaceous sandstone-hosted U deposits as sug-
gested by the original workers (Tonndorf, 2000;
Surán and Veselý, 2001). Sedimentary Basins men-
tioned in the text: C = Cheb Basin; D = Döhlen
Basin; EB = Erzgebirge Basin; S = Sokolov Basin.
Granitic and volcanic complexes mentioned in the
text: M = Meissen; T = Tharandt. Numbers refer to
U mineralization mentioned in the text: (1) Döhlen;
(2) Hamr; (3) Horní-Slavkov; (4) Jáchymov; (5)
Königstein; (6) Markersbach; (7) Olší; (8) Okrouhlá
Radouň; (9) Osečná; (10) Příbram; (11) Rožná; (12)
Stráž; (13) Zadní-Chodov. Note, mineralization of
the northern part of the Bohemian Massif, i.e., the
Erzgebirge and the Ronneburg ore district are shown
in a more detailed map (Fig. 6). Map based on Žák
et al. (2014), spatial distribution of U mineralization
according to Surán and Veselý (2001) and Dahlkamp
(2016).

482
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

in the Tertiary sandstones of Coutras, France (Meunier et al., 1992). metamorphic rocks and migmatites that have been intruded by
Fig. 3 shows the distribution of pre- and post-Variscan U mineralization, 320–345 Ma granitoids (e.g., Janoušek et al., 2004a). The high-grade
Silurian shales, and Variscan post-kinematic granites to highlight the rocks of the Moldanubian Zone represent mainly metamorphosed vol-
spatial relation between different U-rich lithologies and U mineraliza- canosedimentary rocks of the former post-Cadomian Gondwana shelf
tion. and show great similarity with the rocks of the Saxo-Thuringian Zone,
both with respect to protoliths (e.g., the early Ordovician high Rb/Sr
4.1. Uranium mineralization in the Teplá-Barrandian Unit and the and high K/Na rocks of the Frauenbach Group of the Saxo-Thuringian
Moldanubian Zone Zone and the Gföhl Unit of the Moldanubian Zone; Janoušek et al.,
2004a; Romer and Hahne, 2010), peak metamorphic conditions
The Teplá-Barrandian Unit is a low-strain domain of the Variscan (O'Brien and Rötzler, 2003), as well as age of metamorphism and
orogen that collided around 380 Ma with Laurussia, forming the magmatism (see compilation in Kroner and Romer, 2013). Uranium
Sudetes. It is on top of the high-strain domains of the Saxo-Thuringian mineralization in the Moldanubian Zone is bound to shear zones in the
and the Moldanubian zones (Fig. 5) that have reached their meta- variably metamorphosed Variscan basement. The shear zones have
morphic peak conditions around c. 340 Ma (e.g., O'Brien and Rötzler, been repeatedly reactivated under brittle conditions and show altera-
2003). Note, the Teplá-Barrandian Unit was not affected by this 340 Ma tion that varies in extent, íntensity, and type. The U mineralization at
event. At its southeastern margin, the Teplá-Barrandian Unit was in- Rožná and Olší is hosted in the high-grade metamorphic rocks of the
truded by 354–346 Ma calc-alkaline rocks of the Central Bohemian Gföhl Unit close to the tectonic boundary to the Svratka Crystalline
Plutonic Complex (Janoušek et al., 2000, 2004b). The low-grade me- Unit. This boundary became repeatedly reactivated during the Late
tamorphic rocks of the Teplá-Barrandian Unit and the Saxo-Thuringian Stephanian to Permian reorganization of the stress field, resulting in
and Moldanubian zones are covered by Permian volcanosedimentary pre-U c. 308–304 Ma quartz-sulfide and carbonate-sulfide mineraliza-
and Cretaceous sedimentary rocks. There are two major types of U tion, c. 275–265 Ma U mineralization with uraninite, brannerite, cof-
mineralization, post-Variscan vein-type mineralization and Cretaceous finite, and U-Zr-silicates, and c. 230 Ma post-U quartz-carbonate-sulfide
and Tertiary stratiform lignite-, coal-, and sandstone-hosted miner- mineralization (Kříbek et al., 2009). The associated U-stage alteration
alization (Surán and Veselý, 2001). includes a variably important hematitization and albitization envelope
(Kříbek et al., 2009). The shear-zone hosted Okrouhlá Radouň U mi-
4.1.1. U Mineralization in the Teplá-Barrandian Unit neralization occurs near the contact between Variscan granites and
In the Teplá-Barrandian Unit, U mineralization is restricted to Upper high-grade metasedimentary rocks (Dolníček et al., 2014). The miner-
Proterozoic and Lower Cambrian sedimentary rocks, concentrated in alization has a polystage development with broad, external zones of
two zones, a belt along the intrusive contact of the Central Bohemian pre-ore phyllitic alteration (chloritization, sericitization, and argilliti-
Plutonic Belt and the Příbram area (Kříbek et al., 1999). Locally, mi- zation) and narrow, internal zones of episyenite formation, an ore
neralization is at the contact of diabase dikes and their sedimentary stage, and post-ore quartz and carbonate veins (Dolníček et al., 2014).
wall rocks (Škácha et al., 2009). There are three sets of steeply dipping Main ore mineral is uraninite that is variably altered to coffinite, Y-REE-
veins, one parallel to the NE-striking host-rock foliation, the other two rich coffinite, and secondary U-Ti-Si-Zr minerals (René, 2008; Dolníček
northwest- and north-striking sets cutting the foliation (Kříbek et al., et al., 2014). Uraninite from the ore stage has been dated at
1999). The structurally controlled mineralization is irregularly dis- 280–270 Ma, whereas post-ore U redistributions in quartz-vein may
tributed and typically forms clusters of veins. There seems to be a have occurred around 180–160 Ma (Anderson in Dolníček et al., 2014).
systematic change in mineral association within individual veins from Early Permian shear zones host U mineralization in the western part of
an early polymetallic sulfide assemblage (Pb, Zn, Cu) to younger ur- the Moldanubian Zone (e.g., Zadní-Chodov) contain uraninite with
aninite-calcite and sulfide/selenide -calcite assemblages (Zn, Pb, As, Cu, variable amounts of sulfides and selenides that have been repeatedly
Co, and Ag). Polymetallic sulfide mineralization was formed by remobilized and redistributed within the mineralization (Dill, 1986;
medium- to high-salinity fluids, whereas U mineralization was formed René, 2008). Common features of the different styles of U mineraliza-
by low-salinity, low-temperature (80–140 °C) fluids of possible me- tion in the Moldanubian Zone include (i) a deeply eroded basement
teoric origin (Kříbek et al., 1999). Most veins with U mineralization covered by Permian continental sedimentary and volcanosedimentary
occur in the upper Proterozoic sedimentary units, whereas U-beaing rocks, (ii) multiple Permian and younger reactivation of the faults, (iii)
veins in the upper Proterozoic volcanosedimentary and the lower pre-ore migration of oxidized, alkaline basinal low-temperature fluids
Cambrian sedimentary units are subordinate and corresponding veins that resulted in the chloritization of mafic minerals, albitization of
in the granitic intrusions of the Central Bohemian Plutonic Complex are feldspar, and dissolution of quartz (Kříbek et al., 2009; Dolníček et al.,
negligible (Kříbek et al., 1999). Even thought the U veins occur pre- 2014), and (iv) post-ore redistribution of U within the mineralization.
dominantly in the exocontact of the Central Bohemian Plutonic Com-
plex, the U veins – in contrast to the c. 340 Ma quartz-Au veins – are not 4.1.3. U mineralization in the post-Permian sedimentary rocks of the
related to the emplacement of the igneous rocks (Škácha et al., 2009), Bohemian Massif
as indicated by the uraninite U-Pb ages of 270–280 Ma (Kříbek et al., Uranium mineralization in the Cretaceous sandstones, overlying
1999; Škácha et al., 2009). There seems to have been multiple later parts of the Teplá-Barrandian Unit and the Moldanubian Zone, occurs
events of U mobilization and redistribution as demonstrated by U-Pb mainly in Cenomanian units that have been deposited on the crystalline
ages in the range of 150–190 Ma (Škácha et al., 2009). Uranium mi- basement and that are overlain by impermeable Turonian sedimentary
neralization in the Teplá-Barrandian Unit is particular because of its rocks. Mineralization occurs at Hamr and Osečná, as well as at Stráž, in
high content of bitumen that formed in the upper Proterozoic sedi- the NE part of the Cretaceous basin and has been exploited by mining
mentary rocks by pyrolysis during the emplacement of the Central and in-situ leaching, respectively (Norman, 1993; Ekert and Mužák,
Bohemian Plutonic Complex. The pyrolysis products seem to have mi- 2010). This sandstone-hosted mineralization seems to be comparable to
grated away from the intrusive rocks and cumulated in structural traps U mineralization at Königstein, which is bound to basal units of Cen-
(Kříbek et al., 1999). Uranium mineralization is closely associated with omanian sandstones that have been deposited on the crystalline base-
the introduction of bitumen into the mineralization (Kříbek et al., ment. At Königstein, U distribution is controlled by organic-rich sand-
1999). stones and shales with U derived from the leaching of the U-fertile
Markersbach granite in the eastern Erzgebirge (Tonndorf, 2000). For
4.1.2. U mineralization in the Moldanubian Zone the sandstone-hosted U mineralization of Hamr, Osečná, and Stráž,
The Moldanubian Zone is dominated by Variscan high-grade however, U may be derived from leaching of U-fertile granites of the

483
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Sudetes. age of primary mineralization has been debated for some time, largely
Numerous small stratiform U accumulations are also known from due to the small grain size of the dated material and related open
Tertiary lignite-coal-sandstone formations in the Eger graben, in par- system behavior, a Triassic age with a Jurassic thermal overprint in
ticular in the Sokolov and Cheb basins (Dahlkamp, 2016). The source of preferred now (Brockamp et al., 1987; Holliger et al., 1989).
U for this mineralization could be fertile granites and older miner-
alization within the present Erzgebirge that became uplifted during the
4.2. Uranium mineralization in the Saxo-Thuringian Zone
development of the Eger rift.
The westward continuation of the Bohemian Massif is hidden be-
The Saxo-Thuringian Zone is the Variscan segment with by far the
neath Permian and younger sedimentary rocks. Exposures of Variscan
most prominent and most variable U mineralization, including the U
basement in the Vosges and the Schwarzwald-Odenwald-Spessart,
districts of Jáchymov, Schlema-Alberoda, and Ronneburg that are
however, contain magmatic and metamorphic units that are thought to
aligned along the Gera-Jáchymov Zone (Fig. 6). Among the pre-Var-
be correlative with the Mid-German-Crystalline Zone, the Saxo-
iscan rocks of the Saxo-Thuringian Zone, sedimentary rocks of the
Thuringian Zone, and the Moldanubian Zone (Altherr et al., 2000;
glaciomarine Lederschiefer and the Unterer Graptolithenschiefer (black
Kober et al., 2004). Uranium mineralization is subordinate and forms
shales) formations have the highest U content (Fig. 7). During the
two different types, post-Variscan vein mineralization and Permian
Variscan orogeny, these rocks became contrastingly overprinted, de-
sandstone-hosted mineralization. At Menzenschwand and Wittichen
pending on whether they were part of the low-strain (Authochtonous
(Schwarzwald) and Kruth (Vosges), 310–280 Ma vein-type U miner-
Domain of Kroner et al. 2010) or the high-strain (Allochthonous Do-
alization in the Variscan basement (Hofmann and Eikenberg, 1991;
main) domains or the Wrench-and-Thrust Zone, along which the rocks
Meshik et al., 2000) is more than 40 Ma younger than Variscan granitic
of the low-strain and the high-strain domains were tectonically juxta-
magmatism (e.g., Hofmann and Eikenberg, 1991). There is textural
posed against each other during the post-340 Ma emplacement of the
evidence for multiple later remobilization (Staude et al., 2007), which
metamorphic rocks of the high-strain domain.
is in line with isotopic ages for these later events of 165–150 Ma,
In the low-strain domain, the original stratigraphic relations es-
80–90 Ma, and 65–50 Ma (Hofmann, 1989; Hofmann and Eikenberg,
tablished on the former Gondwana shelf are largely preserved. The
1991; Wernicke and Lippolt, 1997; Pfaff et al., 2009). Sediment-hosted
Silurian sedimentary rocks are deposited concordantly on glaciomarine
U mineralization has been described from Müllenbach (Schwarzwald)
diamectites (Lederschiefer) of the Hirnantian glaciation (Fig. 7). The
with mainly stratabound mineralization in fine-grained clastic units
Silurian rocks represent a very condensed sedimentary section of
rich in organic matter, but also as roll-type mineralization in coarse-
60–90 m thickness, whereof the lowermost 30–40 m are dominated by
grained carbonaceous arkosic units (Holliger et al., 1989). Although the
the organic-rich sedimentary rocks of the Unterer Graptolithenschiefer

Fig. 6. Simplified map of the Saxo-Thuringian Zone (inset) showing the spatial distribution of the three major, structurally defined tectonic, i.e., the Authochronous
Domain, the Allochthonous Domain, and the Wrench-and-Thrust Zone along which the Authochronous and Allochthonous domains are juxtaposed (Kroner et al.,
2010). Uranium mineralization is hosted in the Allochthonous Domain and the Wrench-and-Thrust Zone. The primary sources of U are the Silurian shales that host
the Ronneburg district (Wrench-and-Thrust Zone) and the post-kinematic Variscan granites of the Erzgebirge-Vogtland (that may have recycled Silurian rocks
included in the stacked metamorphic units of the Allochronous Domain). Most U mineralization within the Allochronous Domain formed at c. 280 Ma during the
development of the Central European Extensional Province, is structurally controlled, and obtained its U from leaching of U-fertile granites and possibly Silurian
black shales/slates. Most U mineralization shows evidence for multiple metal redistribution during later tectonic processes, which may have added Co, Ni, Bi, As, and
Ag, and during tectonically and climatically controlled controlled oxidation of near-surface parts of individual deposits. Major tectonic elements: BA = Berga
Antiform; EB = Erzgebirge Basin; FL = Franconian Line; Müma = Münchberg Massif; MS = Mehltheuer Synform; SGM = Saxon Granulite Massif; ZTS = Ziegen-
rück-Teuschnitz Syncline. Granites between the Gera-Jáchymov (GJZ) and Zeulenroda-Zobes/Bergen-Mariánské Láznĕ (ZZBMLZ) zones: A = Aue granite;
B = Burkersdorf granite; BB = Bernbach/Beierfeld granite; Ber = Bergen granite; Eib = Eibenstock-Nejdek pluton; Kir = Kirchberg granite; S = Schwarzenberg
granite. Numbers refer to U mineralization mentioned in the text: (1) Aue-Niederschlema; (2) Beerwalde; (3) Bergen; (4) Culmitsch; (5) Gauern; (6) Gottesberg-
Grummelstock; (7) Horní-Slavkov (8) Jáchymov; (9) Korbußen; (10) Lichtenberg; (11) Pöhla/Tellerhäuser; (12) Reust; (13) Ronneburg area; (14) Schirmichau; (15)
Schelma-Alberoda; (16) Schneckenstein; (17) Schneeberg; (18) Schwarzenberg/Johanngeorgenstadt; (19) Sorge; (20) Stolzenberg; (21) Zobes. Cities shown for
orientation: C = Chemitz; G = Gera; J = Jáchymov; ML = Mariánské Láznĕ. Map modified from Hiller and Schuppan (2016) complemented with information from
Förster, 1999; Hahn et al. (2010), and Kroner and Romer (2010). Spatial distribution of U mineralization according to Wismut SDAG (e.g., Runge, 2001), Hiller and
Schuppan (2008, 2016) and Dahlkamp (2016).

484
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Fig. 7. The 250 Ma record of Paleozoic sedimentary rocks of the Schwarzburg section (Thuringia, Fig. 6) reflects changes in the drainage area and the depositional
environment, most prominently, the opening of the Rheic Ocean in the Cambrian to Early Ordovician and the closure of the Rheic Ocean during the Variscan
Orogeny. The most distinctive chemical fingreprints include the temporal variation of Na2O/K2O (black line; averages) and Rb/Sr (symbols), but also the peaks in U,
V, and Sb. Note, high Na2O/K2O and Rb/Sr are chemical fingerprints of intense chemical alteration on the stable continent. The large volume of 500–490 Ma old
sedimentary rocks (up to 3000 m thick) carrying this chemical signature reflects the redistribution of these sediments from the continent to evolving basins during
peri-Gondwana rifting and incipient opening of the Rheic ocean. Note, this signature reappears in the voluminous sediments (after c. 380 and c. 340 Ma) related to
the erosion of Variscan metamorphic rocks exhumed during the various accretion and collision stages (Romer and Hahne, 2010). Note, the Hirnantian glaciation
(dark grey band at c. 445 Ma) largely removes chemically intensely altered material and brings fresh material to the surface. The melting of the Hirnantian ice sheet
leads to anoxic conditions in the ocean and makes these fresh surfaces available for leaching. Anomalous high U, V, and Sb contents in the lowermost unit of Silurian
black shales reflect the entrapment of material leached from the fresh surface rocks. Enhanced contents of U and Sb in younger sedimentary rocks are not correlated
and possibly do not correspond to an original signature of the deposition environment, but reflect redistribution during later events (see text). Light grey bands: (1)
Extension of the peri-Gondwana margin eventually resulting in the opening of the Rheic ocean. (2) First collision of the leading parts of the Armorican Spur
(Armorican Massif and Teplá-Barrandian Unit) with Laurussia. (3) Final collision between Gondwana and Laurussia with subduction of continental crust in the area
of the former Armorican Spur, leading to Variscan UHP and UHT metamorphism. Chemical data from Romer and Hahne (2010) and Romer et al. (2014a). (See above-
mentioned references for further information.)

Formation with thin radiolarite beds and local phosphorite concretions. mineralization occurs in the Ronneburg area (Fig. 6) and has been
These black shales are overlain by the 15–50 m thick Ockerkalk For- mined in open pits (Lichtenberg, Stolzenberg, Ronneburg) and in un-
mation, which mainly includes limestones with intercalations of marl derground operations (Schirmichau, Lichtenberg, Reust, Beerwalde,
and clay, and the 15–20 m thick Oberer Graptolithenschiefer Formation Korbußen). The largest and economically most important deposit
that reaches into the Lower Devonian (Schauer, 1971; Seidel, 2003; (Lichtenberg) had been mined in a 2000 m by 850 m large, 230 m deep
Verniers et al., 2008; Linnemann et al., 2010a,b). The two organic-rich open pit (Runge, 2001). Mineralization is restricted to the Unterer
units have enhanced U contents. It is noteworthy that the Unterer Graptolithenschiefer Formation and bordering lithologies (in particular
Graptolithenschiefer Formation has markedly higher U contents (up to the underlying glaciomarine Lederschiefer Formation) below the base
70 ppm) than the Oberer Graptolithenschiefer Formation (up to of oxidation zones (Runge, 2001). The distribution of U is irregular due
20 ppm; Romer and Hahne, 2010). to redistributions during the Variscan orogeny and multiple tectonic
In the Wrench-and-Thrust Zone (Fig. 6), the Palaeozoic sedimentary reactivation of the Gera-Jáchymov Zone, as well as post-Variscan oxi-
rocks are folded and stacked during the late stages of the emplacement dation of the upper parts of the mineralization (Bolonin and Gradovsky,
of the Variscan nappes of the high-strain domain (e.g., Maletz and 2012).
Katzung, 2003; Heuse et al., 2006; Hahn et al., 2010). Folding and The high-strain rocks of the Allochthonous Domain have been me-
thrusting of the erosional debris (conglomerates, sandstones, and tur- tamorphosed during the Variscan orogeny. With the exception of the
bidites) of the evolving Variscan orogen demonstrates that deformation gneiss domes, the various metamorphic nappes (Phyllite, Garnet-
in the Wrench-and-Thrust Zone was poly-stage, started around 340 Ma Phyllite, and Mica schist – Eclogite units) are dominated by Ordovician
ago and may have lasted at least to 330 Ma (cf. Hahn et al., 2010). sedimentary rocks of the former Gondwana shelf. The lower-grade
Uranium mineralization in Silurian and Devonian black shales occur metamorphic units, which are mainly exposed in the western part of the
only in the Wrench-and-Thrust Zone (Fig. 6), in particular in areas Erzgebirge and Vogtland, also include Silurian lithologies (cf. Mingram,
where these sedimentary rocks have been stacked during the late stages 1998; Rötzler and Plessen, 2010). Erosion of the metamorphic rocks of
of the Variscan Orogeny (Schauer, 1971; Hahn et al., 2010) and/or the high-strain domain and deposition of the erosional debris without
have been cut by late- and post-Variscan faults. The most important U erosional gap in the Autochthonous Domain and the future Wrench-

