<<

*Revision with No Changes Marked Click here to view linked References

1 Continental weathering and redox conditions during the early Oceanic

2 Anoxic Event in the northwestern Tethys: insight from the Posidonia section

3 in the Swiss Jura Mountains

4

5 Jean-Carlos Montero-Serrano 1,2,3,*, Karl B. Föllmi 1, Thierry Adatte 1, Jorge E. Spangenberg 1,4, Nicolas

6 Tribovillard 5, Alicia Fantasia 1, Guillaume Suan 6

7

8 1 Institut des sciences de la Terre, Université de Lausanne, Bâtiment Géopolis, CH-1015 Lausanne, Switzerland

9 2 Institut des sciences de la mer de Rimouski, Université du Québec à Rimouski, 310 allée des Ursulines, Rimouski, Quebec G5L 3A1,

10 Canada

11 3 Geotop Research Centre, Université du Québec à Montréal, Montréal, Quebec H3C 3P8, Canada

12 4 Institut des dynamiques de la surface terrestre, Université de Lausanne, Bâtiment Géopolis, CH-1015 Lausanne, Switzerland

13 5 Laboratoire Géosystèmes, UMR-8217, Université Lille 1, CNRS, Villeneuve d’Ascq, France

14 6 UMR CNRS 5276 LGLTPE, Université Lyon 1, Campus de la Doua, Bâtiment Géode, F-69622 Villeurbanne Cedex, France 15

16 * Corresponding author: J.-C. Montero-Serrano. Phone: +1.418.723.1986 Ext: 1139; Fax: +1.418.724.1842; E-mail addresse:

17 [email protected] (J-C Montero-Serrano), [email protected] (K. Föllmi), [email protected] (T. Adatte),

18 [email protected] (J. Spangenberg), [email protected] (N. Tribovillard), [email protected] (A. Fantasia),

19 [email protected] (G. Suan) 20

21 Highlights

22 -The T-OAE negative CIE reflects a 13C-depleted carbon injection to ocean-atmosphere

23 -Detrital proxies suggest an increase in continental weathering rates during the T-OAE

24 -Redox proxies suggest moderate anoxic bottom-water conditions during the T-OAE

25 -Dysoxic-to-anoxic conditions prevailed until at least to the bifrons Zone

26 -P-enrichments are observed during the T-OAE

27

1

28 Abstract

29 The geochemistry and mineralogy of organic-rich sediments of the Rietheim succession in

30 northern Switzerland were studied to evaluate the main impacts of the early Toarcian Oceanic Anoxic

31 Event (T-OAE; ~183 Ma ago) on the depositional environment in the NW Tethys Ocean. The

13 13 32 geochemical data obtained (notably, G Corg, G CCaCO3, Rock-Eval pyrolysis, TOCcff) support the

33 hypothesis that the negative C-isotope excursion at the onset of the T-OAE mainly reflects a major

34 perturbation in the global carbon cycle rather than local environmental changes (e.g., basinal restriction,

13 15 35 productivity). Rock-Eval pyrolysis and G Corg-G Ntotal crossplots indicate that the organic matter was

36 primarily of marine origin during the T-OAE, and deposited in an epicontinental sea setting. Detrital

37 proxies [notably, detritus index, siliciclastic balance and Ln(Al2O3/Na2O)] suggest that the elevated

38 pCO2 levels and greenhouse climate prevailing during the T-OAE induced significant acceleration of the

39 hydrologic cycle and an increase in continental chemical weathering rates. Redox-sensitive proxies

40 [relative pyrite contents, organic and sulfide balances, V/(V+Ni) ratios, trace elements enrichment

41 factors, trace-element-TOC co-variation patterns and TOC/Ptotal molar ratios] indicate that the Rietheim

42 sediments accumulated under oxic-to-dysoxic bottom-water conditions, with more

43 reducing conditions during the T-OAE (anoxic and possibly euxinic). These changes in redox conditions

44 were probably induced by substantial thermohaline stratification driven by enhanced freshwater input

45 and the recurrent influx of brackish Arctic seawater into the NW Tethys through the Viking Corridor, in

46 agreement with previous sedimentological and geochemical evidences from the NW Tethys Ocean. In

47 the studied succession, such conditions were perhaps also responsible for sequestration of most of the

48 regenerated P in the deeper water column by adsorption and/or co-precipitation into authigenic phases.

49 Overall, the major palaeoenvironmental changes documented in this study provide a robust example of

50 Earth’s possible responses to the rapid onset of extreme greenhouse conditions.

2

51 Keywords: Early Toarcian Oceanic Anoxic Event; Posidonia Shale; Geochemistry; Mineralogy;

52 Paleoenvironment; Switzerland.

53

54 1. Introduction

55 The early Toarcian (~ 183 Ma ago) was a brief period in the Early marked by an intense

56 global warming to about 5–10°C warmer than at present (e.g., Chandler et al., 1992; Rosales et al., 2004;

57 Dera et al., 2011). The lower Toarcian strata are characterized by a pronounced (4–7‰) near-global

58 negative carbon isotope (G13C) excursion recorded in organic matter, carbonates and fossil wood from

59 open- and marginal-marine sections (e.g., Jenkyns et al., 2002; Hesselbo et al., 2007; Al-Suwaidi et al.,

60 2010; Caruthers et al., 2011; Suan et al., 2011; Izumi et al., 2012; Reolid, 2014). It is currently thought

61 that this intense global warming period and its associated environmental changes were the result of a

62 major perturbation in the global carbon cycle, likely triggered by the combination of several geological

63 processes, such as: (1) greenhouse gas inputs resulting from intensified rift-related tectonics (Bernoulli

64 and Jenkyns, 2009) and large-scale eruptions in the Karoo-Ferrar large igneous province (Duncan et al.,

65 1997; Pálfy and Smith, 2000), (2) thermogenic methane release related with magmatic intrusions and

66 metamorphic alterations of Gondwanan coals and organic-rich in the Karoo-Ferrar province (e.g.,

67 Hesselbo et al., 2000; McElwain et al., 2005; Wignall et al., 2006; Svensen et al., 2007), (3) sudden

68 dissociation of large amounts (~2500 Gt; Beerling and Brentnall, 2007) of continental-shelf methane

69 hydrate deposits (e.g., Hesselbo et al., 2000; Kemp et al., 2005), possibly modulated by orbital forcing

70 (Kemp et al., 2005, 2011). Note that thermogenic and biogenic methane are both powerful greenhouse

13 71 gas enriched in isotopically-light carbon (G Cmethane < -60‰; e.g., Hesselbo et al., 2000), whereby a

72 massive liberation of these gases may rapidly disrupt the global carbon cycle.

73 The extreme greenhouse conditions prevailing during early Toarcian also favoured a period of

74 enhanced marine organic carbon deposition termed the early Toarcian Oceanic Anoxic Event or T-OAE

3

75 (e.g., Jenkyns, 1988; Baudin et al., 1990a). However, the expression of the early T-OAE has a wide

76 geographic variability in terms of organic matter content and degree of oxygenation of bottom waters.

77 Indeed, in many epicontinental sections from NW Europe and N Siberia (Fig. 1a), the T-OAE favoured a

78 widespread deposition of organic-rich lithologies, such as black shales (e.g., Jenkyns, 1988; Baudin et

79 al., 1990a,b; Hesselbo et al., 2000; Röhl et al., 2001; Cohen et al., 2004; Pearce et al., 2008; McArthur et

80 al., 2008; Hermoso et al., 2009a,b; Suan et al., 2011). However, in more open marine sections from S

81 Europe (Portugal - Suan et al., 2008; S Spain - Rodríguez-Tovar and Reolid, 2013), northern margin of

82 Gondwana (Oman - Immenhauser et al., 1998; Morocco - Bodin et al., 2010; Algeria - Reolid et al.,

83 2012, 2014a), as well as northern and southern Panthalassa Ocean (W Canada - Caruthers et al., 2011;

84 NE Japan - Gröcke et al., 2011; NW Argentina - Al-Suwaidi et al., 2010), these organic-rich lithologies

85 are absent or extremely condensed, suggesting that intensity of the T-OAE may have been strongly

86 modulated by local and regional factors. In fact, as suggested in McArthur et al. (2008), the degree of

87 water restriction and oxygen depletion varied regionally in Europe, increasing from South to North along

88 the European Epicontinental Seaway. Overall, many of these studies quoted above have led to debate

89 concerning the intensity of marine anoxia and geographic extent of the T-OAE.

90 In this context and in order to improve the characterization of the T-OAE within the NW Tethys

91 Ocean, we present here a new high-resolution multi-proxy record [including the analysis of stable

13 18 13 15 92 isotopes (G CCaCO3, G OCaCO3, G Corg, G Ntotal), bulk mineralogy, elemental geochemistry, and Rock-

93 Eval pyrolysis] of an early Toarcian section located in the northern Swiss Jura Mountains (Rietheim

94 village, Canton Aargau; Fig. 1). In this study, these new geochemical and mineralogical data have been

95 used not only to explore the expression of the early T-OAE in an area not explored yet for its

96 geochemistry, but also to better document the main impacts of this event on the depositional

97 environment in the NW Tethys Ocean.

98

4

99 2. Geological setting

100 2.1. Depositional setting

101 During the Early to Middle Jurassic, the breakup of Pangaea led to the opening of the Central

102 Atlantic and Tethys oceans (Fig. 1a), accompanied by widespread tectonic extension in what is now the

103 Alpine-Mediterranean region, as well as in northern Europe (e.g., Bernoulli and Jenkyns, 2009). In

104 western Europe, the early Toarcian is characterized by widespread organic-rich lithologies deposited in

105 semi-restricted basins surrounded by shallow shelf seas extending across the West European platform to

106 the northern Tethys Ocean (Fig. 1a; Jenkyns et al., 2002; Wignall et al., 2005; McArthur et al., 2008;

107 Suan et al., 2008; Hermoso et al., 2009a,b, 2013). McArthur et al. (2008) suggested that water restriction

108 varied regionally, being greatest in areas deeper and more distant from the Tethys Ocean (e.g., the

109 Cleveland Basin; United Kingdom) and less restricted towards the south and east across the West

110 European platform (e.g., the Paris Basin and SW Germany).

111 The Lower Jurassic stratigraphy in northern Switzerland (tabular and folded Jura; Fig. 1b) has

112 been summarized recently by Reisdorf et al. (2011). The Swiss deposits accumulated in

113 the relatively slowly subsiding transition zone between the SW Germany Basin and the eastern Paris

114 Basin under fully marine conditions (Reisdorf et al., 2011). Therefore, the lithology and fossil content of

115 the Lower Jurassic deposits in northern Switzerland (such as the Posidonia Shale, renamed as Rietheim

116 Member) are similar to those in SW Germany (e.g., Kuhn and Etter, 1994; Röhl et al., 2001; Röhl and

117 Schmid-Röhl, 2005; Reisdorf et al., 2011).

118

119 2.2. Early Toarcian Posidonia Shale in northern Switzerland

120 The studied lower Toarcian Posidonia Shale is located in the Southern Tabular Jura,

121 approximately 400 m SW of the village of Rietheim (Fig. 1b; Canton Aargau; 47°35’48.75’’N,

122 8°16’39.12’’E; cropping out in the bed of the Steigstrasse Stream). This Posidonia Shale succession

5

123 (~790 cm thick) has been lithostratigraphically and biostratigraphically well studied (Kuhn and Etter,

124 1994) including, in fact, three main ammonite biozones (tenuicostatum, falciferum, and bifrons) and

125 several subzones (exaratum, elegans, falciferum, and commune) (Fig. 2). The transition between the

126 tenuicostatum and falciferum zones is not precisely defined due to a scarcity of index fossils and/or to

127 the presence of a stratigraphic gap in the Pliensbachian-Toarcian transition interval (Kuhn and Etter,

128 1994).

129 The bottom of the section, a glauconite-rich marl layer (~4 cm thick) and light-grey bioturbated

130 marls, attributed to the spinatum Zone from the upper Pliensbachian, are intercalated with two compact

131 marly- beds about 10 and 18 cm thick, respectively (Fig. 2). The light-grey bioturbated marls

132 contain diverse benthic micro- and macrofauna (including ammonites, belemnites and shell coatings) as

133 well as phosphatic fish debris, disseminated framboidal pyrite, and pyritized fossils. The tenuicostatum

134 Zone is represented by the transition of light-grey to dark-grey marly clays with generally low organic

135 contents (<1%) and relatively abundant benthic fauna (notably bivalves and brachiopods). These marly

136 clays (~ 30 cm thick) become more laminated and dark towards the upper part of the tenuicostatum

137 Zone, where continuous organic-rich mudstone sedimentation started (up to the bifrons Zone). Indeed,

138 black to brownish laminated bituminous marly clays (“Kartonschiefer-Fazies”; with more than 6%

139 organic carbon) including fish and wood debris, pyritized ammonites, and disseminated framboidal

140 pyrite prevail at the base of falciferum Zone. In contrast, weakly laminated black to grey bituminous

141 marly clays (“Tonmergel-Fazies”; 1 to 6 wt.% organic carbon) dominate from the mid falciferum to the

142 bifrons zones (Fig. 2). Silt-sized quartz grains are present in all types of laminations. Phosphate is

143 commonly in the form of skeletal apatite at the base of falciferum Zone (as seen in thin sections; Kuhn

144 and Etter 1994). Likewise, several thin bioturbated layers with moderately diverse benthic fauna occur

145 within this bituminous marly-clayey facies. In fact, in comparison to the sediments at the base of

146 falciferum Zone, the abundance of benthic fauna (e.g., Pseudomytiloides dubius) and trace fossils (e.g.,

6

147 Chondrites sp. and Planolites sp.) increases in the upper part of the section (falciferum to commune

148 subzone).

149 The organic-rich mudstones are separated by several finely laminated limestone beds

150 (Spinatumbank, Unterer Stein, Homogene Kalkbänke, Oberer Stein and Monotisbank) into a lower,

151 middle, and upper part (Fig. 2). These limestone beds are likely of early diagenetic origin (Röhl et al.,

152 2001; Röhl and Schmid-Röhl, 2005). The well-laminated limestone beds Unterer Stein and Oberer Stein

153 can be traced over many kilometres within the epicontinental basin and are important marker beds (e.g.,

154 Kuhn and Etter 1994; Röhl et al., 2001). The micro-laminations are composed of alternating

155 argillaceous, organic-rich, and fecal pellet/coccolith-rich layers (Kuhn and Etter 1994; Röhl et al., 2001).

