• Aucun résultat trouvé

The significance of galena Pb model ages and the formation of large Pb-Zn sedimentary deposits

N/A
N/A
Protected

Academic year: 2022

Partager "The significance of galena Pb model ages and the formation of large Pb-Zn sedimentary deposits"

Copied!
60
0
0

Texte intégral

(1)

HAL Id: hal-03318657

https://hal.archives-ouvertes.fr/hal-03318657

Submitted on 10 Aug 2021

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires

The significance of galena Pb model ages and the formation of large Pb-Zn sedimentary deposits

Jean Milot, Janne Blichert-Toft, Mariano Ayarzagüena Sanz, Nadège Fetter, Philippe Télouk, Francis Albarede

To cite this version:

Jean Milot, Janne Blichert-Toft, Mariano Ayarzagüena Sanz, Nadège Fetter, Philippe Télouk, et al..

The significance of galena Pb model ages and the formation of large Pb-Zn sedimentary deposits.

Chemical Geology, Elsevier, 2021, �10.1016/j.chemgeo.2021.120444�. �hal-03318657�

(2)

The significance of galena Pb model ages

1

and the formation of large Pb-Zn sedimentary deposits

2 3

Jean Milot1*, Janne Blichert-Toft1,Mariano Ayarzagüena Sanz2, 4

Nadège Fetter1, Philippe Télouk1, Francis Albarede1 5

6

1Ecole Normale Supérieure de Lyon, CNRS, Université de Lyon, 69007 Lyon, France 7

2Sociedad Española de Historia de la Arqueología, 28350 Ciempozuelos (Madrid), Spain 8

9

*Corresponding author: jean.milot@ens-lyon.fr 10

11

Abstract 12

13

In an attempt to clarify the significance of Pb model ages in Pb-Zn sedimentary deposits, 14

we report high-precision Pb isotopic compositions for 64 galenas and 52 K-feldspars, the 15

former from ores and the latter separated from granites. All samples are from Spain and 16

the French Pyrenees. Lead from galena ores is of unequivocal continental origin. With 17

few exceptions, Pb model ages systematically exceed emplacement ages by up to 400 Ma, 18

a gap which is well outside the uncertainties of ~30 Ma assigned to the model. The histo- 19

gram of the new high-precision Pb isotope data shows prominent peaks of galena Pb 20

model ages at 94±38 Ma and 392±39 Ma. When the data are consolidated with literature 21

data and examined in 3-dimensional Pb isotope space, cluster analysis identifies five 22

groups. The model ages of the peaks occur, in order of decreasing peak intensity, at 23

395±40 (Middle Devonian), 90±34 Ma (Middle Cretaceous), and 613±42 Ma (Neoprotero- 24

zoic), with two minor peaks at 185+26 Ma (Jurassic) and 313±41 (Upper Carboniferous).

25

To a large extent, the model ages centered around these peaks correspond to distinct lo- 26

calities. The ages of the peaks do not coincide with any of the Betic, Variscan, or Pan- 27

African tectonic events, which are the main tectonic episodes that shaped Iberian geology, 28

but rather match well-known global oceanic anoxic events. It is argued that surges of 29

metals weathered from continental surfaces scorched during anoxic events accumulated 30

and combined in anoxic water masses with unoxidized marine sulfide released by subma- 31

rine hydrothermal activity to precipitate the primary Pb-Zn stock. Frozen Pb isotope 32

compositions require that galenas from black shales are the source of the final ores. The 33

Revised manuscript with no changes marked Click here to view linked References

(3)

sulfides were later remobilized by large-scale convective circulation of basinal and hydro- 34

thermal fluids. The peaks of K-feldspar Pb model ages are distinct from those of galenas 35

and do not correlate with magmatic emplacement ages. It is suggested that they instead 36

reflect local circulation in Paleozoic sediments surrounding individual plutons. While Pb 37

isotopes can be used as a regional provenance tool, such an approach requires that the 38

data are considered in a fully 3-dimensional space.

39 40

1. Introduction 41

Stanton and Russell (1959) were the first to point out that the Pb isotope compositions of some 42

‘conformable’ strata-bound Pb-Zn 0-3.1 Ga old ore deposits could be accounted for by a single- 43

stage growth curve from the formation of the Earth to the known emplacement age of the ores.

44

Oversby (1974), however, reviewed the data existing at the time and concluded that a single- 45

stage closed-system evolution model was untenable. Using the concept of frozen lead, i.e. the 46

isotope composition of Pb locked into minerals with very low parent/daughter ratios, such as 47

galena in sulfide ores and K-feldspars, at the time of their formation, this issue received two 48

initially successful answers. One model uses a fit of conformable galena data with a two-stage 49

model in which the second stage starts at 3.8 Ga (Stacey and Kramers, 1975), while another 50

model uses a fit by a growth curve with a linear change in mantle U/Pb (Cumming and Richards, 51

1975), which is inconsistent with the evidence of constant U/Pb as demonstrated later by Alba- 52

rède and Juteau (1984) . The determination of Pb isotope compositions of K-feldspars from the 53

3.59 Ga old Amîtsoq gneisses in Greenland likewise led Gancarz and Wasserburg (1977) to 54

support a two-stage model. This model was strongly reinforced by evidence that, when the 238U- 55

206Pb and 235U-207Pb chronometers are considered one by one, the derived model 238U/204Pb () 56

values remain constant and are mutually consistent (Albarède and Juteau, 1984). The misfit 57

between the different models should not, however, be overemphasized. A model age is defined 58

as an apparent age calculated from measured isotopic abundances using simple assumptions, 59

typically closed-system behavior over one or two stages. The model ages calculated for a par- 60

ticular conformable galena from all the models discussed above and their deviation from the 61

respective emplacement ages agree in most cases to within ~30 Ma.

62 63

Conformable ores, however, only represent a small fraction of strata-bound Pb-Zn deposits.

64

Typically, major sedimentary exhalative deposits (SedEx) are syngenetic, i.e. formed at the 65

(4)

epigenetic, i.e. formed significantly after the deposition of their host rock (Wilkinson, 2014).

67

Leach et al. (2005, Appendix D) pointed out that the model ages of seven out of nine major 68

SedEx deposits are slightly to much older than their host rock ages. Metalliferous sediments 69

from the Atlantis II Deep in the Red Sea, which are considered an active equivalent of SedEx 70

deposits, derive their lead from underlying sediments, not from basalts (Dupré et al., 1988).

71

Leaching of Proterozoic lead from the basement is also the interpretation adopted for the MVT 72

deposits of the Phanerozoic Mississipi Valley ores (Doe and Delevaux, 1972; Goldhaber et al., 73

1995; Heyl et al., 1974; Paradis et al., 2007). It has been suggested that SedEx deposits of the 74

Phanerozoic reflect periods of global ocean anoxia (Goodfellow, 2004, 1987; Goodfellow et 75

al., 1993). The discordance between ‘frozen’ model ages and host rock ages means that ore 76

deposits did not form directly from a crustal source in a single stage, implying that a separate 77

stage of Pb-Zn preconcentration is a pre-requisite for Pb-Zn ore genesis, regardless of the ore 78

type and the environment involved.

79 80

Understanding large ore deposits therefore requires to dissociate the primary segregation of the 81

mined metals and the mechanisms by which these metals were transported to their current lo- 82

cations. Of course, a simple two-step representation of ore genesis does not do justice to the 83

idiosyncrasies of the mechanisms of ore deposition either. A two-step approach can, however, 84

be efficiently modeled and robustly tested by the U-Th-Pb system which is only determined by 85

two parameters, radiogenic ingrowth and U/Th/Pb fractionation events. The long-lived Pb- 86

based chronometers uses the rapidly changing proportions of the three isotopes 206Pb, 207Pb, and 87

208Pb upon radioactive decay of their parent elements U and Th.