485
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

and-Thrust Zone results in a repetition of the geochemical signatures of movement between Gondwana and Laurussia after the continental
the sedimentary rocks of the Frauenbach Group (i.e., high Rb/Sr, high collision of these two blocks and the related new stress pattern, as well
87
Sr/86Sr; Romer and Hahne, 2010). Once the Variscan granites of the as the development of the Central European Extensional Province (cf.
Erzgebirge became exposed, U also became available for redistribution Kroner et al., 2016).
(Fig. 7). Most U deposits in the Saxo-Thuringian Zone occur in the high- The tectonic reactivation of Variscan structures not only controls
strain rocks of the Allochthonous Domain. All U mineralization in the the distribution and shape of the sedimentary basins, but also the
Allochthonous Domain post-dates Variscan metamorphism and nappe spatial distribution and orientation of vein-type U mineralization. The
emplacement. most prominent fault zone with vein-type U mineralization is the Gera-
The 327–318 Ma granites of the Erzgebirge-Fichtelgebirge formed Jáchymov Zone that runs from the Autochthonous Domain in the
after the metamorphic nappes were emplaced and are derived from northwest across the Wrench-and-Thrust Zone and the Allochthonous
melting of the orogenically thickened crust (e.g., Förster et al., 1999; Domain to the Tertiary Eger rift in the southeast (Hiller and Schuppan,
Förster and Romer, 2010). Depending on the nature of the protoliths, 2008, 2016). The U deposits of the Ronneburg district, Schlema-Al-
the granites form several chemically and mineralogically contrasting beroda, Schneeberg, Tellerhäuser, Johanngeorgenstadt, Jáchymov, and
suites (low-F biotite granite; low-F two-mica granite; medium-F, low- Příbram – to name only the most important ones – are located along this
P2O5, A-I-type biotite granite; high-F, low-P2O5, A-type Li-mica granite; zone. The Zeulenroda-Zobes/Bergen-Mariánské Láznĕ Zone is a deep
high-F, high-P2O5, S-type Li-mica granite). Most prominent among fault zone parallel to the Gera-Jáchymov Zone (Figs. 5 and 6). It hosts
them are the Sn-specialized granites of the Eastern and Western Erz- among others the U deposits of Zobes, Bergen, Schneckenstein, Got-
gebirge (e.g., Förster et al., 1999; Förster and Romer, 2010). Some of tesberg-Grummetstock, and Horní Slavkov (Hiller and Schuppan,
these 327–318 Ma granites have high U contents (up to 50 ppm U) that 2016). Uranium mineralization of the western Erzgebirge and bor-
is mostly hosted in uraninite, thorite, and to a lesser extent in monazite, dering areas is related to these two fault systems or to smaller faults
xenotime, apatite, and zircon (e.g., Förster, 2010). Therefore, in some systems between them (Fig. 6). Within the Allochthonous Domain, the
of these granites U is readily available for mobilization and con- two fault zones roughly constrain the Eibenstock granite to the NE and
centration to form U deposits (e.g., Förster and Förster, 2000; Förster SW. Furthermore, there occur several granite intrusions (Aue, Bur-
et al., 2008, 2009). Thus, these granites represent a major U source for kersdorf, Bernsbach/Beierfeld, Lauter, Schwarzenberg) along the Gera-
younger vein-type (e.g., Schlema-Alberoda, Schneeberg), sandstone- Jáchymov Zone (Förster et al., 1999, 2009). Uranium mineralization
hosted (e.g., Königstein), and coal-hosted (e.g., Döhlen) U mineraliza- normally occurs in the outer margins of the contact metamorphic aur-
tion (e.g., Tonndorf, 2000; Reichel and Schauer, 2006; Förster, 2010). eole of the granites rather than in the granites (e.g., Hiller and
Exhumation of the high-strain domains and the granites emplaced Schuppan, 2008). Generally, the veins are hosted in metamorphosed
into them, as well as the folding and thrusting of the rocks in the Ordovician to Silurian sedimentary rocks and seem to be particularly U-
Wrench-and-Thurst Zone, provide erosional debris for redistribution rich in pyrite-bearing phyllites and amphibolites (Runge, 2001). The
into Late Carboniferous to Permian basins, as for instance the decimeter to meter wide steeply dipping vein-type mineralization has
Erzgebirge, NW Saxon, and Döhlen basins (Roscher and Schneider, been extensively mined since the 15th century (originally mainly for Ag
2006; Schneider et al., 2006). Uranium in the Silurian black shales and and locally also for Sn and Fe). The five-metal (Ag-Bi-Co-Ni ± U) as-
their metamorphic equivalents, as well as U in readily altered minerals semblage, which is characteristic for these veins, is the result of re-
in the Variscan granites, became available for redistribution by oxidized peated metal redistribution. For instance, hydrothermal redistribution
fluids. Some of this leached U is entrapped in reduced units of these occurred at 180, 150, and 120 Ma and is related to changes in the
Late Carboniferous to Permian basins. Scavenging of U by organic spreading rate of the North Atlantic Ocean and the resulting changes of
material in fluvial and lacustrine environments seems to have been the stress pattern in the continent (e.g., Romer et al., 2010). Typically,
particularly important in the Döhlen Basin (Figs. 5 and 6), possibly native Ag is late in each phase of hydrothermal remobilization and may
favored by the closeness of several high-U granites and volcanic com- be locally associated with native As, Bi, and Sb, and/or Bi-Sb and As-Sb
plexes (e.g., Meissen, Tharandt; Reichel and Schauer, 2006). The ash- alloys (e.g., Ondruš et al., 2003b). The most prominent mineralization
rich (15–50%) and sulfur-rich (2–13%) coals of the Döhlen Basin ty- of this type occurs in the districts of Schneeberg/Schlema/Alberoda,
pically have 10–80 ppm U, but locally show very distinct U enrichment Pöhla/Tellerhäuser, Schwarzenberg/Johanngeorgenstadt, Zobes,
that seems to depend on coal type, reaching average U contents of Bergen, and Jáchymov (Runge, 2001; Ondruš et al., 2003a; Hiller and
0.12 wt% (carbargillite), 0.20 wt% (boghead and channel), 0.30 wt% Schuppan, 2008, 2016).
(silicate and sulfur-rich sapropelic), and 0.49 wt% (gélosic; Reichel and There are two U sources for the vein-type mineralization, the
Schauer, 2006). Typical coal seams have a thickness of 0.4–0.6 m, but granites (Tischendorf and Förster, 1994; Förster, 1999; Förster et al.,
may reach up to 6 m in topographic depressions. The sediments of the 2009) and the Silurian organic-rich sedimentary rocks that additionally
Döhlen basin have been mined for more than 500 years, originally for may have served as sources for Ni, Co, As, Ag, Mo, and Se (e.g., Ondruš
Cu and clay, building stones, and fossils, later mainly for coal (until et al., 2003a; Romer et al., 2010). The importance of these two lithol-
1967), and eventually for U (1963–1989; Reichel and Schauer, 2006). ogies as U source changed with time, as Ni and Co are typically only
Coal seams from other Carboniferous and Early Permian basins of the found in late redistributions that are unrelated with the Permian for-
Variscan orogen (e.g., Stockheim basin; Dill, 1987) are not known to mation of the mineralization (B. Förster, 1996; Ondruš et al., 2003a;
have similarly high U contents as those of the Döhlen basin, which Hiller and Schuppan, 2008). Redistribution of U by oxidizing meteoric
implies that the availability of a U-source is a critical factor to form U- fluids typically reached 100–200 m, but locally as far as 500 m, below
rich peat. the surface (Runge, 2001). Other fault zones parallel to the Gera-Já-
The Permian basin development in the Saxo-Thuringian Zone is chymov and the Zeulenroda-Zobes/Bergen-Mariánské Láznĕ zones in-
controlled by fault zones parallel to the basin axes and transverse to the clude the Franconian Line with minor Permian vein-type U miner-
basins and shows associated mainly felsic volcanism (e.g., Breitkreuz alization (Dill, 1986) and the Elbe Zone with no known Permian vein-
and Kennedy, 1999; Schneider and Romer, 2010). These fault zones type U mineralization. The contrasting importance of U mineralization
have the same orientation as major structural elements of the Variscan along these faults zones seems to reflect the distribution of U-rich
orogen and, thus, represent the reactivation of older structures, which protoliths and their availability for leaching and U redistribution, but
began after c. 305–300 Ma and lasted until about 275 Ma (e.g., also whether these U-rich protoliths have been leached earlier and
Breitkreuz and Kennedy, 1999; Lützner et al., 2006; Schneider et al., whether erosion and tectonic movements brought fertile protoliths into
2006). Reactivation and basin formation is coeval within the entire the oxidation zone and made them available for leaching.
Variscan orogen and, therefore, is related to the change of relative Within the Wrench-and-Thrust Zone, the Gera-Jáchymov and the

486
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Zeulenroda-Zobes/Bergen-Mariánské Láznĕ zones cut through Silurian reactivation of fault zones due to the changing stress regime in Europe,
rocks of the pre-Variscan basement (e.g., Hiller and Schuppan, 2016). which is closely related to the opening of the Tethys and the Atlantic
In the Ronneburg area, vein-type mineralization formed during the oceans and later with the Alpine orogeny (cf. Romer et al., 2010), re-
Permian reactivation of the Gera-Jáchymov Zone. Furthermore, intense sults in the uplift of blocks along these reactivated faults and inversion
Permian weathering (Rotliegend) resulted in the redistribution of U of basins with (i) erosion of uplifted blocks and redistribution of U from
within the lower Silurian black shales in particular along faults of the U-rich basement rocks (e.g., Silurian sedimentary rocks and late-Var-
Gera-Jáchymov Zone that allowed for deeper penetration of oxidized iscan granites) and remobilization of U in existing deposits, (ii) changes
meteoric waters and that may have been reactivated repeatedly (Runge, in the hydrologic setting with redistribution of U from reduced lithol-
2001; Bolonin and Gradovsky, 2012). In the Sorge-Culmitzsch-Gauern ogies that become oxidized, and (iii) fluid circulation along the faults
area, the reactivation of faults of the Zeulenroda-Zobes/Bergen-Mar- with mobilization of U from older U mineralization and leaching of U
iánské Láznĕ Zone resulted in the formation of halfgraben structures from U-rich wall rocks. Thus, exogenic redistribution of U in many cases
that contain conglomerates, red and grey sandstones, shales, and do- are the indirect result of endogenic processes. Although there are dis-
lomites. Marine and terrestrial sedimentary rocks interfinger as in- tinct changes in the European stress field related to changes in
dicated by local ripple and current marks as well as desiccation cracks spreading rate and position of the spreading axis of the Atlantic Ocean
and pseudomorphs after salt crystals (Runge, 2001). Uranium miner- at c. 180, 150, and 120 Ma and the Alpine Orogeny (cf. compilation in
alization is related to sulfide- and plant-rich shales. Locally U and PbS Romer et al., 2010), the most prominent inversion occurred in the
replace plant remains that range from chaff to entire tree trunks Cretaceous (Kley et al., 2008; Al Hseinat and Hübscher, 2017). Erosion
(Runge, 2001). Note, the Kupferschiefer, a bituminous shale that de- related U redistribution resulted in the formation of U mineralization in
veloped at the base of the Zechstein, seems to have its Cu-Ag-richest Cretaceous sandstones, most prominently in Königstein (Tonndorf,
sections related to NW-SE striking fault zones, which highlights the role 2000). The U mineralization occurs mostly in sediments deposited di-
of late structures intersecting redox traps for the enrichment (and rectly above the Variscan basement and only locally in weathered
possibly introduction) of Cu and Ag in this type of stratabound deposit basement rocks (Tonndorf, 2000). The exposed Markersbach granite is
(Borg et al., 2012). Similar to the shale-bound U deposits of the Ron- supposed to be the source of U for the Königstein deposit. Uranium was
neburg and Sorge-Culmitzsch-Gauern areas, there occur late Ni-Co transported by ground waters and trapped in organic-rich intercalations
veins (Runge, 2001; Bolonin and Gradovsky, 2012; Borg et al., 2012). In in terrestrial sandstones and in glauconitic and pyritic layers of silt and
contrast to the coeval U mineralization in the Sorge-Culmitzsch-Gauern clayey silt in Cenomanian sandstones (Tonndorf, 2000). The uplift of
area, the Early Zechstein Kupferschiefer does not show U mineraliza- blocks eventually resulted in changes of the hydrologic setting with
tion (Borg et al., 2012), which indicates that the availibility of leach- oxidation of previously reduced units, which in turn resulted in the
able source rocks for U (and also Cu) is likely to control the contrasting mobilization of U from oxidized units and precipitation of U in deeper,
metal associations in similar lithologies. Thus, tectonic processes ex- still reduced units. This redistribution locally enhanced U contents in
posing potential source rocks and large-scale glaciations (like the Hir- existing U mineralization. This process also operated in the Silurian
nantian glaciation) removing U depleted (as already leached) source shales of the Ronneburg area and the late Permian sedimentary rocks of
rocks from the surface may be the key processes controlling both the the Sorge-Culmitzsch area (e.g., Lange and Freyhoff, 1991; Runge,
spatial distribution and the timing of U mineralization in black shales. 2001; Bolonin and Gradovsky, 2012), as well as in the near-surface
Post-Permian U-redistribution was controlled by both endogenic parts of vein-type mineralization (e.g., Schlema-Alberoda, Zobes-
factors (tectonic reactivation of fault zones with associated hydro- Bergen, Jáchymov, Horní Halže; Plášil et al., 2014; Sejkora et al., 2007;
thermal fluid flow) and exogenic factors (erosion and climate). The Hiller and Schuppan, 2008, 2016). Such changes in the hydrologic

Fig. 8. Simplified map of the Armorican Massif and


the French Massif Central, showing the distribution
of major crustal scale-shear zones, early Paleozoic
sedimentary rocks (generalized), Carboniferous fer-
tile granites, and U mineralization. MC = Margeride
Complex; SSM = Saint Sylvestre Massif; VD = Velay
Dome. Numbers refer to U mineralization and sedi-
mentary basins mentioned in the text: (1) Aumance
Basin; (2) Cérilly Basin; (3) Coutras; (4) Guérande;
(5) Lodève Basin; (6) Mortagne; (7) Pontivy; (8)
Rodez Basin; (9) Saint-Affrique Basin; (10) Saint
Pierre du Cantal. Map from Romer and Kroner
(2018), spatial distribution of U mineralization
based on IRSN (2007) and Ballouard et al. (2017).

487
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

situation with the extension of the oxidation zone also may be related to highly fractionated rare-metal granites (Stussi, 1989; Le Carlier et al.,
climatic conditions that favor intense chemical alteration of surface 2013; Cuney et al. 2001; Cuney and Barbey, 2014; Chelle-Michou et al.,
sediments and may have occurred in the Saxo-Thuringian Zone in the 2017). Granites of the first two groups formed in an overall compres-
Permian, the Cretaceous, and the Tertiary (Roscher and Schneider, sional setting, whereas granites of the third age group were emplaced in
2006; Schwerdtner et al., 2007). a late-orogenic extensional setting (Roig et al., 2002; Le Carlier et al.,
Isotopic dating of different parageneses in vein mineralization de- 2013). There are two events of late orogenic crustal extension, sepa-
monstrates that the primary mineralization formed around rated by the youngest group of granites. Crustal extension was largely
270–280 Ma, and has been redistributed at 180, 150, and 120 Ma and controlled by major, predominantly ENE-WSW and NW-SE oriented
more recently during the Alpine orogeny (Förster, 1996; Romer et al., fault systems. The older extensional event is diachronous across the
2010). The various younger remobilizations have distinctive mineral French Massif Central. It started at c. 330 Ma in the north and c. 320 Ma
assemblages, (i) quartz-polymetallic-sulfides, quartz-hematite, quartz- in the south and lasted to c. 315 Ma. The younger extensional event
carbonate-uraninite, (ii) polymetallic-sulfides-sliver-antimony, dolo- encompasses the age range 310–290 Ma (Joly et al., 2008) and is lar-
mite-uraninite-fluorite, fluorite-quartz, (iii) quartz-hematite-baryte, gely coeval with the development of volcanosedimentary basins in the
and (iv) BiCoNi-uraninite, and Fe-Mn-quartz (e.g., Büder and Schuppan, Central European Sedimentary Province (Roscher and Schneider,
1992; Förster, 1996; Ondruš et al., 2003a; Hiller and Schuppan, 2008, 2006), which by and large is synonymous to the Central European
2016). The occurrence of Bi, Cu, and Ni, along with native As and Ag in Extensional Province of Kroner and Romer (2013).
late mineral parageneses, but not in early ones, indicates that these Uranium mineralization in the French Massif Central occurs in
metals do not represent deposit internal redistributions, but late addi- episyenites and veins within granites, rarely as veins in the meta-
tions. The closeness to Silurian black shales and slates may indicate that morphic rocks near the contact to granites, and in Permian sedimentary
these rocks were the source of these late metal additions. It is note- rocks (e.g., Poty et al., 1986; Cathelineau, 1986; Pagel et al., 1990;
worthy that late metal additions to the sediment hosted U mineraliza- Cuney, 1990; Patrier et al., 1997; Mathis et al., 1990). In the north-
tions in the Ronneburg and Sorge-Culmitzsch areas also include As, Sb, western part of the Massif Central, in the Saint Sylvestre Massif, isotopic
Cu, Ni, and Mo (Lange and Freyhoff, 1991; Runge, 2001; Bolonin and dating demonstrated that formation of episyenite is distinctly older
Gradovsky, 2012). than the formation of U mineralization (e.g., Leroy and Turpin, 1988;
Turpin et al., 1990; Scaillet et al., 1996). The sources of U for episye-
4.3. Uranium mineralization in the French Massif Central nite, vein, and sediment U mineralization are the granites, and in par-
ticular the uraninite bearing peraluminous granites that formed by
Among the Variscan Massifs, the French Massif Central is particular, partial melting of Early Paleozoic orthogneisses (Turpin et al., 1989) or
as it does not expose rocks of a Cadomian basement, which might be the granulitized metasedimentary lower crust (Williamson et al., 1996).
absent. Instead, the French Massif Central represents a stack of meta- Uranium mineralization of the French Massif Central falls in four sub-
morphic nappes that are dominated by Paleozoic volcanosedimentary areas that reflect the regional distribution of U-fertile granites
rocks originally deposited on the Gondwana shelf. The metamorphic (Cathelineau et al., 1990; Cuney et al., 1990; Bril et al., 1994). Typi-
nappes are grouped into three major units (Fig. 8; Ledru et al., 1989; cally, U mineralization in the Saint Sylvestre granitic complex is bound
Faure et al., 2009): the Upper Gneiss Unit, the Lower Gneiss Unit, and to magmatic shear zones that have been intruded by highly fractionated
the Para-Autochtonous Unit. The major difference among these three leucogranites and that have acted as pathway for hydrothermal fluids
units is the contrasting metamorphic peak conditions: (i) the Upper during reactivation of these structures under brittle conditions. During
Gneiss Unit to the north is dominated by a bimodal association of a phase of uplift associated to crustal extension, meteoric fluids that had
mafic/ultramafic rocks with felsic rocks, migmatitic ortho- and para- infiltrated into the basement, were heated, migrated upward, and
gneisses (the youngest detrital zircon of the metasediments has a U-Pb eventually underwent boiling and phase separation. The condensation
age of 523 ± 4 Ma; Melleton et al., 2010), eclogites, and high-pressure of the vapour phase resulted in silica undersaturated fluids that dis-
granulites (Lardeaux et al., 2001), mafic Early Ordovician (ca. 480 Ma) solved selectively quartz from the granites leading to the formation of
magmatism and Late Silurian (ca. 420–410 Ma) high-pressure meta- porous granite bodies: the episyenites (Pécher et al., 1985; El Jarray
morphism (Ducrot et al., 1983); (ii) the Lower Gneiss Unit has medium- et al., 1994). The leucogranites represent the local U sources and the
temperature / medium pressure ortho- and para-gneisses (the youngest episyenites with their high porosity eventually acted as traps for U
detrital zircon has an age of 573 ± 12 Ma; Melleton et al., 2010), some mineralization in areas that have experienced brittle deformation (e.g.,
of these enriched granitic orthogneisses have up to 20 ppm U (Cuney Friedrich et al., 1987; Cuney et al., 1990; Cuney, 1990). It is important
et al., 1990); and (iii) the Para-Autochthonous Unit is dominated by to note that granite emplacement (including the leucogranites), for-
metasedimentary rocks that were intruded by Ordovician plutons and mation of episyenite, and the U mineralization are three events sepa-
generally experienced Variscan low-grade metamorphism. The rated by several tens of million years. Whereas episyenites formed in
youngest detrital zircon from the metasedimentary series has an age of the late Carboniferous (c. 305 Ma; Scaillet et al., 1996), the U miner-
604 ± 16 Ma (Melleton et al., 2010). These main tectonic units of the alization formed by low-temperature hydrothermal fluids in the Per-
French Massif Central are overthrusted to the ENE and WSW by al- mian, essentially coeval with a first U enrichment in the Permian se-
lochthonous low-grade units and are thrust onto non-metamorphic dimentary basins (e.g., Cathelineau, 1986; Leroy and Turpin, 1988;
rocks of the external zone with its fold-and-thrust belt and its foreland Turpin et al., 1990; Cathelineau et al., 1990, 2004).
basin to the south (Ledru et al., 1989; Faure et al., 2009). There are Uranium mineralization is known for a number of Permian (e.g.,
three age groups of Variscan granitoids that intrude the rocks of the Lodève, Cérilly, Rodez, Saint-Affrique; e.g., Lancelot et al., 1984; Pagel,
metamorphic nappes of the French Massif Central: (i) ca. 360–350 Ma 1990; Mathis et al., 1990; Marignac and Cuney, 1999) and Tertiary
peraluminous cordierite-biotite granites and low to highly potassic calc- (e.g., Coutras, Saint Pierre du Cantal; Meunier et al., 1989; Marignac
alkaline gabbro-dioritic to granitic rocks; (ii) ca. 340–325 Ma mainly and Cuney, 1999) sedimentary basins. Most of these basins are bound
two–mica peraluminous leucogranites in the north-western part of the by faults that were active during the development of the Central Eur-
French Massif Central, whereas some cordierite-biotite peraluminous opean Extensional Province and later tectonic reactivation of the same
granites such as the Margeride complex, high-K calc-alkaline granites, structures (e.g., Lévêque et al., 1988; Pagel, 1990). Uranium was lea-
and subordinate low-potassic calc-alkaline granitic rocks occur in the ched by oxidizing meteoric water in the subsurface and became en-
eastern part of the FMC (Chelle-Michou et al., 2017); and (iii) ca. trapped in reduced units within the continental sedimentary rocks of
310 Ma predominantly peraluminous granites, such as the cordierite – the various basins (Pagel, 1990; Marignac and Cuney, 1999). The
biotite granites of the Velay dome (SE Massif Central) that also include basement, in particular its Carboniferous peraluminous leucogranites,

488
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

may be not only the source of U for these sediments. Instead, intense granites (Ballouard et al., 2018). Uranium mineralization is mainly
alteration of rocks at or near the surface also may have resulted in re- restricted to Carboniferous peraluminous leucogranites that are related
distribution of U in vein-type mineralization due to dissolution of mi- to the various branches of the South Armorican Shear Zone and to a
nerals in the oxidized zone and redeposition of U in the reduced zone of lesser extent to Mg-K biotite-hornblende granites along the North Ar-
the same vein (e.g., Cuney, 1978; Schmitt and Thiry, 1987, Condomines morican Shear Zone (Fig. 8; Chauris, 1984; Rolin et al., 2009; Ballouard
et al., 2007), and the entrapment of U mobilized in the exogenic en- et al., 2018). Uranium mineralization seems to be restricted to those
vironment along fractures reaching deep into the basement (Barbier, granites that are derived from a U-enriched source by partial melting
1974). (e.g., Vigneresse et al., 1989; Cuney, 2014). Typically, U deposits re-
lated to the peraluminous two–mica leucogranites forms quartz-ur-
aninite veins within the granites, but locally also at and near the con-
4.4. Uranium mineralization in the Armorican Massif tacts of the granites to Ordovician black shales (Ballouard et al., 2017).
The quartz-uraninite veins show coffinite at depth and secondary U
The Armorican Massif is subdivided into three segments of con- minerals, such as autunite and torbernite, near the surface where oxi-
trasting Variscan history by the roughly E-W striking North Armorican dized meteoric water variably altered the original mineralization and
Shear Zone and the NW-SE striking South Armorican Shear Zone redistributed U (e.g., Chauris, 1977; Ballouard et al., 2018). Vein-type
(Fig. 8). The northern segment, the North Armorican Domain, did not U mineralization is related to deep fluid circulation during the Permian
experience significant Variscan metamorphism. It consists of rocks of extensional event that affected entire Western Europe. The per-
the deeply eroded Cadomian magmatic arc overlain by Cambrian and aluminous two–mica leucogranites are characterized by magmatic en-
Ordovician sandstones. The oldest sediments deposited during the post- richment of U to levels of several tens of ppm, which is mostly hosted in
Hirnantian transgression are Silurian pyritic quartzites and black shales uraninite (Cuney, 2014). Oxidized fluids leaching the granite mobilize
that have 5–15 (35) wt% organic material and enhanced contents of Ni and redistribute U to form quartz-uraninite veins (Ballouard et al.,
(up to 130 ppm) and V (up to 5000 ppm; Dabard and Paris, 1986). 2017, 2018). This low-temperature hydrothermal redistribution oc-
Ordovician and Silurian black shales have the highest U contents curred between 300 and 275 Ma (Ballouard et al., 2018). Uranium
(Cuney et al., 1990). The Silurian sediments are overlain by Devonian mineralization is the result of a polystage process involving magmatic U
platform-type sediments with siltstones, sandstones, and limestones enrichment and U sequestration in minerals that later were easily lea-
(Robardet et al., 1994). The Paleozoic sedimentary cover in the central ched, followed by shearing and late fracturing of the contact zone with
segment, the Central Armorican Domain, has been only moderately deposition of uraninite in quartz veins together with Fe-sulfides
affected by Variscan deformation and metamorphism in the east and (Chauris, 1984). Crystallization of the granite and formation of the
has been stacked by thin-skin tectonics to packages reaching several quartz veins represent two separated processes that may have operated
thousand meters thickness in the west. The southern segment, the South in different tectonic regimes (Chauris, 1984).
Armorican Domain, includes three major tectono-metamorphic units
dominated by Paleozoic sedimentary and magmatic rocks of contrasting
peak metamorphic conditions (e.g., Pitra et al., 2010; Ballouard et al., 4.5. Uranium mineralization in Iberia
2017) that have been intruded by late- and post-kinematic Mg-K biotite-
hornblende calc-alkaline and peraluminous two–mica granites and The pre-Permian basement of Iberia includes five zones (Fig. 9) of
subordinate cordierite-biotite granies between 320 and 300 Ma contrasting pre-Variscan stratigraphic record and contrasting Variscan
(Ballouard et al., 2017). structural and metamorphic inventory. Significant U mineralization
Rocks with anomalous U contents include Silurian sedimentary only occurs in the Central Iberian and Ossa Morena zones (Fig. 9),
rocks, in particular the black shales, and post-kinematic Variscan which include rocks of Gondwana provenance (Noblet and LeFort,

Fig. 9. Simplified map of Iberia, showing the five


structural zones (Cantabrian Zone, West Asturian –
Leonese Zone, Central Iberian Zone, Central Iberian
Zone, Ossa Morena Zone, and South Portuguese
Zone) of contrasting Variscan deformation and me-
tamorphic overprint. The Central Iberian Zone is
commonly distinguished into an “allochthonous”
(i.e., Galicia – Tras-os-Montes Zone) and an “au-
thochtonous” domain. Note, most U mineralization
is hosted in the “authochtonous” domain of the
Central Iberian Zone. The Rheic suture separates the
Ossa Morena Zone from the South Portuguese Zone.
P = Pedroches Batholith; B = Burguillos Pluton;
CB = Calaf Basin; MB = Mequinenza Basin;
S = Salamanca. Numbers refer to U mineralization
mentioned in the text: (1) Mina Fe; (2) Retortillo-
Santidad. Map from Romer and Kroner (2018),
spatial distribution of U mineralization based on
Dahlkamp (2016).