156 Finally, the top of the section (uppermost commune subzone) is marked by a 10–12 cm thick limestone

157 bed (Monotisbank) with abundant shell debris (mainly the bivalve Pseudomonotis substriata) and

158 ammonites (especially Dactylioceratidae).

159

160 3. Material and methods

161 3.1. Samples

162 The lower Toarcian Posidonia Shale succession of Rietheim was sampled systematically along

163 the approximately 790 cm of the stratigraphic log (Fig. 2). A total of 226 samples were collected: 155

164 samples (vertical sample spacing, 2 cm) from a drill core in the first 316 cm of the stratigraphic log (Fig.

165 2), and 71 outcrop samples (vertical sample spacing, 5–10 cm) in the last 470 cm (Fig. 2). An aliquot

166 (about 30 g) of each sample was crushed and pulverized using a Shatterbox with agate mills. Then, each

167 sample was homogenized by sieving through a 100 mesh screen.

168 3.2. Analytical procedure

169 3.2.1. Stable isotope analyses (carbon, oxygen, and nitrogen)

7

13 18 170 Bulk carbonate carbon (G CCaCO3) and oxygen (G OCaCO3) isotope analyses were carried out on

171 powdered bulk-rock samples using a Thermo Fisher Scientific GasBench II and carbonate preparation

172 device interfaced to a Thermo Fisher Scientific Delta Plus XL continuous-flow isotope ratio mass

13 15 173 spectrometer. Organic carbon (G Corg) and total nitrogen (G Ntotal) isotopes were also measured on the

174 organic fraction of decarbonated (10% HCl treatment) samples using a Carlo Erba 1108 elemental

175 analyser connected to a Thermo Fisher Scientific Delta Plus V isotope ratio mass spectrometer via a

176 Confo III interface. The stable carbon and oxygen isotope ratios are reported in the delta notation as the

177 per mil (‰) deviation relative to the Vienna Pee Dee Belemnite (VPDB) standard. The total nitrogen

178 isotope ratios are also reported in the delta notation as the per mil (‰) deviation relative to atmospheric

13 12 18 16 15 14 179 nitrogen (air N2) standard. δ = [(Rsample/Rstandard) - 1] x 1000 with R= C/ C, O/ and N/ N.

180 Analytical precision and accuracy were determined by replicate analyses and by comparison with

181 international and in-house standards (Carrara marble, USGS-24, USGS-32, IAEA-N3, UREA, glycine,

182 and pyridine), and were better than ±0.1‰ (1σ) for carbon, oxygen, and nitrogen.

183 3.2.2. Rock-Eval pyrolysis

184 The type and thermal maturity of the bulk organic matter (OM) was established by Rock-Eval

185 pyrolysis using a model 6 device (Vinci Technologies) under standard conditions (see Béhar et al., 2001

186 for procedural details). The total organic carbon content (TOC, wt.%), Tmax index (°C; OM thermal

187 maturity indicators), Hydrogen Index (HI, mg HC/g TOC), and Oxygen Index (OI, mg CO2/g TOC)

188 were also determined. The HI and OI signatures allow a rough distinction between algal/bacterial (types

189 I and II kerogen) and terrestrial plant (types III and IV kerogen) OM, as well as the degree of

190 degradation and oxidation. Analytical precision was better than 0.05 wt.% (1σ) for TOC, 1.5°C (1σ) for

191 Tmax, 10 mg HC/g TOC (1σ) for HI, and 10 mg CO2/g TOC (1σ) for OI.

192 3.2.3. Total organic carbon on the carbonate-free fraction

8

193 In order to complement Rock-Eval data on the OM characteristics, we also quantified TOC

194 concentrations on the carbonate-free fraction (cff) using a CHN Elemental Analyser (Carlo Erba Flash

195 EA 1112 CHNS/MAS200). The decarbonated fraction was obtained by a double 10% HCl treatment.

196 Precision was better than 1% based on an internal standard (methionine) and replicate samples. The

197 TOCcff allows a robust signature of the organic matter flux regardless of dilution due to carbonate input.

198 3.2.4. Bulk-rock mineralogy

199 The bulk-rock mineralogy was analysed by X-ray diffraction using a Thermo Scientific ARL

200 X’TRA powder diffractometer. This instrument is fitted with a copper tube (Cu K-alpha = 1.540562Å),

201 operating at 40 kV and 30 mA and a post-diffraction graphite monochromator. The random powdered

202 rock samples were scanned from 2° to 65° two-theta in steps of 0.02° two-theta and a counting time of 1

203 second per step, resulting in 3151 data points. For the semi-quantification of the major mineralogical

204 components, the bulk-rock XRD scans obtained were converted into mineral weight percents (wt.%)

205 using the RockJock computer program (Eberl, 2003). This RockJock technique uses a full-pattern fitting

206 method that permits the semi-quantification of whole-sediment mineralogy with a precision of 1-2%. To

207 check the quality of this fitting procedure, a degree-of-fit (DOF = minimum absolute difference) statistic

208 is calculated between the measured and simulated XRD patterns. The DOF obtained with our samples

209 was good with values ranging from as low as 0.131 to as high as 0.402 and averaging 0.202 for the 187

210 samples. For this study the wt.% of the following minerals were calculated (Fig. 6): quartz, potassium

211 (K) feldspar (microcline + orthoclase), plagioclase feldspar (albite), phyllosilicates (biotite, muscovite,

212 illite, chlorite, and kaolinite), pyrite, dolomite, ankerite, and carbonate fluorapatite (CFA).

213 Because the quartz grains are principally detrital in the whole section (as seen in thin sections;

214 Kuhn and Etter 1994), we determined a detritus index, calculated by summing quartz, phyllosilicates, K-

215 feldspar, and Na-plagioclase contents, to observe changes in detrital influx (e.g., Keller et al., 2004;

216 Westermann et al., 2010; Stein et al., 2012; Bomou et al., 2013). High detritus index values may indicate

9

217 higher delivery of terrigenous material from continental sources and/or decreased carbonate productivity

218 and dissolution.

219 3.2.5. Major, minor and trace element analysis

220 Loss on ignition (LOI) was determined gravimetrically by heating the dried samples up to

221 1050°C for two hours. Major and trace-element concentrations were determined by X-ray fluorescence

222 (XRF) spectrometry using a PANalytical PW2400 with an RX tube (Rh anode). Major (and minor)

223 elements (Si, Ti, Al, Fe, Mn, Mg, Ca, Na, K, P) were measured after sample fusion with tetraborate, and

224 trace elements (Nb, Y, Rb, Zn, Cu, Ni, Co, Cr, V, Nd, La, As) after mixing the samples with Mowiol

225 polyvinyl alcohol (2%) and pressed to pellets. The detection limit for major elements is better than 0.01

226 wt.% and for trace elements is generally 1 to 4 ppm. Analytical reproducibility monitored by replicate

227 analyses of selected samples was lower than ±5% for major, minor, and trace elements. Analysis

228 accuracy was assessed by analyses of international and in-house standard reference materials (TS1-

229 Cement, TS3-Clay, TS4-Limestone, TS5-Marlstone, TS7-Sandstone, 372-Portland cement, 368-

230 Dolomite).

231 A compositional Q-mode cluster analysis (e.g., Egozcue and Pawlowsky-Glahn, 2005;

232 Pawlowsky-Glahn and Egozcue, 2011) was performed on our elemental geochemical dataset with the

233 goal of finding elemental associations with similar relative variation patterns that may be interpreted

234 from a palaeoenvironmental standpoint (e.g., Jaminski et al., 1998; Hoffman et al., 1998; Reátegui et al.,

235 2005; Montero-Serrano et al., 2010). This analysis was carried out using a log-ratio approach (Aitchison,

236 1986; Egozcue and Pawlowsky-Glahn 2005). Likewise, as measure of dissimilarity we use the variation

237 array, and as clustering criterion the Ward method. Importantly, in order to isolate the siliciclastic

238 component of the marls and marly clays, calcium-rich samples (Ca > 25 %, ~71 samples) are excluded

239 from the Q-mode cluster analysis. All statistical calculations were conducted with “R” software using the

10

240 packages “StatDA” (Reimann et al., 2008) and “compositions” (van den Boogaart and Tolosana-

241 Delgado, 2008).

242 In order to infer palaeoenvironmental changes (weathering, sediment supply, redox conditions)

243 during the early T-OAE, the stratigraphic distribution of the elemental geochemical associations

244 obtained by the compositional Q-mode cluster analysis were represented as elemental balances.

245 Elemental balances are log-contrasts resulting from a log-ratio of two geometric means of two non-

246 overlapping elemental groups (Egozcue and Pawlowsky-Glahn, 2005). The general expression of an

247 elemental balance is:

248 b = [√ (r x s / r + s)] * Log[gm(A)/gm(B)] [1]

249 where A, B are two non-overlapping elemental groups of a complete composition x of D elements (e.g.,

250 Nb, Rb, La, K, Ti, Al, Si,…, xD), r and s (r + s ≤ D), are the number of elements in A, B respectively,

251 and gm(.) denotes the geometric mean of the elemental groups A and B.

252 Alternatively, the chemical index of alteration (CIA), defined as: CIA = 100 × Al2O3/(Al2O3 + CaO +

253 Na2O + K2O) in molar proportions, may also be used as a proxy to the degree of weathering in the source

254 areas with higher values suggesting more intense chemical weathering (Nesbitt and Young, 1982).

255 However, samples that considerably vary in CaO due to presence of carbonate grains or cements (calcite

256 or dolomite) and phosphates, such as samples studied here, may suggest misleading conclusions if the

257 CIA is used to infer the degree of weathering (e.g., Cullers, 2000; Goldberg and Humayun, 2010). Under

258 this context and based on a statistically model of linear compositional and weathering trends, von

259 Eynatten et al. (2003) suggest that Ln(Al2O3/Na2O) (in molar proportions) may be used as a weathering

260 index similar to CIA. Therefore, we prefer to use the Ln(Al2O3/Na2O) rather than CIA to infer changes

261 in the degree of chemical weathering during deposition of the Rietheim Posidonia Shale succession.

262 Using Ln(Al2O3/Na2O) has the advantage of avoiding uncertainties concerning the correction for CaO

263 associated with phosphate and carbonate phases (e.g., Cullers, 2000; von Eynatten et al., 2003).

11

264 In addition, to compare the relative enrichment of redox-sensitive trace elements (RSTE: Cu, Ni,

265 Zn, V, Co, As), we calculated Enrichment Factors (EF) by comparing Al-normalized metal concentration

266 to those of average shale (Wedepohl, 1991): XEF = [(X/Al)sample / (X/Al)average shale]. In practical terms, EF

267 > 3 represents a detectable authigenic enrichment of an element over average shale concentrations,

268 whereas EF > 10 represents a moderate to strong degree of authigenic enrichment (e.g., Tribovillard et

269 al., 2006; Algeo and Tribovillard, 2009).

270 All analytical data presented are available electronically in the Supplementary Appendix A.

271

272 4. Results

273 4.1. Stable isotopes

13 274 Carbon isotope values of bulk OM (G Corg ~ -32.6 to -25.3‰; Fig. 3a) and the carbonate

13 275 (G CCaCO3 ~ 1.2 to -2.6‰; Fig. 3b) of marls and marly clays show comparable variations throughout the

276 section, with a pronounced negative carbon isotope excursion (or CIE) within the exaratum subzone

13 13 277 (Fig. 3a-b). The magnitude of this negative CIE varies from ~3‰ in G CCaCO3 to ~5‰ in G Corg records.

13 13 278 On average, the G Corg values are in the range of G Corg records of marine OM deposited during

13 279 pronounced greenhouse phases (e.g., Dean et al., 1986; Jenkyns, 2010), whereas G CCaCO3 values are

13 280 within the range of early Toarcian G CCaCO3 records from epicontinental seas of the Peritethyan Realm

281 (e.g., Röhl and Schmid-Röhl, 2005; van de Schootbrugge et al., 2005; Hesselbo et al., 2007; Suan et la.,

282 2008; Hermoso et al., 2009b).

13 283 Extremely light G CCaCO3 values are found within limestone beds and bioturbated marls (e.g., -

13 13 284 4.57‰ in the Unterer Stein; Fig. 3b).G Corg and G CCaCO3 values decrease slightly up section. Oxygen-

18 285 isotope values of the bulk carbonate (G OCaCO3) shows higher values in the upper Pliensbachian interval

286 (-6 to -2.7‰) compared to the lower Toarcian (-8 to -7‰) (Fig. 3c). However, in the Spinatumbank,

18 287 Unterer Stein, and Homogene Kalkbänke limestone beds, the G OCaCO3 values are significantly higher (- 12

15 288 4 to -2.7‰). Nitrogen-isotope values of OM (G Ntotal) range between +1 and +3‰ throughout the

15 289 section (Fig. 3d), with lower values (~ +0.4‰) in the limestone bed Unterer Stein. These G Ntotal values

290 are within the range of values that characterized the Jurassic- interval (Algeo et al., 2014).

291 4.2. Organic matter characterization

292 Total organic carbon (TOC), total organic carbon from the carbonate-free fraction (TOCcff), and

293 the hydrogen index (HI) show lower values (TOC < 0.5 wt.%; TOCcff < 2 wt.%; HI < 70 mg HC/g TOC)

294 in the upper Pliensbachian interval compared to the lower Toarcian (Fig. 3e-g), whereas oxygen index

295 (OI) show higher values (up to 422 mg CO2/g TOC; Fig. 4a). In fact, in the lower Toarcian interval,

296 TOC, TOCcff, HI and OI values range from 2 to 11.8 wt.%, 5 to 18 wt.%, 400 to 794 mg HC/g TOC, 49

297 to 422 mg CO2/g TOC, respectively. High TOC (6 to 11.8 wt.%) and HI (600 to 794 mg HC/g TOC) as

298 well as low OI (<10 mg CO2/g TOC) values characterize the black to brownish laminated bituminous

299 marly clays (exaratum subzone) recording the early Toarcian CIE, whereas lower TOC (< 2 wt.%) and

300 HI (HI < 360 mg HC/g TOC) values characterize the early-diagenetic limestone beds as well as the

301 bioturbated marl intervals. Rock-Eval Tmax values average 427°C without any significant variation (Fig.

302 3h).