88 89

The present work is mostly about what Pb model ages of galena and K-feldspar tell us about 90

the formation age and genetic environment of these minerals Although many model ages seem 91

compatible with geological evidence within the often generally large uncertainties on ore ages, 92

it is not always possible to assess a priori if a particular set of Pb isotope data is revealing the 93

genesis of the Pb carrier or the precise timing of its emplacement. These uncertainties have 94

direct implications on the field of archeology, which uses Pb isotopes for provenancing. For 95

decades, comparison of Pb isotope ratios measured on coins and other artefacts with Pb isotope 96

ratios measured on ores has been used to constrain the provenance of the metals used for pro- 97

ducing the objects in question (Stos-Gale and Gale, 2009). Model ages (Tm) and their associated 98

parameters ( and ), also derived from measured Pb isotope ratios, allow the blind comparison 99

(5)

of abstract Pb isotope compositions of ores and artefacts to be replaced by meaningful geolog- 100

ically identifiable environments (Albarède et al., 2012).

101 102

In the present study, we use a large set of high-precision Pb isotope data from the Iberian Pen- 103

insula and adjacent regions, gathered initially to establish the provenance of metals used for 104

silver artefacts and coins (Albarède et al., 2020), to propose an improved geological interpreta- 105

tion of model ages for galena ores and granitic K-feldspars. The motivation behind this work is 106

not to expand an already well-populated database of Pb isotopes on Iberian galenas and K- 107

feldspars, but to use a data set of uniform analytical quality to discuss how model ages can be 108

interpreted. This new framework places Pb isotope data onto a geological background and in 109

particular ties them to events of Pb-Zn ore formation.

110 111

2. What is a Pb model age?

112

A two-stage model of Pb isotope evolution can be summarized by the following equations:

113

𝑥0 = (206Pb

204Pb)

0

= 𝑥𝑇+ 𝜇(𝑒𝜆238U𝑇− 𝑒𝜆238U𝑇𝑚𝑜𝑑) 114

𝑦0 = (207Pb

204Pb)

0

= 𝑦𝑇+ 𝜇

137.79(𝑒𝜆235U𝑇− 𝑒𝜆235U𝑇𝑚𝑜𝑑) 115

𝑧0 = (208Pb

204Pb)

0

= 𝑧𝑇+ 𝜇𝜅(𝑒𝜆232Th𝑇− 𝑒𝜆232𝑇ℎ𝑇𝑚𝑜𝑑) 116

with 117

𝑥𝑇 = 𝑥0− 𝜇(𝑒𝜆238U𝑇− 1) 118

𝑦𝑇 = 𝑦0− 𝜇

137.79(𝑒𝜆235U𝑇− 1) 119

𝑧𝑇 = 𝑧0− 𝜇𝜅(𝑒𝜆238U𝑇− 1) 120

In these equations, subscript 0 indicates ratios measured today, T* is the age at which the second 121

stage began, and i is the decay constant of nuclide i. The asterisk represents parameters of the 122

reference models discussed below. The two parameters µ=(238U/204Pb)0 (or, short-hand, U/Pb) 123

and =(232Th/238U)0 (or, short-hand, Th/U) of the ore progenitor are time-invariant. This model 124

assumes that apparent radiogenic ingrowth is brought to a halt at Tmod, which is a standard 125

(6)

as sphalerite, pyrite, and chalcopyrite, also is low but not nearly as low as that of galenas. K- 127

feldspars contain significant amounts of initial lead but only little uranium, the latter of which 128

resides mostly in cracks in the form of oxides and phosphates, which, together with ingrown 129

radiogenic lead, can be efficiently removed by aggressive leaching (McNamara et al., 2017).

130

Each triplet (Tm,  ) corresponds to a triplet of isotopic ratios. The Tmod, ,  representation 131

is nothing more than an alternative representation of the 3-dimensional Pb isotope space. Other 132

ratios with a different normalization isotope (e.g. 206Pb instead of the more standard 204Pb) can 133

be used, but the triplet (Tmod,  ) remains unchanged.

134 135

The values 𝑥0, 𝑦0, and 𝑧0 of the so-called modern lead may be obtained in different ways.

136

Stacey and Kramers (1975) computed the intersection of isochrons from Archean terranes in 137

the conventional plot 207Pb/204Pb vs 206Pb/204Pb and then refined T* using the Pb isotope com- 138

position and geological ages of 13 ‘conformable’ galenas. One of the merits of this visionary 139

work was to validate the relevance of the two-stage evolution models. Albarède and Juteau, 140

(1984) took a different approach by considering the evolution of isotopic ratios independently 141

for the 13 galenas listed by Stacey and Kramers (1975), to which they further added the Pb 142

isotopic composition of an Early Archean galena from Isua, Greenland (Appel et al., 1978) and 143

plotted its value against eiTm1, where Tm is the accepted geological age of the sample, and 144

obtained excellent alignments. The intercepts provide the modern values 𝑥0, 𝑦0, and 𝑧0 of the 145

corresponding ratios. In addition, the slopes of the 206Pb/204Pb and 207Pb/204Pb alignments give 146

=(238U/204Pb)0 in the galena protolith, whereas the slope of the 208Pb/204Pb alignment gives the 147

associated =(232Th/238U)0. The  and  values obtained in this way are independent of the 148

(238U/235U)0 ratio of the Earth, a canonical value that was recently revised (Andersen et al., 149

2017; Goldmann et al., 2015). The value of =(232Th/238U)0 derived in this way is indistinguish- 150

able from the planetary value inferred by Blichert-Toft et al. (2010). The parameters derived 151

by Stacey and Kramers (1975) and Albarède and Juteau (1984) are listed in Table 1 and are 152

similar although they were derived in different ways. We assessed by Monte-Carlo error prop- 153

agation that model ages calculated using Stacey and Kramers (1975) and Albarède and Juteau 154

(1984) may differ by up to 30 Ma and the  values by up to 0.3 units. Studies of deep-sea FeMn 155

nodules (Abouchami and Goldstein, 1995; Vlastélic et al., 2001; von Blanckenburg and Igel, 156

1999) and Quaternary loess (Sun and Zhu, 2010), two proxies of the ‘common lead’ of modern 157

upper crust, show that Albarède and Juteau’s (1984) model is a valid representation of the upper 158

(7)

for mantle-derived magmas plot to the right of the modern-lead secondary isochrons of Stacey 160

and Kramers (1975) and Albarède and Juteau (1984), which is tantamount to future model ages.

161

The conundrum of short Pb residence in the deep ocean (~80 y, Nozaki, 1986; Schaule and 162

Patterson, 1981) contrasting with the relative homogeneity of Pb isotopes in FeMn nodules (von 163

Blanckenburg and Igel, 1999) reflects that airborne particle deposition accounts for most of the 164

Pb input into the ocean (Schlosser et al., 2019). As for hydrothermal input from the ridge sys- 165

tem, the low Pb concentration of the oceanic asthenosphere and the rapid precipitation of Pb 166

next to oceanic spreading centers (Dasch, 1981) makes this source a minor contributor to the 167

sedimentary record.

168 169

More complex models, such as the popular Plumbotectonics model of Zartman and Doe (1981) 170

and Zartman and Haines (1988) and the detailed analysis by Kramers and Tolstikhin (1997), 171

provide a powerful illustration of possible Pb isotope evolution in the crust, but the multiple 172

stages and under-constrained parameters they entail make it difficult to assess the reliability of 173

the model ages inferred from these models.

174 175

3. The Iberian Peninsula: Geology and mining districts 176

Given the number and regional distribution of the samples in this study, only a short summary 177

of their Iberian background geology is presented here. Digitized geological maps of Spain, Por- 178

tugal, and France are freely available on the portals of the Geological and Mining Institute of 179

Spain (IGME), the National Laboratory of Energy and Geology of Portugal (LNEG), and the 180

French Geological Survey (BRGM).