489
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

1990; Robardet, 2002; Abalos et al., 2002) and were part of the Ar- et al., 2015; Pérez del Villar et al., 1991), and the Ossa Morena
morican Spur sensu Kroner and Romer (2013). The Central Iberian Zone Zone (Burguillos pluton, Pedroches batholith; Locutura et al.,
is dominated by Ordovician to Carboniferous sedimentary rocks that 1990). Typically U mineralization forms veins and skarns within
commonly are distinguished into Autochthonous and Allochthonous the granites or the sedimentary wall rocks with variable mineral
sequences (Galicia and Trás-os-Montes area) on basis of their con- assemblages including pitchblende, Fe-sulfides, quartz, carbo-
trasting Variscan metamorphic overprint and deformation style. The nates, and/or REE minerals, as well as As, Ni, Co, or Bi minerals
domain of the Autochthonous sequences is characterized by a pre-Or- (e.g., Locutura et al., 1990; Pérez del Villar et al., 1991; Marcuello
dovician metamorphic basement, which is discordantly overlain by et al., 2006).
Ordovician fluviatile, tidal, and shallow marine sandstones and quart- (III) Late-Variscan NW-SE striking shear zones locally induced im-
zites and younger sedimentary rocks. Zircon provenance analysis of portant loss of quartz in granites. This type of U mineralization is
Ordovician sandstones indicates that the Central Iberian Zone had a found in the Central Iberian Zone. The resulting episyenites have a
more eastward position in peri-Gondwana than the other blocks of the high permeability and typically were mineralized during later
Armorican Spur (Stephan et al., in press), whereas the enhanced U events with W, U, and Au (López-Moro et al., 2013). Geochemi-
contents of Silurian black shales indicates that the Central Iberian Zone cally and texturally different types of uraninite commonly are as-
still was part of the extended Gondwana shelf at that time. There are sociated with the various mineralization stages, typically yielding
several groups of Variscan granites, starting with voluminous chemical ages ranging from c. 300 to c. 275 Ma (Timón-Sánchez
325–320 Ma migmatic granites and ending with 310–300 Ma per- et al., in press), implying most importantly that episyenite for-
aluminous leucogranites and vaugnerites. (Martínez et al., 1990; Ugidos mation post-dates granite emplacement and that there are several
et al., 2003; López-Moro et al., 2013, 2017). Minor granites and – possibly unrelated – stages of mineralization.
300–270 Ma old episyenites along late-Variscan strike-slip faults show
polyphase W-U-Au mineralization (e.g., Timón-Sánchez et al., in press). 4.6. Uranium mineralization in Morocco, Nova Scotia, and Cornwall
The Ossa Morena Zone (Fig. 9) was in a passive margin setting during
the Ordovician to Devonian and was variably segmented along Variscan Morocco was part of mainland Gondwana during the Variscan or-
NW-SE oriented transpressional fault zones that control both the em- ogeny and has a similar geological history as those Gondwana frag-
placement of several groups of 340–280 Ma structurally bound granites ments that formed the Armorican Spur and remained attached to
(Tornos et al., 2002). Uranium mineralization in the Ossa Morena Zone Gondwana during the opening of the Paleotethys. Silurian hot shales
is spatially related to these Variscan granites. (e.g., Lüning et al., 2000b; Le Heron et al., 2008) deposited on the
There are three groups of U mineralization, (i) mineralization bound Armorican Spur were overprinted by the Variscan orogeny, possibly
to sedimentary rocks rich in organic material, (ii) mineralization in resulting in the redistribution of U hosted in these black shales by
veins bound to granite intrusions, and (iii) mineralization related to Variscan granites and metamorphic fluids. Uranium mineralization re-
episyenite within granite intrusions. There is rare U mineralization in lated with the Variscan orogenic overprint, however, is only minor (Idir
Permian to Triassic sandstones. Uranium may have been leached from and Renard, 2002). Uranium mineralization is known from Jurassic
Variscan granites, some of which contain uraninite, secondary U- (Hauterivian) plant-rich pelitic sandstones in Jebbel Wafagga (Idir and
phosphates, and allanite as major U hosts (e.g., Mohamud et al., 2015), Renard, 2002), whereas there seems to be no U mineralization related
after Cretaceous rifting (Marfil et al., 2013). Oligocene lignites, which to the U-rich sedimentary rocks in the Cretaceous Timahdit and Oulad
formed in shallow marsh-swamp-lake systems in the Calaf and Mequi- Abdoun basins (e.g., Benalioulhaj and Trichet, 1990). In the Timahdit
nenza basins, are characterized by distinctly enhanced Mo and U con- basin, U is mainly related to black shales, whereas in the Oulad Abdoun
tents in comparison with comparable lignites from other areas (e.g., basin, U is related to phosphorites (Benalioulhaj and Trichet, 1990) and
Querol et al., 1996). may be recovered from them as byproduct of fertilizer production (e.g.,
Owen, 1992; Cuney, 2010).
(I) Uranium mineralization bound to sedimentary rocks rich in or- Basement areas that have separated from Gondwana during the
ganic material are known from the Central Iberian Zone. Cambrian opening of the Rheic Ocean, i.e., may show late Cambrian and possibly
black shales (Mina Fe, Alameda, Fuented de Oñoro; Both et al., early Ordovician sedimentary rocks of Gondwana provenance, were not
1994; Martín-Izard et al., 2002; Pérez del Villar et al., 2002) and covered by the Hirnantian ice sheet and, thus, did not have near-surface
late Ordovician black shales (Retortillo-Santidad) show significant rocks that have not been depleted in U before. Therefore, these areas
U mineralization that typically forms in breccias, veins, and frac- did not represent a source of leachable U in the early Silurian when
tures in the host rocks. The mineral association and truncation extensive black shales could have trapped and concentrated U leached
relations among the veins indicate that pitchblende and uraninite from the near-surface basement. Black shales on Avalonia (Nova Scotia,
as main U minerals had been repeatedly redistributed. Deep per- British Isles) are likely to have lower U contents than those deposited
colating meteoric water leach the sediments and mobilize U in areas that remained with mainland Gondwana. Typical U contents of
particular from the black shales and nearby peraluminous leuco- such black shales may be less than 10 ppm (e.g., Wignall and Myers,
granites and eventually precipitate U in veins within the black 1988; Lu et al., 2017). Therefore, U deposits on Avalonia are rare and
shales (e.g., Mina Fe, Retortillo-Santidad; Both et al., 1994; Martin- relatively small, as for instance roll-front type mineralization in the
Izard et al., 2002; Huertas et al., 2015). Multiple redistribution of Carboniferous sedimentary rocks of the Horton Group (Nova Scotia)
U during the Variscan and Alpine orogeny, which in part is re- and vein-type mineralization associated with the Permian granites of
flected in apparently conflicting ages derived from field relations Cornwall.
and U-Pb dating (Both et al., 1994; Martin-Izard et al., 2002). Most On Nova Scotia, U mineralization is documented from
of the various mineralization show important near-surface redis- Carboniferous sandstones of the Horton Group that seem to have con-
tribution of U by leaching of U in the oxidation zone and re- centrated U derived from deeply weathered Devonian granites in roll-
precipitation in reducing deeper parts of the shales (Pérez del front type deposits, either associated to carbon-pyrite-rich sandstone of
Villar et al., 2002, Huertas et al., 2013, 2015). The lack of deep the Horton Bluff Formation or to hematized arkosic sandstone of the
bore holes or mines reaching into non-remobilized mineralization Cheverie Formation (Ryan and O'Beirne-Ryan, 2009). Sedimentary
prevented the determination of the age of the original U deposi- rocks of the Horton group extend into New Brunswick. There, U
tion. showings are known from a belt that has been overprinted by the
(II) Granite related U mineralization occur in the Central Iberian Zone, Acadian orogeny and hosts Devonian granites (Hassan et al., 1987).
which hosts a large number of fertile granites (e.g., Mohamud Common to these areas with U mineralization or U showings is that

490
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

they have Ordovician to Devonian clastic sedimentary rocks with in- alteration resulted in sedimentary rocks characterized by a very high
tercalations of black shale. There is U mineralization (mainly the Millet Chemical Index of Alteration (CIA) and very low contents of Ca, Na, Sr,
Brook deposit) associated with Devonian-Carboniferous two–mica and Pb (these elements are typically feldspar-bound) and redox-sensi-
peraluminous granites, in particular the 370 Ma South Mountain tive elements, such as Mo, V, and U. During the early Ordovician
Batholith where episyenites have been recognized (Kontak and opening of the Rheic Ocean, the Cambrian sedimentary rocks were
Cormier, 1991; Chatterjee et al., 1982). The main U-bearing minerals in redistributed and their geochemical signature was inherited by younger
the veins are pitchblende and torbenite at depth and autunite and Pb- sedimentary rocks. Because of the intense chemical alteration of the
meta-autunite in the near-surface (Chatterjee et al., 1982). These Gondwana surface, pre-Hirnantian near-surface rocks do not represent
granites served also as U sources for later U redistribution. a likely U source for U mineralization within the Variscan belt of
In Cornwall, U mineralization typically is closely related to the Europe.
Cornish granites, well known for their Sn mineralization. The richest U The Hirnantian glaciation (444.7–443.4 Ma; Brenchley et al., 1994;
mineralization includes structurally late N-S striking veins at South Rohrssen et al. 2013) removed deeply altered basement and sedimen-
Terras Mine (LeBoutillier, 2002) hosted in grey and brown slate of the tary rocks from large parts of Gondwana, leaving behind a surface with
Lower Devonian Meadfoot Group to the south of the metamorphic rocks that had not yet been exposed to chemical alteration. These newly
aureole of the St. Austell granite (Smale, 1993). Main U minerals are exposed rocks at the post-Hirnantian surface had not yet lost their U
uraninite, pitchblende, and coffinite. Isotopic dating yields a wide range due to surficial alteration processes and, therefore, represent a much
of ages from c. 280 Ma to c. 50 Ma, indicating that there was not one more fertile U source than pre-Hirnanatian surface rocks. Chemical
single event of mineralization, but there were several events of mi- alteration may mobilize U by oxidation from these rocks and transport
neralization or redistribution (Darnley et al., 1965). The oldest ur- U toward the oceans. Major source minerals for U mobilizations may be
aninite typically is associated with Sn and shows high Th contents old high-U zircon, allanite, and U-thorite that have acquired over time
(Darnley et al., 1965). This uraninite may be related to the emplace- significant α-recoil damage, which facilitates the liberation of U (e.g.,
ment of the Cornish granites and the formation of Sn mineralization. Poitrasson, 2002; Geisler et al., 2003; Romer, 2003). Furthermore,
Uraninite of the second age group (c. 220–c. 160 Ma) is associated with some U-rich granites may have a significant portion of their U hosted in
sulfide mineralization of Co, Ni, Cu, Fe, As, Bi, and/or Ag. Uraninite of uraninite (e.g., Cuney, 2014). Whether the post-Hirnantian land surface
the youngest group yields ages around 50 Ma (Darnley et al., 1965; represented a major source of U does not only depend on the nature of
LeBoutillier, 2002). the minerals that host U, but resulted from the fact that these rocks had
not been leached before and readily mobilized U still was available for
5. Formation of the source rocks for U mineralization: pre- leaching and redistribution. Note, sediments deposited by the Hirnan-
Variscan U-enrichment in Paleozoic sediments tian ice sheet are dominated by sands and show significant regional
facies changes (Le Heron et al., 2009). Because of the absence of vas-
Paleozoic rocks that have been reworked during the Variscan or- cular plants in the Ordovician and early Silurian, these sedimentary
ogeny and that may be source rocks for the high U-granites of the rocks are unlikely to host redox-controlled U deposits, whereas calcrete-
Variscan belt include remains of the Cadomian magmatic arc (e.g., related U mineralization may have formed locally in closed basins.
Linnemann et al., 2008), Lower Paleozoic orthogneisses, and latest Thus, the sedimentary cover on Gondwana generally did not act as trap
Precambrian to Silurian (and possibly Devonian) sedimentary rocks. for U when the Hirantian ice sheet melted. Instead, U leached from the
Rocks of the Cadomian magmatic arc have low U contents (Linnemann newly exposed rocks was dominantly transported by surface runoff to
and Romer, 2002; Linnemann et al., 2004) as have most Paleozoic se- the sea.
dimentary rocks, in contrast to U-rich Lower Paleozoic orthogneisses On Gondwana, there is a close spatial and temporal relation be-
and organic-rich shales (Fig. 7 and Romer and Hahne, 2010; Romer tween sources and traps in the Silurian: (i) Melting of the Hirnantian ice
et al., 2014a). The latter have been deposited under anoxic conditions, sheet resulted in a global rise of sea level and transgression on the shelf
which resulted in the enrichment of redox-sensitive elements like U, Ni, area. Paleotopography on the shelf resulted in the development of ba-
Co, Mo, and V (Fig. 7). There is no direct correlation between organic sins with restricted circulation. These basins, which correspond to
content and U content (Lüning and Kolonic, 2003). Instead, the low- paleo-depressions such as former glacial outwash valleys and structural
ermost units of Silurian black shales have considerably higher U con- intrashelf basins (Lüning et al., 2000a; Le Heron et al., 2008, 2009),
tents than similar black shales in stratigraphically higher positions. developed anoxic conditions that reflect the effects of high organic
Anomalously high U content in the lowermost Silurian black shales is production and restricted circulation, which may be favored by strati-
not restricted to the Schwarzburg profile (Fig. 7), but is also well known fication of the water masses due to freshwater influx from the melting
from black shales of northern Gondwana and the Armorican Spur, ice sheet (Lüning et al., 2000a). Continued rise of the sea level moved
Central Asia, and southern China (Melchin et al., 2013; Fig. 4). The high the coastline inland, which implies that coarse detritus was trapped in
organic content of these shales makes them to very important source retrograding river mouths, resulting in less dilution of the black shales
rocks for the formation of oil and gas reservoirs in northern Africa, by coarse material and eventually in better water-circulation on the
Arabia, and Iran (e.g., Lüning et al., 2000a; Le Heron et al., 2009). shelf, which in turn resulted in less anoxic conditions and in lower
Therefore, these shales have been extensively drilled. Gamma-logs de- organic contents ins stratigraphically higher shales (Lüning et al.,
monstrate that the lowermost Silurian black shales of northern Africa, 2000a). Thus, anoxic conditions to trap dissolved U on the shelf per-
Arabia, and Iran commonly have markedly higher U contents than sisted only for a relatively short period (Lüning et al., 2000a). (ii)
stratigraphically higher black shales (e.g., Lüning et al., 2000b; Al- Leachable U source rocks become available on Gondwana as the Hir-
Juboury and Al-Hadidy, 2009; Yan et al., 2015; Saberi et al., 2016). antian ice sheet melted (melting of the ice sheets does not affect the
The formation of U-rich early Silurian shales requires a U source and availability of leachable U on the continents that have not been gla-
an efficient, spatially restricted trap. In the following, we argue that ciated). The melt water represents both the leaching and the transport
earlier glaciated Gondwana is the U source and the extended shelf is the agent. Leaching of U will deplete the source rocks and eventually less U
trap. will be available for leaching and redistribution, explaining why stra-
Large parts of Gondwana were extensively eroded after the forma- tigraphically higher black shales have lower U contents than the stra-
tion of the 620–600 Ma old West Gondwana Orogen and the tigraphically lowermost black shales deposited during the post-glacial
520–500 Ma old East Gondwana Orogen (Avigad et al., 2005; transgression. Uranium-rich black shales may develop only on the shelf
Hauzenberger et al., 2004; Hawkesworth et al., 2013; Ganade de Araujo over a short period, in part as the U source rocks become readily de-
et al., 2014). In particular during the Cambrian, extensive chemical pleted in leachable U (unless there is active tectonics making fresh

491
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

rocks available for leaching), in part as U delivered from the continent shelfs that had been transgressed during the melting of the Hirnantian
will not be added to the U budget of the ocean, but is efficiently pre- ice sheet, but only those that received U from the drainage of previously
cipitated on the shelf (in analogy with Mo, cf. McManus et al., 2006; glaciated areas are likely to be particularly U-rich. For instance in South
Gilleaudeau and Kah, 2013). Black shales deposited on the outer shelf Africa, shales deposited on the diamictites of the Table Mountain Group
and in a deep anoxic ocean will not develop particularly high contents
of U, as U coming from the continent is precipitated in black shales on
the anoxic shelf and little U remained in solution to reach the open
ocean. (iii) The development of anoxic conditions alone is not sufficient
for black shales to develop high U contents as there is no one-to-one
relation between U-rich shales and glaciations: the source of U must not
have been depleted by U leaching in the time between the generation of
a fresh surface and the development of anoxic sedimentation environ-
ment on the shelf. Otherwise, the U source may have become depleted
before trapping conditions were established. Furthermore, there are
also U-anomalous black shales without preceding large-scale glaciation
in the drainage area that represents the U source area (e.g., Sun et al.,
2005), which implies that major glaciation does not represent the only
mechanism to make U sources available for U leaching and redis-
tribution. Thus, for the development of U-rich black shales, the avail-
ability of leachable U source rocks seems to be the restricting require-
ment.
The Silurian black shales of Gondwana, in particular the strati-
graphically lower black shale units, are characterized by high U-con-
tents (e.g., Fig. 7). These U-rich shales formed on the inundated shelves
of the formerly glaciated continent. Silurian shales of Northern Africa,
Arabia, and Iran have not been involved in later orogenic events and
represent now a major source lithology for the oil occurrences of these
regions (e.g., Lüning et al., 2000b; Al-Juboury and Al-Hadidy, 2009; Le
Heron et al., 2009). In contrast, Silurian black shales of Iberia, Ar-
morica, the French Massif Central and various tectonic elements of the
Bohemian Massif have been involved in the Variscan orogeny and have
experienced regionally contrasting Variscan metamorphism. These
rocks do not host hydrocarbon resources. Instead, these Silurian shales
show close spatial association with U mineralization. For instance,
major U deposits in the Ronneburg area are largely hosted in Silurian
black shales. Within the Variscan Massifs, Silurian black shales host
vein-type U mineralization and the sedimentary rocks that melted to
form the Variscan granites also included Silurian strata. At some loca-
tions, the relation of U mineralization to Silurian black shales of the
former Gondwana shelf may be indirect. For instance, post-Variscan
sediment-hosted occurrences (Königstein, Hamr, Osečná, Stráž in Cre-
taceous sandstones; Döhlen, Stockheim, Lodève, and Aumance basins
with Carboniferous to Permian continental sandstones and coal seams)
derive their U from altered granites (e.g, Meunier et al., 1989;
Tonndorf, 2000; Reichel and Schauer, 2006; Cuney et al., 2014). These
granites, however, may derive part of their U budget from Silurian
Fig. 10. Spatial relation between glaciations with major ice sheets on
rocks that melted in the source of the granites.
Gondwana, black shales deposited during the deglaciation, and post-glacial U
Note, the presence of Silurian black shales seems not to be the sole mineralization. (a) Hirnantian glaciation; the maximum extent of the ice sheet
requirement for later U mineralization. Even though there are Silurian according to Le Heron et al. (2009), distribution of Silurian hot shales from
black shales on Avalonia (e.g., Verniers et al., 2008), which had not Lüning et al. (2000a,b) and Le Heron et al. (2009), distribution of Silurian black
been covered by the Hirnantian ice sheet, U mineralization on Avalonia shales (and their metamorphic equivalents) on fragments of the Armorican Spur
is small and rare, possibly indicating that (i) black shales deposited on and Avalonia according to Dabard and Paris (1986), Wignall and Myers (1988),
the shelf (and in particular in the coastal zone) scavenge U from the Romer and Hahne (2010), Trela et al. (2016), and Lu et al. (2017). (b) Karoo
continental runoff, therefore, are more fertile than deep-water black glaciation; the extent of the Karoo glaciation, which includes at least three
shales and (ii) large-scale glaciations in the drainage area make U stages is disputed. The extent of the ice sheet shown here follows Isbell et al.
(2003), distribution of post-glacial Permian bituminous shales and coal deposits
sources available for leaching by removing previously leached rocks
according to Cairncross (2001), Chakraborty et al. (2003), Catuneanu et al.
and bringing rocks to the surface that have not been depleted in U
(2005), and Veevers et al. (2006). Spatial distribution of U mineralization
earlier. schematically according to Toens et al., (1980), Toens and Le Roux (1981),
A link between glaciation and U-rich black shales may not be re- Kreuzer et al. (2010), and Penney (2012). Numbers refer to Permian sedi-
stricted to Silurian rocks of northern Gondwana, but may also be pre- mentary basins mentioned in the text: (1) Botswana Basin; (2) Callingasta-Us-
sent in other parts of Gondwana and apply to glaciations of different pallata Basin; (3) Canning Basin; (4) Carnarvon Basin; (5) Collie Basin; (6) East-
age, such as the Quaternary, Karoo (Permian), and Andean-Saharan Africa Basin; (7) Eromanga Basin; (8) Mid-Zambesis Basin; (9) Parana-Karoo
(Hirnantian) glaciations, but possibly also the Sturtian-Varangian Basin; (10) Satpura Gondwana Basin; (11) Springbokflats Basin. Note, Europe
(Cryogenian, late Neoproterozoic) and Huronian (early was not affected by the Karoo glaciation (with the possible exception of topo-
Paleoproterozoic) glaciations. For instance, the Hirnantian glaciation graphically high areas). The source for Late Carboniferous to Permian sediment-
covered part of Gondwana (Fig. 10). Silurian black shales occur on all hosted U mineralization is not related to deglaciation, but to tectonic processes.

492
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

have higher U contents than shales deposited before the diamictites importance of leachable U sources.
(e.g., Young et al., 2004). The unusual chemical composition of these
shales is attributed to a reducing environment during rapid sea level 6. Uranium mobilization during the Variscan orogeny
rise that accompanied the melting of the late Ordovician glaciers
(Young et al., 2004). Partial melting of continental crust is the dominant process of U
An example for the interplay of sea level rise and a stratified water redistribution during the Variscan orogeny. The major importance of
column in generating anoxic conditions is the Black Sea. Sea level rise crustal melting is not the formation of U mineralization, as there is no
resulted in the flooding of a former sweet water lake by sea water in- granite-bound magmatic U mineralization in Variscan Europe, but the
vading through the Dardanelles and the Bosporus (e.g., Ryan et al., transport of U from the lower crust to the middle and upper crust and to
2003; Lericolais et al., 2011), which resulted in a stratified water body make U available for leaching and further redistribution by tectonic and
in which deeper parts became anoxic during initial oxidation of organic exogenic processes. The spatial association of structurally controlled U
material. Freshwater discharge supports this stratification and supplies mineralization with granites, either in the immediate wallrocks or
redox sensitive elements. Sediments deposited in the Black Sea and in within the granites, is not related to the cooling of the granites and does
an area with anoxic upwelling water show broadly similar trace metal not imply that this type of mineralization is related to granitic mag-
contents, which implies that the trace metal content is largely domi- matism, in particular, as isotopic dating reveals that U mineralization
nated by the availability of the material in the water column and, thus, commonly is 20–50 Ma younger than the emplacement of the granites
the input from surface runoff (e.g., Brumsack, 2006). Uranium contents (e.g., Turpin et al., 1990; Scaillet et al., 1996; Kříbek et al., 1999;
in black shales of the Black Sea reach 15–20 ppm (Rolison et al., 2017). Cathelineau et al., 2004; Hiller and Schuppan, 2008, 2016; Škácha
The highest authigenic U contents occur in abyssal sediments and the et al., 2009; Ballouard et al., 2018). Therefore, structurally controlled
lowest ones in sediments of the outer shelf (Anderson and Fleisher, mineralization it treated here as post-Variscan tectonic mineralization.
1991). Sediments of the inner shelf have intermediate U contents. Au- There are three types of Variscan granites that differ in age, tectonic
thigenic U is inversely related to sediment accumulation rate (Anderson setting, and heat source, (i) syn-kinematic granites emplaced along
and Fleisher, 1991). crustal-scale structures at the time of thrusting of Variscan meta-
The Permian glaciation covered a large area of southern Africa and morphic nappes, (ii) post-kinematic granites that represent crustal
South America. Tillites left behind after this glaciation are covered by melts that largely form by dehydration melting, and (iii) post-kinematic
fluvial and lacustrine sandy deposits with regionally widespread and granites and related rocks of the Central European Extensional Province
locally significant U mineralization. The most important of these post- (e.g., Förster and Romer, 2010, Romer et al., 2014b; Couzinié et al.,
glacial sedimentary basins is the Parana-Karoo basin of South Africa, 2014; Laurent et al., 2017). The syn-kinematic granites are regionally
Angola, and Brazil. Smaller basins in similar position are known from diachronous as different crustal blocks within the Armorican Spur
Argentina, Zambia, Tanzania, and Kenya (e.g., Callingasta-Uspallata, collided at different times. They include mantle-derived melts, such as
Botswana, Mid-Zambezi, and East African basins; Toens et al., 1980; durbachites, shoshonites, and monzonites (e.g., Wenzel et al., 1997;
Toens and Le Roux, 1981). In these post-glacial sedimentary rocks, U Janoušek and Holub, 2007), and melts that formed in the roof of UHT
mineralization is present in high and low grade coal seams and black metamorphic units during their emplacement (e.g., Förster and Romer,
shales (Toens et al., 1980). Note, within these basins, there is also U 2010). The post-kinematic anatectic melts form in stacks of high-grade
mineralization in paleochannels in sandstones and delta fans that are metamorphic rocks, post-dating nappe emplacement by some
not necessarily related to glaciation and also in higher strata (Toens 10–30 m.y. as in the Erzgebirge, the French Massif Central, or the
et al., 1980) that do not show a relation with glaciation. Uranium of Central Iberian Massif (Förster and Romer, 2010; Couzinié et al., 2014;
these settings and strata may represent redistributed material from the Pascual et al., 2013; Denèle et al., 2014; Laurent et al., 2017). Melting
older U-rich strata that became available for redistribution by tectonic may have been induced by emplacement of UHT metamorphic rocks at
processes leading to uplift and increased erosion. In South Africa, the depth or CO2-influx into the lower crust (e.g., Cuney and Barbey, 2014;
richest U-bearing sediments are Late Permian carbonaceous shale and Romer and Kroner, 2016). Post-kinematic magmatism in the Central
coal measures of the Hammanskraal Formation within the Springbok European Extensional Province is the result of the final closure of the
Flats Basin, with 82,000 t U at about 400 ppm (Coles, 1998). The soure Rheic between Laurussia and Africa along the Appalachian-Mauritanide
is the U-rich red granite at the top of the Bushveld complex. belt (Kroner et al., 2016) that results in mantle upwelling with crustal
Direct evidence for early Paleozoic glaciations on the Baltic Shield is dehydration melting or decompression melting of metasomatized
lacking as glacial sands and scratched pebbles that occur locally (e.g., mantle. Granites of Cornwall and the Central Iberian Zone (e.g., Müller
Kulling, 1964; Lindström, 1972) also might reflect high altitude. Fur- et al., 2006; Carvalho et al., 2012), widespread volcanism in Permian
thermore, direct evidence for early Paleozoic glaciation may have been basins (e.g., Timmerman, 2004; Breitkreuz and Mock, 2004), and local
destroyed as Cambrian and Ordovician rocks have been reworked vaugnerite plugs (Couzinié et al., 2014; López-Moro et al., 2017) are
during the Caledonian orogeny or have been eroded from the shield different expressions of the same large-scale tectonic process. These
areas and are only locally preserved (e.g., Lidmar-Bergström, 1996; granites differ in source lithology and melting conditions, but also
Lidmar-Bergström and Näslund, 2002). In the Cambrian, Baltica was at chemical composition and U content. The chemical composition con-
temperate latitudes, but was affected by permafrost as indicated by ice- trols the crystallization behavior of the melt, in particular whether U
wedge casts, tapering fissures, and collapse structures (Cherns and becomes enriched in late melts or is lost to early crystallizing minerals
Wheeley, 2009). These sedimentary rocks with permafrost indications that eventually are dispersed in the magmatic rocks (cf. Cuney, 2014;
are overlain by Alum Shales that are preserved in sediments overthrust Cuney and Kyser, 2015; Ballouard et al., 2018). The granite composi-
by the Caledonian nappes and local remains of the former Phanerozoic tion also has a fundamental control on the availability of U for future
sediment cover of the Baltic shield (e.g., Andersson et al., 1982). The leaching, as the U host determines whether U is readily available for
Scandinavian Alum Shales are characterized by enhanced contents of U, redistribution (e.g., U bound in uraninite) or becomes available only
Mo, and V (Andersson et al., 1982; Berry et al., 1986; Schovsbo, 2003). after a significant period to time due to increased metamictization of
Uranium contents vary considerably, being higher shorewards the U-host (e.g., allanite; Cuney, 2014; Cuney and Kyser, 2015;
(Schovsbo, 2002). They may locally reach several hundred ppm Ballouard et al., 2018).
(Andersson et al., 1982). High U contents, which are characteristic for The granite type controls the nature of the U and Th host minerals.
the Scandinavian Alum Shales have not been observed in coeval black For instance, the dominant U host in peraluminous granites is uraninite
shales from Llyn Peninsula (Wales), Levis (Quebec), Schaghticoke (New (e.g., leucogranites of the Limousin and the Erzgebirge; Thomas et al.,
York), and the Cordillera of Bolivia (Berry et al., 1986), highlighting the 2006; Cuney, 2014). In high-K metaluminous granites, U and Th may be