303 4.3. Bulk mineralogical composition

304 The bulk mineralogy in the Rietheim Posidonia Shale is dominated by calcite, phyllosilicates, and

305 quartz, and lower proportions of K-feldspar, Na-plagioclase, pyrite, ankerite, dolomite, and carbonate

306 fluorapatite or CFA (Fig. 6 and 7). An unidentified group of organic material and some poorly

307 crystallized minerals (e.g., iron oxide) are unquantified and not considered here (Fig. 7j). In general,

308 calcite distribution is inversely correlated to phyllosilicates, quartz, K-feldspar and Na-plagioclase (Fig.

309 7a-e). Calcite is the dominant mineral (60–96%) from the finely laminated limestone beds (e.g., Unterer

310 Stein, Homogene Kalkbänke, Oberer Stein, and Monotisbank), whereas phyllosilicates (19–53%) and

311 quartz (5–18%) are the dominant minerals in the marl and marly clay intervals (Fig. 7a-e). A decrease in

13

312 calcite content from 60% to 25% is observed at the base of the exaratum subzone (Fig. 7e). The upper

313 Pliensbachian (spinatum Zone) interval is characterized by high ankerite (up to 18%) and dolomite (up to

314 7.5%) contents (Fig. 7f-g). Indeed, the vertical distribution of dolomite mimics the ankerite record. The

315 CFA ranges between 0% to 1.6%, with higher values at the base of falciferum and bifrons zones (Fig.

316 7h). The pyrite content increases (up to 6.9%) in exaratum subzone (broadly coeval with the negative

317 CIE) and then slowly decrease upsection (Fig. 7i).

318 4.4. Major, minor and trace element analysis

319 The compositional Q-mode cluster analysis of the Rietheim Posidonia Shale succession samples

320 revealed three major elemental geochemical associations (Fig. 8):

321 (1) Siliciclastics (Nb, Rb, La, K, Ti, Al, Si, Cr, Na, Y, Nd, Mg): this geochemical group consists of

322 elements associated with clay and silt particles, mainly as chemical constituents of the detrital

323 mineral grains (e.g., phyllosilicates, quartz, Na-plagioclases, K-feldspar) or adsorbed on their

324 surface (e.g., Calvert and Pedersen, 2007; Montero-Serrano et al., 2010). However, Mg shows a

325 different vertical distribution with respect to the other siliciclastics elements (Fig. 8a), likely

326 suggesting that this element is also associated with dolomite and ankerite (Fig. 7f-g). Overall, the

327 variability of riverine and aeolian inputs related to continental climatic fluctuations as well as the

328 hydraulic partitioning of the sediments mostly control the distribution and accumulation of these

329 elements.

330 (2) Organic (P, As, TOC, Ni, V) – sulfide (Zn, Cu, Fe, Co): these elements are commonly associated

331 with the OM and the authigenic sulfides (e.g., pyrite) fraction, whose accumulation is largely

332 controlled by the oxic-anoxic conditions of the sedimentary environment (e.g., Jaminski et al.,

333 1998; Algeo and Maynard, 2004; Tribovillard et al., 2006; Calvert and Pedersen, 2007). Among

334 these elements, V/(V+Ni) ratio and TOC/Ptotal molar ratios are reputed as paleoredox-indices

335 (e.g., Hatch and Leventhal, 1992; Jones and Manning, 1994; Algeo and Maynard, 2004;

14

336 Tribovillard et al., 2006; Algeo and Ingall, 2007). Indeed, V/(V+Ni) ratios lower than <0.46,

337 0.46–0.60, 0.54–0.82, greater than 0.82 may suggest oxic, dysoxic, anoxic and euxinic conditions

338 in the water column., respectively (Hatch and Leventhal, 1992).

339 (3) Carbonates (Ca, Mn): this geochemical group consists of elements related with calcium carbonate

340 (mainly calcite) associated with calcareous cement as well as benthic fauna (including

341 ammonites, belemnites, shell coatings) and coccoliths, which agrees with the petrographic

342 interpretations (Kuhn and Etter 1994). The carbonate-associated elements correlate negatively

343 with the other two elemental associations, and thus likely represent a dilution fraction.

344 Based on these geochemical associations, elemental balances are performed (e.g.,

345 b{siliciclastic/carbonate}, b{sulphide/carbonate} and b{organic/carbonate}). The vertical distribution of siliciclastic (Nb-Rb-

346 La-K-Ti-Al-Si-Cr-Na-Y-Nd), sulfide (Fe-Cu-Zn-Co) and organic (TOC-P-As-V-Ni) balances show

347 similar patterns across the Rietheim Posidonia Shale succession (Fig. 8a-d; Fig. 10c; Fig. 11b-c), with

348 minimum values recorded in the spinatum Zone and maximum values in the negative CIE (exaratum

349 subzone) that then slowly decrease upsection. Similarly, Ln(Al2O3/Na2O) values also increase

350 significantly during the negative CIE (exaratum subzone), then slowly decrease upsection and increasing

351 again during the upper part of commune subzone (Fig. 10b). The V/(V+Ni) ratio range between 0.50 to

352 0.85 with maximum values recorded also during the negative CIE that then slowly decrease upsection

353 (Fig. 11d). RSTE enrichment factors (Cu, Ni, Zn, V, Co, As) show a modest enrichment (2 to 9) within

354 the negative CIE compared to average shale values (Fig. 11e-f). Finally, a good correspondence is

355 observed between the different detrital [detritus index, siliciclastic balance, Ln(Al2O3/Na2O); Fig. 10]

356 and redox [pyrite content, sulfide and organic balances, V/(V+Ni); Fig. 11a-f, i] proxies used here.

357 4.5. Phosphorus contents and TOC/Ptotal molar ratios

358 In the Rietheim succession, profiles of total P contents mimic the stratigraphic trend of the CFA,

359 with higher values during the negative CIE and early bifrons Zone (Figs. 8d and 11g). Furthermore,

15

360 higher P values correspond also to higher TOC values during the negative CIE (Fig. 8d), whereas lower

361 total P contents are characterized in the upper part of exaratum and at the base of elegans subzones.

362 We also determined Corg/Porg molar ratios to evaluate the burial efficiency of P relative to total

363 organic-matter preservation in the sediments (e.g., Algeo and Ingall, 2007; Westermann et al., 2013).

364 Indeed, Corg/Porg molar ratios measured in organic-rich sediments may reflect the importance of

365 preferential P release under dysoxic-to-anoxic bottom-water conditions and as such serve as a

366 paleoredox proxy (e.g., Bodin et al., 2006; Algeo and Ingall, 2007; Mort et al., 2007; Bodin et al., 2010;

367 Stein et al., 2011; Westermann et al., 2013). In this study, we use TOC/Ptotal molar ratios (Ptotal = Preactive

368 + Pdetritic; assuming the Pdetritic is likely to represent well-crystallized authigenic carbonate fluorapatite)

369 rather than Corg/Porg or Corg/Preactive (Preactive = Poxide-associated + Pauthigenic+ Porg; Anderson and Delaney,

370 2001), since as discussed in Algeo and Ingall (2007), the Pdetritic is generally a minor component (mean

371 ~13%) of Ptotal in black shales units. This assumption is in agreement with the compositional Q-mode

372 cluster analysis, where Ptotal is associated with TOC (Fig. 8d-e). The TOC/Ptotal molar ratios calculated

373 for the Rietheim Posidonia Shale succession range from 37 to 383 (median ~ 167) at the base of

374 falciferum Zone, with maximum values within the upper part of the exaratum subzone (Fig. 11h).

375 Conversely, in the spinatum, upper part of the falciferum and base of commune subzones, TOC/Ptotal

376 molar ratios are lower, ranging from 33 to 160 (median ~ 78). In general, the increases in the TOC/Ptotal

377 molar ratios are associated to relatively low P concentrations (Fig. 8d) and CFA contents (Fig. 11g).

378

379 5. Interpretation and discussion

380 5.1. Diagenetic effects

13 18 381 The anomalous G CCaCO3 and G OCaCO3 values observed within limestone beds and bioturbated

382 marls in the Rietheim Posidonia Shale succession likely reflect early diagenetic carbonate precipitation

383 (e.g., Algeo et al., 1992; Röhl and Schmid-Röhl, 2005). Therefore, in order to improve the

16

384 palaeoenvironmental signature of our geochemical records, these limestone beds and bioturbated marls

385 (~ 71 samples of the global geochemical dataset) were excluded from the subsequent interpretation and

386 discussion.

387 Stable oxygen isotopes in carbonates are more affected by diagenesis compared to carbon

18 388 isotopes (Schrag et al., 1995). At the Rietheim Posidonia Shale succession, the observed G OCaCO3

389 pattern (higher values in limestone beds and lower in marls; Fig. 3c) most likely reflects the timing of

390 burial diagenesis, rather than changes in the palaeoceanic isotope signal (such as salinity and

391 temperature). Indeed, during burial diagenesis, limestone cement forms early, in shallow depths and

392 lower burial temperatures, and thus acquires higher G18O values (-5 to -2‰) compared to marls (-9 to -

393 6‰), which cement later (e.g., Algeo et al., 1992; Schrag et al., 1995; Rosales et al., 2001).

18 13 394 In this context and despite the diagenetic alteration of the G OCaCO3 signature, the G CCaCO3 as

13 395 well as the G Corg records probably represent a mainly palaeoenvironmental origin because they are in

396 the same range of values and have similar rates of changes to those found in the Tethys, subboreal and

397 Proto-Atlantic Oceans (e.g., Jenkyns, 1988; Hesselbo and Jenkyns, 1998; Hesselbo et al., 2000; Röhl et

398 al., 2001; Röhl and Schmid-Röhl, 2005; Suan et al., 2008; Hermoso et al., 2009a,b; and references

399 therein).

400 5.2. Maturity, sources and preservation of the organic matter

401 Rock-Eval Tmax values average 427°C without any significant variation, indicating that the OM in

402 the Rietheim Posidonia Shale succession is immature with regards to hydrocarbon generation (Fig. 3h).

403 Therefore, variations in HI and OI cannot be attributed to changes in thermal maturity. The sudden shifts

404 towards lower HI values (<450 mg HC/g TOC) observed through the lower Toarcian interval (e.g., in the

405 thin bioturbated horizons between the falciferum and commune subzones; Fig. 3g and Fig. 4a) likely

406 suggest enhanced OM degradation during intermittent periods of oxygenation (e.g., Röhl and Schmid-

407 Röhl, 2005). The vertical distribution of HI (Fig. 3g) together with HI-OI crossplots (Fig. 4a) indicate

17

408 that OM in the upper Pliensbachian interval mainly consists of recycled and/or terrestrial material (type

409 III/IV kerogen), whereas in the lower Toarcian interval, it is mainly dominated by marine phytoplankton

410 and algal material (type I/II kerogen). Similar OM characteristics have also been reported in nearby

411 lower Toarcian successions from SW Germany (e.g., Röhl et al., 2001; Röhl and Schmid-Röhl, 2005).

13 15 412 The G Corg-G Ntotal crossplots (Algeo et al., 2014) suggest that the Rietheim Posidonia Shale

413 succession was deposited in an open epicontinental sea setting (Fig. 4c). In this depositional setting the

414 variability in sedimentary dynamics might have been driven by fluctuations in sea level (Reisdorf et al.,

415 2011). Indeed, sequence stratigraphic studies from NW Tethys have provided evidence for sea-level fall

416 close to the onset of the early Toarcian negative CIE (e.g., Perilli, 2000; Wignall et al., 2005; Pittet et al.,

417 2014). Accordingly, the OM source changes observed in the Rietheim Posidonia Shale succession may

418 have been related to eustatic sea-level changes. A sea-level lowstand during the upper Pliensbachian

419 interval may explain the increase in terrestrial-derived OM, whereas the enhanced input of marine-

420 derived OM in the lower Toarcian interval may correspond to a rising sea-level trend. Likewise, the

421 relatively high TOC (up to 11.8 wt.%) and TOCcff (up to 18 wt.%) recorded in the lower Toarcian

422 interval (Fig. 3e-f) suggests that likely elevated surface productivity and/or oxygen-depleted bottom-

423 water conditions prevailed in the NW Tethys during the T-OAE. These depositional conditions were

424 very favourable to OM preservation in the Rietheim Posidonia Shale succession. Similar results have

425 been reported by Röhl et al. (2001) and Röhl and Schmid-Röhl (2005) in the Posidonia Shale succession

426 from SW Germany.

427 5.3. Link between organic matter source and organic carbon isotopes

13 13 428 The larger amplitude of the negative CIE observed in G Corg records relative to G CCaCO3 in the

429 Rietheim Posidonia Shale might be related by changes in OM sourcing (e.g., van de Schootbrugge et al.,

13 430 2013). The influence of changes in the type of preserved OM on the G Corg records may be explored by

13 431 comparing pre- and post-excursion G Corg values with coeval HI values (e.g., van de Schootbrugge et

18

13 432 al., 2013). In the Rietheim Posidonia Shale succession,G Corg-HI crossplots data from the spinatum and

13 433 bifrons zones show a significant negative correlation (Fig. 4b), suggesting that part of the bulk G Corg

434 records likely reflect changes in OM sourcing. Following the approach proposed by van de

435 Schootbrugge et al. (2013), the residual changes that are not explained by HI variations may be

13 13 13 436 calculated by subtracting the G Corg values reconstructed from theG Corg-HI crossplots from the G Corg

13 13 437 data. For the studied succession, the obtained residual G Corg record (hereinafter G Corg HI-corr) show an

13 438 overall shape and negative CIE magnitude (~ 3‰) similar to the bulk G CCaCO3 record (Fig. 5b-c). These

13 439 results support the idea that the raw G Corg records cannot be used for reconstructing global carbon cycle

440 perturbations without careful consideration of the OM sources (van de Schootbrugge et al. 2013).

