Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events

Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events

Accepted Manuscript Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events J.P. Navarro-Ramirez, S. B...

18MB Sizes 1 Downloads 35 Views

Accepted Manuscript Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events J.P. Navarro-Ramirez, S. Bodin, L. Consorti, A. Immenhauser PII:

S0195-6671(16)30203-8

DOI:

10.1016/j.cretres.2017.03.009

Reference:

YCRES 3555

To appear in:

Cretaceous Research

Received Date: 9 September 2016 Accepted Date: 9 March 2017

Please cite this article as: Navarro-Ramirez, J.P., Bodin, S., Consorti, L., Immenhauser, A., Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events, Cretaceous Research (2017), doi: 10.1016/j.cretres.2017.03.009. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT 1 2

Response of western South American epeiric-neritic ecosystem to middle Cretaceous Oceanic Anoxic Events

3 4

Navarro-Ramirez J.P.1, *, Bodin S.2, Consorti L.3, Immenhauser A.1

5

1

6

Germany

7

2

Aarhus University, Department of Geoscience, Høegh-Guldbergs Gade 2, 8000 Aarhus C, Denmark

8

3

Departament de Geologia (Paleontologia), Universitat Autònoma de Barcelona, 08193 Bellaterra,

9

Spain

M AN U

SC

RI PT

Ruhr-Universität Bochum, Institut für Geologie, Mineralogie und Geophysik, D-44870 Bochum,

10

* Corresponding author. Tel.: +49-234-32-2325.

12

E-mail address: [email protected]

13

Revised version for the Journal Cretaceous Research

16 17 18

EP

15

AC C

14

TE D

11

Keywords: Peru, mid-Cretaceous, Pacific, OAE1d, OAE2, chemostratigraphy, carbonates

1

ACCEPTED MANUSCRIPT 19

Abstract Little is known about the impact of the mid-Cretaceous Oceanic Anoxic Events (OAEs) on the

21

neritic carbonate systems in South America. In order to fill this knowledge gap, the present paper

22

reports on the record of environmental changes in the Albian–Turonian neritic carbonates from the

23

western South American domain in Peru. Owing to the very expanded and well-exposed sections in

24

the Oyon region of central Peru, the OAE 1d and 2 intervals were sampled at high temporal resolution

25

for both bulk micrite and bulk organic matter carbon isotopes, allowing us to compare the fingerprint

26

of these two events between the northern and central Peruvian regions. This suggests the installation of

27

two marked depositional modes: 1) the Albian–Turonian formation of a regional facies belt constituted

28

by oyster-rich mixed siliciclastic-carbonate deposition along the western South America platform; 2) a

29

restricted oligotrophic environment, characterized by the mass occurrence of Perouvianella peruviana

30

and associated miliolids in central Peru during the late Cenomanian–Turonian. These observation

31

advocate for the following scenario: Global warming during the late Albian–early Turonian resulted in

32

humid climate on the western platform. This in turn caused enhanced chemical weathering rates on the

33

Brazilian Shield, resulting in high runoff of nutrients onto the western platform. Nutrient runoff

34

promoted the diversification of benthic oyster communities. Due to the uplift of the Marañon Massif

35

and the installation of the Huarmey Trough, central Peru was isolated from the Pacific and from

36

eastern deltaic influx of the Brazilian continental basement, allowing the local development of

37

oligotrophic conditions during OAE 2. Furthermore, an increased influx of argillaceous sediment and

38

reduced carbonate production is recorded in northern Peru at the onset of OAE 2, marked by a

39

prominent negative shift in δ13C. This negative carbon-isotope excursion has also been identified in

40

other sections in the Pacific domain and can be linked to an increase in isotopically light pCO2 induced

41

by the formation of the Caribbean large igneous province.