181 182

The geology of the Iberian Peninsula was affected by Neoproterozoic to Neogene geodynamic 183

activity (Quesada and Oliveira, 2019) and involved three major orogenic cycles: the Cadomian 184

(Pan-African) events spanning the Ediacaran to the Cambrian, the Variscan (Hercynian) events, 185

and the Alpine events. The Pan-African (Cadomian) series of tectonic events occurred during 186

the late Neoproterozoic, about 650–550 Ma ago and led to the formation of the continent and 187

its collision with island arcs and other Proterozoic terranes. During the Variscan orogeny, 188

Gondwana collided with the northern Laurussia continent to form Pangaea. The resulting su- 189

percontinent gradually broke up during the Mesozoic Era and various fragments collided to 190

form Alpine orogens.

191

(8)

Figure 1 depicts the main features of Iberian geology while Fig. 2 positions the most commonly 193

used geological units. The bulk of Iberia was assembled and deformed during the Variscan 194

orogeny and constitutes the western expansion of the West European Variscides. The main 195

tectonic provinces underlain by pre-Variscan and Variscan basement are the South Portuguese, 196

Ossa Morena, Central Iberia, and West Asturian Leonese Zones, to which the internal zones of 197

the Pyrenees, the Betics, and the Catalan Coastal Range must be added. The pre-Variscan (Ne- 198

oproterozoic, Cadomian) basement includes fragments from the Pan-African, notably in west- 199

ern and northwestern Spain, the eastern Pyrenees, and the whole of the Ossa Morena Zone. The 200

Pyrenees is a prominent mountain range resulting from the Mid-Cretaceous and Eocene colli- 201

sion between Iberia and western Europe.

202 203

Variscan granites represent the most striking marker of the collision between Gondwana frag- 204

ments and Laurussia. Early petrological and chronological literature was reviewed by Castro et 205

al. (2002) and Paleozoic intrusive magmatism was recently re-assessed by Ribeiro et al. (2019).

206

The latter occupies a large surface area in northwestern and central Iberia, Catalonia, and the 207

Pyrenees. High 87Sr/86Sr assigns to all these rocks an unmistakably crustal origin (Castro et al., 208

2002). Most granitic rocks seem to have formed by reworking of Early- to Mid-Proterozoic 209

protoliths and were emplaced during the Variscan orogeny. Mesozoic and post-collisional Ne- 210

ogene volcanism from the Betic Cordilleras was reviewed by Gómez-Pugnaire et al. (2019).

211 212

In order to see through these overarching major tectonic events and grasp the relative im- 213

portance of other less imposing, but not necessarily less critical, geodynamic events, as well as 214

the significance of geological unit subdivision, we made use of zircon U-Pb geochronology of 215

detrital rocks of different stratigraphic ages, largely Paleozoic. In northern, northwestern, and 216

central Iberia and in the Ossa Morena Zone, the dominant age peaks are Ediacaran and Cam- 217

brian (Cadomian) (Ábalos et al., 2012; Fernández-Suárez et al., 2014; Martínez Catalán et al., 218

2008; Pastor-Galán et al., 2013; Silva et al., 2014; Talavera et al., 2012; Villaseca et al., 2016).

219

The same observation was made in the Betics (Esteban et al., 2017; Jabaloy-Sánchez et al., 220

2020; Santamaría-López and Sanz de Galdeano, 2018) and the Pyrenees (Filleaudeau et al., 221

2012). In contrast, the South Portuguese Zone south of the Rheic suture shows evidence of a 222

Silurian to Mid-Devonian event (380-420 Ma) superimposed on the dominant Cadomian event 223

(Pereira et al., 2017, 2014, 2012; Pérez-Cáceres et al., 2017). No zircon population with clearly 224

Alpine ages has been identified so far.

225

(9)

Overall, most of Iberia is a Variscan crustal segment with a strong heritage of the Pan-African 227

Gondwana supercontinent, which itself was built by the destruction of Proterozoic cratons. The 228

Variscan collision strongly deformed and metamorphosed this material with little addition of 229

juvenile magmas (Ribeiro et al., 2019). Only in the South Portuguese Zone is a Silurian-Devo- 230

nian geodynamic event interpreted as a short-lived intra-oceanic arc (Pereira et al., 2012) un- 231

questionably different from the rest of the Variscan domain. Alpine tectonics is restricted to the 232

boundaries between the microplates inherited from the breakup of Pangea.

233 234

A recent database (www.ehu.eus/ibercron/iberlid) collects Pb isotope compositions measured 235

by a variety of techniques over several decades on nearly 3000 geological and archeological 236

samples (de Madinabeitia et al., 2021). The work of these authors and their database can be 237

used to acquire some background on Iberian Pb-Zn mines and guide the reader through the 238

various geological environments of specific samples.

239 240

3.1. The mines of southeast Iberia 241

The southeastern region of the Iberian Peninsula can be divided into two major lithological 242

groups: (1) the Paleozoic and Mesozoic (Late Cretaceous to Paleogene) meta-sedimentary 243

rocks formed before the Betic Alpine orogeny under a subduction-related compressive regime 244

due to the convergence of European and African plates, and (2) the Miocene igneous rocks of 245

the Cabo de Gata-Cartagena volcanic belt associated with post-collisional extension (e.g. Arri- 246

bas and Tosdal, 1994; Esteban-Arispe et al., 2016; Turner et al., 1999). This Miocene volcanic 247

field results from the fusion of mantle materials metasomatized by subduction processes (e.g.

248

Benito et al., 1999). From southwest to northeast, interaction of rising melts with crustal rocks 249

led to the formation of calc-alkaline, high-K calc-alkaline and shoshonitic, ultrapotassic, and 250

rare alkaline basaltic rocks (Esteban-Arispe et al., 2016). Hydrothermal systems associated with 251

the calc-alkaline, high-K calc-alkaline, and shoshonitic rocks altered the volcanic rocks and 252

formed epithermal precious and base-metal deposits such as the San José, Rodalquilar, Carbon- 253

eras, Mazarrón, and Cartagena-La Unión deposits (e.g. Arribas et al., 1995; Arribas and Tosdal, 254

1994; Ruano et al., 2000).

255 256

(10)

3.1.1. The Cartagena-La Unión Range 257

The mining district of La Unión is located in the Sierra de Cartagena, which forms the eastern 258

end of the Betic Cordilleras (e.g. Oen et al., 1975; Sabaté et al., 2015). Depending on the loca- 259

tion, the mineralization consists mainly of Pb-Zn veins or stratabound deposits, associated with 260

hydrothermal alteration or carbonate replacement, hosted by volcanic, sub-volcanic, or meta- 261

morphic rocks (Sabaté et al., 2015). Galena is abundant in both vein and stratabound deposits 262

and has been intensively exploited since the Carthaginian expansion and subsequent Roman 263

times (e.g. Sanmarti et al., 2013; Soler et al., 2013). The Ag-bearing galena located in the Cabo 264

de Palos were, in ancient times, the richest in Hispania, containing between 5 and 6 kilograms 265

of silver per ton of lead.

266 267

3.1.2. Mazarrón 268

The Mazarrón mining district is located in the northern part of the Cabo de Gata-Cartagena 269

volcanic belt, at the eastern end of the Betic Cordilleras. This Zn-Pb-Cu-Ag-Fe sulfide deposit 270

is one of the best examples of epithermal mineralization associated with high-K calc-alcaline 271

and shoshonitic volcanics (Esteban-Arispe et al., 2016). Lead-bearing ores are represented by 272

lead sulfides and argentiferous galena, accompanied by sphalerite, pyrite, antimony and arsenic 273

minerals, and gypsum. The grade range between 3000 and 6000 grams per ton of ore, in some 274

places reaching 10,000 grams.

275 276

3.1.3. Sierra de Almagrera 277

The mining district of the Sierra de Almagrera is located in the central part of the Cabo de Gata- 278

Cartagena volcanic belt (e.g. Martinez Frias et al., 1992; Navarro et al., 2008; Suárez, 2016).