441 5.2. Palaeoenvironmental changes related to the Toarcian Oceanic Anoxic Event

442 The variations observed in the mineralogical and geochemical record of the Rietheim Posidonia

443 Shale section are likely related to the onset of the T-OAE during the Early Jurassic. Hence, we discuss

444 our results in terms of the major palaeoenvironmental changes, which fostered deposition of the

445 Rietheim Posidonia Shale in the NW Tethys Ocean during the early T-OAE.

446 5.2.1. Origin of the negative G13C excursion and global carbon cycle perturbations

13 13 447 G Corg HI-corr and G CCaCO3 records of marls and marly clays show a parallel temporal evolution

448 throughout the Rietheim Posidonia Shale succession, with a pronounced negative CIE (~ 3‰) within the

449 exaratum subzone (Fig. 5). This parallel trend, together with the magnitude of the negative CIE, supports

450 the hypothesis that the early Toarcian negative CIE mainly reflects a major perturbation in the global

451 carbon cycle, likely associated both with an massive output of volcanogenic CO2 and thermogenic

452 methane (CH4) from OM sources in the Karoo-Ferrar large igneous province (e.g., Hesselbo et al., 2000;

453 Svensen et al., 2007; Hesselbo et al., 2007; Suan et al., 2008).

454 However, the magnitude of this negative CIE in the Rietheim Posidonia Shale succession is

455 slightly reduced (~3‰) compared, for example, to lower Toarcian strata from SW Germany

19

456 (Dotternhausen; Röhl and Schmid-Röhl, 2005), the UK (Yorkshire; Hesselbo et al., 2000; Kemp et al.,

457 2005), and northern Siberia (Kelimyar River; Suan et al., 2011) (Fig. 9). In agreement with Kuhn and

458 Etter (1994) and Reisdorf et al. (2011), we attribute this to the presence of a stratigraphic gap in the

13 459 Pliensbachian-Toarcian transition interval. Indeed, most of the tenuicostatum Zone, where G Corg values

460 up to -26‰ are frequently recorded (Röhl and Schmid-Röhl, 2005; Fig. 9b), is missing in the Rietheim

461 Posidonia Shale succession (Kuhn and Etter, 1994). This interpretation is supported by the fact that

13 462 G Corg values recorded just below the limestone bed Unterer Stein in Dotternhausen (Röhl et al., 2001;

463 Röhl and Schmid-Röhl, 2005) and Rietheim are almost identical (~ -32.6‰; Fig. 9a-b). Similar

464 interpretations have also been performed by Röhl and Schmid-Röhl (2005) in the Franconian Posidonia

465 Shale succession from Schesslitz in SW Germany, as well as by Pittet et al. (2014) in the Lusitanian

466 Basin (Portugal) and Ruebsam et al. (2014) in the Lorraine Sub-Basin (NE Paris Basin, France).

467 Therefore, we hypothesize that lower Toarcian strata deposited in a basin-margin setting, such as the

468 Rietheim in northern Switzerland, were probably affected by processes of submarine erosion, reworking,

469 and/or non-deposition due to a fast drop in relative sea-level during the Pliensbachian–Toarcian

470 transition (e.g., Kuhn and Etter, 1994; Hardenbol et al., 1998; Röhl and Schmid-Röhl, 2005; Reisdorf et

471 al., 2011; Pittet et al., 2014; Ruebsam et al., 2014).

472 5.2.3. Continental weathering

473 Changes in the weathering at the source areas and clastic influx that prevailed during deposition

474 of the Rietheim Posidonia Shale succession can be inferred using the mineralogical and geochemical

475 data here obtained (Fig. 10). Indeed, the detritus index, siliciclastic balance and Ln(Al2O3/Na2O) are

476 based on minerals (quartz, phyllosilicates, K-feldspar, Na-plagioclase) and chemical elements, which

477 mainly constitute the detrital component (Nb-Rb-La-K-Ti-Al-Si-Cr-Na-Y-Nd). Consequently, higher

478 values in these detrital proxies could be interpreted here as periods of: (1) stronger rainfall, enhanced

479 chemical weathering and continental runoff, and/or (2) lowered carbonate production in surface water.

20

480 In the Rietheim Posidonia Shale succession, the meaningful increase in the detritus index,

481 siliciclastic balance and Ln(Al2O3/Na2O) observed during the early Toarcian negative CIE (exaratum

482 subzone; Fig. 10) suggest a greater increase in detrital input during this interval relative to the rest of the

483 succession, likely due to an enhanced terrestrial discharge resulting from a strongly accelerated

484 hydrological cycle and higher continental weathering rates induced by intensification of greenhouse

485 conditions during the early T-OAE. Seawater osmium-isotope excursion to more radiogenic values

486 (Cohen et al., 2004) in the NW Tethys Ocean also suggest a warm and humid climate as well as a

487 transient increase in continental weathering rates during the early T-OAE. Likewise, numerical models

488 on regional hydrologic responses to large-scale atmospheric circulation patterns during the Early Jurassic

489 indicate that strong monsoonal circulation dominated along the mid-latitude coasts of Tethys and

490 Panthalassa during the Early Jurassic, including the early Toarcian (e.g., Parrish and Curtis, 1982;

491 Chandler et al., 1992; Dera and Donnadieu, 2012). These monsoonal systems are likely associated with

492 localized pressure cells whose positions are controlled by topography and coastal geography (Chandler

493 et al., 1992). In fact, the large continental areas were mainly located between 30° and 60° latitude in both

494 hemispheres during the Early Jurassic (Fig. 1a), which led to a meridional trade-wind/monsoon-wind

495 circulation pattern (e.g., Parrish and Curtis, 1982; Chandler et al., 1992; Röhl et al., 2001).

496 However, the increase in the detritus index and siliciclastic balance during the negative CIE

497 likely cannot be entirely attributed to a strongly enhanced detrital discharge. The pronounced decrease in

498 both calcite (from 60% down to 14%; Fig. 7e) and Ca (from 20 down to 5%; Fig. 8b) contents at the

499 onset of the negative CIE instead suggest that the increase in these proxies may also reflect a relative

500 decrease in carbonate productivity. Thus, in agreement with several studies on continental weathering,

501 carbonate production and ocean acidification during the early Toarcian (e.g., Röhl et al., 2001; Cohen et

502 al., 2004; Suan et al., 2008; Mattioli et al., 2008, 2009; Hermoso et al., 2009a,b; Dera et al., 2009;

503 Trecalli et al., 2012; Hermoso and Pellenard, 2014; Reolid et al., 2014b), we hypothesize that this drop

21

504 in carbonate content recorded at the onset of the negative CIE in the Rietheim Posidonia Shale

505 succession (Fig. 8b) may likely be the result of a combination of three main factors: (1) dilution by

506 higher detrital input resulting from an accelerated hydrological cycle and enhanced continental runoff,

507 (2) decreasing in carbonate production associated with an enhanced continental runoff and southwards

508 flux of brackish Arctic seawater which controlled turbidity and salinity in surface waters within the NW

509 Tethys Ocean, and/or (3) enhanced carbonate dissolution related to strong seawater acidification

510 conditions driven by a massive injection of CO2 into the atmosphere and then into the ocean.

511 Overall, our mineralogical and geochemical proxies support the hypothesis that the massive

512 injection of CH4/CO2 into the ocean-atmosphere system and subsequent global warming induced during

513 the Early T-OAE an acceleration of the hydrologic cycle and an increase in global chemical weathering

514 rates as well as possibly also decreasing in carbonate production (e.g., Mattioli et al., 2009; Trecalli et

515 al., 2012). However, further investigations are required to improve this latter hypothesis in the Rietheim

516 Posidonia Shale section.

517 5.2.4. Redox conditions

518 In this study, redox conditions that prevailed during the deposition of the Rietheim Posidonia

519 Shale succession are investigated using several proxies (Fig. 11a-f and Fig. 12): pyrite content, sulfide

520 (Fe-Cu-Zn-Co) and organic (TOC-P-As-V-Ni) balances, the V/(V+Ni) ratio, RSTE enrichment factors,

521 relationships between Al-normalized RSTE and TOC and the TOC/Ptotal molar ratios.

522 -Inferences from pyrite, redox-sensitive elements and V/(V+Ni) ratio

523 The stratigraphic distribution of relative contents of pyrite (%), sulfide (Fe-Cu-Zn-Co) and

524 organic components (TOC-P-As-V-Ni) as well the V/(V+Ni) ratio show similar trends across the

525 Rietheim Posidonia Shale succession, with higher values recorded during the negative CIE (exaratum

526 subzone) that then slowly decrease upsection (Fig. 11a-d). These stratigraphic trends, together with

527 threshold values proposed by Hatch and Leventhal (1992) for the V/(V+Ni) ratio (see Section 4.4),

22

528 suggest that sediments in the exaratum subzone represent the onset as well as peak of the T-OAE

529 episode where anoxic (and possibly euxinic) conditions prevailed at the sediment-water interface,

530 whereas the remainder of the falciferum Zone and the overlying bifrons Zone represent a continuation of

531 oxic-to-dysoxic conditions. The modest enrichment (2 to 9), which exhibit the RSTE (Cu, Ni, Zn, V, Co,

532 As) compared to average shale values within the negative CIE also supports the interpretation that

533 anoxic bottom-water conditions developed during the exaratum subzone (Fig. 11e-f). Similar results

534 have been obtained by Brumsack (1991) in the nearby Posidonia Shale succession from the SW

535 Germany.

536 The redox conditions, which prevailed in the sedimentary environment, can also be inferred

537 through the co-variation patterns between the Al-normalized RSTE and TOC (Fig. 12a; Algeo and

538 Maynard, 2004; Tribovillard et al., 2006). Note in Fig. 12 that Al-normalized RSTE (V, Ni, As and Cu)

539 plotted versus TOC allows the discrimination of different redox facies (dysoxic, anoxic and euxinic) in

540 both modern (Irino and Pedersen, 2000; Tada et al., 2000; Piper and Dean, 2002) and ancient (Arthur et

541 al., 1990; Brumsack, 1991; Warning and Brumsack, 2000; Algeo and Maynard, 2004) marine

542 depositional environments. In the Rietheim Posidonia Shale succession, these Al-normalized RSTE-TOC

543 co-variation patterns (Fig. 12b-e) suggest that the predominating palaeo-oxygenation conditions during

544 the deposition of these sediments were (1) oxic during the spinatum subzone, (2) anoxic (possibly

545 euxinic) during the exaratum subzone, and (3) dysoxic, with intermittent anoxia, during the elegans to

546 commune subzones.

547 Based on the similar covariation between the detritus index, siliciclastics, organic and sulfide

548 balances during the negative CIE (Fig. 10), we hypothesize that detrital inputs may indirectly represent

549 the main cause for the modest excess of trace-metal content in the exaratum subzone. Indeed, in

550 marginal environments, enhanced continental runoff during the early T-OAE likely provided high

551 amounts of nutrients, which lead to increased productivity and greater settling of OM. In addition,

23

552 according to Mattioli et al. (2009) and Dera and Donnadieu (2012), enhanced freshwater input in

553 conjunction with the recurrent influx of brackish Arctic seawater through the Viking Corridor could have

554 increased the themohaline stratification within the NW Tethys Ocean during the early T-OAE. Overall,

555 the combined effects of these processes likely induced depletion in bottom-water oxygen levels and thus

556 a modest RSTE enrichment. Consequently, from a geochemical point of view, the organic-rich sediments

557 of the exaratum subzone were likely deposited in a sedimentary environment very similar to the more

558 recent Plio-Pleistocene Black Sea and/or Mediterranean sapropels (e.g., Brumsack, 1991). That is, a

559 stratification of the water column due to salinity differences and the prevalence of anoxic to euxinic

560 bottom-water conditions (e.g., Murray et al., 1989; Brumsack, 1991).

561 -Inferences from phosphorus contents and TOC/Ptotal molar ratios

562 P is an important limiting nutrient in the ocean, driving primary productivity (e.g., Tyrrell, 1999;

563 Benitez-Nelson, 2000). The flux of dissolved P into the ocean is mainly controlled by continental runoff

564 (Föllmi, 1996), and the transfer of P into sediments occurs by incorporation into OM and fish debris or

565 by aggregation onto clay particles, iron, and manganese oxyhydroxides (Delaney, 1998). However, P

566 burial within sediments is strongly influenced by changes in the benthic redox conditions (e.g., van

567 Cappellen and Ingall, 1994; Ingall et al., 2005; Algeo and Ingall, 2007; Mort et al., 2007). Indeed,

568 dysoxic-to-anoxic bottom-water conditions not only promote an enhanced preservation of OM but also a

569 remobilization and release of mineral-associated phosphate and organic P into the uppermost sediments

570 and overlying water column (van Cappellen and Ingall, 1994; Ingall et al., 2005; Mort et al., 2007).

571 Under such conditions, the increased regeneration of P in sediments may lead to both enhanced

572 phosphate recycling (low P burial efficiency) and increased primary productivity in a positive feedback

573 loop (van Cappellen and Ingall, 1994; Ingall et al., 2005), resulting in sedimentary Corg/Porg ratios higher

574 than the Redfield ratio of 106 (Redfield, 1958, Algeo and Ingall, 2007).

24

575 In Rietheim Posidonia Shale succession, TOC/Ptotal molar ratios higher than the Redfield ratio

576 (106 to 383; Fig. 11h) at the base of falciferum Zone suggest that the deposition of organic-rich

577 sediments during the early Toarcian may have occurred under more oxygen-depleted conditions. These

578 interpretations are in agreement with the stratigraphic trends of all RSTE proxies (Fig. 11a-f and Fig.

579 12), which suggested a moderate shift toward anoxic conditions. Conversely, TOC/Ptotal molar ratios

580 lower than the Redfield ratio (<106) associated to relatively high P and CFA contents in the spinatum,

581 upper part of the falciferum and base of commune subzones suggest that the depositional environment

582 evolved towards more oxic-to-dysoxic bottom waters. Note that the presence of oxygen not only

583 promotes the oxidative destruction of OM but also retention of remineralized organic P within the

584 sediment (Algeo and Ingall, 2007). The predominance of bioturbated layers in the upper part of the

585 Rietheim Posidonia Shale succession supports this interpretation (Fig. 2). However, because high P and

586 CFA contents are not coincident with maxima in TOC content in the spinatum, upper part of the

587 falciferum and base of commune subzones, these P-rich intervals might also be related with a high

588 abundance of preserved fish debris within the sediment (see Section 2.2). Large amount of fish debris

589 have also been reported in the falciferum-bifrons Zone boundary from lower Toarcian sections of the SW

590 Germany (Rohl and Schmid-Rohl, 2005) and SE France (Suan et al., 2013).

591 On the other hand, the high P contents associated with the high TOC values during the negative

592 CIE (Fig. 8d) suggest that a part of the remobilized P during redox recycling was likely trapped within

593 the sediments in authigenic P bearing minerals (e.g., authigenic apatite). Indeed, we hypothesize that

594 oxygen-depleted bottom-water condition in conjunction with enhanced surface water stratification led to

595 sequestration of a large part of the regenerated P in the deeper water column by adsorption and/or co-

596 precipitation into authigenic phases (e.g., apatite). This has also been observed in Mediterranean

597 sapropels (e.g., Slomp et al., 2004) and organic-rich sediments from the early Aptian OAE (Stein et al.,

598 2011; Westermann et al., 2013) as well as Cenomanian/Turonian OAE (Mort et al., 2007; Kraal et al.,

25

599 2010), which received a sudden increase of OM. However, as discussed above, a part of the total P

600 contents in sediments deposited under anoxic conditions might also be driven by the abundance of

601 phosphatic fish debris (Schenau and De Lange, 2001), which may either accumulate in sediments due to

602 combination of several environmental factors such as high production of fish debris in productive surface

603 waters, abrupt fish mortality and efficient preservation under anoxic conditions. Indeed, according to

604 selective extraction data from eastern Mediterranean (Slomp et al., 2004) and Arabian Sea sediments

605 (Schenau and de Lange, 2000, 2001) anoxic bottom waters increase the preservation of phosphatic fish

606 hard parts, becoming an important reactive P sink within the sediments (Slomp and Van Cappellen,

607 2007; Paytan and McLaughlin, 2007; März et al., 2014). In the Rietheim Posidonia Shale succession,

608 thin sections analysis of black shale samples show the presence of phosphatic fish debris (Kuhn and Etter

609 1994). This would explain the good correlation between Ptotal and CFA contents through the entire

610 section (Fig. 8d and Fig. 11g-h). Overall, our results suggest that changes in TOC/Ptotal molar ratios

611 cannot be entirely attributed to changes in redox conditions in the Rietheim Posidonia Shale succession.