493
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

hosted in a wide range of minerals, including allanite, U-thorite, Nb- Förster, 2000; Tabaud et al., 2015). There is minor U mineralization in
oxides, and uraninite (e.g., monzodiorites and monzogranites of the the Vosges and the Schwarzwald and major U mineralization in the
Bretagne and Erzgebirge; Förster et al., 2000; Cuney, 2014; Ballouard Erzgebirge (Figs. 2, 4, and 6). Silurian black shales and HHP granites
et al., 2018), whereas in peraluminous granites, U and Th are dom- with spatially associated U mineralization are also known from Iberia
inantly hosted in zircon, U-Zr-Th-silicophosphates, and niobates (e.g., Martin-Izard et al., 2002; Mohamud et al., 2015). HHP granites
(Cuney, 2014). The nature of U and Th host minerals determines from Cornwall and Nova Scotia (e.g., Chatterjee et al., 1983; Marshall
whether U and Th eventually become available for redistribution by et al., 2010) show only minor U mineralization (e.g., LeBoutillier, 2002;
leaching. In most peraluminous granites, as much as 90% of U and Th Ryan and O'Beirne-Ryan, 2009), possibly reflecting that Avalonia that
reside in accessory minerals. Among the accessory minerals in per- had separated from Gondwana in the early Ordovician and was not
aluminous granites, uraninite hosts 30–90% of the available U and glaciated during the late Ordovician (e.g., McKerrow et al., 2000; Cocks
monazite hosts most Th (Cuney and Friedrich, 1987; Förster and and Fortey, 2009; Pollock et al., 2009; Delabroye and Vecoli, 2010),
Förster, 2000). In rocks without monazite, allanite and thorite may be which resulted in black shales that had lower U contents than those on
important hosts for both Th and U (Cuney, 2014). In volcanic rock, a Gondwana with its vaste deglaciated U source area. Thus on Avalonia,
significant portion of U and Th may be in the glassy matrix (Cuney, HHP granites may have overall less U available for leaching than
2014). Whether U and Th become available for redistribution depends comparable rocks from the heavily mineralized Variscan areas of the
on the chemical stability of their host. For instance, uraninite and Armorican Spur, such as the Erzgebirge, the Armorican Massif, or the
volcanic glass are not very stable und U and Th may become readily French Massif Central.
leached in presence of oxidized fluids (e.g., Langmuir, 1978; Cuney,
2010). Thus, if oxidized fluids are available, U and Th may become 7. Tectonic controls of post-Variscan uranium mobilization in
mobilized from uraninite or volcanic rocks soon after crystallization. In Europe
contrast, zircon, allanite, monazite, xenotime, and niobates are rather
stable and will not release U and Th before the crystal lattice of the host There seems to be no magmatic, metamorphic, or hydrothermal U
mineral has become strongly metamict. The destruction of the crystal mineralization related to the Variscan compressional stress regime in
lattice by cumulation of α-recoil damages depends on the contents of U the interior of the orogenic belt. In contrast, there is abundant and lo-
and Th, as well as time. The higher the U and/or Th contents, the cally important U mineralization associated with the extensional re-
shorter the period of time is needed to damage the crystal lattice suf- gime that prevailed throughout Western Europe after c. 300 Ma. There
ficiently (domains of overlapping α -recoil damages form a network are three types of U mineralization: (i) episyenites, (ii) quartz-pitch-
that connects with the mineral surface) to make U and Th mobile. The blende veins, and (iii) sediment-hosted mineralization. Mineralization
cummulation of α -recoil damage is counteracted by thermal annealing, of all three types is directly or indirectly tectonically controlled.
which tends to occur at lower temperatures for phosphates than for Mineralization in episyenites and veins is structurally controlled by
silicates (e.g., Meldrum et al., 1998). Thus, U and Th hosted in zircon, fault and shear zones that acted as transport channels for the miner-
allanite, monazite, xenotime, and niobates become available for alizing fluids and are the location of wallrock alteration and metal
leaching at different times after crystallization. The fertility of a granite, precipitation. Typically, the formation of episyenite and U mineraliza-
thus, may vary over time, depending on the nature of U hosts, thermal tion are two separate events (cf. Leroy and Turpin, 1988). For sediment-
history, and leaching history. hosted U mineralization, the tectonic control is more diverse: (a) Faults
The spatial distribution of particular granite types depends on are the transport channels for U or the fluid that induced precipitation
source lithologies and melting conditions, which together control of U, whereas the sediments are the traps for U at the unconformity or
whether U and Th partition during partial melting of the soure rocks within the basin. (b) The sediments are both the transport channels and
into the melt or the restite. For instance, the spatial distribution of traps for U. The role of tectonic processes for the formation of sediment-
High-Heat Producing (HHP) granites, which represent fertile source hosted mineralization is to develop topography that enhances erosion,
rocks for U mineralization (e.g., Cuney and Kyser, 2015), in belts re- which removes U-depleted near-surface rocks and makes undepleted U
flects (i) the spatial distribution of sedimentary protoliths and their source rocks available for leaching. This link between tectonic pro-
metamorphic equivalents (e.g., Bernard-Griffith et al., 1985; Cuney cesses and U mobilization explains why different types of U miner-
et al., 1990; Förster and Romer, 2010) and (ii) that the protoliths may alization that form in contrasting environments are essentially coeval
have been enriched in U (Cuney et al., 1990). In the Armorican Massif over vast areas. At Lodève, there was a so-called U preconcentration
and the Saxo-Thuringian Zone, the only rocks within the various li- during the Permian, whereas economic U mineralization formed in the
thologic units (Cadomian basement with arc rocks and related sedi- Jurassic and Cretaceous (Lancelot and Vella, 1989).
mentary rocks and its cover of Cambrian to Silurian sedimentary and Post-Variscan tectonic activity includes (i) the development of the
volcanic rocks) that have enhanced U contents are black shales (e.g., Central European Extensional Province with late Carboniferous to
Cuney, 1990; Romer and Hahne, 2010). In these areas, Silurian black Permian sedimentary basins and rifting (Oslo Rift, Southern Permian
shales are preserved on tectonic blocks that did not experience ex- Basin, Northern Permian Basin) during the final stages of the closure of
tensive Variscan tectonic reworking (e.g., Dabard and Paris, 1986; the Rheic ocean along the Appalachian-Mauritanide belt (e.g., Ziegler,
Seidel, 2003; Verniers et al., 2008; Romer and Hahne, 2010). As the 1990; Roscher and Schneider, 2006; Kroner et al., 2016), (ii) opening of
Silurian black shales had regional distribution on the Gondwana shelf the Tethys (e.g., Ziegler, 1988; Gómez and Goy, 2005; Kroner et al.,
(e.g., Lüning et al., 2000a; Le Heron et al., 2009), the reworking of 2016), (iii), opening of the Central Atlantic (e.g., Vogt and Tucholke,
Gondwana shelf sediments during the Variscan orogeny also involved 1989; Nomade et al., 2007), (iv) opening of the Northern Atlantic with
Silurian rocks. Thus, it is quite possible that the distribution of HHP establishing of new rifting zones and changes in spreading rate at c.
granites is controlled by the distribution of (i) Silurian black shales at 180 Ma, 150 Ma, 120 Ma, and 80 Ma (Vogt and Tucholke, 1989), and
depth and (ii) heat sources resulting in crustal melting. In France, such (v) Cretaceous basin inversion during Iberia-Africa convergence and the
granites define a belt that extends along different branches of the South initial stages of the Alpine orogeny (e.g., Cuney, 1978; Voigt et al.,
Armorican Shear Zone into the western French Massif Central (Fig. 8). 2008; Kley and Voigt, 2008). Even though the various events of tectonic
All U districts of the Armorican Massif (Pontivy, Guérande, Mortagne) reactivation are listed as separate events, some of them overlap or show
and the western French Massif Central (Western Marche, Limousin, continuous transitions from one to the other. Common to these tectonic
Millevaches, Margeride) occur along this belt (Cathelineau et al., 1990; events is that they mobilized brines in Mezozoic basins, which resulted
Turpin et al., 1990). HHP granites also occur in the Vosges, Schwarz- in the formation of fault-bound vein-type mineralization typically
wald, and Erzgebirge (Figs. 2, 4, and 6; Förster et al., 1999; Förster and containing Pb, Zn, Ba, and/or F (e.g., Patrier et al., 1997; Boiron et al.,

494
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

2002; Romer et al., 2010), as well as in the redistribution of U (Cuney, assemblages, which not only redistributed U within the veins, but
1978; Förster, 1996). The economically most important period of U introduced other metals, most prominently Ag-Sb and Bi-Co-Ni
mobilization and mineralization in Central and Western Europe is the (e.g., Hiller and Schuppan, 2008; Ondruš et al., 2003a,b). In con-
late Carboniferous and especially the early Permian, when U-fertile trast, vein-type U mineralization on Armorica, in the French Massif
granites and volcanic rocks became available for leaching. (e.g., Leroy Central, and in Iberia does not have such complex metal associa-
and Holliger, 1984; Förster, 1996; Cuney and Kyser, 2015; Ballouard tions, even though Bi is locally associated with U (e.g., Chauris,
et al., 2018). 1977; Cuney, 1978). Most vein-type U mineralization shows in
addition to this tectonically controlled hydrothermal redistribution
(i) Uranium mineralization related to episyenites. This type of mi- also low-temperature U redistribution via oxidative fluids that lo-
neralization is particularly important in the French Massif Central cally have redistributed U to several hundred meters depth (e.g.,
and to a lesser extent in the Central Iberian Zone (e.g., Cathelineau, Cuney, 1978; Runge, 2001; Hiller and Schuppan, 2008).
1986; Patrier et al., 1997; López-Moro et al., 2013). Episyenites (iii) Black shale- and sandstone-hosted U mineralization. The sedi-
typically form at the expense of (two mica) leucogranites and felsic mentary rocks represent the U trap. Tectonic processes make U
gneisses by dequartzification occasionally associated with the available for transport, both by opening the transport channels or
formation of secondary albite or K-feldspar. This type of alteration by tectonic uplift, which turn results in increased erosion and,
typically occurs in the temperature range of 450–350 °C by low- thus, brings rocks to the surface that have not loss U during earlier
salinity aqueous fluids. As these altered rocks are highly perme- leaching, but still respresent a source of leachable U. Thus, tectonic
able, early high-temperature alteration may be followed by a series uplift eventually results in an increase of dissolved U surface
of hydrothermal alteration stages at increasingly lower tempera- runoff.
ture, including precipitation of adularia and quartz, silicification, (a) Faults in the basement cut through organic- and/or sulfide-rich
argillitization, and kaolinitization, and supergene alteration strata overlying the basement. Uranium leached from the
(Leroy, 1978, 1984; Cathelineau, 1986; Leroy and Turpin, 1988; basement or mobilized from U mineralization in the basement
Patrier et al., 1997). Episyenites are structurally controlled and is precipitated as U-bearing fluids migrating along the faults
occur along fault and shear zones or form pipes (Pécher et al., interact with these overlying reducing sedimentary rocks.
1985; Lespinasse and Pécher, 1986). There are barren and miner- (b) Uranium is transported within permeable layers of a sedi-
alized episyenites. Fluid inclusions from primary uraninite within mentary unit and is precipitated at locations with organic-
episyenite dominantly fall in the range from 250 to 200 °C, which and/or sufide-rich layers in sandstones and shales. Commonly,
is distintly lower than fluid inclusions in episyenite (450–350 °C; fluid migration and U precipitation occurs in the basal sedi-
Cathelineau, 1986). Fluid temperatures in the range 150–100 °C mentary units above the unconformity. Locally, however, fluid
seem to be related to later redistributions (e.g., Lespinasse and flow may have also permeated into the heavily altered, sa-
Cathelineau, 1990; El Jarray et al., 1994). Isotopic dating indicates propelic basement to form U mineralization (e.g., Tonndorf,
that episyenite formation and U mineralization is related to two 2000).
separate events, as most episyenites yield ages around 305 Ma and
U mineralization typically yields ages around 290–280 Ma or The importance of tectonic processes on sediment-hosted U miner-
younger (e.g., Turpin et al., 1990). Isotopic dating also indicates alization possibly is best illustrated by the Cretaceous and Tertiary
that there are several periods of U mineralization and that there tectonic reactivation of old structures in Western Europe. Cretaceous
may be multiple events of mineralization or redistribution (e.g., tectonic reactivation resulted in basin inversion and uplift of basement
Lévêque et al., 1988). blocks and is recorded in erosional discordances in the sedimentary
(ii) Quartz-(sulfide)-uraninite-pitchblende veins. Vein-hosted U mi- record (e.g., Peyaud et al., 2005), the abundance of Cretaceous apatite
neralization is particularly important in the Erzgebirge, French fission-track (120–80 Ma) and apatite and fluorite (U-Th-Sm)-He ages
Massif Central, and Armorica (Figs. 6 and 8). This type of U mi- (110–80 Ma, e.g., Barbarand et al., 2001; Danišík et al., 2012; Vamvaka
neralization formed during rifting in the Central European Exten- et al., 2014; Wolff et al., 2015), locally important hydrothermal struc-
sional Province around 290–270 Ma (e.g., Förster, 1996; Ballouard turally-controlled Ba-F and Pb-Zn mineralization (e.g., Boiron et al.,
et al., 2018). Mineralization is spatially closely associated with U- 2002; Schmidt Mumm and Wolfgramm, 2004), the redistribution of U
fertile granites (U commonly hosted in uraninite or allanite) that in older vein-type U mineralization (e.g., Lévêque et al., 1988; Förster,
acted as source for U leaching. The granites are 30–50 m.y. older 1996), and the deposition of regionally important continental sand-
than the mineralization (e.g., Cathelineau et al., 1990; Förster, stones (e.g., Uličný, 2001; Uličný et al., 2009). Tectonic inversion
1996; Romer et al., 2010; Ballouard et al., 2018). Mineralization makes U sources available for leaching and generates the siliciclastic
may be particularly rich in segments that cut sufide- or organic- debris that eventually represents the U trap and, thus, the host rock of U
rich sedimentary or metamorphic rocks that may have acted as mineralization. Thus, both mineralization and host rocks result from
traps (e.g., Runge, 2001). Major tectonic lineaments influence the tectonic inversion. Uranium in the Cenomanian sandstones at König-
distribution of vein type mineralization in double sense: these stein (Fig. 6) possibly is derived from the Markersbach granite (e.g.,
structures controlled the emplacement and shape of the granites Tonndorf, 2000), whereas U for the mineralization in the Cenomanian
and their brittle reactivation formed the mineralized structures, sandstones at Hamr, Osečná, and Stráž may have been leached from the
such as Riedel faults to the major lineament and faults along granitic rocks of the Sudetes. Similarly, the uplift of the Erzgebirge in
contacts between rocks of contrasting rheological properties. There the Tertiary resulted in the formation of U mineralization in lignites in
was repeated U redistribution that seems to fall in the same age the Eger Graben (Fig. 6). The spatial closeness of U-bearing lignites to
ranges as distinct large-scale tectonic events at c. 180, 150, and the Jáchymov ore district highlights the importance of the distribution
120 Ma, and more recently during the Alpine orogeny (e.g., Romer of fertile source rocks for the distribution of younger sediment-bound
et al., 2010; Ballouard et al., 2018). There are also regional dif- mineralization.
ferences in the mineral and metal assemblages, which indicates
that the local wallrocks played an important role as metal source. 8. Exogenic controls of post-Variscan uranium mobilization in
For instance, the oldest (Permian) vein-type U mineralization in Europe
the Erzgebirge is characterized by quartz-sulfide (Fe, As, Cu, Zn,
Pb, and Se), quartz-hematite, and U-quartz-calcite associations that Leachable U is readily lost from oxidized rocks at surface. Extreme
have been overprinted by multiple younger hydrothermal mineral environmental conditions are relevant for U mineralization by making

495
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

U available for leaching (for instance removing earlier leached rocks by mineralization may have formed. During the extension of the northern
glacial erosion), controlling the distribution of areas of particular in- margin of Gondwana, which eventually resulted in the opening of the
tense leaching, and controlling the distribution of traps for U pre- Rheic Ocean, Cambrian and early Ordovician sediments on Gondwana
cipitation (e.g., anoxic conditions on the shelf and in lakes or super- were largely eroded and redistributed to the shelf areas of Gondwana.
saturation of groundwater in arid basins). For exogenic U Therefore, Cambrian and Ordovician rocks of Gondwana are poor
mineralization to form, a U source and a U trap, generally a reducing source rocks for younger U mineralization. The situation changed with
environment, have to occur at the same time and have to be in close the Hirnantian glaciation that resulted in the exposure of basement
spatial context in the sense that transport of U from the source to the rocks that did not lose U during earlier weathering and alteration (see
sink has to be possible. A source without a trap will become U-depleted above).
without forming U mineralization, whereas a trap without an un- The Carboniferous to Permian assembly of Pangea eventually re-
depleted U source in the drainage area has little U to cumulate and, sulted in a large-scale desertification of Gondwana and Variscan Europe
therefore, is not likely to concentrate significant U mineralization. (Michel et al., 2015), largely due to the closure of the Rheic Ocean and
In Western Europe, the Hirnantian glaciation was important for U the overall northdrift of Pangea (Roscher and Schneider, 2006;
mineralization, as glacial erosion brought rocks to the surface that had Schneider et al., 2006). The trend of increasing desertification was in-
not yet lost their leachable U and melting of the ice sheets eventually terrupted by several phases of more humid climate. In each phase, the
led to the Silurian transgression with black shale deposits on the increase of humidity occurs toward the top of the corresponding glacial
Gondwana shelf (Fig. 10; see Section 5). Later glaciations were not as cycle, which implies the humid phases on northern Gondwana are
important for U mineralization in Variscan Europe, but have resulted in coupled with the growth of ice sheets on southern Gondwana (Roscher
U-rich black shale in other areas. For instance, the late Carboniferous to and Schneider, 2006; Schneider et al., 2006). Both plate drift and cli-
Permian Karoo glaciation did not affect present Europe (Fig. 10) except matic changes influence the nature of U trap on northern Gondwana
for sealevel changes. Thus, rocks that already had lost their leachable U with its marine and lacustrine coastal swamps in the Carboniferous
remained at the surface. The Quaternary glaciation apparently did not (e.g., coal-bearing Döhlen and Lodève basins) to the predominance of
develop major traps, even though anomalous enrichment of U in local desert basins in the Permian. The different nature of traps in turn also
post-glacial wetland, bogs, and swamps is well documented (e.g., would result in different types of U mineralization, whereas the coeval
Wilson, 1984; Regenspurg et al., 2010). Furthermore, the distribution availability of leachable U determines the importance of U miner-
of U contents in European Quaternary soil and stream sediments does alization. The contrasting availability of U for leaching in the Carbo-
not follow the distribution of glaciated areas (Plant et al., 2003). In- niferous may account for rich U mineralization the Döhlen basin and
stead, the highest U contents are found (i) in the Variscan Massifs that only minor U mineralization in the Stockheim basins (e.g., Dill, 1987;
have important U-fertile granites, a history of U mining, and periglacial Reichel and Schauer, 2006).
deposits and (ii) in Alpine areas (which include reworked Variscan Periods of intense chemical weathering due to a particularly ag-
basement). In contrast, stream sediments from Scandinavia, which have gressive atmosphere (high CO2) or high temperatures are important for
been covered by an ice sheet, generally have relatively low U contents U mineralization because of the following reasons: (i) intense chemical
(Plant et al., 2003) except for stream sediments (and to a lesser extent in weathering results in extreme source depletion and younger traps are
soil) from areas in southern Norway and Finland that have importat unlikely to receive surface U; (ii) in semi-arid and arid environments, U
rare element pegmatites and related granites. Thus, glaciation may may concentrate locally in closed basins (roll-front type and calcrete-
rejuvenate an old land surface, but whether a formerly glaciated area hosted U mineralization); (iii) the oxidized zone may penetrate deeper,
represents an important U source largely depends on the lithology and resulting in the redistribution of U to deeper, still reduced levels.
the nature of U-hosting minerals at the exposed surface. The important feature of deep alteration and oxidation of near-
There is significant U mineralization in areas that have been directly surface rocks is that these rocks lose their leachable U during this
affected by the Karoo glaciation (Fig. 10). In Europe, however, the in- process and thereafter do not represent a U source in continuation.
fluence of the Karoo glaciation was restricted. Eustatic sealevel fluc- Furthermore, U mobilization occurs only over a short period until near-
tuation (e.g., Catuneanu et al., 2005) locally resulted in the formation surface rocks are depleted in leachable U and stop to be U sources,
of organic-rich sediments. With the possible exception of high-altitude independent whether environmental conditions are favorable for U
areas, there was no Karoo glaciation in Europe. Thus in coastal areas of mobilization. Existing near-surface U mineralization and U-rich rocks
Europe, the Karoo glaciation may have resulted in the formation of U may be affected in two different ways: (i) oxidation of U mineralization,
traps, but there were no leachable U sources at the surface available. U-fertile granites, or black shales may result in the mobilization and
Only in high-altitude areas, local glaciations may have produced re- dispersion of U eventually resulting in the destruction of older U mi-
stricted U sources that eventually might have given rise to local Alpine- neralization or depletion of potential U source rocks (e.g., Fischer et al.,
type peat and bog mineralization, which has a low preservation po- 2009). (ii) Mobilization of U by deeply penetrating oxidized meteoric
tential. The Permian represents the most important period of U mi- fluids results in loss of U in the oxidation zone and cumulation of U in
neralization in Western and Central Europe. Permian U deposits are not the reduction zone. In particualr for older U mineralization, this may
related to glaciation in the U source. Instead, Permian U mineralization result in strong local U enrichment (e.g., Hiller and Schuppan, 2008;
in Western and Central Europe is related to tectonic processes. Bolonin and Gradovsky, 2012).
Intense chemical alteration with the complete decomposition of
feldspar is enhanced by both elevated temperatures and high atmo- 9. Uranium mineralization in Variscan Europe – the result of
spheric CO2 (e.g., Avigad et al., 2005), but also requires a gentle to- coupled endogenic and exogenic processes
pography that allows for a long residence of the sediments on the
continent. The intense chemical alteration during the Cambrian (and The spatial distribution of U mineralization within the Variscan
Ordovician) that eventually resulted in the formation of mature sand- orogeny is in a first-order approach dominantly controlled by the dis-
stones all over Gondwana (e.g., Avigad et al., 2005; Meinhold et al., tribution of Silurian hot shales. These shales formed on the proximal
2013) not only decomposed feldspar, but also removed readily oxidized shelf on the northern part of Gondwana after the Hirnantian glaciation.
minerals, such as Mo- and U-bearing minerals. Thus, intense chemical The particular feature of these black shales and their high U content is
alteration is an efficient process to mobilize U and to deplete a source their causal relation with the Hirnantian glaciation that (i) created the
lithology in U. During the Cambrian and Ordovician, there were no U source by glacial erosion of the old land surface; (ii) made undepleted
traps for U in areas of future Variscan Europe, with the possible ex- source rocks available for leaching during and after the melting of the
ception of continental basins where sandstone-hosted and surficial U ice sheet; and (iii) created the sedimentary trap due to sea level rise