279

Ore deposits mainly consist of Ag-rich veins or stratabound ore bodies (Herrerias deposit) 280

hosted by Paleozoic schists (graphitic and quartz-rich phyllites) and quartzite of the Alpujárride 281

Complex (Internal Betic Zone). The major veins contain an assemblage of base metal sulfides 282

(Ag-bearing galena, sphalerite, pyrite, chalcopyrite, and marcasite) and Pb-Sb-Cu-Ag sulfosalts 283

(Arribas and Tosdal, 1994; Martínez-Frías et al., 1989). Locally, the strong supergene alteration 284

and secondary enrichment resulted in abundant native Ag and secondary minerals such as jar- 285

osite (Arribas and Tosdal, 1994).

286 287

3.1.4. The Gádor Range 288

The Sierra de Gádor is located in the western part of the Betic Cordilleras and belong to the 289

(11)

Gádor Range, with an extent of some twenty square kilometers. The mineralization consists of 291

lenticular and stratabound F-(Pb-Zn) deposits hosted by Middle Triassic carbonate rocks of the 292

lower Alpujárride Complex (Arribas and Tosdal, 1994; Fontboté et al., 1983). Mineral assem- 293

blages generally consist of fluorine, galena, sphalerite, pyrite, and bornite (Arribas and Tosdal, 294

1994). Although minerals occasionally appeared at the surface, the most important concentra- 295

tions of lead were located between 60 and 100 m depths.

296 297

3.2. The mines of the South Central Iberian Zone 298

Ore deposits from this area are associated with various lithological units, including slates and 299

shales of the Rumblar basin and Variscan granites. The igneous rocks are differentiated to var- 300

iable extents and include diorites, granodiorites, and porphyry. The veins in this east Lusitan- 301

Marianic region contain Cu-Pb, Pb, and Pb-Ag mineralizations, as well as some Sn-W and 302

epithermal deposits (Arboledas et al., 2014).

303 304

3.2.1. The Alcudia Valley 305

The Alcudia Valley ore district is located in the southwest of the Ciudad Real province and 306

belongs to a large metallogenic province known as the Sierra Morena (e.g. Palero-Fernández 307

and Martín-Izard, 2005; Santos Zalduegui et al., 2004). The most important geological feature 308

of this district is a succession of anticlines and synclines affecting sedimentary Neoproterozoic 309

and Ordovician-Early Carboniferous rocks. All the deposits of this district are related to Var- 310

iscan granite magmatism of Carboniferous age, in particular the intrusion of the Los Pedroches 311

Batholith in the southern part of the district (Palero et al., 2003; Palero-Fernández and Martín- 312

Izard, 2005). Ore deposits from the Alcudia Valley were classified according to their morphol- 313

ogy, host rock relations, structural setting, mineral assemblages, and geochemical parameters 314

(Palero et al., 1992, 1991). Most of the deposits of economic importance are hydrothermal Pb- 315

Zn-Ag veins with breccia textures located in tensional fractures within Neoproterozoic rocks 316

(Palero-Fernández and Martín-Izard, 2005). The ore mineral assemblage mainly consists of ga- 317

lena, sphalerite, chalcopyrite, pyrite, marcasite, and Ag-sulfosalts.

318 319

3.2.2. Linares-La Carolina 320

The mining district of Linares-La Carolina is located east of the Alcudia Valley, in the north- 321

western part of the Jaén province. This district is divided into two mineralized sectors, Linares 322

(12)

granite emplacement (e.g. Palero-Fernández and Martín-Izard, 2005; Santos Zalduegui et al., 324

2004). Deposits in Linares consist of hydrothermal veins of base-metal sulfides hosted mainly 325

by granites from the Linares pluton, and less commonly by Paleozoic rocks. The mineral as- 326

semblage contains Ag-poor galena, sphalerite, pyrite, and chalcopyrite in a quartz-barite-car- 327

bonate gangue. In La Carolina, deposits are base-metal sulfide veins hosted by Ordovician to 328

Silurian meta-sedimentary rocks, and more rarely by granites from the Santa Elena pluton. The 329

mineral assemblage contains Ag-bearing galena, sphalerite, and pyrite in a quartz gangue, 330

sometimes associated with ankerite and barite (Garcia De Madinabeitia, 2003).

331 332

3.3. The mines of southwest Iberia 333

The Iberian Pyrite Belt, located in the southwest of the Iberian Peninsula comprising part of 334

Portugal, Huelva, and the Seville province (Spain), contains diverse mining districts that have 335

been major sources of metals through centuries. As the Romans were unable to extract silver 336

from sulfides, silver was instead extracted from jarosites, which lie beneath the gossan, the 337

oxidised zone of sulfides. The ores are hosted by two geological zones: the Ossa Morena and 338

South Portuguese zones.

339 340

3.3.1. The Ossa Morena Zone 341

The Ossa Morena Zone (OMZ) is located south of the Central Iberian Zone and to the northeast 342

of the South Portuguese Zone. The geology of this tectonostratigraphic unit, described in detail 343

by Tornos and Chiaradia (2004), is particularly complex. It resulted from the accretion of ter- 344

ranes ranging from Late Riphean to Late Carboniferous in age. In short, the OMZ encompasses 345

(1) disseminated pre-Cadomian sequences of high-grade metamorphic and siliciclastic rocks, 346

(2) syn-orogenic Cadomian units of Neoproterozoic-Early Cambrian backarc to intra-arc se- 347

quences, (3) a volcano-sedimentary unit formed during Paleozoic intracontinental rifting, (4) 348

an Ordivician to Early Devonian passive margin sequence, and (5) syn-Variscan sedimentary 349

rocks (Tornos and Chiaradia, 2004). Ore deposits in the OMZ are related to three major oro- 350

genic and sedimentary events (Marcoux et al., 2002). First, during the Cadomian orogeny, small 351

volcanogenic massive sulfide (VMS) deposits of Upper Riphean to Lower Neoproterozoic age 352

were associated with calk-alcaline volcanism. Second, sedimentary-exhalative (SedEx) ore 353

lenses associated with alkaline volcanism are hosted by late Cadomian (Neoproterozoic-Cam- 354

brian) marbles. Third, Zn-Cu-Pb massive sulfide deposits of Upper Devonian to Permian age 355

(13)

are coeval with the giant deposits of the South Portuguese Zone described below and associated 356

with alkaline volcanism (Baeza-Rojano et al., 1981).

357 358

3.3.2. The South Portuguese Zone 359

The South Portuguese Zone (SPZ) counts many world-class giant and supergiant VMS depos- 360

its, mainly in the province of Huelva. This region includes, among others, the mines of Riotinto, 361

Tharsis, Cueva de la Mora, Sotiel Coronada, Castillo de Buitron, and La Zarza. Aznalcollar in 362

Sevilla, Aljustrel in the district of Beja, and Neves Corvo in southern Portugal are also consid- 363

erable in size. These deposits all belong to the Iberian Pyrite Belt that occupies the southwest 364

corner of the Iberian Peninsula (Almodóvar et al., 1997; Relvas et al., 2001; Sáez et al., 1996;

365

Tornos, 2006). The SPZ is interpreted to have been sutured to the Iberian Massif during the 366

Middle Carboniferous. The local stratigraphy includes three main units, from top to bottom: (1) 367

the Phyllite-Quartzite group (PQ), (2) the Volcanic-Siliceous (VS) Complex, and (3) the Culm 368

group (Sáez et al., 1996). The VMS deposits are hosted in the VS Complex which consists of 369

successive mafic and felsic volcanic sequences interbedded with mudstone and chemical sedi- 370

mentary rocks of Late Famennian to Early Late Visean age (Oliveira, 1990; Tornos, 2006). The 371

stratigraphy is locally modified by igneous intrusions and thrust faults (Tornos, 2006). Overall, 372

the major ore deposits hosted by the VS Complex consist of massive sulfide lenses that contain 373

up to 170 Mt of ore. Most of the deposits show a direct relationship with black shales that form 374

a significant part of the VS Complex (Tornos, 2006).