612 Alternatively, high P contents in organic-rich sediments deposited under anoxic conditions may

613 also be related to the development of benthic microbial mats, such as coccoid bacteria and cryptalgal

614 (Tribovillard et al., 2010; Westermann et al., 2013). In fact, the presence of abundant biomass in the

615 sediment together with the development of microbial mats at the sediment–water interface might have

616 limited the exchanges between the water column and the pore space within sediment, hampering pore

617 water renewal and promoting the quantitative trapping of remobilized P in the sediment during redox

618 recycling. This environmental process has been recently proposed to operate during the deposition of

619 Jurassic organic-rich carbonates from the French Jura Mountains (Tribovillard et al., 2010).

620

621 6. Conclusions

26

622 The results of this high-resolution geochemical-mineralogical research allow for the following

623 generalizations and conclusions:

13 13 624 1. Carbon isotope (G Corg-corr, G CCaCO3), Rock-Eval pyrolysis, and TOCcff data provide clear

625 evidence that the pronounced ~3‰ negative CIE recorded in the Rietheim Posidonia Shale

626 succession during the onset of the T-OAE mainly reflects a major perturbation in the global

627 carbon cycle, likely induced by a massive injection of 13C-depleted carbon into the whole

628 exchangeable carbon reservoir, rather than changes in the type of OM and/or degree of

629 stratification of bottom waters.

13 15 630 2. The Rock-Eval pyrolysis data and G Corg-G Ntotal crossplots suggest that the OM in the upper

631 Pliensbachian and lower Toarcian interval is primarily of terrestrial and marine origin,

632 respectively, and deposited in an epicontinental sea setting.

633 3. The detrital proxies [notably, detritus index, siliciclastic balance and Ln(Al2O3/Na2O)]

634 corroborate that the elevated pCO2 levels and greenhouse climate prevailing during the T-OAE

635 induced an acceleration of the hydrologic cycle and an increase in chemical weathering rates.

636 4. The redox-sensitive proxies [notably, relative pyrite contents, organic and sulfide balances,

637 V/(V+Ni) ratios, RSTE enrichment factors, Al-normalized RSTE versus TOC co-variation

638 patterns, and TOC/Ptotal molar ratios] suggest that the sedimentary deposition of the Rietheim

639 Posidonia Shale succession took place under oxic-to-dysoxic bottom-water conditions, slightly

640 shifting toward anoxic and possibly euxinic conditions during the negative CIE. The T-OAE

641 influence on the sedimentary environment prevailed until at least to the bifrons Zone. Overall,

642 our results further support the idea that the influx of brackish Arctic seawater through the Viking

643 Corridor could have facilitated the thermohaline stratification and subsequent dysoxic-to-anoxic

644 bottom-water conditions within the NW Tethys Ocean (Dera and Donnadieu, 2012).

27

645 5. Enhanced rainfall and river freshwater inputs together with recurrent discharges of boreal

646 brackish seawaters into the NW Tethys through the Viking Corridor (Dera and Donnadieu,

647 2012), appear to be a plausible scenario for explaining the P-enrichments (mainly authigenic P)

648 in the Rietheim Posidonia Shale sediments during the early Toarcian warming event.

649 6. The geochemical characteristics of organic-rich sediments from the exaratum subzone suggest

650 they were likely deposited in a sedimentary environment very similar to the Black Sea and/or

651 Mediterranean sapropels.

652 Taken as a whole, our high-resolution geochemical and mineralogical data provide a solid record

653 of the major palaeoenvironmental changes that fostered deposition of the Rietheim Posidonia Shale

654 succession within the NW Tethys Ocean during the development of the T-OAE, and corroborate that

655 deposition of the lower Toarcian black shales was mainly driven by a global climate change even though

656 it was influenced by local environmental changes.

657

658 Acknowledgements

659 This research was funded through the GeoNova project and Université de Lausanne, with the

660 financial support of the Swiss University Conference (CUS). We acknowledge Jean-Claude Lavanchy

661 (Centre d’Analyse Minérale, UNIL), Hassan Saleh Khozyem/Tiffany Monnier (ISTE-UNIL) and

662 Brahimsamba Bomou (ISTE-UNIL) for their support with the XRF, Rock-Eval and elemental analyses,

663 respectively. We thank Thomas Algeo (University of Cincinnati) for his comments and suggestions,

664 which significantly helped to improve the early version of this work. We further thank Christine Laurin

665 for reviewing the grammar. Lastly, we specially thank the editor Finn Surlyk and the reviewer Matías

666 Reolid (Universidad de Jaén) for their constructive and helpful comments, which significantly improved

667 this work.

668

28

669 Appendix A. Supplementary material

670 Supplementary data associated with this article can be found in the online version at http…

671 File Excel. Rietheim Posidonia Shale dataset.exe

672

673 References

674 Anderson, L.D., Delaney, M.L., 2001. Carbon to phosphorus ratios in sediments: implications for

675 nutrient cycling. Global Biogeochemical Cycles 15, 65-79.

676 Aitchison, J., 1986. The statistical analysis of compositional data. Monographs on statistics and applied

677 probability. Chapman & Hall Ltd., London; reprinted in 2003 by the Blackburn Press, New Jersey,

678 416 p.

679 Algeo, T.J., Wilkinson, B.H., Lohmann, K.C, 1992. Meteoric-burial diagenesis of Pennsylvanian

680 carbonate: water/rock interactions and basin geothermics. Journal of Sedimentary Petrology 62, 652-

681 670.

682 Algeo, T.J., Maynard, J.B., 2004. Trace-element behavior and redox facies in core shales of Upper

683 Pennsylvanian Kansas-type cyclothems. Chemical Geology 206, 289-318.

684 Algeo, T.J., Ingall, E., 2007. Sedimentary Corg:P ratios, paleocean ventilation, and Phanerozoic

685 atmospheric pO2. Palaeogeography, Palaeoclimatology, Palaeoecology, 256, 130-155.

686 Algeo, T.J., Tribovillard, N., 2009. Environmental analysis of paleoceanographic systems based on

687 molybdenum–uranium covariation. Chemical Geology 268, 211-225.

688 Algeo, T. J., Meyers, P. A., Robinson, R. S., Rowe, H., and Jiang, G. Q., 2014. Icehouse–greenhouse

689 variations in marine denitrification. Biogeosciences 11, 1273-1295.

690 Al-Suwaidi, A.H., Angelozzi, G.N., Baudin, F., Damborenea, S.E., Hesselbo, S.P., Jenkyns, H.C.,

691 Manceñido, M.O., Riccardi, A.C., 2010. First record of the Early Toarcian Oceanic Anoxic Event

29

692 from the Southern Hemisphere, Neuquén Basin, Argentina. Journal of the Geological Society of

693 London 167, 633-636.

694 Arthur, M.A., Jenkyns, H.C., Brumsack, H.-J., Schlanger, S.O., 1990. Stratigraphy, geochemistry, and

695 paleoceanography of organic-carbon rich Cretaceous sequences. In: Ginsburg, R.N., Beaudoin, B.

696 (Eds.), Cretaceous Research. Events and Rhythms Kluwer Acad, Norwell, pp. 75-119.

697 Baudin, F., Herbin, J.-P., Vandenbroucke, M., 1990a. Mapping and geochemical characterization of

698 Toarcian organic matter in the Mediterranean Tethys. Organic Geochemistry 16, 677-687.

699 Baudin, F., Herbin, J.-P., Bassoullet, J.-P., Dercourt, J., Lachkar, G., Manivit, H. and Renard, M. 1990b.

700 Distribution of organic matter during the Toarcian in the Mediterranean Tethys and Middle East.

701 73–92. In HUC, A. Y. (ed.). Deposition of organic facies. Vol. 30. American Association of

702 Petroleum Geologists, Tulsa, 234 pp.

703 Beerling, D. J., Brentnall, S. J., 2007. Numerical evaluation of mechanisms driving Early Jurassic

704 changes in global carbon cycling. Geology 35, 247-250.

705 Benitez-Nelson, C.R., 2000. The biogeochemical cycling of phosphorus in marine systems. Earth-

706 Science Reviews 51, 109-135.

707 Bernoulli, D., Jenkyns, H.C., 2009. Ancient oceans and continental margins of the Alpine–

708 Mediterranean Tethys: deciphering clues from Mesozoic pelagic sediments and ophiolites.

709 Sedimentology 56, 149-190.

710 Behar, F., Beaumont, V., Penteado, H.L.D., 2001. Rock-Eval 6 technology: performances and

711 developments. Oil and Gas Science and Technology 56, 111-134.

712 Bodin, S., Godet, A., Matera, V., Steinmann, P., Vermeulen, J., Gardin, S., Adatte, T., Coccioni, R.,

713 Föllmi, K.B., 2006. Enrichment of redox-sensitive trace metals (U, V, Mo, As) associated with the

714 late Hauterivian Faraoni oceanic anoxic event. International Journal of Earth Sciences (Geologische

715 Rundschau) 96, 327-341.

30

716 Bodin, S., Mattioli, E., Fröhlich, S., Marshall, J.D., Boutib, L., Lahsini, S., Redfern, J., 2010. Toarcian

717 carbon isotope shifts and nutrient changes from the Northern margin of Gondwana (High Atlas,

718 Morocco, Jurassic): palaeoenvironmental implications. Palaeogeography, Palaeoclimatology,

719 Palaeoecology 297, 377-390.

720 Bomou, B., Adatte, T., Tantawy, A. A., Mort, H., Fleitmann, D., Huang, Y., Föllmi, K.B., 2013. The

721 expression of the Cenomanian–Turonian oceanic anoxic event in Tibet. Palaeogeography,

722 Palaeoclimatology, Palaeoecology 369, 466-481.

723 Brumsack, H.-J., 1991. Inorganic geochemistry of the German Posidonia Shale: paleoenvironmental

724 consequences. In: Modern and Ancient Continental Shelf Anoxia (R.V. Tyson & T.H. Pearson,

725 Edts.), Geological Society, London, Special Publications 58, 353-362.

726 Calvert, S.E., Pedersen, T.F., 2007. Elemental proxies for palaeoclimatic and palaeoceanographic

727 variability in marine sediments: interpretation and application. In: Hillaire-Marcel, C., Vernal, A.D.

728 (Eds.), Proxies in Late Cenozoic Paleoceanography. Elsevier, Amsterdam, pp. 567–644.

729 Caruthers, A.H., Gröcke, D.R., Smith, P.L., 2011. The significance of an early Jurassic (Toarcian)

730 carbon isotope excursion in Haida Gwaii (Queen Charlotte Islands), British Columbia, Canada.

731 Earth and Planetary Science Letters 307, 19-26.

732 Chandler, M.A., Rind, D., Ruedy, R., 1992. Pangaean climate during the Early Jurassic: GCM

733 simulations and the sedimentary record of paleoclimate. Geological Society of America Bulletin

734 104, 543-559.

735 Cohen, A.S., Coe, A.L., Harding, S.M., Schwark, L., 2004. Osmium isotope evidence for the regulation

736 of atmospheric CO2 by continental weathering. Geology 32, 157-160.

737 Cullers, R. L., 2000. The geochemistry of shales, siltstones and sandstones of Pennsylvian- age,

738 Colorado, USA: implications for provenance and metamorphic studies. Lithos 51, 181–203.

31

739 Dean, W.E., Arthur, M.A., Claypool, G.E., 1986. Depletion of 13C in Cretaceous marine organic matter:

740 source, diagenetic, or environmental signal? Marine Geology 70, 119-157.

741 Delaney, M.L., 1998. Phosphorus accumulation in marine sediments and the oceanic phosphorus cycle.

742 Global Biogeochemical Cycles 12, 563-572.

743 Dera, G., Pellenard, P., Neige, P., Deconinck, J.-F., Pucéat, E., Dommergues, J.-L., 2009. Distribution of

744 clay minerals in Early Jurassic Peritethyan seas: palaeoclimatic significance inferred from

745 multiproxy comparisons. Palaeogeography, Palaeoclimatology, Palaeoecology 271, 39-51.

746 Dera, G., Brigaud, B., Monna, F., Laffont, R., Pucéat, E., Deconinck, J.-F., Pellenard, P., Joachimski,

747 M.M., Durlet, C., 2011. Climatic ups and downs in disturbed Jurassic world. Geology 39, 215-218.

748 Dera, G., Donnadieu, Y., 2012. Modeling evidences for global warming, Arctic seawater freshening, and

749 sluggish oceanic circulation during the Early Toarcian anoxic event. Paleoceanography 27, PA2211,

750 doi:10.1029/2012PA002283.

751 Duncan, R.A., Hooper, P.R., Rahacek, J., Marsh, J.S. & Duncan, A.R., 1997. The timing and duration of

752 the Karoo igneous event, southern Gondwana. Journal of Geophysical Research, 102, 18127-18138.

753 Eberl, D.D., 2003. User guide to RockJock—a program for determining quantitative mineralogy from X-

754 ray diffraction data. USGS Open File Report OF 03-78. 40 pp.

755 Egozcue, J.J., Pawlowsky-Glahn, V., 2005. Group of parts and their balances in compositional data

756 analysis. Mathematical Geology 37, 795-828.