496
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

during melting that in combination with the stratified water body cooling history of their host rock is essential for making U available for
(sweet water on salt water) and high organic productivity eventually leaching. Heating induces recrystallization and may lead to the in-
resulted in anoxic conditions. Enhanced U contents probably are re- corporation of U into the crystal lattice of the host mineral, which
stricted to the inner shelf and are confined to a short period of time, as makes U unavailable for mobilization and redistribution. Long re-
leaching of the glacially modeled land surface rather rapidly depleted sidence at low temperature, which eventually results in the acquisition
the U source. of significant radiation damage and turns the host mineral of U and Th
Redistribution of U from the Silurian black shales took a wide range increasingly metamict, makes U available for leaching and redistribu-
of different paths that are linked to endogenic and exogenic processes tion. It should be noted that volcanic rocks and uraninite are “instant” U
that made U available for redistribution and created traps for con- sources as their U content is available for redistribution under oxidizing
centration of U. conditions right after the rock or mineral formed, whereas silicate,
niobate, and phosphate minerals that substitute U into their crystal
(i) Stacking of Paleozoic sedimentary packages locally repeated U- lattice are “delayed” U sources as their U only may become mobilized
rich lithologies. Although this process is unlikely to result in mi- after the host mineral has become metamict. If the U content of a rock is
neralization, it increases the amount of U locally available for later distributed among different accessory minerals with appreciable con-
redistribution. In the Ronneburg area, the stacking of Silurian tents of U and/or Th, the various accessory minerals may become
black shales may have been an essential feature for this district to “sufficient” metamictized to allow for U mobilization at different time,
develop subsequently major U mineralization. depending on the contents of U and Th and the temperature at which
(ii) Partial melting of high-grade metamorphic rocks after the empla- annealing of radiation damage takes place. Such samples may be re-
cement of the metamorphic nappes resulted in the partitioning of U peatedly fertile, even though the leachable U in the various leaching
from the metamorphic rocks into the melts. The emplacement of events is likely to be contributed from different accessory minerals.
these melts in the upper crust redistributes U from the lower crust The formation of exogenic mineralization is strongly controlled by
to the middle and upper crust. Within the Variscan orogen, most the coupled availability of U sources and traps, as a source generally
high-grade nappe complexes are dominated by metamorphosed becomes depleted. If there is no trap available at that time, leached U
sedimentary rocks of the former Gondwana shelf. Among these becomes dispersed, no mineralization forms. Similarly, a trap will not
sedimentary rocks, the Silurian black shales are particularly U-rich. be mineralized if the source already is depleted. Active tectonic pro-
Partial melting of metamorphosed Silurian black shales not only cesses may make undepleted U sources available for erosion and
may enhance the U contents in the melt, but also may contribute to leaching. Thus, periods of tectonic activity correspond to periods of
the reduced character of the melts. Thus, assimilation and partial exogenic U redistribution and U eventually mineralization such that
melting of Silurian black shales may explain, why some Variscan endogenic processes indirectly control exogenic U mineralization.
two–mica leucogranites carry accessory uraninite. Permian vol-
canism reflecting large-scale crustal extension with mantle up- Acknowledgments
welling and crustal melting reflects a second magmatic event that
may have redistributed U to the upper crust and the surface. We thank Uwe Kroner (TU Freiberg) for discussion and the tem-
(iii) Late- and post-Variscan hydrothermal U mineralization involved plates to several maps. We greatfully appreciate the very careful and
the leaching of granitic (in particular two-mica leucogranites with detailed review of an anonymous reviewer and the thoughtful editorial
accessory uraninite) and sedimentary source rocks (in particular handling by Franco Pirajno.
Silurian black shales). The spatial distribution of hydrothermal U
mineralization is controlled by the spatial distribution of U-fertile References
two-mica leucogranites (possibly reflecting the distribution of
Silurian black shales at depth) and Silurian black shales at the level Abalos, B., Carreras, J., Druguet, E., Viruete, J.E., Pugnaire, M.T.G., Quesada, C.,
of hydrothermal fluid circulation along fault zones. The miner- Fernandez, R., Gil Ibarguchi, J.I., 2002. Variscan and pre-Variscan Tectonics. In:
Gibbons, W., Moreno, C. (Eds.), The Geology of Spain. Geol. Soc., Lond, pp. 155–183.
alized structures are spatially closely related to the granites as the Al Hseinat, M., Hübscher, C., 2017. Late Cretaceous to recent tectonic evolution of the
emplacement of the granites is controlled by older shear and fault North German Basin and the transition zone to the Baltic Shield/southwest Baltic Sea.
zones and later stress on the continental crust reactivated favor- Tectonophysics 708, 28–55.
Al-Juboury, A.I., Al-Hadidy, A.H., 2009. Petrology and depositional evolution of the
ably oriented old structures. This includes the same structures al- Paleozoic rocks of Iraq. Marine Petrol. Geol. 26, 208–231.
ready exploited by the granites, but also rheological contrasts Albarède, F., Michard, A., 1986. Transfer of continental Mg, S, O and U to the mantle
along the granite contact or within the contact metamorphic aur- through hydrothermal alteration of the oceanic crust. Chem. Geol. 57, 1–15.
Alexandre, P., 2010. Mineralogy and geochemistry of the sodium metasomatism-related
eole. Such repeated reactivation of old structures explains both the uranium occurrence of Aricheng South. Guyana. Mineral. Depos. 45, 351–367.
spatial coincidence of episyenite and U mineralization, which Alexandre, P., Kyser, K., Polito, P., Thomas, D., 2005. Alteration mineralogy and stable
formed at different temperatures and different time (e.g., isotope geochemistry of Paleoproterozoic basement-hosted unconformity-type ur-
anium deposits in the Athabasca Basin. Canada. Econ. Geol. 100, 1547–1563.
Cathelineau, 1986; Turpin et al., 1990), and the multiple episodes
Alexandre, P., Kyser, K., Jiricka, D., 2009. Critical geochemical and mineralogical factors
of hydrothermal redistribution within vein-type U mineralization for the formation of unconformity-related uranium deposits: comparison between
(e.g., Ondruš et al., 2003a; Hiller and Schuppan, 2008). barren and mineralized systems in the Athabasca Basin, Canada. Econ. Geol. 104,
(iv) Post-Variscan U mobilization within the Variscan belt mainly is 413–435.
Altherr, R., Holl, A., Hegner, E., Langer, C., Kreuzer, H., 2000. High-potassium, cal-
controlled by weathering and hydrothermal mobilization. The calkaline I-type plutonism in the European Variscides: Northern Vosges (France) and
most important periods of U mineralization are the Permian, which northern Schwarzwald (Germany). Lithos 501, 51–73.
in Europe mainly is linked to the extensional event of the Central Anderson, R.F., Fleisher, M.Q., 1991. Uranium Precipitation in Black Sea Sediments. In:
Izdar, E., Murray, J.W. (Eds.) Black Sea Oceanography. NATO ASI Series., Ser. C 351,
European Extensional Province, and later basin inversions related 443–458.
to the spreading of the Atlantic Ocean and/or to the Alpine colli- Andersson, A., Dahlman, B., Gee, D.G., 1982. Kerogen and uranium resources in the
sion. Cambrian alum shales of the Billingen-Falbygden and Närke areas, Sweden. Geol.
Fören, Stockholm Förh. 104, 197–209.
Anderson, J.R., Goodknight, C.S., Sewell, J.M., Riley, J.K., 1987. Favorability of
The formation of endogenic mineralization is largely controlled by Precambrian quartz-pebble conglomerates in the United States as uranium hosts. In:
the availability of source rocks and a heat source that mobilizes U into Uranium deposits in Proterozoic quartz-pebble conglomerates. IAEA, Vienna, pp.
7–40 IAEA-TECDOC-427.
melts or hydrothermal fluids. Source depletion occurs when U hosted in Avigad, D., Sandler, A., Kolodner, K., Stern, R.J., McWilliams, M., Miller, N., Beyth, M.,
volcanic glass, uraninite, or strongly metamict U-bearing minerals is 2005. Mass-production of Cambro-Ordovician quartz-rich sandstone as a con-
mobilized. If metamict minerals represent an important U host, the sequence of chemical weathering of Pan-African terranes: environmental

497
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

implications. Earth Planet. Sci. Lett. 240, 818–826. Bruneton, P., Cuney, M., 2016. Geology of uranium deposits. In: Hore-Lacy, I. (Ed.),
Ballèvre, M., Bosse, V., Ducassou, C., Pitra, P., 2009. Palaeozoic history of the Armorican Uranium for Nuclear Power: Resources, Mining and Transformation to Fuel. Elsevier,
Massif: Models for the tectonic evolution of the suture zones. C. R. Geosci. Paris. 341, Amsterdam, pp. 11–52.
174–201. Büder, W., Schuppan, W., 1992. Zum Uranerzbergbau im Lagerstättenfeld Schlema-
Ballouard, C., Poujol, M., Boulvais, P., Mercadier, J., Tartèse, R., Venneman, T., Deloule, Alberoda im Westerzgebirge. Schr. GDMB 64, 203–221.
E., Jolivet, M., Kéré, I., Cathelineau, M., Cuney, M., 2017. Magmatic and hydro- Butt, C.R.M., Mann, A.W., Horwitz, R.C., 1984. Regional Setting, Distribution and Genesis
thermal behavior of uranium in syntectonic leucogranites: the uranium mineraliza- of Surficial Uranium Deposits in Calcretes and Associated Sediments in Western
tion associated with the Hercynian Guérande granite (Armorican Massif, France). Ore Australia. TECDOC-322. IAEA, Vienna, pp. 121–127.
Geol. Rev. 80, 309–331. Cai, C., Li, H., Qin, M., Luo, X., Wang, F., Ou, G., 2007. Biogenic and petroleum-related
Ballouard, C., Poujol, M., Mercadier, J., Deloule, E., Boulvais, P., Baele, J.M., Cuney, M., ore-forming processes in Dongsheng uranium deposit, NW China. Ore Geol. Rev. 32,
Cathelineau, M., 2018. Uranium metallogenesis in the peraluminous leucogranites 262–274.
from the Pontivy-Rostrenen magmatic complex (French Armorican Variscan Belt): Cairncross, B., 2001. An overview of the Permian (Karoo) coal deposits of southern Africa.
the result of long term oxidized hydrothermal alteration during strike-slip deforma- J. Afr. Earth Sci. 33, 529–562.
tion. Mineral. Depos. 53, 601–628. Castiñeiras, P., Villaseca, C., Barbero, L., Romera, C.M., 2008. SHRIMP U-Pb zircon dating
Barbarand, J., Lucazeau, F., Pagel, M., Séranne, M., 2001. Burial and exhumation history of anatexis in high-grade migmatite complexes of Central Spain: implications in the
of the south-eastern Massif Central (France) constrained by apatite fission-track Hercynian evolution of Central Iberia. Int. J. Earth Sci. 97, 35–50.
thermochronology. Tectonophysics 335, 275–290. Carvalho, P.C.S., Neiva, A.M.R., Silva, M.M.V.G., Corfu, F., 2012. A unique sequential
Barbier, M.J., 1974. Continental weathering as a possible origin of vein-type uranium melting mechanism for the generation of anatectic granitic rocks from the Penafiel
deposits. Mineral. Depos. 9, 271–288. area, northern Portugal. Lithos 155, 110–124.
Barthel, F.H., 1974. Review of uranium occurrences in Permian sediments in Europe, with Castro, A., Douce, A.E.P., Corretgé, L.G., Jesús, D., El-Biad, M., El-Hmidi, H., 1999. Origin
special reference to uranium mineralizations in Permian sandstone. In: Formation of of peraluminous granites and granodiorites, Iberian massif, Spain: an experimental
Uranium Ore Deposits. IAEA, STI/PUB/374, Vienna, 277–289. test of granite petrogenesis. Contrib. Mineral. Petrol. 135, 255–276.
Basson, I.J., Greenway, G., 2004. The Rössing uranium deposit: a product of late-kine- Catuneanu, O., Wopfner, H., Eriksson, P.G., Cairncross, B., Rubidge, B.S., Smith, R.M.H.,
matic localization of uraniferous granites in the Central Zone of the Damara Orogen, Hancox, P.J., 2005. The Karoo basins of south-central Africa. J. Afr. Earth Sci. 43,
Namibia. J. Afr. Earth Sci. 38, 413–435. 211–253.
Benalioulhaj, S., Trichet, J., 1990. Comparative study by infrared spectroscopy of the Cathelineau, M., 1986. The hydrothermal alkali metasomatism effects on granitic rocks:
organic matter of phosphate-rich (Oulad Abdoun basin) and black shale (Timahdit quartz dissolution and related subsolidus changes. J. Petrol. 27, 945–965.
basin) series (Morocco). Org. Geochem. 16, 649–660. Cathelineau, M., Boiron, M.C., Holliger, P., Poty, B., 1990. Metallogenesis of the French
Bernard-Griffith, J., Peucat, J.J., Sheppard, S.M.F., Vidal, P., 1985. Petrogenesis of part of the Variscan orogen. Part II: Time-space relationships between U, Au and Sn-
Hercynian leucogranites from the southern Armorican Massif: contribution of REE W ore deposition and geodynamic events – mineralogical and U-Pb data.
and isotopic (Sr, Nd, Pb and O) geochemical data to the study of source rock char- Tectonophysics 177, 59–79.
acteristics and ages. Earth Planet. Sci. Lett. 74, 235–250. Cathelineau, M., Fourcade, S., Clauer, N., Buschaert, S., Rousset, D., Boiron, M.C.,
Berry, W.B.N., Wilde, P., Quinby-Hunt, M.S., Orth, C., 1986. Trace element signatures in Meunier, A., Lavastre, V., Javoy, M., 2004. Dating multistage paleofluid percolations:
Dictyonema Shales and their geochemical and stratigraphic significance. Norsk Geol. a K-Ar and 18O/16O study of fracture illites from altered Hercynian plutonites at the
Tidsskr. 66, 45–51. basement/cover interface (Poitou High, France). Geochim. Cosmochim. Acta 68,
Boberg, W.W., 2010. The Nature and Development of the Wyoming Uranium Province. In: 2529–2542.
Goldfarb, R.J., Marsh, E.E., Monecke, T. (Eds.), The Challenge of Finding New Chakraborty, C., Ghosh, S.K., Chakraborty, T., 2003. Depositional record of tidal-flat
Mineral Resources: Global Metallogeny, Innovative Exploration, and New sedimentation in the Permian coal measures of central India: Barakar Formation,
Discoveries. Econ. Geol. Spec. Publ, pp. 653–674. Mohpani Coalfield, Satpura Gondwana Basin. Gondwana Res. 6, 817–828.
Boiron, M.C., Cathelineau, M., Banks, D.A., Buschaert, S., Fourcade, S., Coulibaly, Y., Chelle-Michou, C., Laurent, O., Moyen, J.F., Block, S., Paquette, J.L., Couzinié, S.,
Michelot, J.L., Boyce, A., 2002. Fluid transfers at a basement/cover interface: Part II. Gardien, V., Vanderhaeghe, O., Villaros, A., Zeh, A., 2017. Pre-Cadomian to late-
Large-scale introduction of chlorine into the basement by Mesozoic basinal brines. Variscan odyssey of the eastern Massif Central, France: Formation of the West
Chem. Geol. 192, 121–140. European crust in a nutshell. Gondwana Res. 46, 170–190.
Bolonin, A.V., Gradovsky, I.F., 2012. Supergene processes and uranium ore formation in Chatterjee, A.K., Robertson, J., Pollock, D., 1982. A summary on the petrometal1ogenesis
the Ronneburg ore field, Germany. Geol. Ore Deposits 54, 122–131. of the uranium mineralization, Millet Brook, South Mountain Batholith, Nova Scotia.
Bonnetti, C., Cuney, M., Malartre, F., Michels, R., Liu, X., Peng, Y., 2015. The Nuheting Nova Scotia Dept. Mines Energy Rept. 782–1, 57–67.
deposit, Erlian Basin, NE China: synsedimentary to diagenetic uranium mineraliza- Chatterjee, A.K., Strong, D.F., Muecke, G.K., 1983. A multivariate approach to geo-
tion. Ore Geol. Rev. 69, 118–139. chemical distinction between tin-specialized and uranium-specialized granites of
Bonnetti, C., Cuney, M., Bourlange, S., Deloule, E., Poujol, M., Liu, X., Peng, Y., Yang, J., southern Nova Scotia. Can. J. Earth Sci. 20, 420–430.
2017a. Primary uranium sources for sedimentary-hosted uranium deposits in NE Chauris, L., 1977. Les associations paragénétiques dans la métallogenie varisque du
China: insight from basement igneous rocks of the Erlian Basin. Mineral. Depos. 52, massif armoricain. Mineral. Depos. 12, 353–371.
297–315. Chauris, L., 1984. Accidents linéamentaires et minéralisations uranifères; L’exemple de la
Bonnetti, C., Liu, X., Zhaobin, Y., Cuney, M., Michels, R., Malartre, F., Mercadier, Cai, J., ceinture batholitique hercynienne médio-armoricaine (France). Bull. Soc. géol.
2017b. Coupled uranium mineralisation and bacterial sulphate reduction for the France 26, 1375–1380.
genesis of the Baxingtu sandstone-hosted U deposit, SW Songliao Basin, NE China. Cherns, L., Wheeley, J.R., 2009. Early Palaeozoic cooling events: peri-Gondwana and
Ore Geol. Rev. 82, 108–129. beyond. In: Bassett, M.G. (Ed.), Early Palaeozoic Peri-Gondwana Terranes: New
Borg, G., Piestrzýnski, A., Bachmann, G.H., Püttmann, W., Walther, S., Fiedler, M., 2012. Insights from Tectonics and Biogeography. Geol. Soc., Lond., Spec. Publ., pp.
An overview of the European Kupferschiefer deposits. Econ. Geol. Spec. Publ. 16, 257–278.
455–486. Chi, G., Haid, T., Quirt, D., Fayek, M., Blamey, N., Chu, H., 2017. Petrography, fluid
Both, R., Arribas, A., Saint-André, B., 1994. The origin of breccia-hosted uranium deposits inclusion analysis, and geochronology of the End uranium deposit, Kiggavik,
in carbonaceous metasediments of the Iberian Peninsula: U-Pb geochronology and Nunavut, Canada. Mineral. Depos. 52, 211–232.
stable isotope studies of the Fé deposit, Salamanca province, Spain. Econ. Geol. 89, Church, T.M., Sarin, M.M., Fleisher, M.Q., Ferdelman, T.G., 1996. Salt marshes: an im-
584–601. portant coastal sink for dissolved uranium. Geochim. Cosmochim. Acta 60,
Bowell, R.J., Davies, A.A., 2017. Assessment of supergene uranium-vanadium anomalies, 3879–3887.
Meob Bay deposit, Namibia. Geochem.: Exploration Environ., Anal. 17, 101–112. Cole, D.I., 1998. Uranium. In: Wilson, M.G.C., Anhaeusser, C.R. (Eds.), The Mineral
Breitkreuz, C., Kennedy, A., 1999. Magmatic flare-up at the Carboniferous/Permian Resources of South Africa. Council for Geoscience, Handbook, pp. 642–658.
boundary in the NE German Basin revealed by SHRIMP zircon ages. Tectonophysics Condomines, M., Loubeau, O., Patrier, P., 2007. Recent mobilization of U-series radio-
302, 307–326. nuclides in the Bernardan U deposit (French Massif Central). Chem. Geol. 244,
Breitkreuz, C., Mock, A., 2004. Are laccolith complexes characteristic of transtensional 304–315.
basin systems? Examples from the Permo-Carboniferous of Central Europe. In: Corvino, A.F., Pretorius, L.E., 2013. Uraniferous leucogranites south of Ida Dome, central
Breitkreuz, C., Petford, N. (Eds.), Physical Geology of High-Level Magmatic Systems. Damara Belt, Namibia: morphology, distribution and mineralization. J. Afr. Earth Sci.
Geol. Soc., Lond., Spec. Publ., pp. 13–31. 80, 60–73.
Brenchley, P.J., Marshall, J.D., Carden, G.A.F., Robertson, D.B.R., Long, D.G.F., Meidla, Couzinié, S., Moyen, J.F., Villaros, A., Paquette, J.L., Scarrow, J.H., Marignac, C., 2014.
T., Hints, L., Anderson, T.F., 1994. Bathymetric and isotopic evidence for a short- Temporal relationships between Mg-K mafic magmatism and catastrophic melting of
lived Late Ordovician glaciation in a greenhouse period. Geology 22, 295–298. the Variscan crust in the southern part of Velay Complex (Massif Central, France). J.
Bril, H., Marignac, C., Cathelineau, M., Tollon, F., Cuney, M., Boiron, M.C., 1994. Geosci. 59, 1–18.
Metallogenesis of the French Massif Central: time-space relationships between ore Cocks, L.R.M., Fortey, R.A., 2009. Avalonia: a long-lived terrane in the Lower Palaeozoic?
deposition and tectono-magmatic events. In: Keppie, J.D., Chantraine, J., Rolet, J., In: Bassett, M.G. (Ed.), Early Palaeozoic Peri-Gondwana Terranes: New Insights from
Santallier, D.S., Piquet, A. (Eds.), Pre-Mesozoic Geology in France and Related Areas. Tectonics and Biogeography. Geol. Soc., Lond., Spec. Publ., pp. 141–155.
Springer, Heidelberg, pp. 379–402. Cuney, M., 1978. Geologic environment, mineralogy, and fluid inclusions of the Bois
Brockamp, O., Zuther, M., Clauer, N., 1987. Epigenetic-hydrothermal origin of the sedi- Noirs-Limouzat uranium vein, Forez, France. Econ. Geol. 73, 1567–1610.
ment-hosted Müllenbach uranium deposit, Baden-Baden, W-Germany. Monogr. Ser. Cuney, M., 1990. Contrôles magmatiques et structuraux de la métallogenèse uranifère
Min. Dep. 27. Bornträger, Berlin, pp. 87–98. tardi-hercynienne; exemple du district de la Crouzille (Haute-Vienne). Chron. Rech.
Brumsack, H.-J., 2006. The trace metal content of recent organic carbon-rich sediments: Min. 499, 9–17.
implications for Cretaceous black shale formation. Palaeogeogr., Palaeoclimat., Cuney, M., 2009. The extreme diversity of uranium deposits. Mineral. Depos. 44, 3–9.
Palaeoecol. 232, 344–361. Cuney, M., 2010. Evolution of uranium fractionation processes through time: driving the