375 376

3.4. The mines of northern Iberia 377

The density of ore deposits in the northern part of the Iberian Peninsula is lower than in the 378

south. However, several mining districts are noteworthy in the Catalan Coastal Ranges (north- 379

eastern Iberia), the Pyrenees (northern edge of the Iberian Peninsula), the Basque-Cantabrian 380

Zone (north-central Iberia), and the West Asturian Leonese Zone (northwestern Iberia).

381 382

3.4.1. The Pyrenees 383

The Pyrennean chain, between France and Spain, formed during the Alpine orogeny which was 384

surperimposed on Paleozoic terranes and Precambrian basement previously deformed by the 385

Variscan orogeny (Munoz et al., 2016). The Alpine deformation of the Pyrenees is mostly lim- 386

ited to thrusting and faulting of the Paleozoic rock series. The Variscan phase of the Pyrenees 387

(14)

dates from the Late Carboniferous (Namuro-Wesphalian) and was accompanied by calc-alka- 388

line magmatism and high-temperature–low-pressure metamorphism (Denele et al., 2009;

389

Laumonier et al., 2010). The Paleozoic formations of the Axial Zone of the Pyrenees hosted 390

many SedEx Pb-Zn stratiform deposits. The more economically significant occur in Lower 391

Paleozoic meta-sedimentary rocks of Ordovician to Lower Devonian ages, such as the 392

Pierrefitte, Bentaillou, Aran Valley, and Benasque Pass deposits (e.g. Cardellach et al., 1996;

393

García-Sansegundo et al., 2014; Nicol et al., 1997), and are considered to have formed during 394

Early Paleozoic rifting. Additionally, Pb-Zn vein deposits with Ag-rich tetrahedrite, Les Ar- 395

gentières and Lacore deposits next to Aulus-Les-Bains (Munoz et al., 2016), are found in De- 396

vonian terranes, but formed by Mesozoic hydrothermal fluids circulating during the post-Var- 397

iscan extensional regime.

398 399

3.4.2. The Catalan Coastal Range 400

The Catalan Coastal Range is made of folded Variscan basement overlain by a Mesozoic to 401

Cenozoic sedimentary cover (e.g. Canals and Cardellach, 1997; Canet et al., 2003; Edel et al., 402

2015; Parviainen et al., 2008). The Variscan basement consists of Paleozoic sedimentary rocks 403

and Late-Variscan granites peneplaned during pre-Triassic erosion, then covered by Mesozoic 404

sedimentary sequences (Canals and Cardellach, 1997). Numerous mineralized veins including 405

barite, fluorite, and base-metal sulfides in variable proportions are hosted by the Paleozoic 406

basement and locally cut across Triassic red beds. A detailed classification of the different types 407

of deposits is given by Canals and Cardellach (1997). The overall relationships between the 408

type of enclosing rock and vein mineralogy are: deposits enclosed in Paleozoic meta-sedimen- 409

tary rocks have a complex mineralogy including Co, Ni, and Ag arsenides and sulfides, whereas 410

deposits hosted by granites are rich in fluorite and/or barite and depleted in base metal sulfides 411

412

3.4.3. The Basque-Cantabrian Zone 413

The Basque-Cantabrian basin is located west of the Pyrenees, in central northern Spain. This 414

region roughly consists of Mesozoic to Tertiary sediments overlying the Paleozoic Asturian 415

basement faulted and deformed during the Variscan orogeny (e.g. Velasco et al., 2003). Sedi- 416

ment-hosted Pb-Zn deposits are found in two geological settings (Velasco et al., 1996). The 417

first is Carboniferous clastic meta-sediments of the Cinco Villas massif (east of the Basque- 418

Cantabrian basin) which includes SedEx ore deposits affected by Variscan deformation and 419

metamorphism and associated remobilization (Pesquera and Velasco, 1993, 1989). The second 420

(15)

ore deposits (Herrero, 1989; Velasco et al., 1996). South of the Basque-Cantabrian zone, Pb- 422

Zn deposits also occur in the Paleozoic Sierra de la Demanda. Several deposits from this region, 423

such as Udias, Novale, La Florida, Mercadal, Comillas, Punta Calderon, and Reocin have been 424

historically exploited for Pb and Zn (e.g. Pašava et al., 2014; Velasco et al., 2003).

425 426

3.4.4. The West Asturian Leonese Zone 427

The West Asturian-Leonese Zone (WALZ) is located northwest of the Iberian Peninsula. The 428

Zn-Pb deposits are found in the calcareous Vegadeo Formation in the westernmost outcrops of 429

the WALZ. This formation constitutes a major carbonate intercalation in the mainly siliciclastic 430

Lower Cambrian sedimentary succession in this part of the Variscan Belt of Spain (Tornos et 431

al., 1996). In short, the Vegadeo Formation consists of a thick bed of carbonates with rare in- 432

tercalations of shales and sandstones of variable thickness. Two types of ore deposits are hosted 433

by this formation at different stratigraphic positions. Ores from the first type are disseminated 434

stratiform ores with thin layers of sphalerite and galena replacing earlier pyrite and are located 435

in the Lower Member of the Vegadeo formation. These deposits are thought to have a pre- 436

metamorphic origin with sulfur and fluids being derived from the host carbonates. The second 437

type, of greater economic importance, is located on top of the Vegadeo Formation always in 438

contact with the overlaying shales and sandstones. The mineral assemblage of this type is com- 439

posed of sphalerite and galena, with minor amounts of chalcopyrite, pyrite, Co-Ni-As sulfides, 440

bismuthinite, tetrahedrite, and Pb-Bi sulfosalts. Several indices suggest that these last Variscan 441

mineralizations result from the remobilization of the pre-metamorphic stratiform ores (Tornos 442

et al., 1996).

443 444

4. Analytical techniques 445

The high-precision Pb isotope compositions of the 64 galena and 52 K-feldspar analyzed in the 446

present study, all from Spain and the French Pyrenees, are listed in Tables S1 and S2, respec- 447

tively, of the Supplementary Material. In the following, we briefly summarize the analytical 448

techniques used at ENS Lyon for separating Pb from galena and K-feldspar and measuring the 449

Pb isotopic compositions my multiple-collector inductively-coupled plasma mass spectrometry 450

(MC-ICP-MS; Nu Plasma 500 HR).

451 452

(16)

4.1. Galena 453

Galena ores were crushed and powdered in a clean pre-conditioned agate mortar. About 0.1 g 454

of each sample was weighed into clean Savillex beakers for acid digestion. Samples were di- 455

gested in a wet chemistry laboratory at ENS Lyon (so as not to contaminate the low-blank clean 456

laboratory at ENS Lyon with the high Pb concentrations of galena) with 10 ml distilled con- 457

centrated HNO3 on a hotplate at 110°C for a several hours. Galena readily dissolves according 458

to the reaction:

459

PbS + 8 HNO3 = PbSO4 + 8 NO2 + 4 H2O 460

The NO2 gas was regularly evacuated from the closed beakers by slightly unscrewing their lids 461

as pressure built up until complete sample digestion was achieved. Once the reaction complete, 462

samples were evaporated to dryness on a hotplate at 90°C. The dry PbSO4 residue was then 463

dissolved in 20 ml distilled 3M HNO3 on a hotplate at 110°C and repeatedly placed in an ultra- 464

sonic bath until total dissolution. The elemental composition of the digested galena samples 465

was measured on an aliquot of known volume of the total dissolved sample using a Thermo 466

Scientific ICAP-Q quadrupole ICP-MS (inductively-coupled plasma mass spectrometer) at 467

ENS Lyon. The total procedural Pb blank of the wet laboratory part of the procedure was < 20 468

ng, which is negligible relative to the amount of Pb in the samples of the order of 10-100s of 469

mg. For each sample, an aliquot containing about 1 µg of Pb was taken for subsequent Pb 470

separation and purification under clean laboratory conditions for high-precision Pb isotopic 471

analysis.