757 Föllmi, K.B., 1996. The phosphorus cycle, phosphogenesis and marine phosphate-rich deposits. Earth-

758 Science Reviews 40, 55-124.

759 Goldberg, K., Humayun, M., 2010. The applicability of the chemical index of alteration as a

760 paleoclimatic indicator: An example from the Permian of the Paraná Basin, Brazil.

761 Palaeogeography, Palaeoclimatology, Palaeoecology 293, 175–183.

32

762 Gröcke, D.R., Hori, R. S., Trabucho-Alexandre, J., Kemp, D. B., Schwark, L., 2011. An open ocean

763 record of the Toarcian oceanic anoxic event. Solid Earth 2, 245-257.

764 Hatch, J.R., Leventhal, J.S., 1992. Relationship between inferred redox potential of the depositional

765 environment and geochemistry of the Upper Pennsylvanian (Missourian) Stark Shale Member of the

766 Dennis Limestone, Wabaunsee County, Kansas, U.S.A. Chemical Geology 99, 65–82.

767 Hallam, A., 2001. A review of broad pattern of the Jurassic sea-level changes and their possible causes

768 in the light of current knowledge. Palaeogeography, Palaeoclimatology, Palaeoecology167, 23-37

769 Hardenbol, J., Thierry, J., Farley, M.B., de Graciansky, P.-C., Vail, P.R., 1998. Mesozoic and Cenozoic

770 sequence chronostratigraphic framework of European basins. In: de Graciansky, P.-C., Hardenbol,

771 J., Jacquin, T., Vail, P.R. (Eds.), Mesozoic and Cenozoic Sequence Stratigraphy of European basins.

772 Special Publication, Society for Sedimentary Geology, vol. 60, pp. 3-13.

773 Hermoso, M., Le Callonnec, L., Minoletti, F., Renard, M., Hesselbo, S.P., 2009a. Expression of the

774 Early Toarcian negative carbon-isotope excursion in separated carbonate microfractions (Jurassic,

775 Paris Basin). Earth and Planetary Science Letters 277, 194-203.

776 Hermoso, M., Minoletti, F., Le Callonnec, L., Jenkyns, H.C., Hesselbo, S.P., Rickaby, R.E.M., Renard,

777 M., de Rafélis, M., Emmanuel, L., 2009b. Global and local forcing of Early Toarcian seawater

778 chemistry: a comparative study of different paleoceanographic settings (Paris and Lusitanian

779 basins). Paleoceanography 24, PA4208, doi:10.1029/2009PA001764.

780 Hermoso, M., Minoletti, F., Pellenard, P., 2013. Black shale deposition during Toarcian super-

781 greenhouse driven by sea level. Climate of the Past 9, 2703-2712.

782 Hermoso, M., Pellenard, P., 2014. Continental weathering and climatic changes inferred from clay

783 mineralogy and paired carbon isotopes across the early to middle Toarcian in the Paris Basin.

784 Palaeogeography, Palaeoclimatology, Palaeoecology 399, 385–393.

33

785 Hesselbo, S.P., Jenkyns, H.C., 1998. British Lower Jurassic sequence stratigraphy. In: DeGraciansky, P.-

786 C., Hardenbol, J., Jacquin, T., Vail, P.R., Farley, M.B. (Eds.), Sequence Stratigraphy of European

787 Basins: Special Publication of the SEPM, 60, pp. 561-581.

788 Hesselbo, S.P., Gröcke, D.R., Jenkyns, H.C., Bjerrum, C.J., Farrimond, P., Morgans Bell, H.S., Green,

789 O.R., 2000. Massive dissociation of gas hydrate during a Jurassic oceanic anoxic event. Nature 406,

790 392-395.

791 Hesselbo, S.P., Jenkyns, H.C., Duarte, L.V., Oliveira, L.C.V., 2007. Carbon-isotope record of the Early

792 Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine carbonate (Lusitanian

793 Basin, Portugal). Earth and Planetary Science Letters 253, 455-470.

794 Hoffman, D.L., Algeo, T.J., Maynard, J.B., Joachimski, M.M., Hower, J.C., and Jaminski, J., 1998.

795 Regional and stratigraphic variation in bottomwater anoxia in offshore core shales of Upper

796 Pennsylvanian cyclothems from the Eastern Midcontinent Shelf (Kansas), USA. In: Schieber, J.,

797 Zimmerle, W., and Sethi, P.S., eds., Shales and Mudstones, v. 1: Stuttgart, Schweizerbart'sche, p.

798 243-269.

799 Immenhauser, A., Schreurs, G., Peters, T., Matter, A., Hauser, M., Dumitrica, P., 1998. Stratigraphy,

800 sedimentology, and depositional environments of the Permian to uppermost Cretaceous Batain

801 Group, eastern Oman. Eclogae geologicae Helvetiae 91, 217–235.

802 Ingall, E., Kolowith, L., Lyons, T., Hurtgen, M., 2005. Sediment carbon, nitrogen and phosphorus

803 cycling in an anoxic fjord, Effingham Inlet, British Columbia. American Journal of Science 305,

804 240-258.

805 Irino, T., Pedersen, T. 2000. Geochemical character of glacial to interglacial sediments at Site 1017,

806 southern Californian Margin: Minor and trace elements. Proc. Ocean Drill. Program Sci. Results

807 167, 263-271.

34

808 Izumi , K., Miyaji, T., Tanabe, K., 2012. Early Toarcian (Early Jurassic) oceanic anoxic event recorded

809 in the shelf deposits in the northwestern Panthalassa: evidence from the Nishinakayama Formation

810 in the Toyora area, west Japan. Palaeogeography, Palaeoclimatology, Palaeoecology 315-316, 100-

811 108.

812 Jaminski, J., Algeo, T.J., Maynard, J.B., and Hower, J.C., 1998. Climatic origin of dm-scale

813 compositional cyclicity in the Cleveland Member of the Ohio Shale (Upper ), Central

814 Appalachian Basin, USA. In: Schieber, J., Zimmerle, W., and Sethi, P.S., eds., Shales and

815 Mudstones, v. 1: Stuttgart, Schweizerbart'sche, p. 217-242.

816 Jenkyns, H.C., 1988. The early Toarcian (Jurassic) event: stratigraphy, sedimentary, and geochemical

817 evidence. American Journal of Science 288, 101-151.

818 Jenkyns, H.C., Gröcke, D., Hesselbo, S.P., 2001. Nitrogen isotope evidence for water mass

819 denitrification during the early Toarcian Oceanic Anoxic Event. Paleoceanography, 16, 593-603.

820 Jenkyns, H.C., Jones, C.E., Gröcke, D.R., Hesselbo, S.P., Parkinson, D.N., 2002. Chemostratigraphy of

821 the Jurassic system: applications, limitations and implication for palaeoceanography. Journal of the

822 Geological Society 159, 351-378.

823 Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochemistry, Geophysics, Geosystems

824 11, Q03004, doi:10.1029/2009GC002788.

825 Jones, B., Manning, D.A.C., 1994. Comparison of geochemical indices used for the interpretation of

826 palaeoredox conditions in ancient mudstones. Chemical Geology 111, 111–129.

827 Keller, G., Pardo, A., 2004. Age and paleoenvironment of the Cenomanian–Turonian global stratotype

828 section and point at Pueblo, Colorado. Marine Micropaleontology 51, 95-128.

829 Kemp, D.B., Coe, A.L., Cohen, A.S., Schwark, L., 2005. Astronomical pacing of methane release in the

830 Early Jurassic period. Nature 437, 396-399.

35

831 Kraal, P., Slomp, C.P., Forster, A., Kuypers, M.M.M., 2010. Phosphorus cycling from the margin to

832 abyssal depths in the proto-Atlantic during oceanic anoxic event 2. Palaeogeography,

833 Palaeoclimatology, Palaeoecology 295, 42-54.

834 Kuhn, O., Etter, W., 1994. Der Posidonienschiefer der Nordschweiz: Lithostratigraphie, Biostratigraphie

835 und Fazies. Eclogae Geologicae Helvetiae 87, 113-138.

836 Küspert, W., 1982. Environmental change during oil shale deposition as deduced from stable isotope

837 ratios. In: Einsele, S., Seilacher, A. (Eds.), Cyclic and event stratification. Springer, New York, pp.

838 482-501.

839 März C., Poulton, S.W., Wagner, T., Schnetger, B., Brumsack, H-J., 2014. Phosphorus burial and

840 diagenesis in the central Bering Sea (Bowers Ridge, IODP Site U1341): Perspectives on the marine

841 P cycle. Chemical Geology 363, 270-282.

842 Mattioli, E., Pittet, B., Suan, G., Mailliot, S., 2008. Calcareous nannoplankton across the Early Toarcian

843 Anoxic Event: implications for paleoceanography within the western tethys. Paleoceanography 23,

844 PA3208, DOI: 10.1029/2007PA001435.

845 Mattioli, E., Pittet, B., Petitpierre, L., Mailliot, S., 2009. Dramatic decrease of pelagic carbonate

846 production by nannoplankton across the Early Toarcian anoxic event (T-OAE). Global and

847 Planetary Change 65, 134-145.

848 McArthur, J.M., Algeo, T.J., van de Schootbrugge, B., Li, Q., Howarth, R.J., 2008. Basinal restriction,

849 black shales, Re–Os dating, and the Early Toarcian (Jurassic) oceanic anoxic event.

850 Paleoceanography 23, PA4217.

851 McElwain, J.C.,Wade-Murphy, J., Hesselbo, S.P., 2005. Changes in carbon dioxide during an oceanic

852 anoxic event linked to intrusion into Gondwana coals. Nature 435, 479-482.

36

853 Montero-Serrano, J.C., Palarea-Albaladejo, J., Martín-Fernández, J.A., Martínez-Santana, M., Gutiérrez-

854 Martín, J.V., 2010. Sedimentary chemofacies characterization by means of multivariate analysis.

855 Sedimentary Geology 228, 218-228.

856 Mort, H., Adatte, T., Föllmi, K., Keller,, Steinmann, P., Matera V., Berner, Z., And Stüben, D., 2007.

857 Phosphorus and the roles of productivity and nutrient recycling during Oceanic Anoxic Event 2.

858 Geology 35-6, 483-486.

859 Murray, J. W., Jannasch, H. W., Honjo, S., Anderson, R. F., Reeavrcn, W. S., Top, Z., Friederich, G. E.,

860 Codispoti, L. A., Izdar, E., 1989. Unexpected changes in the oxic/anoxic interface in the Black Sea.

861 Nature 338, 411-413.

862 Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred from major

863 element chemistry of lutites. Nature 299, 715–717.

864 Pálfy, J., Smith, P.L., 2000. Synchrony between Early Jurassic extinction, oceanic anoxic event, and the

865 Karoo-Ferrar flood basalt volcanism. Geology 28, 747-750.

866 Parrish, J.T., Curtis, R.L., 1982. Atmospheric circulation, upwelling, and organic-rich rocks in the

867 Mesozoic and Cenozoic areas. Palaeogeography, Palaeoclimatology, Palaeoecology 40, 31-66.

868 Paytan, A. and McLaughlin, K., 2007. The oceanic phosphorus cycle. Chemical Review 107, 563–576.

869 Pawlowsky-Glahn, V., Egozcue, J., 2011. Exploring compositional data with the CoDa-dendrogram.

870 Austrian Journal of Statistics 40, 103-113.

871 Pearce, C.R., Cohen, A.S., Coe, A.L., Burton, K.W., 2008. Molybdenum isotope evidence for global

872 ocean anoxia coupled with perturbations to the carbon cycle during the Early Jurassic. Geology 36,

873 231-234.

874 Piper, D.Z., Dean, W.E., 2002. Trace-element deposition in the Cariaco Basin under sulfate reducing

875 conditions - a history of the local hydrography and global climate, 20 ka to the present. US

876 Geological Survey Professional Paper 1670, 41pp.

37

877 Perilli, N., 2000. Calibration of early–middle Toarcian nannofossil events based on high resolution

878 ammonite biostratigraphy in two expanded sections from the Iberian Range (East Spain). Marine

879 Micropaleontology 39, 293–308.

880 Pittet, B., Suan, G., Lenoir, F., Duarte, L.V., Mattioli, E., 2014. Carbon isotope evidence for sedimentary

881 discontinuities in the lower Toarcian of the Lusitanian Basin (Portugal): Sea level change at the

882 onset of the Oceanic Anoxic Event. Sedimentary Geology 303, 1–14.

883 Prauss, M., Ligouis, B., Luterbacher, H., 1991. Organic matter and palynomorphs in the

884 ‘Posidonienschiefer’ (Toarcian, Lower Jurassic) of southern Germany. Journal Geological Society

885 London Special Publications 58, 335-351.

886 Reátegui, K., Martínez, M., Esteves, I., Gutiérrez, J.V., Martínez, A., Meléndez, W., Urbani, F., 2005.

887 Geochemistry of the Mirador Formation (Late Eocene-Early Oligocene), south western Venezuela:

888 chemostratigraphic constraints on provenance and the influence of the sea level. Geochemical

889 Journal 39 (3), 213-226.

890 Redfield, A.C., 1958. The biological control of chemical factors in the environment. American Scientist

891 46, 205-222.

892 Reimann, C., Filzmoser, P., Garrett, R.G., Dutter, R., 2008. Statistical Data Analysis Explained -

893 Applied environmental statistics with R. John Wiley, Chichester, 362 pp.

894 Reisdorf, A. G., Wetzel, A., Schlatter, R., & Jordan, P., 2011. The Staffelegg Formation: a new

895 stratigraphic scheme for the Early Jurassic of northern Switzerland. Swiss Journal of Geosciences

896 104 , 97-146.

897 Reolid, M., Rodríguez-Tovar, F.J., Marok, A., Sebane, A., 2012. The Toarcian Oceanic Anoxic Event in

898 the Western Saharan Atlas, Algeria (North African Paleomargin): role of anoxia and productivity.

899 Geological Society of America Bulletin 124, 1646-1664.

38

900 Reolid, M., Marok, A., Sebane, A., 2014a. Foraminiferal assemblages and geochemistry for interpreting

901 the incidence of Early Toarcian environmental changes in North Gondwana palaeomargin (Traras

902 Mountains, Algeria). Journal of African Earth Sciences 95, 105-122.

903 Reolid, M., Mattioli, E., Nieto, L.M., Rodríguez-Tovar, F.J., 2014b. The Early Toarcian Ocanic Anoxic

904 Event in the External Subbetic (Southiberian Palaeomargin, Westernmost Tethys): geochemistry,

905 nannofossils and ichnology. Palaeogeography, Palaeoclimatology, Palaeoecology 411, 79-94.