498
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

secular variation of uranium deposit types. Econ. Geol. 105, 553–569. II: Xenotime. Am. Mineral. 83, 1302–1315.
Cuney, M., 2012. Uranium and thorium: the extreme diversity of the resources of the Förster, H.-J., 1999. The chemical composition of uraninite in Variscan granites of the
world’s energy minerals. In: Sindling-Larsen, R., Wellmer, F.-.W. (Eds.), Non-re- Erzgebirge, Germany. Miner. Mag. 63, 239–252.
newable Resource Issues: Geoscientific and Societal Challenges. International Year of Förster, H.-J., 2010. Late-Variscan granites of the Aue – Schwarzenberg Zone (western
the Planet. Springer, pp. 91–129. Erz- gebirge, Germany): Composition of accessory minerals and mineralogical mass
Cuney, M., 2014. Felsic magmatism and uranium deposits. Bull. Soc. géol. France 185, balance of the lanthanides and actinides. Z. Geol. Wiss. 38, 125–144.
75–92. Förster, A., Förster, H.-J., 2000. Crustal composition and mantle heat flow: Implications
Cuney, M., Barbey, P., 2014. Uranium, rare metals, and granulite-facies metamorphism. from surface heat flow and radiogenic heat production in the Variscan Erzgebirge
Geosci. Frontiers 5, 729–745. (Germany). J. Geophys. Res. 105, 27917–27938.
Cuney, M., Friedrich, 1987. Physicochemical and crystal-chemical controls on accessory Förster, B., Haack, U., 1995. U/Pb-Datierungen von Pechblenden und die hydrothermale
mineral paragenesis in granitoïds. Implications on uranium metallogenesis. Bull. Entwicklung der U-Lagerstätte Aue-Niederschlema (Erzgebirge). Z. Geol. Wiss. 23,
Minéral. 110, 235–247. 581–588.
Cuney, M., Kyser, K., 2009. Recent and not-so-recent developments in uranium deposits Förster, H.-J., Romer, R.L., 2010. Carboniferous magmatism. In: Linnemann, U., Romer,
and implications for exploration. Mineral. Assoc. Canada, Short Course Ser. 39, 257. R.L. (Eds.), Pre-Mesozoic Geology of Saxo-Thuringia – From the Cadomian Active
Cuney, M., Kyser, K., 2015. Geology and geochemistry of uranium and thorium deposits. Margin to the Variscan Orogen. Schweizerbart Sci. Publ, Stuttgart, pp. 287–308.
Mineral. Assoc. Canada, Short Course Ser. 46, 345 pp. Förster, H.-J., Tischendorf, G., Trumbull, R.B., Gottesmann, B., 1999. Late-collisional
Cuney, M., Friedrich, M., Blumenfeld, P., Bourguignon, A., Boiron, M.C., Vigneresse, J.L., granites in the Variscan Erzgebirge (Germany). J. Petrol. 40, 1613–1645.
Poty, B., 1990. Metallogenesis in the French part of the Variscan orogen. Part I: U Förster, H.-J., Rhede, D., Hecht, L., 2008. Chemical composition of radioactive accessory
preconcentrations in pre-Variscan and Variscan formations – a comparison with Sn, minerals: implications to the evolution, alteration, age, and uranium fertility of the
W and Au. Tectonophysics 177, 39–57. Fichtelgebirge granites (NE Bavaria, Germany). N. Jb. Mineral. Abh. 185, 161–182.
Cuney, M., Brouand, M., Stussi, J.M., Virlojeux, D., 2001. Le complexe plutonique de Förster, H.-J., Romer, R.L., Gottesman, B., Tischendorf, G., Rhede, D., 2009. Are the
Charroux-Civray (Vienne): témoin du magmatisme infra-carbonifère dans le segment granites of the Aue-Schwarzenberg Zone (western Erzgebirge, Germany) a major
occidental de la chaîne varisque européenne. Géologie de la France 1–2, 143–166. source for metalliferous ore deposits? a geochemical, Sr–Nd–Pb isotopic, and geo-
Cuney, M., Emetz, A., Mercadier, J., Mykchaylov, V., Shunko, V., Yuslenko, A., 2012. chronological study. N. Jb. Mineral. Abh. 186, 163–184.
Uranium deposits associated with Na-metasomatism from central Ukraine: a review Friedrich, M., Cuney, M., Poty, B., 1987. Uranium geochemistry in peraluminous leuco-
of some of the major deposits and genetic constraints. Ore Geol. Rev. 44, 82–106. granites. Uranium 3, 353–385.
Dabard, M.-P., Paris, F., 1986. Palaeontological and geochemical characteristics of Fyfe, W.S., Brown, G.C., 1979. The geochemical cycle of uranium. Phil. Trans. R. Soc.
Silurian black shale formations from the Central Brittany Domain of the Armorican Lond., Ser. A, Math. Phys. Sci. 433–445.
Massif (northwest France). Chem. Geol. 55, 17–29. Ganade de Araujo C.E., Rubatto, D., Hermann, J., Cordani, U.G., Caby, R., Basei, M.A.S.,
Dahlkamp, F.J., 1993. Uranium Ore Deposits. Springer-Verlag, Berlin, pp. 460. 2014. Ediacaran 2,500-km-long synchronous deep continental subduction in the West
Dahlkamp, F.J., 2016. Uranium Deposits of the World – Europe. Springer-Verlag, Berlin, Gondwana Orogen. Nature Communications 5, Article number: 5198, doi: 10.1038/
pp. 792. ncomms6198.
Danišík, M., Štěpančíková, P., Evans, N.J., 2012. Constraining long‐term denudation and Gébelin, A., Roger, F., Brunel, M., 2009. Syntectonic crustal melting and high-grade
faulting history in intraplate regions by multisystem thermochronology: an example metamorphism in a transpressional regime, Variscan Massif Central, France.
of the Sudetic Marginal Fault (Bohemian Massif, central Europe). Tectonics 31, 19. Tectonophysics 477, 229–243.
https://doi.org/10.1029/2011TC003012. Geisler, T., Pidgeon, R.T., Kurtz, R., Bronswijk, W.V., Schleicher, H., 2003. Experimental
Darnley, A.G., English, T.H., Sprake, O., Preece, T.H., Avery, D., 1965. Ages of uraninite hydrothermal alteration of partially metamict zircon. Am. Mineral. 88, 1496–1513.
and coffinite from south-west England. Mineral. Mag. 34, 159–176. Gilleaudeau, G.J., Kah, L.C., 2013. Oceanic molybdenum drawdown by epeiric sea ex-
Delabroye, A., Vecoli, M., 2010. The end-Ordovician glaciation and the Hirnantian Stage: pansion in the Mesoproterozoic. Chem. Geol. 356, 21–37.
a global review and questions about Late Ordovician event stratigraphy. Earth-Sci. Gómez, J.J., Goy, A., 2005. Late Triassic and Early Jurassic palaeogeographic evolution
Rev. 98, 269–282. and depositional cycles of the Western Tethys Iberian platform system (Eastern
Denèle, Y., Laumonier, B., Paquette, J.L., Olivier, P., Gleizes, G., Barbey, P., 2014. Timing Spain). Palaeogeogr., Palaeoclimat., Palaeoecol. 222, 77–94.
of granite emplacement, crustal flow and gneiss dome formation in the Variscan Hahn, T., Kroner, U., Melzer, P., 2010. Early Carboniferous synorogenic sedimentation in
segment of the Pyrenees. Geol. Soc., Lond Spec. Publ. 405, 265–287. the Saxo-Thuringian Basin and the adjacent Allochthonous Domain. In: Linnemann,
Derome, D., Cathelineau, M., Cuney, M., Fabre, C., Lhomme, T., Banks, D.A., 2005. U., Romer, R.L. (Eds.), Pre-Mesozoic Geology of Saxo-Thuringia – From the Cadomian
Mixing of sodic and calcic brines and uranium deposition at McArthur River, Active Margin to the Variscan Orogen. Schweizerbart, Stuttgart, pp. 171–192.
Saskatchewan, Canada: a Raman and laser-induced breakdown spectroscopic study of Hall, S.M., Mihalasky, M.J., Tureck, K.R., Hammarstrom, J.M., Hannon, M.T., 2017.
fluid inclusions. Econ. Geol. 100, 1529–1545. Genetic and grade and tonnage models for sandstone-hosted roll-type uranium de-
Dill, H.G., 1986. Fault-controlled uranium black ore mineralization from the Western posits, Texas Coastal Plain, USA. Ore Geol. Rev. 80, 716–753.
edge of the Bohemian Massif (NE Bavaria, FR Germany). TECDOC-361 In: Fuchs, H. Hassan, H.H., Hale, W.E., Chrzanowski, M., 1987. Geology of uranium and associated
(Ed.), Vein Type Uranium Deposits. IAEA, Vienna, pp. 275–291. elements in New Brunswick. Geol. Surv. Can., Open File 1769, 40 pp.
Dill, H., 1987. Environmental and diagenetic analyses of Lower Permian epiclastic and Hauzenberger, C.A., Bauernhofer, A.H., Hoinkes, G., Wallbrecher, E., Mathu, E.M., 2004.
pyroclastic fan deposits – their role for coal formation and uranium metallogeny in Pan-African high pressure granulites from SE-Kenya: petrological and geothermo-
the Stockheim Trough (FRG). Sed. Geol. 52, 1–26. barometric evidence for a polycyclic evolution in the Mozambique belt. J. Afr. Earth
Dolníček, Z., René, M., Hermannová, S., Prochaska, W., 2014. Origin of the Okrouhlá Sci 40, 245–268.
Radouň episyenite-hosted uranium deposit, Bohemian Massif, Czech Republic: fluid Hawkesworth, C., Cawood, P., Dhuime, B., 2013. Continental growth and the crustal
inclusion and stable isotope constraints. Mineral. Depos. 49, 409–425. record. Tectonophysics 609, 651–660.
Dostal, D., Kontak, D.J., Karl, S.M., 2014. The Early Jurassic Bokan Mountain peralkaline Heuse, T., Schauer, M., Maletz, J., Meisel, S., 2006. Thüringisch-Fränkisches
granitic complex (southeastern Alaska): geochemistry, petrogenesis and rare-metal Schiefergebirge. In: Heuse, T., Leonhardt, D. (Eds.), Stratigraphie von Deutschland
mineralization. Lithos 202–203, 395–412. VII. Silur. Schriftenr. Deut. Ges. Geowiss, pp. 85–91.
Ducrot, J., Lancelot, J.R., Marchand, J., 1983. Datation U-Pb sur zircons de l'éclogite de la Herrmann, S., Seifert, T., Baumann, L., 1995. Zur Ausbildung von spätvariszischen Quarz-
Borie (Haut-Allier, France) et conséquences sur l'évolution anté-hercynienne de Polymetall-Mineralisationen im Lagerstättenrevier Schneeberg. Z. geol. Wiss. 23,
l'Europe Occidentale. Earth Planet. Sci. Lett. 18, 97–113. 573–580.
El Jarray, A., Boiron, M.-C., Cathelineau, M., 1994. Percolation microfissurale de vapeurs Hiller, A., Schuppan, W., 2008. Geologie und Uranbergbau im Revier Schlema-Alberoda.
aqueuses dans le granite de Pény (Massif de Saint-Sylvestre, Massif Central): relation Bergbau in Sachsen 14, 171.
avec la dissolution du quartz. C. R. Sci. Paris 318, 1095–1102. Hiller, A., Schuppan, W., 2016. Das Lagerstättengebiet Zobes-Bergen im Vogtland und
Ekert, V., Mužák, J., 2010. Mining and remediation at the Stráž pod Ralskem uranium benachbarte Uranvorkommen im Bereich des Bergener Granitmassivs. Bergbau in
deposit. GeoScience Eng. 61 (3), 1–6. Sachsen 18, 160.
Evert, L., Shock, D.C., Sassani, C., Betz, H., 1997. Uranium in geologic fluids: estimates of Hofmann, B., 1989. Genese, Alteration und rezentes Fliess-System der Uranlagerstiitte
standard partial molar properties, oxidation potentials, and hydrolysis constants at Krunkelbach (Menzenschwand, Südschwarzwald). Natl. Coop. Storage Radioactive
high temperatures and pressures. Geochim. Cosmochim. Acta 61, 4245–4266. Waste Tech. Rept. 88–30, 195.
Faure, M., Lardeaux, J.M., Ledru, P., 2009. A review of the pre-Permian geology of the Hofmann, B., Eikenberg, J., 1991. The Krunkelbach uranium deposit, Schwarzwald,
Variscan French Massif Central. C. R. Geosci. Paris 341, 202–213. Germany: correlation of radiometric ages (U-Pb, U-Xe-Kr, K-Ar, 230Th-234U) with
Fayek, M., Kyser, K., 1997. Characterization of multiple fluid-flow events and rare-earth- mineralogical stages and fluid inclusions. Econ. Geol. 86, 1031–1049.
element mobility associated with the formation of unconformity-type uranium de- Holliger, P., Pagel, M., Pironon, J., 1989. A model for 238U radioactive daughter loss from
posits in the Athabasca Basin. Saskatchewan. Can. Mineral. 35, 627–658. sediment-hosted pitchblende deposits and the Late Permian-Early Triassic deposi-
Fischer, C., Schmidt, C., Bauer, A., Gaupp, R., Heide, K., 2009. Mineralogical and geo- tional U-Pb age of the Müllenbach uranium ore (Baden-Württemberg, F.R.G.). Chem.
chemical alteration of low-grade metamorphic black slates due to oxidative weath- Geol. 80, 45–53.
ering. Chem. Erde 69, 127–142. Hou, B., Keeling, J., Li, Z., 2017. Paleovalley-related uranium deposits in Australia and
Förster, B., 1996. U/Pb Datierung an Pechblenden der U-Lagerstätte Aue-Niederschlema China: a review of geological and exploration models and methods. Ore Geol. Rev.
(Erzgebirge). PhD thesis Univ. Giessen, Germany, 212 pp + appendices. 88, 201–234.
Förster, H.-J., 1998a. The chemical composition of REE–Y–Th–U-rich accessory minerals Huertas, F.J., Gervilla, F., Gwatkin, C., 2013. Uranium mineralization in the
from peraluminous granites of the Erzgebirge-Fichtelgebirge region, Germany. Part I: Retortillo–Santidad Area (Salamanca, Spain): role of late alteration. In: Jonsson, E.
The monazite-(Ce) – brabantite solid solution series. Am. Mineral. 83, 259–272. et al. (Eds.) Mineral Deposit Research for a High-Tech World. Proc. 12th Biennial SGA
Förster, H.-J., 1998b. The chemical composition of REE–Y–Th–U-rich accessory minerals Meeting, 1594–1597.
from peraluminous granites of the Erzgebirge-Fichtelgebirge region, Germany. Part Huertas, F.J., Gervilla, F., Butcher, A.R., 2015. Mineralogy of the Uranium Ore in the

499
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Retortillo-Santidad Deposit (Salamanca, Spain): a QEMSCAN Study. In: André-Meyer, Kyser, K., Hiatt, E., Renac, C., Durocher, K., Holk, G., Deckart, K., 2000. Diagenetic fluids
A.-S. et al. (Eds.) Mineral Resources in a Sustainable World. 13th Biennial SGA in Paleo-and Meso-Proterozoic sedimentary basins and their implications for long
Meeting, 1811–1814. protracted fluid histories. In: Kyser, T.K. (Ed.) Fluids and Basin Evolution. Min. Assoc.
IAEA, 1984. Surficial Uranium Deposits. IAEA, Vienn, pp. 252 IEAEA-TECDOC-322. Can. Short Course 28, 225–262.
IAEA, 2018. Geological Classification of Uranium Deposits with Description of Selected Lancelot, J., Vella, V., 1989. Liassic U-Pb age of the Rabejac pitchblende - isotopic ar-
Examples. IAEA, Vienna, pp. 428 IAEA-TECDOC-1842. guments for a Permian uraniferous preconcentration in the Lodeve basin (Herault).
IRSN, 2007. Inventaire national des sites miniers d’uranium. Institut de Radioprotection Bull. Soc. Géol. France 5, 309–315.
et de Sûreté Nucléaire, Fontenay-aux-Roses, pp. 315. Lancelot, J.R., de Saint André, B., de La Boisse, H., 1984. Systematique U-Pb et évolution
Idir, E.H., Renard, J.P., 2002. Synthèse des résultats de l'exploration de l'uranium dans du gisement d'uranium de Lodève (France). Mineral. Depos. 19, 44–53.
l'Anti-Atlas (Maroc). Mines Géologie et Energie Rabat 58, 69–94. Lange, G., Freyhoff, G., 1991. Geologie und Bergbau in der Uranlagerstätte Ronneburg/
Ingham, E.S., Cook, N.J., Cliff, J., Ciobanu, C.L., Huddleston, A., 2014. A combined Thüringen. Erzmetall 44, 264–269.
chemical, isotopic and microstructural study of pyrite from roll-front uranium de- Langmuir, D., 1978. Uranium solution-mineral equilibria at low temperatures with ap-
posits, Lake Eyre Basin, South Australia. Geochim. Cosmochim. Acta 125, 440–465. plications to sedimentary ore deposits. Geochim. Cosmochim. Acta 42, 547–569.
Isbell, J.L., Lenaker, P.A., Askin, R.A., Miller, M.F., Babcock, L.E., 2003. Reevaluation of Lardeaux, J.M., Ledru, P., Daniel, I., Duchène, S., 2001. The Variscan French Massif
the timing and extent of late Paleozoic glaciation in Gondwana: role of the trans- Central – a new addition to the ultra-high pressure metamorphic “club”: exhumation
antarctic mountains. Geology 31, 977–980. processes and geodynamic consequences. Tectonophysics 323, 143–167.
Jaireth, S., McKay, A., Lambert, I., 2008. Association of large sandstone uranium deposits Laurent, O., Couzinié, S., Zeh, A., Vanderhaeghe, O., Moyen, J.F., Villaros, A., Gardien,
with hydrocarbons. AusGeo News 89, 8–12. V., Chelle-Michou, C., 2017. Protracted, coeval crust and mantle melting during
Jaireth, S., Roach, I.C., Bastrakov, E., Liu, S., 2016. Basin-related uranium mineral sys- Variscan late-orogenic evolution: U-Pb dating in the eastern French Massif Central.
tems in Australia: a review of critical features. Ore Geol. Rev. 76, 360–394. Int. J. Earth Sci. 106, 421–451.
Janoušek, V., Holub, F.V., 2007. The causal link between HP-HT metamorphism and Carlier, Le, de Veslud, C., Alexandre, P., Ruffet, G., Cuney, M., Cheilletz, A., 2013. A two-
ultrapotassic magmatism in collisional orogens: case study from the Moldanubian stage exhumation in Western French Massif Central: new geochronological evidences
Zone of the Bohemian Massif. Proc. Geologists' Assoc. 118, 75–86. of syn-collisional extension. Lithos 175–176, 1–15.
Janoušek, V., Bowes, D.R., Rogers, G., Farrow, C.M., Jelínek, E., 2000. Modelling diverse Le Heron, D.P., Khoukhi, Y., Paris, F., Ghienne, J.F., Le Herissé, A., 2008. Black shale,
processes in the petrogenesis of a composite batholith: the Central Bohemian Pluton, grey shale, fossils and glaciers: anatomy of the Upper Ordovician-Silurian succession
Central European Hercynides. J. Petrol. 41, 511–543. in the Tazzeka Massif of eastern Morocco. Gondwana Res. 14, 483–496.
Janoušek, V., Finger, F., Roberts, M., Frýda, J., Pin, C., Dolejš, D., 2004a. Deciphering the Le Heron, D.P., Craig, J., Etienne, J.L., 2009. Ancient glaciations and hydrocarbon ac-
petrogenesis of deeply buried granites: whole-rock geochemical constraints on the cumulations in North Africa and the Middle East. Earth-Sci. Rev. 93, 47–76.
origin of largely undepleted felsic granulites from the Moldanubian Zone of the LeBoutillier, N.G., 2002. The tectonics of Variscan magmatism and mineralisation in
Bohemian Massif. Trans. R. Soc. Edinburgh: Earth Sci. 95, 141–159. South West England. Ph.D. Thesis. University of Exeter, pp. 712.
Janoušek, V., Braithwaite, C.J.R., Bowes, D.R., Gerdes, A., 2004b. Magma-mixing in the Lecomte, A., Cathelineau, M., Deloule, E., Brouand, M., Peiffert, C., Loukola-Ruskeeniemi,
genesis of Hercynian calc-alkaline granitoids: an integrated petrographic and geo- K., Pohjolainen, E., Lahtinen, H., 2014. Uraniferous bitumen nodules in the
chemical study of the Sázava intrusion, Central Bohemian Pluton, Czech Republic. Talvivaara Ni–Zn–Cu–Co deposit (Finland): influence of metamorphism on uranium
Lithos 78, 67–99. mineralization in black shales. Mineral. Depos. 49, 513–533.
Joly, A., Martelet, G., Chen, Y., Faure, M., 2008. A multidisciplinary study of a syntec- Ledru, P., Lardeaux, J.M., Santallier, D., Autran, A., Quenardel, J.M., Floc’h, J.P.,
tonic pluton close to a major lithospheric-scale fault—Relationships between the Lerouge, G., Maillet, N., Marchand, J., Ploquin, A., 1989. Où sont les nappes dans le
Montmarault granitic massif and the Sillon Houiller Fault in the Variscan French Massif Central Français? Bull. Soc. Geol. France 3, 605–618.
Massif Central: 2. Gravity, aeromagnetic investigations, and 3-D geologic modeling. Ledru, P., Courrioux, G., Dallain, C., Lardeaux, J.M., Montel, J.M., Vanderhaeghe, O.,
J. Geophys. Res. 113, B01404. https://doi.org/10.1029/2006JB004744. Vitel, G., 2001. The Velay Dome (French Massif Central): melt generation and granite
Kley, J., Voigt, T., 2008. Late Cretaceous intraplate thrusting in central Europe: effect of emplacement during orogenic evolution. Tectonophysics 342, 207–237.
Africa-Iberia-Europe convergence, not Alpine collision. Geology 36, 839–842. Lericolais, G., Guichard, F., Morigi, C., Popescu, I., Bulois, C., Gillet, H., Ryan, W.B.F.,
Kley, J., Franzke, H.J., Jähne, F., Krawczyk, C., Lohr, T., Reicherter, K., Scheck- 2011. Assessment of Black Sea water-level fluctuations since the Last Glacial
Wenderoth, M., Sippe, J., Tanner, D., van Gent, H., the SPP Structural Geology Group, Maximum. Geol. Soc. Am. Spec. Pap. 473, 33–50.
2008. Strain and stress (Chapter 3.3). In: Littke, R., Bayer, U., Gajewski, D., Leroy, J., 1978. The Margnac and Fanay uranium deposits of the La Crouzille District
Nelskamp, S. (Eds.), Dynamics of Complex Intercontinental Basins – The Central (Western Massif Central, France): geologic and fluid inclusion studies. Econ. Geol. 73,
European Basin System. Springer, Berlin-Heidelberg, pp. 234–245. 1611–1634.
Kober, B., Kalt, A., Hanel, M., Pidgeon, R.T., 2004. SHRIMP dating of zircons from high- Leroy, J., 1984. Episyénitisation dans le gisement d'uranium du Bernardan (Marche):
grade metasediments of the Schwarzwald/SW-Germany and implications for the comparaison avec des gisements similaires du nord-ouest du Massif Central français.
evolution of the Moldanubian basement. Contrib. Mineral. Petrol. 147, 330–345. Mineral. Depos 19, 26–35.
Kontak, D.J., Cormier, R.F., 1991. Geochronological evidence for multiple tectono- Leroy, J., Holliger, P., 1984. Mineralogical, chemical and isotopic (U-Pb method) studies
thermal overprinting events in the East Kemptville muscovite–topaz leucogranite, of Hercynian uraniferous mineralizations (Margnac and Fanay mines, Limousin,
Yarmouth County, Nova Scotia, Canada. Can. J. Earth Sci. 28, 209–224. France). Chem. Geol. 45, 121–134.
Kreuzer, O.P., Markwitz, V., Porwal, A.K., McCuaig, T.C., 2010. A continent-wide study of Leroy, J.L., Turpin, L., 1988. REE, Th and U behaviour during hydrothermal and super-
Australia's uranium potential: Part I: GIS-assisted manual prospectivity analysis. Ore. gene processes in a granitic environment. Chem. Geol. 68, 239–251.
Geol. Rev. 38, 334–366. Lespinasse, M., Cathelineau, M., 1990. Fluid percolation in a fault zone: a study of fluid
Kříbek, B., Žák, K., Spangenberg, J., Jehlička, J., Prokeš, S., Komínek, J., 1999. Bitumens inclusion planes in the St. Sylvestre granite, north west Massif Central, France.
in the Late Variscan hydrothermal vein-type uranium deposit of Příbram, Czech Tectonophysics 184, 173–187.
Republic: sources, radiation-induced alteration, and relation to mineralization. Econ. Lespinasse, M., Pécher, A., 1986. Microfracturing and regional stress field: a study of
Geol. 94, 1093–1114. preferred orientations of fluid inclusion planes in a granite from the Massif Central,
Kříbek, B., Žák, K., Dobeš, P., Leichmann, J., Pudilová, M., René, M., Scharm, B., France. J. Struct. Geol. 8, 169–180.
Scharmová, M., Hájek, A., Holeczy, D., Hein, U.F., Lehmann, B., 2009. The Rožná Leventhal, J.S., 1991. Comparison of organic geochemistry and metal enrichment in two
uranium deposit (Bohemian Massif, Czech Republic): shear zone-hosted, late black shales: Cambrian Alum Shale of Sweden and Devonian Chattanooga Shale of
Variscan and post-Variscan hydrothermal mineralization. Miner. Depos. 44, 99–128. United States. Mineral. Depos. 26, 104–112.
Kroner, U., Romer, R.L., 2010. The Saxo-Thuringian Zone – tip of the Armorican spur and Lévêque, M.H., Lancelot, J.R., George, E., 1988. The Bertholène uranium deposit –
part of the Gondwana plate. In: Linnemann, U., Romer, R.L. (Eds.), Pre-Mesozoic Mineralogical characteristics and U-Pb dating of the primary U mineralization and its
Geology of Saxo-Thuringia – From the Cadomian Active Margin to the Variscan subsequent remobilization: consequences upon the evolution of the U deposits of the
Orogen. Schweizerbart, Stuttgart, pp. 371–394. Massif Central, France. Chem. Geol. 69, 147–163.
Kroner, U., Romer, R.L., 2013. Two plates - many subduction zones: the Variscan orogeny Lidmar-Bergström, K., 1996. Long term morphotectonic evolution in Sweden.
reconsidered. Gondw. Res. 24, 298–329. Geomorphology 16, 33–59.
Kroner, U., Hahn, T., Romer, R.L., Linnemann, U., 2007. The Variscan orogeny in the Lidmar-Bergström, K., Näslund, J.O., 2002. Landforms and uplift in Scandinavia. In: Doré,
Saxo-Thuringian zone – heterogenous overprint of Cadomian/Palaeozoic Peri- A.G., Cartwright, J.A., Stoker, M.S., Turner, J.R., White, N. (Eds.), Exhumation of the
Gondwana Crust. In: Linnemann, U., Nance, D., Kraft, P., Zulauf, G. (Eds.), The North Atlantic Margin: Timing, Mechanisms and Implications for Petroleum
Evolution of the Rheic Ocean: From Avalonian-Cadomian active margin to Exploration. Geol. Soc., Lond., Spec. Publ., pp. 103–116.
Alleghenian-Variscan collision. Geol. Soc. Am. Spec. Pap., pp. 153–172. Lindström, M., 1972. Ice-marked sand grains in the Lower Ordovician. Geol. et Paleont. 6,
Kroner, U., Romer, R.L., Linnemann, U., 2010. The Saxo-Thuringian Zone of the Variscan 25–31.
Orogen as part of Pangea. In: Linnemann, U., Romer, R.L. (Eds.), Pre-Mesozoic Linnemann, U., Romer, R.L., 2002. The Cadomian orogeny in Saxo-Thuringia, Germany:
Geology of Saxo-Thuringia – From the Cadomian Active Margin to the Variscan geochemical and Nd-Sr-Pb isotopic characterization of marginal basins with con-
Orogen. Schweizerbart, Stuttgart, pp. 3–16. straints to geotectonic setting and provenance. Tectonophysics 352, 33–64.
Kroner, U., Roscher, M., Romer, R.L., 2016. Ancient plate kinematics derived from the Linnemann, U., Gehmlich, M., Tichomirowa, M., Buschmann, B., Nasdala, L., Jonas, P.,
deformation pattern of continental crust: Paleo- and Neo-Tethys opening coeval with Lützner, H., Bombach, K., 2000. From Cadomian Subduction to Early Palaeozoic
prolonged Gondwana-Laurussia convergence. Tectonophysics 681, 220–233. Rifting: The Evolution of Saxo-Thuringia at the margin of Gondwana in the light of
Kulling, O., 1964. Översikt över norra Norrbottensfjällens kaledonidberggrund. Sver. single zircon geochronology and basin development (Central European Variscides,
Geol. Ser. Ba 19, 166. Germany). In: Franke, W., Haak, V., Oncken, O., Tanner, D. (Eds.), Orogenic
Kuschka, E., 2002. Die Uranerz-Baryt-Fluorit-Lagerstätte Niederschlag bei Bärenstein und Processes – Quantification and Modelling in the Variscan Belt of Central Europe.
benachbarte Erzvorkommen. Bergbau in Sachsen 6, 219 pp. Geol. Soc., Lond., Spec. Publ., pp. 131–153.
Kyser, T.K., 2007. Fluids, basin analysis, and mineral deposits. Geofluids 7, 238–257. Linnemann, U., McNaughton, N.J., Romer, R.L., Gehmlich, M., Drost, K., Tonk, C., 2004.