472 473

Lead was eluted at ENS Lyon following a one-step column chromatography procedure that 474

uses microcolumns filled with 500 µL anion-exchange resin (AG1-X8, 100-200 mesh). The 475

sample aliquots containing 1 µg of Pb dissolved in distilled 3M HNO3 were first evaporated to 476

dryness on a hotplate at 110°C. The residues were then taken up in 1 mL distilled 6M HCl and 477

dried down to convert the samples to chloride form. The samples where thereafter dissolved in 478

1 mL double-distilled 1M HBr on a hotplate at 110°C and, once cooled down, loaded onto the 479

anion-exchange columns. Sample matrices were eluted with double-distilled 1M HBr and Pb 480

was collected with distilled 6M HCl. After evaporation to dryness of the pure Pb fractions, a 481

few drops of concentrated distilled HNO3 were added and the samples dried down again prior 482

(17)

to Pb isotopic analysis, ensuring that any traces of potentially interfering Br2 and organic ma- 483

terial were driven off. The total procedural Pb blank of the clean laboratory part of the procedure 484

was < 20 pg, which again is negligible compared to the starting amount of this step of 1 µg Pb.

485 486

The Pb isotopic compositions of the galena samples were measured on a Nu Plasma 500 HR 487

MC-ICP-MS at ENS Lyon following the procedures of Blichert-Toft et al. (2003) and White et 488

al. (2000). Instrumental mass bias was corrected using added Tl, and sample-standard bracket- 489

ing referred to the values of Eisele et al. (2003) for the NIST 981 Pb reference material. Re- 490

peated measurements of NIST 981, analyzed every second sample throughout the individual 491

analytical sessions, yielded an external reproducibility < 100 ppm (0.01%) for 204Pb-normalized 492

ratios and < 50 ppm (0.005%) for 207Pb/206Pb and 208Pb/206Pb. In-run analytical errors (2-sigma 493

mean, based on the measurement of 60 isotope ratios with 10 seconds integration time each for 494

each sample and standard run) being systematically smaller than the external reproducibility 495

for all samples, they are not reported with the data in Tables S1 (galenas) and S2 (K-feldspars) 496

so as to render the data tables more manageable and readable.

497 498

4.2. K-feldspar 499

The K-feldspars were separated from their respective granites via grinding, sieving, and hand- 500

picking under a binocular microscope. All subsequent leaching, dissolution, and Pb separation 501

procedures were done in the clean laboratory at ENS Lyon. The separated K-feldspars were 502

leached in hot distilled 6M HCl and rinsed in distilled water prior to dissolution in either a 503

3:1:0.5 mixture of double-distilled concentrated HF:HNO3:HClO4 or partially dissolved in cold 504

double-distilled 30% HF depending on how pure the K-feldspar separates were (i.e. to avoid 505

high-(U-Th)/Pb inclusions potentially present in trace amounts of quartz sometimes left behind 506

with the K-feldspar). After evaporation to dryness, the samples were taken up in distilled 6M 507

HCl to ensure complete dissolution, dried down again, then taken up in double-distilled 1M 508

HBr and run through the same Pb columns as described above for galena. The total procedural 509

Pb blank also was the same as for galena, < 20 pg. The Pb isotopic compositions of the K- 510

feldspars were measured as described above for galena.

511 512

(18)

5. Results 513

The new Pb isotope data for galena and K-feldspar are listed in Tables S1 and S2, respectively, 514

together with the relevant literature data. The new data and the complete data set built by con- 515

solidation of the new data with literature data are considered separately. The reason for this 516

distinction is that, since the early 2000s, MC-ICP-MS has significantly improved the accuracy 517

and precision of Pb isotope data. While data quality therefore may vary among studies in the 518

published literature, MC-ICP-MS data are consistently superior owing to the excellent control 519

of analytical mass bias allowed by this technique.

520 521

Although only few absolute datings of Pb-Zn mineralizations are available in the literature, 522

emplacement ages have been estimated for most of the ore deposits considered here. These ages 523

are reported in Table S1. For each mining region, several ore deposition events may have oc- 524

curred. As shown in Fig. 3, with rare exceptions, model ages are systematically older than em- 525

placement ages by up to 400 Ma, a time difference that largely exceeds the uncertainties on the 526

model ages.

527 528

We highlight this observation with a few examples, such as the Alcudia-Valley-los Pedroches 529

area, where five types of Pb-Zn deposits have been defined corresponding to distinct ore form- 530

ing events (Palero-Fernandez et al., 2003; Palero-Fernandez and Martin-Izard, 2005). The 531

model ages of the first four ore types which formed from the Silurian (450-420 Ma) to the 532

Upper Carboniferous (320-300 Ma) are Neoproterozoic (about 600 Ma). In contrast, the model 533

ages of galenas from the last ore-forming event of Permian to Triassic age (260-240 Ma) are 534

Devonian (380-400 Ma) (see Table S1). A similar pattern is observed in the Betic Cordilleras, 535

where three phases of ore formation have been identified (Arribas and Tosdal, 1994). A first 536

mineralisation type of Late Permian to Lower Triassic age (240-230 Ma) hosts galenas with 537

Devonian model ages (380-400 Ma), whereas the two Miocene types host galenas with Creta- 538

ceous to Tertiary model ages. Likewise, galenas from the Catalan Coastal Range have Devonian 539

model ages, whereas the ore-forming events occurred from Trias to the Lower Jurassic (225- 540

195 Ma) (Canals and Cardellach, 1997, 1993). Overall, with few exceptions, galenas from the 541

whole Iberian Peninsula have model ages significantly older than their host deposits (Fig. 3), 542

regardless of ore type deposit (i.e. SedEx, VHMS, MVT).

543 544

(19)

The maps of Fig. 4 (206Pb/204Pb (a), 207Pb/206Pb (b),208Pb/206Pb (c), Tmod (Pb model age) (d), 545

(apparent 238U/204Pb) (e) and (apparent 232Th/238U) (f)) reveal that these variables are 546

strongly controlled by geography. With the exception of the OMZ and the South Central Iberian 547

Zone, the data are consistent within the same region. 206Pb/204Pb and Tmod are negatively corre- 548

lated, whereas 208Pb/206Pb and  are positively correlated. The  map separates a province with 549

high values, the Pyrenees and the Catalan Coastal Range, from the rest of Iberia. Galenas with 550

the highest 206Pb/204Pb and the youngest Tmod are restricted to the Eastern Betic realm, whereas 551

galenas with the lowest 206Pb/204Pb, and hence the oldest, are found in León (northwestern 552

Spain), the South Central Iberian Zone, the OMZ (Extremadure), and the northern Pyrenees.

553

Other galenas from the OMZ, the South Central Iberian Zone, the Catalonia Batholith, and the 554

Pyrenees fall within an intermediate 206Pb/204Pb range. 207Pb/204Pb and Tmod mirror this distri- 555

bution and 208Pb/204Pb also shows the same strong contrast between the Betic realm, the OMZ, 556

and the Pyrenee-Catalonia domain.

557 558

As mentioned above, the present study focuses on the geological interpretation of Pb isotope 559

data using the model parameters Tmod, , and  calculated from the equations given by Albarède 560

et al. (2012) and the parameters determined by Albarède and Juteau (1984). We used the re- 561

cently redetermined value of 137.79 for 238U/235U (Andersen et al., 2017; Goldmann et al., 562

2015). The provinciality of calculated model ages is particularly sharp (Fig. 4). The Tmod data 563

are also presented in 1-dimensional histogram form (Figs. 5 and 6), which will be discussed 564

below. In the 3-dimensional Tmod, , and space, the datawere treated as representing mixtures 565

of stochastic variables with normal distributions. The peaks are identified by a code (which can 566

be obtained from the senior author upon request) written in Matlab using the Statistical Toolbox 567

and the results are presented in Table 2.