906 Reolid, M., 2014. Stable isotopes on foraminifera and ostracods for interpreting incidence of the

907 Toarcian Oceanic Anoxic Event in Westernmost Tethys: role of water stagnation and productivity.

908 Palaeogeography, Palaeoclimatology, Palaeoecology 395, 77-91.

909 Rodríguez-Tovar, F.J., Reolid, M., 2013. Environmental conditions during the Toarcian Oceanic Anoxic

910 Event (T-OAE) in the westernmost Tethys: influence of the regional context on a global

911 phenomenon. Bulletin of Geosciences 88-4, 697–712.

912 Röhl, H.J., Schmid-Röhl, A., Oschmann, W., Frimmel, A., Schwark, L., 2001. The Posidonia Shale

913 (Lower Toarcian) of SW-Germany: an oxygen-depleted ecosystem controlled by sea level and

914 palaeoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 165, 27-52.

915 Röhl, H.-J., Schmid-Röhl, A., 2005. Lower Toarcian (Upper Liassic) black shales of the Central

916 European Epicontinental Basin: a sequence stratigraphic case study from the SW German Posidonia

917 Shale (Lower Toarcian). SEPM. Special Publication 82, 165-189.

918 Rosales, I., Quesada, S., Robles, S., 2004. Paleotemperature variations of Early Jurassic seawater

919 recorded in geochemical trends of belemnites from the Basque–Cantabrian basin, northern Spain.

920 Palaeogeography, Palaeoclimatology, Palaeoecology 203, 253–275.

921 Ruebsam, W., Münzbergerb, P., Schwarka, L., 2014. Chronology of the Early Toarcian environmental

922 crisis in the Lorraine Sub-Basin (NE Paris Basin). Earth and Planetary Science Letters 404, 273–

923 282.

39

924 Schenau, S.J. and de Lange, G.J., 2000. A novel chemical method to quantify fish debris in marine

925 sediments. Limnology and Oceanography 45, 963-971.

926 Schenau, S.J. and de Lange, G.J., 2001. Phosphorus regeneration vs. burial in sediments of the Arabian

927 Sea. Marine Chemistry 75, 201–207.

928 Schrag, D.P., DePaolo, D.J., Richter, F.M., 1995. Reconstructing past sea surface temperatures from

929 oxygen isotope measurements of bulk carbonate. Geochimica et Cosmochimica Acta 59, 2265-2278.

930 Slomp, C.P., Thomson, J., De Lange, G., 2004. Controls on phosphorus regeneration and burial during

931 formation of eastern Mediterranean sapropels. Marine Geology 203,141–159.

932 Slomp, C.P. and Van Cappellen, P., 2007. The global marine phosphorus cycle: sensitivity to oceanic

933 circulation. Biogeosciences 4, 155-171.

934 Stein, M., Föllmi, K.B., Westermann, S., Godet, A., Adatte, T., Matera, V., Fleitmann, D., Berner, Z.,

935 2011. Progressive palaeoenvironmental change during the Late BarremianeEarly Aptian as prelude

936 to Oceanic Anoxic Event 1a: evidence from the Gorgo a Cerbara section (Umbria-Marche Basin,

937 central Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 302, 396-406.

938 Suan, G., Pittet, B., Bour, I., Mattioli, E., Duarte, L.V., Mailliot, S., 2008. Duration of the Early Toarcian

939 carbon isotope excursion deduced from spectral analysis: consequence for its possible causes. Earth

940 and Planetary Science Letters 267, 666-679.

941 Suan, G., Mattioli, E., Pittet, B., Lécuyer, C., Suchéras-Marx, B., Duarte, L.V., Philippe, M., Reggiani,

942 M.L., Martineau, F., 2010. Secular environmental precursors to Early Toarcian (Jurassic) extreme

943 climate changes. Earth and Planetary Science Letters 290, 448-458.

944 Suan, G., Nikitenko, B. L., Rogov, M. A., Baudin, F., Spangenberg, J. E., Knyazev, V. G., Glinskikh, L.

945 A., Goryacheva, A. A., Adatte, T., Riding, B. J., Föllmi, K. B., Pittet, B., Mattioli, E., Lecuyer, C.,

946 2011. Polar record of Early Jurassic massive carbon injection. Earth and Planetary Science Letters

947 312, 102-113.

40

948 Svensen, H., Planke, S., Chevallier, L., Malthe-Sørenssen, A., Corfu, F., Jamtveit, B., 2007.

949 Hydrothermal venting of greenhouse gases triggering Early Jurassic global warming. Earth Planet.

950 Sci. Lett. 256, 554–566.

951 Tada, R., Sato, S., Irino, T., Matsui, H., Kennett, J.P., 2000. Millennial-scale compositional variations in

952 late Quaternary sediments at Site 1017, southern California. Proceedings of the Ocean Drilling

953 Program - Scientific Results 167, 277-296.

954 Trecalli, A., Spangenberg, J.E., Adatte, T., Föllmi, K.B., Parente, M., 2012. Carbonate platform evidence

955 of ocean acidification at the onset of the early Toarcian oceanic anoxic event. Earth and Planetary

956 Science Letters 357-358, 214-225.

957 Tribovillard, N., Recourt, P., Trentesaux, A., 2010. Bacterial calcification as a possible trigger for

958 francolite precipitation under sulfidic conditions. Comptes Rendus Geoscience 342-1, 27–35.

959 Tribovillard, N., Algeo, T.J., Lyons, T.W., Riboulleau, A., 2006. Trace metals as paleoredox and

960 paleoproductivity proxies: an update. Chemical Geology 232, 12-32.

961 Van Cappellen, P., Ingall, E.D., 1994. Benthic phosphorus regeneration, net primary production, and

962 ocean anoxia: a model of the coupled marine biogeochemical cycles of carbon and phosphorus.

963 Paleoceanograpghy 9, 677-692.

964 van de Schootbrugge, B., McArthur, J.M., Bailey, T.R., Rosenthal, Y., Wright, J.D., Miller, K.G., 2005.

965 Toarcian oceanic anoxic event: an assessment of global causes using belemnite C isotope records.

966 Paleoceanography 20, PA3008. doi:10.1029/2004PA001102.

967 van de Schootbrugge, B., Bachan, A., Suan, G., Richoz, S., Payne, J. L., 2013. Microbes, mud and

968 methane: cause and consequence of recurrent Early Jurassic anoxia following the end- mass

969 extinction. Palaeontology 56, 685–709.

970 van den Boogaart, K.G., Tolosana-Delgado, R., 2008. “Compositions”: a unified R package to analyze

971 compositional data. Computers and Geosciences 34, 320-338.

41

972 von Eynatten, H., Barcelo-Vidal, C., Pawlowsky-Glahn, V., 2003. Modelling compositional change: the

973 example of chemical weathering of granitoid rocks. Mathematical Geology 35, 231–251.

974 Vrielynck, B., Bouysse, P., 2003. The Changing Face of the Earth. Commission for the Geological Map

975 of the World, CCGM-UNESCO Publishing. 32 p., 11 pl.

976 Warning, B., Brumsack, H.J., 2000. Trace metal signatures of eastern Mediterranean sapropels.

977 Palaeogeography, Palaeoclimatology, Palaeoecology 158, 293-309.

978 Wedepohl, K.H., 1991. The composition of the upper Earth's crust and the natural cycles of selected

979 elements. Metals in natural raw materials. Natural resources. In: Merian, E. (Ed.), Metals and their

980 Compounds in the Natural Environment. VCH, Weinheim, Germany, pp. 3-17.

981 Westermann, S., Caron, M., Fiet, N., Fleitmann, D., Matera, V., Adatte, T., Föllmi, K.B., 2010. Evidence

982 for oxic conditions during oceanic anoxic event 2 in the northern Tethyan pelagic realm. Cretaceous

983 Research 31, 500-514.

984 Westermann S., Stein M., Matera V., Fiet N., Fleitmann D., Adatte T., Föllmi K.B., 2013. Rapid changes

985 in the redox conditions of the western Tethys Ocean during the OAE 1a interval (early Aptian).

986 Geochimica et Cosmochimica Acta 121, 467-486.

987 Wignall, P.B., Newton, R.J., Little, C.T.S., 2005. The timing of paleoenvironmental change and cause-

988 and-effect relationships during the Early Jurassic mass extinction in Europe. American Journal of

989 Science 305, 1014-1032

990 Wignall, P.B., McArthur, J.M., Little, C.T.S., Hallam, A., 2006. Palaeoceanography: methane release in

991 the Early Jurassic period. Nature 441, E5, doi:10.1038/nature04905.

992

993

994

995

42

996 Figure captions

997

998 Figure 1. (a) Palaeogeographical maps for the early Toarcian (redrawn from Vrielynck and Bouysse,

999 2003) showing the main large igneous provinces (Karoo-Ferrar and North Sea), the Viking Corridor, the

1000 NW Tethyan region, and the Rietheim Posidonia Shale succession studied herein. VC indicates the

1001 location of the Viking Corridor. Locations of the Lower Toarcian successions discussed in this study are

1002 noted as black circles: Dotternhausen, SW Germany (D; Röhl et al., 2001); Schesslitz, SW Germany (S;

1003 Küspert, 1982; Prauss et al., 1991); Yorkshire, England (Y; Cohen et al., 2004; Kemp et al., 2005); and

1004 Kelimyar River, N Siberia (KR; Suan et al., 2011). (b) Map showing the location of the Rietheim

1005 Posidonia Shale succession and geological setting in northern Switzerland.

1006

1007 Figure 2. Biostratigraphy and lithostratigraphy for the upper Pliensbachian to lower Toarcian interval at

1008 the Rietheim Posidonia Shale succession. The ammonite biostratigraphy is from Kuhn and Etter (1994).

1009 Photographs of the drill core and its stratigraphic position are also shown.

1010

13 13 1011 Figure 3. Stratigraphic distribution of the: isotopic compositions of (a-b) carbon (G Corg, G CCaCO3), (c)

18 15 1012 oxygen (G OCaCO3), (d) nitrogen (G Ntotal), (e) total organic carbon (TOC), (f) total organic carbon from

1013 the carbonate-free fraction (TOCcff), (g) hydrogen index (HI), and (h) Rock-Eval Tmax from the Rietheim

1014 Posidonia Shale succession. The shaded interval indicates the negative carbon isotope excursion (CIE)

1015 related to the T-OAE.

1016

1017 Figure 4. (a) Hydrogen Index (HI) vs. Oxygen Index (OI) crossplot illustrating the organic matter origin

1018 of the Rietheim Posidonia Shale samples. Note that the organic matter was primarily of marine origin

13 1019 during the lower Toarcian (exaratum to commune subzones). (b) G Corg-HI crossplots. The least square

43

13 1020 linear regression is based on samples from outside of the T-OAE. Note that G Corg values slightly

13 15 1021 decrease with increasing HI values. (c) G Corg-G Ntotal crossplots showing the depositional settings of

13 15 1022 the Rietheim Posidonia Shale sediments. The average G Corg-G Ntotal values of the five depositional

1023 settings are from Algeo et al. (2014).

1024

13 13 1025 Figure 5. Comparison between the bulk G Corg records (a) with G Corg corrected for changing OM

13 1026 sourcing (b) and G CCaCO3 (c). Note that the bulk organic C-isotope data were corrected using the HI and

1027 raw C-isotope data from the spinatum and bifrons zones. The blue, red and black lines represent the 9-

1028 point moving average of the isotopic data.

1029

1030 Figure 6. X-ray diffractograms of typical bulk rock samples from the early Toarcian negative carbon

1031 isotope excursion (CIE) in the Rietheim Posidonia Shale succession.

1032

1033 Figure 7. Stratigraphic distribution of the bulk mineralogical semiquantification from the Rietheim

1034 Posidonia Shale sediments. The shaded interval indicates the negative carbon isotope excursion (CIE)

1035 related to the T-OAE.

1036

1037 Figure 8. (a-d) Stratigraphic profiles of the major geochemical associations obtained by the

1038 compositional Q-mode cluster analysis. The shaded interval indicates the negative carbon isotope

1039 excursion (CIE) related to the T-OAE. The red lines represent the 9-point moving average of the data. (e)

1040 Dendrogram of association of variables obtained by applying Ward clustering algorithm based on the

1041 variation matrix of elemental geochemical dataset.

1042

44

1043 Figure 9. Overview of upper Pliensbachian to lower Toarcian carbon-isotope records from representative

1044 Tethyan and Boreal successions. Carbon-isotope data from (a) Rietheim Posidonia Shale, N Switzerland

1045 (this study; the black line represent the 9-point moving average of the isotopic data), (b) Dotternhausen

1046 and Zimmern, SW Germany (modifed from Küspert, 1982; Prauss et al., 1991 and Röhl et al., 2001), (c)

1047 Yorkshire, England (Cohen et al., 2004; Kemp et al., 2005) and (d) Kelimyar River, N Siberia (Suan et

1048 al., 2011). See Fig. 1a for the geographical location of these stratigraphic sections. The shaded interval

1049 indicates the negative carbon isotope excursion (CIE) related to the T-OAE. Note that carbon-isotope

1050 profiles in geographically distant sections exhibit a similar trend in the lower Toarcian interval.

1051

1052 Figure 10. Stratigraphic distribution of the main detrital proxies obtained from the Rietheim Posidonia

1053 Shale succession. (a) detritus index (summing of quartz, phyllosilicates, K-feldspars, Na-plagioclases);

1054 (b) Ln(Al2O3/Na2O); (c) XRF siliciclastic balance (Nb, Rb, La, K, Ti, Al, Si, Cr, Na, Y, Nd). Balances

1055 are represents according to equation 1 (see Section 3.2.5). The red lines represent the 9-point moving

1056 average of the data. The shaded interval indicates the negative carbon isotope excursion (CIE) related to

1057 the T-OAE.

1058

1059 Figure 11. Stratigraphic distribution of the main redox proxies obtained from the Rietheim Posidonia

1060 Shale succession. (a) relative content of pyrite (%); (b-c) sulphide (Zn, Cu, Fe, Co) and OM (P, As,

1061 TOC, Ni, V) balances. Balances are represents according to equation 1 (see Section 3.2.5); (d) V/(V+Ni)

1062 ratios. Redox zones are from Hatch and Leventhal (1992); (e-f) elements enrichment factors (EF) for the

1063 OM and sulphide associations; (g) relative content of fluorapatite (%); (h) TOC and total phosphorus

1064 (Ptotal) relationship (TOC/Ptotal). The red lines represent the 9-point moving average of the data. The

1065 shaded interval indicates the negative carbon isotope excursion (CIE) related to the T-OAE.