500
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

West African Provenance for Saxo-Thuringia (Bohemian Massif): Did Armorica ever the Late Ordovician– early Silurian: review and new insights from black shales and
leave pre-Pangean Gondwana? – U/Pb-SHRIMP zircon evidence and the Nd-isotopic nitrogen isotopes. Geol. Soc. Am. Bull. 125, 1635–1670.
record. Int. J. Earth Sci. 93, 683–705. Meldrum, A., Boatner, L.A., Weber, W.J., Ewing, R.C., 1998. Radiation damage in zircon
Linnemann, U., D́Lemos, R., Drost, K., Romer, R.L., Samson, S.D., Strachan, R., 2008. and monazite. Geochim. Cosmochim. Acta 62, 2509–2520.
Cadomian tectonics. In: McCann, T. (Ed.), The Geology of Central Europe. Geol. Soc. Melleton, J., Cocherie, A., Faure, M., Rossi, P., 2010. Precambrian protoliths and Early
London, pp. 103–154. Paleozoic magmatism in the French Massif Central: U-Pb data and the North
Linnemann, U., Romer, R.L., Gerdes, A., Jeffries, T.E., Drost, K., Ulrich, J., 2010a. The Gondwana connection in the west European Variscan belt. Gondw. Res. 17, 13–25.
Cadomian Orogeny in the Saxo-Thuringian Zone. In: Linnemann, U., Romer, R.L. Meshik, A.P., Lippolt, H.J., Dymkov, Yu.M., 2000. Xenon geochronology of Schwarzwald
(Eds.), Pre-Mesozoic Geology of Saxo-Thuringia – From the Cadomian Active Margin pitchblendes. Mineral. Depos. 35, 190–205.
to the Variscan Orogen. Schweizerbart, Stuttgart, pp. 37–58. Meunier, J.D., Brulhet, J., Pagel, M., 1992. Uranium mobility in the sediment-hosted
Linnemann, U., Hofmann, M., Romer, R.L., Gerdes, A., 2010b. Transitional stages be- uranium deposit of Coutras, France. Appl. Geochem. 7, 111–121.
tween the Cadomian and Variscan orogenies: basin development and tectono-mag- Meunier, J.D., Trouiller, A., Brulhet, J., Pagel, M., 1989. Uranium and organic matter in a
matic evolution of the southern margin of the Rheic Ocean at the Saxo-Thuringian paleodeltaic environment: the Coutras deposit (Gironde, France). Econ. Geol. 84,
Zone (North Gondwana shelf). In: Linnemann, U., Romer, R.L. (Eds.), Pre-Mesozoic 1541–1556.
Geology of Saxo-Thuringia – From the Cadomian Active Margin to the Variscan Michel, L.A., Tabor, N.J., Montañez, I.P., Schmitz, M.D., Davydov, V.I., 2015.
Orogen. Schweizerbart, Stuttgart, pp. 59–98. Chronostratigraphy and paleoclimatology of the Lodève Basin, France: evidence for a
Lippmaa, E., Maremäe, E., Pihlak, A.T., 2011. Resources, production and processing of pan-tropical aridification event across the Carboniferous-Permian boundary.
Baltoscandian multimetal black shales. Oil Shale 28, 68–77. Palaeogeogr., Palaeoclimat., Palaeoecol. 430, 118–131.
Locutura, J., Tornos, F., Florido, P., Baeza, L., 1990. Metallogeny. In: Dallmeyer, D., Min, M., Xu, H., Chen, J., Fayek, M., 2005. Evidence of uranium biomineralization in
Martínez Garcia, E. (Eds.), Pre-Mesozoic Geology of Iberia. Springer, pp. 321–330. sandstone-hosted roll-front uranium deposits, northwestern China. Ore Geol. Rev. 26,
López-Moro, F.J., Candelas Moro, M., Timón, S.M., Cembranos, M.L., Cózar, J., 2013. 198–206.
Constraints regarding gold deposition in episyenites: the Permian episyenites asso- Mingram, B., 1998. The Erzgebirge, Germany – a subducted part of northern Gondwana:
ciated with the Villalcampo Shear Zone, central western Spain. Int. J. Earth Sci. 102, geochemical evidence for repetition of early Paleozoic metasedimentary sequences in
721–744. metamorphic thrust units. Geol. Mag. 135, 785–801.
López-Moro, J.J., Romer, R.L., López-Plaza, M., Gózalez Sánchez, M., 2017. Zircon and Mohamud, A.H., Sebastián Cózar, J.S., Rodrigo-Naharro, J., Pérez del Villar, L., 2015.
allanite U-Pb ID-TIMS ages of vaugnerites from the Calzadilla Pluton, Salamanca Distribution of U and Th in an Iberian U-fertile granitic complex (NW, Spain): air-
(Spain): dating mantle-derived magmatism and post-magmatic subsolidus overprint. borne-radiometry, chemical and statistical approaches. J. Geochem. Expl. 148,
Geol. Acta 15, 395–408. 40–55.
Lovley, D.R., Phillips, E.J., Gorby, Y.A., Landa, E.R., 1991. Microbial reduction of ur- Montero, P., Bea, F., Zinger, T.F., Scarrow, J.H., Molina, J.F., Whitehouse, M., 2004. 55
anium. Nature 350, 413–416. million years of continuous anatexis in Central Iberia: single-zircon dating of the Pena
Lu, X., Kendall, B., Stein, H.J., Li, C., Hannah, J.L., Gordon, G.W., Ebbestad, J.O.R., 2017. Negra Complex. J. Geol. Soc. 161, 255–263.
Marine redox conditions during deposition of Late Ordovician and Early Silurian Morford, J.L., Emerson, S.R., Breckel, E.J., Kim, S.H., 2005. Diagenesis of oxyanions (V,
organic-rich mudrocks in the Siljan ring district, central Sweden. Chem. Geol. 457, U, Re, and Mo) in pore waters and sediments from a continental margin. Geochim.
75–94. Cosmochim. Acta 69, 5021–5032.
Lüning, S., Kolonic, S., 2003. Uranium spectral gamma-ray response as a proxy for or- Moyen, J.-F., Laurent, O., Chelle-Michou, C., Couzinié, S., Vanderhaeghe, O., Zeh, A.,
ganic richness in black shales: applicability and limitations. J. Petrol. Geol. 26, Villaros, A., Gardien, V., 2017. Collision vs. subduction-related magmatism: two
153–174. contrasting ways of granite formation and implications for crustal growth. Lithos
Lüning, S., Craig, J., Loydell, D.K., Štorch, P., Fitches, B., 2000a. Lower Silurian ‘hot 277, 154–177.
shales’ in North Africa and Arabia: regional distribution and depositional model. Müller, A., Seltmann, R., Halls, C., Siebel, W., Dulski, P., Jeffries, T., Spratt, J., Kronz, A.,
Earth-Sci. Rev. 49, 121–200. 2006. The magmatic evolution of the Land’s End pluton, Cornwall, and associated
Lüning, S., Loyell, D.K., Sutcliffe, O., Ait Salem, A., Zanella, E., Craig, J., Harper, D.A.T., pre-enrichment of metals. Ore Geol. Rev. 28, 329–367.
2000b. Silurian – Lower Devonian black shales in Morocco: which are the organically Nash, J.T., 1981. Geology and genesis of major world hard-rock uranium deposits. An
richest horizons? J. Petrol. Geol. 23, 293–311. overview. U.S. Geol. Surv. Open-File Rep. 81-166, 123 pp.
Lützner, H., Littmann, S., Mädler, J., Romer, R.L., Schneider, J.W., 2006. Stratigraphic Nash, J.T., 2010. Volcanogenic uranium deposits: Geology, geochemical processes, and
and radiometric age data for the continental Permocarboniferous reference-section criteria for resource assessment. US Geol. Surv. Open-File Rep. 1001, 110 pp.
Thüringer-Wald, Germany. In: Wong, Th.E. (Ed.) Proc. XVth Int. Congr. Nash, J.T., Granger, H.C., Adams, S.S., 1981. Geology and concept of genesis of important
Carboniferous and Permian Stratigraphy, Utrecht 2003. Roy. Netherlands Acad. Arts types of uranium deposits. Econ. Geol. 75th Anniv. Vol, 63–116.
Sci., 161–174, Amsterdam. Naumov, G.B., Vlasov, B.P., Golubev, V.N., Mironova, O.F., 2017. The Schlema-Alberoda
Maletz, J., Katzung, G., 2003. Silur. In: Seidel, G. (Ed.), Geologie von Thüringen. five-element uranium deposit, Germany: an example of self-organizing hydrothermal
Schweizerbart Sci. Publ., pp. 130–139. system. Geol. Ore Dep. 59, 1–13.
Marcuello, A., Gómez, P., Carrera, J., Ayora, C., 2006. Multicomponent reactive transport Neiva, A.M.R., Silva, P.B., Corfu, F., Ramos, J.M.F., 2011. Sequential melting and frac-
modeling at the Ratones uranium mine, Cáceres (Spain). J. Iber. Geol. 32, 133–146. tional crystallization: Granites from Guarda-Sabugal area, central Portugal. Chem.
Marfil, R., La Iglesia, A., Estupiñan, J., 2013. Origin and nature of the aluminium phos- Erde 71, 227–245.
phate-sulfate minerals (APS) associated with uranium mineralization in Triassic red- Neumann, E.-R., Wilson, M., Heeremans, M., Spencer, E.A., Obst, K., Timmerman, M.J.,
beds (Iberian Range, Spain). Estudios Geológicos 69 (1), 21–34. Kirstein, L., 2004. Carboniferous-Permian rifting and magmatism in southern
Marignac, C., Cuney, M., 1999. Ore deposits of the French Massif Central: insight into the Scandinavia, the North Sea and northern Germany: a review. In: Wilson, M.,
metallogenesis of the Variscan collision belt. Mineral. Depos. 34, 472–504. Neumann, E.-R., Davies, G.R., Timmerman, M.J., Heeremans, M., Larsen, B.T. (Eds.),
Marshall, V., van Zyl, J., Bryan, S.E., Uysal, T., Gasparon, M., 2010. Comparative pet- Permo-Carboniferous Magmatism and Rifting in Europe. Geol. Soc., London, Spec.
rology & geochemistry of high heat-producing granites in Australia & Europe. Publ., pp. 11–40.
Australian Geothermal Conference 2010, Adelaide, pp. 41–47. Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred
Martin-Izard, A., Arribas Sr., A., Arias, D., Ruiz, J., Fernández, F.J., 2002. The Fé deposit, from major element chemistry of lutites. Nature 299, 715–717.
west-central Spain: tectonic-hydrothermal uranium mineralization associated with Noblet, C., Lefort, J.P., 1990. Sedimentological evidence for a limited separation between
trasnpressional faulting of Alpine age. Can. Mineral. 40, 1505–1520. Armorica and Gondwana during the Early Ordovician. Geology 18, 303–306.
Martínez, F.J., Corretgé, L.G., Suárez, O., 1990. Central Iberian Zone (autochthonous Norman, R.E., 1993. Uranium Production in Eastern Europe and its Environmental
sequences): distribution, characteristics and evolution of metamorphism. In: Impact: A Literature Survey. ORNL/TM-12240. Oak Ridge National Laboratory, pp.
Dallmeyer, D., Martínez Garcia, E. (Eds.), Pre-Mesozoic Geology of Iberia. Springer, 1–28.
pp. 207–211. Nomade, S., Knight, K.B., Beutel, E., Renne, P.R., Verati, C., Féraud, G., Marzoli, A.,
Mathis, V., Robert, J.P., Saint Martin, J., 1990. Geology and metallogeny of the uranium Youbi, N., Bertrand, H., 2007. Chronology of the Central Atlantic Magmatic Province:
deposits in the Lodeve Permian basin (southern Massif Central, France). Chron. Rech. implications for the Central Atlantic rifting processes and the Triassic-Jurassic biotic
Min. 58, 31–40. crisis. Palaeogeogr., Palaeoclimat., Palaeoecol. 244, 326–344.
McKeough, M.A., Lentz, D.R., McFarlane, C.R., Brown, J., 2013. Geology and evolution of O'Brien, P.J., Rötzler, J., 2003. High-pressure granulites: formation, recovery of peak
pegmatite-hosted U-Th ± REE-Y-Nb Mineralization, Kulyk, Eagle, and Karin Lakes conditions and implications for tectonics. J. metamorphic Geol. 21, 3–20.
region, Wollaston Domain, northern Saskatchewan, Canada: examples of the dual OECD-NEA - IAEA. 1992. Uranium 1991: Resources, Production and Demand. OECD,
role of extreme fractionation and hybridization processes. J. Geosci. 58, 321–346. Paris, 280 pp.
McKerrow, W.S., Mac Niocaill, C., Ahlberg, P.E., Clayton, G., Cleal, C.J., Eagar, R.M.C., Oelsner, O.W., 1958. Die erzgebirgischen Granite, ihre Vererzung und die Stellung der Bi-
2000. The late Palaeozoic relations between Gondwana and Laurussia. In: Franke, W., Co-Ni-Formation innerhalb dieser Vererzung. Geologie 7, 682–701.
Haak, V., Oncken, O., Tanner, D. (Eds.), Orogenic Processes: Quantification and Ondruš, P., Veselovský, F., Gabašová, A., Hloušek, J., Šrein, V., 2003a. Geology and
Modelling in the Variscan Belt. Geol. Soc., Lond., Spec. Publ., pp. 9–20. hydrothermal vein system of the Jáchymov (Joachimsthal) ore district. J. Czech Geol.
McManus, J., Berelson, W.M., Severmann, S., Poulson, R.L., Hammond, D.E., Soc. 48 (3–4), 19–147.
Klinkhammer, G.P., Holm, C., 2006. Molybdenum and uranium geochemistry in Ondruš, P., Veselovský, F., Gabašová, A., Hloušek, J., Šrein, V., Vavřín, I., Skála, R.,
continental margin sediments: paleoproxy potential. Geochim. Cosmochim. Acta 70, Sejkora, J., Drábek, M., 2003b. Primary minerals of the Jáchymov ore district. J.
4643–4662. Czech Geol. Soc. 48 (3–4), 3–18.
Meinhold, G., Morton, A.C., Avigad, D., 2013. New insights into peri-Gondwana paleo- Owen, A.D., 1992. Byproduct uranium. Resources Policy 18, 137–147.
geography and the Gondwana super-fan system from detrital zircon U-Pb ages. Pagel, M., 1990. Permian and uranium metallogeny. Chron. Rech. Min. 58, 57–68.
Gondw. Res. 23, 661–665. Pascual, F.J.R., Arenas, R., Catalán, J.R.M., Fernández, L.R.R., Wijbrans, J.R., 2013.
Melchin, M.J., Mitchell, C.E., Holmden, C., Štorch, P., 2013. Environmental changes in Thickening and exhumation of the Variscan roots in the Iberian Central System:

501
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

tectonothermal processes and 40Ar/39Ar ages. Tectonophysics 587, 207–221. France and Related Areas. Springer, pp. 142–151.
Patrier, P., Beaufort, D., Bril, H., Bonhomme, M., Fouillac, A.M., Aumaître, R., 1997. Robertson, D.S., Tilsley, J.E., Hogg, G.M., 1978. The time-bound character of uranium
Alteration-mineralization at the Bernardan U deposit (Western Marche, France): the deposits. Econ. Geol. 73, 1409–1419.
contribution of alteration petrology and crystal chemistry of secondary phases to a Rohrssen, M., Love, G.D., Fischer, W., Finnegan, S., Fike, D.A., 2013. Lipid biomarkers
new genetic model. Econ. Geol. 92, 448–467. record fundamental changes in the microbial community structure of tropical seas
Pécher, A., Lespinasse, M., Leroy, J., 1985. Relations between fluid inclusion trails and during the Late Ordovician Hirnantian glaciation. Geology 41, 127–130.
regional stress field: a tool for fluid chronology. An example of an intragranitic ur- Roig, J.-Y., Faure, M., Maluski, H., 2002. Superimposed tectonic and hydrothermal events
anium ore deposit (northwest Massif Central, France). Lithos 18, 229–237. during the late-orogenic extension in the Western French Massif Central: a structural
Peiffert, C., Cuney, M., Nguyen-Trung, C., 1994. Uranium in granitic magmas: Part 1. and 40Ar/39Ar study. Terra Nova 14, 25–32.
Experimental determination of uranium solubility and fluid-melt partition coeffi- Rolin, P., Marquer, D., Colchen, M., Cartannaz, C., Cocherie, A., Thiery, V., Quenardel, J.-
cients in the uranium oxide-haplogranite-H2O-Na2CO3 system at 720–770 C, 2 kbar. M., Rossi, P., 2009. Famenno-Carboniferous (370–320 Ma) strike slip tectonics
Geochim. Cosmochim. Acta 58, 2495–2507. monitored by syn-kinematic plutons in the French Variscan belt (Massif Armoricain
Peiffert, C., Nguyen-Trung, C., Cuney, M., 1996. Uranium in granitic magmas: Part 2. and French Massif Central). Bull. Soc. géol. France 180, 231–246.
Experimental determination of uranium solubility and fluid-melt partition coeffi- Rolison, J.M., Stirling, C.H., Middag, R., Rijkenberg, M.J.A., 2017. Uranium stable isotope
cients in the uranium oxide-haplogranite-H2O-NaX (X = Cl, F) system at 770°C, 2 fractionation in the Black Sea: Modern calibration of the 238U/235U paleo-redox
kbar. Geochim. Cosmochim. Acta 60, 1515–1529. proxy. Geochim. Cosmochim. Acta 203, 69–88.
Penney, R., 2012. Australian sandstone-hosted uranium deposits. Appl. Earth Sci. 121, Romberger, S.B., 1984. Transport and deposition of uranium in hydrothermal systems at
65–75. temperatures up to 300°C: geological implications. In: DeVivo, B. (Ed.), Uranium
Pérez del Villar, L., Moro, Y.C., de las Candelas, M., 1991. Las mineralizaciones in- Geochemistry, Mineralogy, Geology, Exploration and Resources. Inst Min. Met., pp.
tragraniticas de uranio en el batolito de Cabeza de Araya (Provincia de Cáceres): El 12–17.
Saguazal, brechas NNE y La Zafrilla. Studia Geológica Salmanticensia 27, 85–112. Romer, R.L., 2003. Alpha-recoil in U-Pb geochronology: effective sample size matters.
Pérez del Villar, L., Bruno, J., Campos, R., Gómez, P., Cózar, J.S., Garralón, A., Buil, B., Contrib. Mineral. Petrol. 145, 481–491.
Arcos, D., Carretero, G., 2002. The uranium ore from Mina Fe (Salamanca, Spain) as a Romer, R.L., Hahne, K., 2010. Life of the Rheic Ocean: scrolling through the shale record.
natural analogue of processes in a spent fuel repository. Chem. Geol. 190, 395–415. Gondw. Res. 17, 236–253.
Peyaud, J.B., Barbarand, J., Carter, A., Pagel, M., 2005. Mid-Cretaceous uplift and erosion Romer, R.L., Kroner, U., 2015. Sediment and weathering control on the distribution of
on the northern margin of the Ligurian Tethys deduced from thermal history re- Paleozoic magmatic tin–tungsten mineralization. Miner. Deposita 50, 327–338.
construction. Int. J. Earth Sci. 94, 462–474. Romer, R.L., Kroner, U., 2016. Phanerozoic tin and tungsten mineralization – tectonic
Pfaff, K., Romer, R.L., Markl, G., 2009. U-Pb ages of ferberite, chalcedony, agate, U-mica controls on the distribution of enriched protoliths and heat sources for crustal
and pitchblende: constraints on the mineralization history of the Schwarzwald ore melting. Gondw. Res. 31, 60–95.
district. Eur. J. Mineral. 21, 817–836. Romer, R.L., Kroner, U., 2018. Paleozoic gold in the Appalachians and Variscides. Ore
Pitra, P., Ballèvre, M., Ruffet, G., 2010. Inverted metamorphic field gradient towards a Geol. Rev. 92, 475–505.
Variscan suture zone (Champtoceaux Complex, Armorican Massif, France). J. Romer, R.L., Heinrich, W., Schröder-Smeibidl, B., Meixner, A., Fischer, C.-O., Schulz, C.,
Metamorphic Geol. 28, 183–208. 2005. Elemental dispersion and stable isotope fractionation during reactive fluid-flow
Plant, J.A., Reeder, S., Salminen, R., Smith, D.B., Tarvainen, T., De Vivo, B., Petterson, and fluid immiscibility in the Bufa del Diente aureole, NE-Mexico: evidence from
M.G., 2003. The distribution of uranium over Europe: geological and environmental radiographies and Li, B, Sr, Nd, and Pb isotope systematics. Contrib. Mineral. Petrol.
significance. App. Earth Sci. 112, 221–238. 149, 400–429.
Plášil, J., Sejkora, J., Škoda, R., Škácha, P., 2014. The recent weathering of uraninite from Romer, R.L., Schneider, J.C., Linnemann, U., 2010. Post-Variscan deformation and hy-
the Červená vein, Jáchymov (Czech Republic): a fingerprint of the primary miner- drothermal mineralization in Saxo-Thuringia and beyond: a geochronological review.
alization geochemistry onto the alteration association. J. Geosci. 59, 223–253. In: Linnemann, U., Romer, R.L. (Eds.), Pre-Mesozoic Geology of Saxo-Thuringia –
Poitrasson, F., 2002. In situ investigations of allanite hydrothermal alteration: examples from the Cadomian Active Margin to the Variscan Orogen. Schweizerbart Sci. Publ,
from calc-alkaline and anorogenic granites of Corsica (southeast France). Contrib. Stuttgart, pp. 347–360.
Mineral. Petrol. 142, 485–500. Romer, R.L., Meixner, A., Hahne, K., 2014a. Lithium and boron isotopic composition of
Polito, P.A., Kyser, T.K., Thomas, D., Marlatt, J., Drever, G., 2005. Re-evaluation of the sedimentary rocks – the role of source history and depositional environment: a 250
petrogenesis of the Proterozoic Jabiluka unconformity-related uranium deposit, Ma record from the Cadomian orogeny to the Variscan orogeny. Gondw. Res. 26,
Northern Territory, Australia. Mineral. Depos. 40, 257–288. 1093–1110.
Polito, P.A., Kyser, T.K., Jackson, M.J., 2006. The role of sandstone diagenesis and aquifer Romer, R.L., Meixner, A., Förster, H.-J., 2014b. Lithium and boron in late-orogenic
evolution in the formation of uranium and zinc-lead deposits, southern McArthur granites – isotopic fingerprints for the source of crustal melts? Geochim. Cosmochim.
Basin, Northern Territory, Australia. Econ. Geol. 101, 1189–1209. Acta 131, 98–114.
Polito, P.A., Kyser, K.T., Stanley, C., 2009. The Proterozoic, albitite-hosted, Valhalla ur- Ronov, A.B., Migdisov, A.A., 1996. Time variations in the abundances of rocks, minerals,
anium deposit, Queensland, Australia, a description of the alteration assemblage and elements in the Russian platform sediment cover. Geochem. Int. 33 (5), 30–63.
associated with uranium mineralisation in diamond drill hole V39. Miner. Depos. 44, Roscher, M., Schneider, J.W., 2006. Permo-Carboniferous climate: Early Pennsylvanian to
11–40. Late Permian climate development of central Europe in a regional and global context.
Pollock, J.C., Hibbard, J.P., Sylvester, P.J., 2009. Early Ordovician rifting of Avalonia and In: Lucas, S.G., Cassinis, G., Schneider, J.W. (Eds.), Non-Marine Permian
birth of the Rheic Ocean: U-Pb detrital zircon constraints from Newfoundland. J. Biostratigraphy and Biochronology. Geol. Soc., Lond., Spec. Publ., pp. 95–136.
Geol. Soc. 166, 501–515. Roscoe, S.R., Minter, W.E.L., 1993. Pyritic paleoplacer gold and uranium deposits. In:
Poty, B., Leroy, J., Cathelineau, M., Poty, M., Friedrich, M., Lespinasse, M., Turpin, L., Kirkham, R.V., Sinclair, W.D., Thorpe, R.I., Duke, J. (Eds.), Mineral deposit modeling.
1986. Uranium deposits spatially related to granites in the French part of the Geological Association of Canada, Special Paper 40, pp. 103–124.
Hercynian orogen. In: Vein type U deposits. IAEA-TECDOC-361, 215–246. Rötzler, K., Plessen, B., 2010. The Erzgebirge: a pile of ultrahigh- to low-pressure nappes
Pretorius, D.A., 2012. The nature of the Witwatersrand gold-uranium deposits. Handbook of Early Palaeozoic rocks and their Cadomian basement. In: Linnemann, U., Romer,
of strata-bound and stratiform ore deposits, 7, 29–88. R.L. (Eds.), Pre-Mesozoic Geology of Saxo-Thuringia – from the Cadomian Active
Puelles, P., Ábalos, B., Gil Ibarguchi, J.I., 2005. Metamorphic evolution and thermobaric Margin to the Variscan Orogen. Schweizerbart Sci. Publ, Stuttgart, pp. 253–270.
structure of the subduction-related Bacariza high-pressure granulite formation (Cabo Runge, W. (Ed.), 2001. Chronik der Wismut. Wismut GmbH, Chemnitz, pp. 2739.
Ortegal Complex, NW Spain). Lithos 84, 125–149. Ryan, W.B.F., Major, C.O., Lericolais, G., Goldstein, S.L., 2003. Catastrophic flooding of
Querol, X., Cabrera, Ll, Pickel, W., López-Soler, A., Hagemann, H.W., Fermández-Turiel, the Black Sea. Annu. Rev. Earth Planet Sci. 31, 525–554.
J.L., 1996. Geological controls on the coal quality of the Mequinenza subbituminous Ryan, R.J., O'Beirne-Ryan, A.M., 2009. Uranium occurrences in the Horton Group of the
coal deposit, northeast Spain. Int. J. Coal Geol. 29, 67–91. Windsor area, Nova Scotia and the environmental implications for the Maritimes
Regenspurg, S., Margot-Roquier, C., Harfouche, M., Froidevaux, P., Steinmann, P., Junier, Basin. Atlantic Geol. 45, 171–190.
P., Bernier-Latmani, R., 2010. Speciation of naturally-accumulated uranium in an Saberi, M.H., Rabbani, A.R., Ghavidel-syooki, M., 2016. Hydrocarbon potential and pa-
organic-rich soil of an alpine region (Switzerland). Geochim. Cosmochim. Acta 74, lynological study of the Latest Ordovician – Earliest Silurian source rock (Sarchahan
2082–2098. Formation) in the Zagros Mountains, southern Iran. Marine Petrol. Geol. 71, 12–25.
Reichel, W., Schauer, M., 2006. Das Döhlener Becken bei Dresden – Geologie und Scaillet, S., Cuney, M., Carlier, Le, de Veslud, C., Cheilletz, A., Royer, J.J., 1996. Cooling
Bergbau. Bergbau in Sachsen, Band 12, 343 pp. pattern and mineralization history of the Saint Sylvestre and western Marche leu-
Renac, C., Kyser, T.K., Durocher, K., Dreaver, G., O'Connor, T., 2002. Comparison of di- cogranite pluton, French Massif Central: II. Thermal modelling and implications for
agenetic fluids in the Proterozoic Thelon and Athabasca Basins, Canada: implications the mechanisms of uranium mineralization. Geochim. Cosmochim. Acta 60,
for protracted fluid histories in stable intracratonic basins. Can. J. Earth Sci. 39, 4673–4688.
113–132. Schaltegger, U., Schneider, J.-L., Maurin, J.-C., Fernando Corfu, F., 1996. Precise U-Pb
René, M., 2008. Anomalous rare earth element, yttrium and zirconium mobility asso- chronometry of 345–340 Ma old magmatism related to syn-convergence extension in
ciated with uranium mineralization. Terra Nova 20, 52–58. the Southern Vosges (Central Variscan Belt). Earth Planet. Sci. Lett. 144, 403–419.
Richard, A., Rozsypal, C., Mercadier, J., Banks, D.A., Cuney, M., Boiron, M.C., Schauer, M., 1971. Biostratigraphie und Taxonomie der Graptolithen des tieferen Silurs
Cathelineau, M., 2012. Giant uranium deposits formed from exceptionally uranium- unter besonderer Berücksichtigung der Deformation. Freib. Forsch. H. C273, 1–185.
rich acidic brines. Nature Geosci. 5, 142–146. Schmidt Mumm, A., Wolfgramm, M., 2004. Fluid systems and mineralization in the north
Robardet, M., 2002. Alternative approach to the Variscan Belt in southwestern Europe: German and Polish basin. Geofluids 4, 315–328.
Preorogenic paleobiogeographical constraints. Geol. Soc. Am. Spec. Pap. 364, 1–15. Schmitt, J.M., Thiry, M., 1987. Uranium behaviour in a gossan-type weathering system:
Robardet, M., Bonjour, J.L., Paris, F., Morzadec, P., Racheboeuf, P.R., 1994. Ordovician, example of the Bertholène deposit (Aveyron, France). Bull. Minéral. 110, 197–208.
Silurian, and Devonian of the Medio-North-Armorican Domain. In: Keppie, J.D., Schneider, J.W., Romer, R.L., 2010. The Late Variscan Molasses (Late Carboniferous –
Chantraine, J., Rolet, J., Santallier, D.S., Piquet, A. (Eds.), Pre-Mesozoic Geology in Late Permian) of the Saxo-Thuringian Zone. In: Linnemann, U., Romer, R.L. (Eds.),

502
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Pre-Mesozoic Geology of Saxo-Thuringia – From the Cadomian Active Margin to the Trela, W., Podhalańska, T., Smolarek, J., Marynowski, L., 2016. Llandovery green/grey
Variscan Orogen. Schweizerbart, Stuttgart, pp. 323–346. and black mudrock facies of the northern Holy Cross Mountains (Poland) and their
Schneider, J.W., Körner, F., Roscher, M., Kroner, U., 2006. Permian climate development relation to early Silurian sea-level changes and benthic oxygen level. Sediment. Geol.
in the northern peri-Tethys area—The Lodève basin, French Massif Central, com- 342, 66–77.
pared in a European and global context. Palaeogeogr., Palaeoclimat.,Palaeoecol. 240, Trumbull, R.B., Förster, H.J., Rhede, D., 2010. REE-Y-Th-U-bearing accessory minerals
161–183. and their contribution to the lanthanide and actinide trace-element budget in an
Schwerdtner, G., Anger, H., Störr, M., 2007. Die Kaolinlagerstätten des Kemmlitzer anorogenic granite from the Erongo Complex, NW Namibia. Z. Geol. Wiss. 38,
Reviers. Bergbau in Sachsen 13, 116 pp. 145–165.
Schovsbo, N.H., 2002. Uranium enrichment shorewards in black shales: a case study from Turner-Peterson, C.E., Fishman, N.S., 1986. Geologic synthesis and genetic models for
the Scandinavian Alum Shale. GFF 124, 107–115. uranium mineralization in the Morrison Formation, Grants Uranium Region, New
Schovsbo, N.H., 2003. The geochemistry of Lower Palaeozoic sediments deposited on the Mexico. Am. Assoc. Petrol. Geol Spec Vol. SG 22, 357–388.
margins of Baltica. Bull. Geol. Soc. Denm. 50, 11–27. Turpin, L., Cuney, M., Friedrich, M., Bouchez, J.L., Aubertin, M., 1989. Meta-igneous
Seidel, G. (Ed.), 2003. Geologie von Thüringen. Schweizerbart Science Publisher, origin of Hercynian peraluminous granites in NW French Massif Central: implications
Stuttgart, pp. 601. for crustal history reconstructions. Contrib. Mineral. Petrol. 104, 163–172.
Sejkora, J., Čejka, J., Šrein, V., 2007. Supergene uranium mineralization from Horní Turpin, L., Leroy, J.L., Sheppard, S.M.F., 1990. Isotopic systematics (O, H, C, Sr, Nd) of
Halže near Měděnec (Krušné hory Mountains), Czech Republic. J. Geosci. 52, superimposed barren and U-bearing hydrothermal systems in a Hercynian granite,
199–210. Massif Central, France. Chem. Geol. 88, 85–98.
Škácha, P., Goliáš, V., Sejkora, J., Plášil, J., Strnad, L., Škoda, R., Ježek, J., 2009. Ugidos, J.M., Valladares, M.I., Barba, P., Ellam, R.M., 2003. The Upper Neoproterozoic-
Hydrothermal uranium-base metal mineralization of the Jánská vein, Březové Hory, Lower Cambrian of the Central Iberian Zone, Spain: chemical and isotopic (Sm-Nd)
Příbram, Czech Republic: lead isotopes and chemical dating of uraninite. J. Geosci. evidence that the sedimentary succession records an inverted stratigraphy of its
54, 1–13. source. Geochim. Cosmochim. Acta 67, 2615–2629.
Skirrow, R.G., Jaireth, S., Huston, D.L., Bastrakov, E.N., Schofield, A., van der Wielen, Uličný, D., 2001. Depositional systems and sequence stratigraphy of coarse-grained deltas
S.E., Barnicoat, A.C., 2009. Uranium mineral systems: processes, exploration criteria in a shallow-marine, strike-slip setting: the Bohemian Cretaceous Basin, Czech
and a New Deposit Framework. Geosci. Australia Rec. 20, 44. Republic. Sedimentology 48, 599–628.
Smale, C.V., 1993. South Terras. Cornwall's premier uranium and radium mine. J. R Inst, Uličný, D., Laurin, J., Čech, S., 2009. Controls on clastic sequence geometries in a
Cornwall New Ser. 1, 304–321. shallow-marine, transtensional basin: the Bohemian Cretaceous Basin, Czech
Sørensen, H., Bailey, J.C., Rose-Hansen, J., 2011. The emplacement and crystallization of Republic. Sedimentology 56, 1077–1114.
the U– Th–REE-rich agpaitic and hyperagpaitic lujavrites at Kvanefjeld, Ilímaussaq Vamvaka, A., Siebel, W., Chen, F., Rohrmüller, J., 2014. Apatite fission-track dating and
alkaline complex, South Greenland. Bull. Geol. Soc. Denm. 59, 69–92. low-temperature history of the Bavarian Forest (southern Bohemian Massif). Int. J.
Spirakis, C.S., 1996. The roles of organic matter in the formation of uranium deposits in Earth Sci. 103, 103–119.
sedimentary rocks. Ore Geol. Rev. 11, 53–69. Veevers, J.J., Belousova, E.A., Saeed, A., Sircombe, K., Cooper, A.F., Read, S.E., 2006.
Staude, S., Wagner, T., Markl, G., 2007. Mineralogy, mineral composition and fluid Pan-Gondwanaland detrital zircons from Australia analysed for Hf-isotopes and trace
evolution at the Wenzel hydrothermal deposit, southern Germany: implications for elements reflect an ice-covered Antarctic provenance of 700–500 Ma age, TDM of
the formation of Kongsberg-type silver deposits. Can. Mineral. 45, 1147–1176. 2.0–1.0 Ga, and alkaline affinity. Earth-Sci. Rev. 76, 135–174.
Stephan, T., Kroner, U., Romer, R.L. (in press) The pre-orogenic detrital zircon record of Veríssimo, C.U.V., Santos, R.V., Parente, C.V., de Oliveira, C.G., Cavalcanti, J.A.D., Neto,
the Peri-Gondwana crust. Geol. Mag. https://doi.org/10.1017/S0016756818000031. J.D.A.N., 2016. The Itataia phosphate-uranium deposit (Ceará, Brazil) new petro-
Stussi, J.M., 1989. Granitoid chemistry and associated mineralization in the French graphic, geochemistry and isotope studies. J. South Am. Earth Sci. 70, 115–144.
Variscan. Econ. Geol. 84, 1363–1381. Verniers, J., Maletz, J., Kříž, J., Žigaitė, Ž., Paris, F., Schönlaub, H.P., Wrona, R., 2008.
Sun, Y.Z., Wang, J., Li, S., Jin, K., Lin, M., 2005. Mechanism of uranium accumulation in Silurian. In: McCann, T. (Ed.), The Geology of Central Europe. Geol. Soc London, pp.
the Kupferschiefer from Poland and Germany. Energy Explor. Exploit. 23, 463–473. 249–302.
Surán, J., Veselý, P., 2001. The uranium industry in the history of the Czech Republic and Vigneresse, J.-L., Cuney, M., Jolivet, J., Bienfait, G., 1989. Selective heat producing
recent developments. IAEA-TECDOC 1258, 45–52. elements enrichment in a crustal segment of Brittany (France). Tectonophysics 159,
Tabaud, A.S., Janoušek, V., Skrzypek, E., Schulmann, K., Rossi, P., Whitechurch, H., 47–60.
Guerrot, C., Paquette, J.L., 2015. Chronology, petrogenesis and heat sources for Vogt, P.R., Tucholke, B.E., 1989. The evolution of the North Atlantic. In: Bally, A.W.,
successive Carboniferous magmatic events in the Southern-Central Variscan Vosges Palmer, A.R. (Eds.), The Geology of North America: an overview. Geol. Soc. Am.,
Mts (NE France). J. Geol. Soc. 172, 87–102. Washington, pp. 53–80.
Taylor, P.N., Moorbath, S., Goodwin, R., Petrykowski, A.C., 1980. Crustal contamination Voigt, T., Reicherter, K., von Eynatten, H., Littke, R., Kley, J., 2008. Sedimentation during
as an indicator of the extent of early Archaean continental crust: Pb isotopic evidence basin inversion. In: Littke, R., Bayer, U., Gajewski, D., Nelskamp, S. (Eds.), Dynamics
from the late Archaean gneisses of West Greenland. Geochim. Cosmochim. Acta 44, of Complex Sedimentary Basins. The Example of the Central European Basin System.
1437–1453. Springer, Heidelberg, pp. 211–232.
Teixeira, R.J.S., Neiva, A.M.R., Gomes, M.E.P., Corfu, F., Cuesta, A., Croudace, I.W., 2012. Waber, N., Schorscher, H.D., Peters, T., 1992. Hydrothermal and supergene uranium
The role of fractional crystallization in the genesis of early syn-D3, tin-mineralized mineralization at the Osamu Utsumi mine, Poços de Caldas, Minas Gerais, Brazil. J.
Variscan two-mica granites from the Carrazeda de Ansiães area, northern Portugal. Geochem. Expl. 45, 53–112.
Lithos 153, 177–191. Wenzel, T., Mertz, D.F., Oberhänsli, R., Becker, T., Renne, P.R., 1997. Age, geodynamic
Thickpenny, A., Leggett, J.K., 1987. Stratigraphic distribution and palaeo-oceanographic setting, and mantle enrichment processes of a K-rich intrusion from the Meissen
significance of European early Palaeozoic organic-rich sediments. In: Brooks, J., massif (northern Bohemian massif) and implications for related occurrences from the
Fleet, A.J. (Eds.), Marine Petroleum Source Rocks. Geol. Soc., Lond., Spec. Publ., pp. mid-European Hercynian. Geol. Rundsch. 86, 556–570.
231–247. Wernicke, R.S., Lippolt, H.J., 1997. Evidence of Mesozoic multiple hydrothermal activity
Thomas, R., Webster, J.D., Rhede, D., Seifert, W., Rickers, K., Förster, H.J., Heinrich, W., in the basement at Nonnenmattweiher (southern Schwarzwald), Germany.
Davidson, P., 2006. The transition from peraluminous to peralkaline granitic melts: Mineralium Deposita 32, 197–200.
evidence from melt inclusions and accessory minerals. Lithos 91, 137–149. Wignall, P.B., Myers, K.J., 1988. Interpreting benthic oxygen levels in mudrocks: a new
Timmerman, M.J., 2004. Timing, geodynamic setting and character of Permo- approach. Geology 16, 452–455.
Carboniferous magmatism in the foreland of the Variscan Orogen, NW Europe. In: Williamson, B.J., Shaw, A., Downes, H., Thirlwall, M.F., 1996. Geochemical constraints
Wilson, M., Neumann, E.-R., Davies, G.R., Timmerman, M.J., Heeremans, M., Larsen, on the genesis of Hercynian two-mica leucogranites from the French Massif Central.
B.T. (Eds.), Permo-Carboniferous Magmatism and Rifting in Europe. Geol. Soc., Chem. Geol. 127, 25–42.
Lond., Spec. Publ., pp. 41–74. Wilson, M.R., 1984. Uranium enrichment in European peat bogs. Surficial Uranium
Timón-Sánchez, S.M., López-Moro, F.J., Romer, R.L., Rhede, D., Fernández, A., Moro Deposits. Int. At. Energy Agency, Tecdoc 322, 197–200.
Benito, C., in press. Compositional and textural variations of uraninite reflecting the Wolff, R., Dunkl, I., Kempe, U., von Eynatten, H., 2015. The age of the latest thermal
development of post-Variscan hydrothermal W-Au deposits in the Western Spanish overprint of tin and polymetallic deposits in the Erzgebirge, Germany: constraints
Central System. Geologica Acta. from fluorite (U-Th-Sm)/He thermochronology. Econ. Geol. 110, 2025–2040.
Tischendorf, G., Förster, H.-J., 1994. Hercynian granite magmatism and related me- Yan, D., Wang, H., Fu, Q., Chen, Z., He, J., Gao, Z., 2015. Geochemical characteristics in
tallogenesis in the Erzgebirge: a status report. In: Gehlen, K.V., Klemm, D.D. (Eds.), the Longmaxi Formation (Early Silurian) of South China: implications for organic
Mineral Deposits of the Erzgebirge/Krušné hory (Germany/Czech Republic). Monogr. matter accumulation. Marine Petrol. Geol. 65, 290–301.
Ser. Mineral Depos., pp. 5–23. Young, G.M., Minter, W.E.L., Theron, J.N., 2004. Geochemistry and palaeogeography of
Toens, P.D., Le Roux, J.P., Hartnady, C.J.H., van Biljon, W.J., 1980. The Uranium Geology upper Ordovician glaciogenic sedimentary rocks in the Table Mountain Group, South
and Tectonic Correlation Between the African and Latin American Continents. South Africa. Palaeogeogr., Palaeoclimat., Palaeoecol. 214, 323–345.
African Atomic Energy Board, pp. 80. Žák, Verner, J., Janoušek, K., Holub, V.F.V., Kachlík, V., Finger, F., Hajná, J., Tomek, F.,
Toens, P.D., Le Roux, J.P., 1981. The Permo-Triassic uranium deposits of Southern Africa Vondrovic, L., Trubac, J., 2014. A plate-kinematic model for the assembly of the
within the African-South-American Gondwana framework. In: Uranium Deposits in Bohemian Massif constrained by structural relationships around granitoid plutons. In:
Latin America: Geology and Exploration. IAEA STI/PUB/505, 25–43. Schulmann, K., Martínez Catalán, J.R., Lardeaux, J.M., Janoušek, V., Oggiano, G.
Tonndorf, H., 2000. Die Uranlagerstätte Königstein. Bergbau in Sachsen, Band 7, 208 pp (Eds.), The Variscan Orogeny: Extent, Timescale and the Formation of the European
+ appendices. Crust. Geol. Soc., Lond., Spec. Publ., pp. 169–196.
Tornos, F., Casquet, C., Relvas, J.M.R.S., Barriga, F.J.A.S., Sáez, R., 2002. The relationship Ziegler, P.A., 1987. Evolution of the Western Approaches Trough. In: Ziegler, P.A. (Ed.),
between ore deposits and oblique tectonics: the SW Iberian Variscan Belt. In: Compressional intra-plate deformations in the Alpine foreland. Tectonophysics, pp.
Blundell, D.J., Neubauer, E., von Quadt, A. (Eds.), The Timing and Location of Major 341–346.
Ore Deposits in an Evolving Orogen. Geol. Soc., Lond., Spec. Publ., pp. 179–198. Ziegler, P.A., 1988. Post-Hercynian plate reorganization in the Tethys and Arctic–North

503
R.L. Romer, M. Cuney Ore Geology Reviews 102 (2018) 474–504

Atlantic domains. In; Manspeizer, W. (Ed.) Triassic-Jurassic Rifting. Developments in Zielinski, R.A., Burruss, R.C., 1991. Petrogenesis and geological history of a uranium
Geotectonics 22, 711–755. source rock: A case study in northeastern Washington, USA. Appl. Geochem. 6,
Ziegler, P.A., 1990. Geological Atlas of Western and Central Europe (2 ed.). Shell Int. 597–612.
Petrol. Mij. B.V., distrib. by Geol. Soc. Publ. House Bath. pp. 239, 56 encl. ISBN 0- Zindler, A., Hart, S., 1986. Chemical geodynamics. Ann. Rev. Earth Planet. Sci. 14,
444-42084-3. 493–571.

504

Das könnte Ihnen auch gefallen