568 569

The one-dimensional Tmod histogram of the new galena Pb isotope data identifies two particu- 570

larly significant peaks at 94±38 Ma and 392±39Ma (Fig. 5) with the significance level assessed 571

by a t-test, and a minor peak at ~614 Ma (Fig. 5). When the database obtained by combing the 572

new Pb isotope data with literature data (N=464) (Table S1), and  and  are added as additional 573

dimensions, the data set can be treated as a 3-dimensional population mixture: a five-component 574

mixture gives a best fit with well-defined model age peaks (Table 2). Two major model age 575

peaks are observed at 395±40 (N=237, Middle Devonian; Table 2) and 90±34 Ma (N=80, Mid- 576

(20)

Additionally, a broad Neoproterozoic to Early Paleozoic peak (N=74, 613±42 Ma; Table 2) and 578

two minor peaks at 185±26 Ma (Jurassic) and 313±41 (Upper Carboniferous) (Table 2) are also 579

identified. The average Th/U ratio of the galenas (3.99±0.13) matches the average value for 580

European and circum-Mediterranean galenas (Blichert-Toft et al., 2016).

581 582

For K-feldspars, emplacement ages and Pb model ages are not correlated (r = 0.14). Most 583

early Variscan granites dated at >300 Ma are located in Northwestern Iberia, the Eastern Pyre- 584

nees, and the Catalan Coastal Range and have model ages clustering between 402 and 449 Ma 585

(Upper Silurian-Devonian). The K-feldspar data from various types of granites together with 586

those from the literature reveal a picture very different from that of galena with well-defined 587

Tmod peaks at 329±24 Ma (Upper Carboniferous) and 417±36 Ma (Fig. 6). The Tmod histogram 588

also hints at a minor peak at 232±14 Ma (Fig. 6). The 3-dimensional (Tmod, , and ) analysis 589

(Table 2) reveals a four-component mixture with two dominant peaks at 336±30 Ma (N=46, 590

Carboniferous; Table 2) and 424±34 Ma (N=38, Devonian; Table 2) and two broad, less signif- 591

icant peaks at 214±53 Ma (N=12) and 185±180 Ma (N=7) (Table 2). The two major peaks de- 592

fined by K-feldspars are significantly different from the two major peaks defined by galenas 593

(Fig. 7). The average Th/U ratio of the K-feldspars (3.93±0.21) also matches the average value 594

for European and circum-Mediterranean galenas (Blichert-Toft et al., 2016), but a t-test 595

(p=310-12) shows that the difference in Th/U between galenas and K-feldspars is significant.

596 597

6. Discussion 598

As explained in Section 2, Pb isotopes do not have enough resolution and robustness to disen- 599

tangle the intricacies of the different mechanisms by which large amounts of metal are accu- 600

mulated in the upper crust but they do respond to the formation of large Pb segregations 601

(U/Th/Pb fractionation) and to the timing of the fractionation events in question. Lead isotopes 602

(Moorbath and Welke, 1968), often complemented by oxygen and strontium isotopes (Michard- 603

Vitrac et al., 1980), clearly help assess the relative contributions of mantle and crust to mag- 604

matic rocks and ore deposits. They also delineate model age provinces reflecting tectonique 605

cycles (Blichert-Toft et al., 2016). However, contrary to the case of VMS deposits (e.g.

606

McCallum et al., 1999; Piercey and Kamber, 2019), they appear of little help in refining the 607

SedEx and MVT genetic models of ore deposits. Due to the enormous fractionation factors of 608

Pb with respect to U and Th during ore formation, Pb isotopes provide testable time constraints 609

(21)

611

The Discussion will first address the issue of Pb model age significance in galenas and K- 612

feldspars and show that this new interpretation fits the Pb isotope data obtained at other locali- 613

ties.

614 615

6.1 A model of early Pb-Zn segregation and a new interpretation of Pb model ages 616

A first-order observation is that the Pb isotope compositions of the present galena and K-feld- 617

spar samples are clearly of crustal origin and distinct from mantle values, e.g. the values meas- 618

ured on basalts and compiled by Hofmann (1997). Mantle Pb sources may be suspected when 619

model ages are negative (Albarède et al., 2020). The origin of the original Pb stock therefore 620

firmly plots in the field of upper crustal rocks. Despite evidence that 207Pb/204Pb and  values 621

are higher in the upper crust than in the mantle, the high 206Pb/204Pb values adopted for the 622

modern Upper Continental Crust by Zartman and Haines (1988) (19.33) and Kramers and 623

Tolstikhin (1997) (19.17) correspond to future model ages, as calculated in Albarède et al.

624

(2012), of 231 and 153 Ma, values that are unreasonably low with respect to observations on 625

galenas, whether conformable or not (Albarède and Juteau, 1984; Stacey and Kramers, 1975).

626

We therefore conclude that Pb model ages calculated from the two-stage modeling of conform- 627

able galenas provide a simple but consistent framework within which to discuss the age and 628

origin of the original Pb source of Pb-Zn ore deposits.

629 630

Lead-zinc strata-bound ores appear to form during certain geological periods. SedEx emplace- 631

ment ages form two clusters, one in the Mesoproterozoic and one in the Paleozoic (Leach et al., 632

2001). Likewise, there appear to be two major episodes of MVT formation, one in the Devo- 633

nian–Permian and the other in the Cretaceous–Tertiary. Even if reliable emplacement ages, and 634

even host rock ages, may not always be available for the Iberian Peninsula, the Pb model ages 635

obtained in the present work provide a different perspective on Pb-Zn pre-concentration. The 636

new galena data define two sharp peaks consistent with Middle to Late Devonian and Early to 637

Middle Cretaceous ages. These ages are regionally very consistent and seem to depend on the 638

size and type of the ore deposits. The former peak dominates the SPZ, the OMZ, the South 639

Central Iberian Zone, and the Catalan Coastal Range. Consolidating the new data and literature 640

data highlights a significant peak in the Neoproterozoic. Adding the and variables does not 641

improve age resolution in 2-dimensional histograms, but confirms the prominence of four 642

(22)

events occurring, in decreasing order of importance, in the Devonian, Mid-Cretaceous, Edia- 643

caran, and Carboniferous. Most samples defining the Ediacaran peak are located in the Pyre- 644

nees, northwestern Spain, and, sparsely, the OMZ and South Central Iberian Zone, which are 645

all localities where Neoproterozoic terranes are present, but predate the Cadomian collision.

646

The Late Devonian peak can exclusively be associated with U-Pb dates from the SPZ arc (Pe- 647

reira et al., 2017, 2012, 2012; Pérez-Cáceres et al., 2017). The case of the Early Cretaceous 648

peak is exemplary because, although this period coincides with the opening of the South Atlan- 649

tic, no major tectonic event or significant magmatic activity are known locally from this period 650

(Martín-Chivelet et al., 2002). Overall, the major peaks of galena model ages do not seem to be 651

associated with tectonic or magmatic events affecting the Iberian domain and emplacement 652

ages, where they can be assessed, are significantly older than the ages of the magmas and fluids 653

that carried the metal load. This case is clear for the Betic belt where Lower Cretaceous Pb 654

model ages are recorded in ores associated with Miocene volcanics.

655 656

It has long been accepted that the source of Pb ores is in the crust, and in particular in sedimen- 657

tary rocks (Heyl et al., 1974). For the Betics, Arribas and Tosdal (1994) suggested that the 658

protolith corresponds to the surrounding Paleozoic metasediments. Clearly, however, not all 659

sedimentary sequences foreshadow the formation of large Pb-Zn deposits. In addition, large- 660

scale circulation of groundwater is expected to average the Pb isotopic composition of the solute 661

to the ‘common lead’ of the time. Selective leaching of lead from some older formation(s) 662

existing at the time of formation of Pb-Zn deposits is therefore needed to ensure that their Pb 663

model ages are older than their emplacement ages.