1066

45

1067 Figure 12. Al-normalized redox-sensitive trace-elements (V, Ni, As and Cu) plotted versus total organic

1068 carbon from the Rietheim Posidonia Shale sediments. This plot is used to discriminate between different

1069 redox facies (dysoxic, anoxic and euxinic; e.g., Algeo and Maynard, 2004; Trivovillard et al., 2006).

1070 Redox-sensitive trace-metal data from the Cariaco Basin (Piper and Dean, 2002), the Southern California

1071 margin (Irino and Pedersen, 2000; Tada et al., 2000), Posidonia Shale from SW Germany (Brumsack,

1072 1991), upper Pennsylvanian black shales (Algeo and Maynard, 2004), and Mediterranean and Black Sea

1073 sapropels (Warning and Brumsack, 2000) are also plotted.

46

*Highlights (for review)

Highlights

-The T-OAE negative CIE reflects a 13C-depleted carbon injection to ocean-atmosphere

-Detrital proxies suggest an increase in continental weathering rates during the T-OAE

-Redox proxies suggest moderate anoxic bottom-water conditions during the T-OAE

-Dysoxic-to-anoxic conditions prevailed until at least to the bifrons Zone

-P-enrichments are observed during the T-OAE Figure 1

Figure 1 (Double column - full page width)

Legend (a)Rietheim (b) Present day coastline Continental/oceanic canton Basel Aargau crust transition N Zürich Y Fault 0 50 km Oceanic ridge Bern S Thrust D Oceanic subduction zone Switzerland Igneous provinces Emergent land Shallow marine Geneva Deep marine N 1000 km Rietheim 7°E 8°E 9°E Posidonia Shale

Rietheim Early Toarcian (~180 Ma) Rhine BOREAL KR ASIA Basel 60°N OCEAN VC EUROPE Zürich LAURASIA North SE ASIA Sea Aar 30°N NORTH AMERICA Mega Lhassa

47°N Aar Bern N 0° PANTHALASSA OCEAN TETHYS OCEAN AFRIQUE GONDWANA Legend SOUTH INDIA Tabular Jura Molasse s.l. in the Jura Mountains 30°S AMERICA Folded Jura Subalpine molasse and folded molasse Karoo AUSTRALIA Tertiary rift basins Variscan basement and Tertiary Ferrar volcanites (Hegau) 60°S Upper freshwater molasse Subalpine chains, Helvetic nappes Upper marine molasse ANTARCTICA and Prealps klippen Faults: Thrusts Major thrusts Figure 2

Figure 2 (1.5 column)

Rietheim Posidonia Shale

tage Height S A-ZoneA-subzone (cm) 480 Monotisbank Legend

440 Black to grey bituminous marly clay ()Tonmergel-Fazies 400 Black to brownish laminated bituminous 360

bifrons marly clay (Kartonschiefer-Fazies )

commune 320 Light-grey to dark-grey marly clay 280 Light-grey bioturbated marl 240 Limestone beds 200 Bioturbated layers

falciferum 160

120 S1 S2 S3 S4 S5 S6 S7 S8 S9 S10 80 0 Top 40 5

Early Toarcian 0

elegans

falciferum

Spinatumbank

10 Oberer Stein -40 Oberer Stein 15

-80 Homogene Kalkbänke

Unterer Stein -120 Homogene 20 Kalkbänke -160 25 Base

exaratum -200 Unterer Stein 30 -240 ? semi-? ten. celatum 35 -280 Spinatumbank

-320 40

Pliens.

spinatum

spinatum

45 cm Homogene Kalkbänke Figure 3

Figure 3 (Double column - full page width)

Rietheim Posidonia Shale 15 d13 d N(‰ N air) C(CaCO3 ‰ VPDB) total 2 TOCcff (wt.%) Tmax (°C) 01234 StageA-ZoneA-subzoneHeight(cm) -5 -4 -3 -2 -1 0 1 2 01015520400 420 440 460 480 440 (a) (b) (c) (d) (g) (h) Monotisbank (e) (f) 400 360

bifrons

commune 320 280 240 200

falciferum 160 120

mature

80 Immature 40

Early Toarcian

elegans

falciferum 0

-40 Stein Oberer -80 -120

Kalkbänke -160 Homogene

exaratum

-200 CIE

Stein

? semi-? -240 Unterer ten celatum -280

spin. spin. -320 Pliens. -34 -32 -30 -28 -26 -24 -10-9-8-7-6-5-4-3-2 024681012 0 200 400 600 800 d13 d18 TOC (wt.%) Hydrogen Index Spinatumbank C(‰ VPDB) O(‰ VPDB) org CaCO3 (mgHC/gTOC) Figure 4 Figure 4 (single column)

(a) 900 Type I Legend marls - marly clays 800 commune falciferum elegans 700 exaratum semicelatum Type II spinatum 600 commune elegans exaratum 500 semicelatum spinatum 400 Mixtures

(mgHC/gTOC) Type II+III

Hydrogen Index 300

200

100 Type III Type IV 0 0 100 200 300 400 500 Oxygen index

(mgCO2 /gTOC) (b) -24 d13C = -0.0016*HI - 27.741 org T-OAE -26 R²=0.42 background -28

‰ VPDB) -30

org

C( -32

13

d -34 0 100 200 300 400 500 600 700 800 Hydrogen Index (mgHC/gTOC) (c) 8

6 Shelf Oceanic

2 4 Upwelling 2

‰ N air)

total 0 Epeiric

N( Oceanic-Mediterranean

15

d -2

-4 -38 -36 -34 -32 -30 -28 -26 -24 -22 -20 -18 d13 C(org ‰ VPDB) Figure 5 Figure 5 (single column)

Rietheim Posidonia Shale

d13 C(org HI-corr ‰ VPDB) tageA-Zone Height(cm) -4 -2 0 2 S A-subzone480 440 (a) (b) (c)

Monotisbank 400 360

bifrons

commune 320 280 240 200

falciferum 160 120 80 40

Early Toarcian 0

elegans

falciferum -40 Stein Oberer -80 -120

Kalkbänke -160 Homogene

exaratum -200 CIE

Stein

Unterer ? semi-? -240 ten celatum -280

spin. spin. -320

Pliens. ~5‰ ~3‰ ~3‰

Spinatumbank -32 -30 -28 -26 -413 -2 0 2 d13 d C(‰ VPDB) C(org ‰ VPDB) CaCO3 Figure 6

Figure 6

Mu K-Feld Bi Ka An MuQz CFA

Illi Chl MuQz CaQz Plag Ca Do CFA CaPy Ca Py Ca CaQz Py Qz 3000

2500 Top -123 cm 2000 -150 cm 1500 -220 cm

CIE 1000

Counts second per -236 cm 500 -246 cm 0 Base 5 101520253035404550556065 2(CuK)θα Figure 7

Figure 7

Rietheim Posidonia Shale Phyllosilicates (%) K-Feldspar (%) Dolomite (%) Fluorapatite (%) Unquantified (%) StageA-ZoneA-subzoneHeight(cm) 02550 02468 02468 0.00.61.21.8 012345 480 440

Monotisbank 400 (a) (b) (c) (d)(e) (f) (g) (h) (i) (j) 360

bifrons

commune 320 280 240 200

falciferum 160 120 80 40

Early Toarcian

elegans 0

falciferum -40 Stein Oberer -80 -120

Kalkbänke -160 Homogene

exaratum -200 CIE

Stein

Unterer ? semi-? -240 ten celatum -280

spin. spin. -320 Pliens. strata 0 6 12 18 01234 0 255075100 0 5 10 15 20 02468

missing Spinatumbank Quartz (%) Na-Plagioclase (%) Calcite (%) Ankerite (%) Pyrite (%) Figure 8

Figure 8

Rietheim Posidonia Shale (a) Siliciclastics group (b) Carbonate group Si (%) K(%) Cr (μg/g) Nb (μg/g) Y(μg/g) Mg (%) Ca (%) Height(cm) 0 5 1015 20 25 0123 40 80 120 160 0 5 1015 20 25 10 20 30 40 50 024 0102030 StageA-ZoneA-subzone480 440

Monotisbank 400 360

bifrons commune 320 280 240 200

falciferum 160 120 80 40 Early Toarcian 0

falciferum

elegans -40 Stein -80 Oberer -120

Kalkbänke

-160 Homogene

CIE exaratum -200

Stein ? ? -240 Unterer ten semi. -280

spin. spin. -320 Pliens. strata 05100.0 0.2 0.4 0.6 0.1 0.2 0.3 0408012010 30 50 70 90 20 40 60 80 0 500 1000 1500 missing Spinatumbank Al (%) Ti (%) Na (%) Rb (μg/g) La (μg/g) Nd (μg/g) Mn (μg/g)

Rietheim (c) (d) (e) Posidonia Shale Sulfide group Organic group Variance ln(xij /x ) Elemental Cu (μg/g) Co (μg/g) P(μg/g) V(μg/g) 0.0 0.2 0.4 0.6 0.8 1.0 associations Height(cm) 050100 10 30 50 0 1000 2000 0200400 StageA-ZoneA-subzone480 P 440 As

Monotisbank 400 TOC 360

bifrons Zn* commune 320 Cu* 280 Fe* 240 Co* 200 rganic + oNi sulfide* falciferum 160 V 120 Mn 80 Ca 40 Early Toarcian 0 Mg

falciferum

elegans -40 carbonate Nd Stein Oberer Y -80 Na -120 Cr

Kalkbänke

-160 Homogene Si

exaratum -200 CIE

Stein

Unterer Al ? ? -240 ten semi. -280 Ti

siliciclastics

spin. spin. -320 K Pliens. strata 1357 20 80 140 200 036912 05010010 70 130 190 La missing Spinatumbank Fe (%) Zn (μg/g) TOC (%) As (μg/g) Ni (μg/g) Rb Nb Figure 9

Figure 9

Rietheim Dotternhausen & Zimmern (c)Yorkshire (d) Kelimyar River (a)N Switzerland (b) SW Germany UK N Siberia

tage Height(cm) Northwestern Tethys Boreal S A-Zone480 440 400 360 Height bifrons StageA-Zone (m)

320 Height(m) Height(m) StageA-Zone StageA-Zone 30 280 U. T. 12 14 240 25 200 10 160 12 braun. 120 20 8 80 bifrons 10 40 15 Lower Toarcian 6 falciferum 0 8

falciferum -40 commune 10 -80 ? 4

6 Lower Toarcian -120 5 -160 Lower Toarcian 2 -200 Lower Toarcian 4

falciferum

falciferum

? -240 0 CIE ten a. 0 -280 strata 2 missing

-320 ten. -5 -34 -32 -30 -28 -26 -24 -2

Pliens.

spinatum 0 d13

C(‰ VPDB) viligaensis org -10 -4

Pliensbachian -2 tenuicostatum -34 -32 -30 -28 -26 -24 -34 -32 -30 -28 -26 -24 13

spinatum -15 d C(‰ VPDB) d13 org

Pliensbachian C(org ‰ VPDB) -34 -32 -30 -28 -26 -24 13

Pliens. d C(org ‰ VPDB)

spinatum Figure 10 Figure 10 (single column)

Detrital proxies Rietheim Posidonia Shale Clastic influx Clastic influx

Detritus (%) b{Siliciclastic/Carbonate} StageA-ZoneA-subzoneHeight(cm) 25 50 75 -2.2 -1.6 -1.0 480 440 (a) (b) (c) 400 360 Monotisbank

bifrons

commune 320 280 240 200

falciferum 160 120 80 40

Early Toarcian

elegans 0

falciferum -40 Stein Oberer -80 -120

-160 Kalkbänke

Homogene

exaratum -200 CIE

? ? -240 Stein semi- Unterer ten celatum -280

spin. spin. -320

Pliens. 2.8 3.0 3.2 3.4 strata missing Ln(Al O /Na O)

Spinatumbank 23 2 chemical weathering Figure 11

Figure 11 (Double column - full page width)

Redox proxies

Rietheim anoxia/euxinia Posidonia Shale OM preservation b Pyrite (%) {Organic/Carbonate} EFaverage shale Fluorapatite (%) StageA-Zone Height(cm) 02468-3.0 -2.4 -1.8 -1.2 -0.6 0.1 1 10 0.0 0.6 1.2 1.8 A-subzone480 440

Monotisbank (a) (b) (c) (d) (e)(f) (g) (h) 400 360

bifrons

commune As Co 320 EF EF VEF CuEF 280 NiEF ZnEF 240 200

falciferum 160

EF = 1 EF = 1 EF = 3 EF = 3 120 Oxic

Anoxic Euxinic

Dysoxic 80

40 Redfield ratio Early Toarcian 0

falciferum

elegans -40 Stein Oberer -80 -120 -160

Kalkbänke

Homogene

CIE exaratum -200

? ? -240 Stein semi- Unterer ten celatum -280

spin.

spin. -320 Pliens. strata -2.5 -2.1 -1.7 -1.3 -0.9 0.4 0.6 0.8 1.0 0.1 1 10 0 100 200 300 400 missing b{Sulfide/Carbonate} V/(V+Ni) EFaverage shale TOC/P

Spinatumbank total

euxinia decreased oxygenation primary productivity Figure 12 Figure 12 (single column)

Anoxic Euxinic euxinic facies threshold threshold (weak TOC-TE correlation) (a)

TE/Al

anoxic facies (strong TOC-TE correlation) dysoxic facies background concentration 2.5 10 40 60 TOC (wt. %) 50 (b)

-4 40 30 20

V/Al (x10 ) 10 0 012345678910111213 40 TOC (wt. %)

30 (c)

-4

20

Ni/Al (x1010 )

0 012345678910111213 TOC (wt. %) 18 15 (d)

-4 12 9 6 As/Al (x103 ) 0 012345678910111213 TOC (wt. %) 25

20 (e)

-4 15

10

Cu/Al (x105 )

0 012345678910111213 TOC (wt. %)

Legend Redox facies Posidonia Shale Posidonia Shale Sapropels Rietheim SW Germany Anoxic - Euxinic semicelatum Base Anoxic Cariaco Basin exaratum Upper Pennsylvanian Anoxic elegans black shales falciferum Southern California Anoxic commune Top margin Dysoxic Dysoxic