664 665

Goodfellow and Jonasson (1986), Goodfellow (1987), and Turner (1992) advocated that sedi- 666

ment-hosted Pb-Zn ore deposits formed in a stratified ocean during anoxic or euxinic events.

667

Sáez et al. (2011) suggested a connection between the local anoxia of the local volcano-sedi- 668

mentary environment and the deposition of massive sulfide ores of Paleozoic age at Rammels- 669

berg (Germany), Tharsis (Southern Spain and Portugal), and Draa Sfar (Morocco). Here, Pb 670

model ages that are older than emplacement ages take the assumption of anoxic conditions one 671

step further by suggesting that if the well-defined peaks of model ages characterize unusual 672

sedimentary events, then the current ore deposits formed in a different environment. Remarka- 673

bly, the sequence of events identified by the five-component breakdown of this work (Table 2) 674

is strongly reminiscent of well-known episodes of anoxia in the Phanerozoic ocean. An appeal- 675

(23)

segregation, therefore is the effect of global oceanic anoxic events. Black shales are a typical 677

rock type deposited during anoxic events but the whole-rocks are not by themselves the likely 678

primary source of Pb and Zn ore deposits. In order for the sources of Pb-Zn ores to freeze in 679

the sharp peaks of model ages observed here, Pb isotope compositions require that the 680

=238U/204Pb) values of the parent rocks must be close to zero. Any source rock with >>0 681

inevitably leads to a broad spread in model ages due to variable radiogenic ingrowth. The lim- 682

ited number of available data on Pb isotope systematics of old black shales are indicative of 683

high values (from 9.6to 1000 and more) (Chen et al., 2009; Fetter et al., 2019; Jiang et al., 684

2006). Such elevated parent/daughter ratios render the black shale whole-rocks themselves un- 685

suitable sources of Pb because they would not ‘freeze’ the source Pb isotope compositions but 686

rather drive radiogenic ingrowth well beyond common lead and induce a broad scatter of iso- 687

topic ratios. Although mineralogical studies of shales are few, finely divided, but abundant ga- 688

lena has been described in black shales, often included in arsenopyrite (Abraitis et al., 2004;

689

Belkin and Luo, 2008), and would be more suitable Pb sources.

690 691

The case for sulfides deposited during these events, or in their immediate aftermath, is stronger.

692

As demonstrated for sedimentary Pb-Zn sulfidic ores deposited during the Silurian and Devo- 693

nian of North America (Goodfellow, 1987), the Neoproterozoic and Early Paleozoic at Imiter, 694

Morocco (Essarraj et al., 2016), and on the Yangtze Platform, China (Chen et al., 2009; Jiang 695

et al., 2006), oceanic anoxic events are commonly associated with the deposition of volcano- 696

sedimentary sulfide-rich ores. Oceanic anoxic events are known to coincide with major reduc- 697

tion and drawdown of marine sulfate, enhanced pyrite burial, and major shifts in the 34S of 698

marine barite (Paytan et al., 2004). Oceanic anoxic events further are associated with increase 699

in global weathering (Blättler et al., 2011; Percival et al., 2019), suggested to be triggered by 700

global wildfires (Brown et al., 2012; Kump, 1988), and development of euxinic conditions over 701

large expanses of the ocean bottom (Owens et al., 2013). A comparison between dates of major 702

anoxic events and the model age peaks observed in the present study shows the following cor- 703

respondences:

704 705

1. The oldest anoxic event relevant to the present study took place during the Cryogenian, 706

possibly at the end of the Marinoan glaciation (Sahoo et al., 2012): the largest silver 707

mine of Morocco, Imiter, was exploited since at least Medieval times and is associated 708

(24)

rather large spread of model ages with a peak at 613±42 Ma (Table 2), however, does 710

not warrant further elaboration.

711

2. The dominant model age peak at 395±40 Ma (Upper Devonian) (Table 2) is consistent 712

with the widespread period of anoxia known as the Kellwasser events (Frasnian-Fame- 713

nian extinction) (Buggisch, 1991) and dated at 372 Ma (Percival et al., 2018), possibly 714

triggered by eruption of the Viluy traps, Eastern Siberia (Courtillot et al., 2010; Percival 715

et al., 2019, 2018).

716

3. The Lower Toarcian global anoxic event (~186 Ma; Jenkyns, 1988) is less conspicuous 717

but still detectable as a model age peak in Iberian galenas (Table 2). Ages consistent 718

with this event are found elsewhere in other Pb-Zn ores, e.g. in Britain, France, and 719

Germany (Bode, 2008; Marcoux, 1987; Rohl, 1996).

720

4. In contrast, the well-documented oceanic anoxic events OAE1a (110 Ma), OAE1b 721

(~111 Ma), and OAE2 (93 Ma) (Jenkyns, 2010) overlap with the sharp Pb model age 722

peak in Iberian galenas (90±34 Ma; Table 2). The resolution and accuracy of Pb model 723

ages is insufficient to allow a more specific identification.

724 725

Although robust genetic models of SedEx and MVT ore deposits have become largely consen- 726

sual (Wilkinson, 2014), the original mechanisms leading to the original accumulation of the 727

Pb-Zn sulfide stock are not entirely clear. Galena is notoriously rare in oceanic hydrothermal 728

sulfides (Oudin, 1983) and Pb concentrations in mid-ocean ridge vent fluids are low (< 100 729

ppb, Chen et al., 1986; Michard et al., 1983), simply because Pb abundances in MORB and in 730

the upper mantle are very low (Albarède, 2005; Salters and Stracke, 2004). While the Pb source 731

of sediments is heavily controlled by erosion of continental surfaces (Percival et al., 2019), any 732

sulfide, whether hydrothermal or biogenic, is rapidly oxidized to sulfate in seawater under pre- 733

vailing periods of oxic conditions but not under anoxic conditions. As shown by 34S values in 734

sedimentary pyrite drawn from the euxinic levels of the Black Sea (Lyons, 1997), reduction of 735

marine sulfate during anoxic events does not account for the source of sulfur from massive 736

sulfide deposits (Seal, 2006). Sources of Pb and S therefore most likely are distinct: whenever 737

hydrothermal sulfide spouted from mid-ocean ridges and back-arc basins form VMS deposits, 738

their 34S falls within a relatively narrow range (1-7‰) close to the mantle value (Hannington, 739

2014; Seal, 2006). Anoxic water masses regularly injected by hydrothermal plumes rich in 740

mantle sulfur offer a potential mechanism by which Pb and Zn from continental runoff will 741

Références

Documents relatifs

However, if we take a non-compact group and con- sider the space of all continuous functions with the topology of convergence uniform on every compact set, a novel situation arises

Le coefficient de luminescence dans plusieurs bandes, obtenu par excitation de la bande A, a été mesuré en fonction de la température pour les deux systèmes crystallins et le rôle

En matière d’entretien pour le conjoint, le juge est donc lié par les conclusions des parties et n’est pas autorisé à accorder plus que demandé, même lorsque le

On the contrary, frequency synthesizers based on fractional-N phase-locked loop (PLL) controlling an LC voltage-controlled oscillator (VCO) have been used in miniature atomic

As seen on profiles 9 to 12, low resistivity and high chargeability values (B anomaly on Figures 6 and 7), which were located on the Eastern part seemed to match with the

Like the gaze of small children you know all about me. You have worn my underwear. You have read my newspaper. You have seen my father whip me. You have seen me stroke my father’s

In order to estimate a degree of influence of Earth mag- netic field on the directional properties of the ionospheric perturbation, generated during the 2008 Wenchuan earth- quake,

Importantly, during the incubation of bacterial strains in the presence of PMBN, no release of ß-lactamase activity was observed in the external medium (Fig.  4S