AC C

EP

TE D

M AN U

SC

RI PT

20

42

2

ACCEPTED MANUSCRIPT 43

1. Introduction The mid-Cretaceous (i.e., the Albian–Turonian interval, ca. 110–90 Ma; Ogg and Hinnov, 2012) is

45

considered as a time governed by intensification of greenhouse conditions due to elevated emission of

46

pCO2 to the atmosphere following massive pulses of ocean crust production and the formation of large

47

igneous provinces (LIPs; Arthur et al., 1985; Barron and Washington 1985; Larson, 1991; Leckie et

48

al., 2002; Snow et al., 2005; Bodin et al., 2015). The intensified greenhouse conditions enhanced the

49

hydrological cycle, which in turn accelerated continental weathering rates (Berner et al., 1983;

50

Weissert, 1990; Weissert et al. 1998; Hay, 1998; Menegatti et al., 1998; Wortmann et al., 2004; van

51

Helmond et al., 2013; Bodin et al., 2015) and led to increased fluxes of continental nutrients to the

52

oceans. This chain of events stimulated biological productivity, leading to widespread organic-carbon

53

deposition and enhanced marine anoxia and euxinia (Manabe and Bryan, 1985; Föllmi et al., 1994,

54

2006; Turgeon and Brumsack, 2006; Owens et al., 2013; Pogge von Strandmann et al., 2013; Lechler

55

et al., 2015). The processes described above probably occurred on some level throughout the

56

Cretaceous, but became particularly enhanced during several geologically short intervals (<1 Myr),

57

known as oceanic anoxic events (OAEs). These latter are best defined based on carbon isotope (δ13C)

58

excursions that reflect a fundamental perturbation in the global carbon cycle (Erbacher et al., 1996;

59

Leckie et al., 2002). The best-studied OAEs in Europe, North America, and from Atlantic ODP sites

60

include the early Aptian OAE1a, the Albian OAE1b, OAE1c and OAE1d, and the Cenomanian–

61

Turonian OAE2. Additional events, such as the Coniacian OAE3 and the mid-Cenomanian event I

62

(MCEI) represent related phenomena that lack widespread organic carbon-rich facies and/or a

63

significant δ13C excursion (e.g., Jarvis et al., 2006; Jenkyns, 2010).

AC C

EP

TE D

M AN U

SC

RI PT

44

64

OAE1a and OAE2 are conventionally interpreted as the result of massive outgassing of the

65

Ontong Java, Manihiki and Hikurangi LIP (Tarduno et al., 1991; Larson and Erba, 1999; Méhay et al.,

66

2009; Tejada et al., 2009) and by the Caribbean LIP (Snow et al., 2005; Du Vivier et al., 2014). At the

67

onset of OAE1a and OAE2, a negative shift in lithium and calcium isotope ratios is observed, most

68

likely associated to weathering of continental crust (Blättler et al., 2011; Pogge von Strandmann et al., 3

ACCEPTED MANUSCRIPT 2013; Lechler et al., 2015). Calcium isotope patterns recorded in the Portland (USA), Pont d’Issole

70

(southeastern France), and Eastbourne (England) sections, suggested that the Caribbean LIP triggered

71

a change in the predominant type of biomineralization (i.e., ocean acidification) at the onset of OAE2

72

(Du Vivier et al., 2015). As recorded in the geological record, transient periods of high atmospheric

73

CO2 led to several mass extinctions, most likely because excess CO2 reduced the ability of carbonate-

74

secreting organisms to secrete their carbonate shells (e.g., Glikson, 2010). In a more qualitative way,

75

sedimentological studies suggest enhanced weathering at the onset of OAE 1a and OAE2, as reflected

76

by the relative abundance of kaolinite (Stein et al., 2012; Gertsch et al., 2010).

SC

RI PT

69

Evidences of a sudden influx of siliciclastic material coeval to OAEs have been reported in the

78

North Atlantic and the Tethys domains (Weissert, 1990; Wortmann et al., 2004) and are indicative of a

79

change to more humid and warmer conditions (Weissert, 1990; Weissert et al., 1998; Wortmann et al.,

80

2004). This concept is exemplified in the evolution of the northern Tethyan platform during the Early

81

Cretaceous, where carbon isotope (δ13C) excursions occurred at times of elevated nutrient levels due to

82

intensified continental weathering rates (Erba, 1994; Weissert et al., 1998; Immenhauser et al., 2005;

83

Bodin et al., 2006; Föllmi et al., 2006; Huck et al., 2010, 2011, 2012, 2013, 2014). This caused a

84

change from oligo- to meso- or eutrophic conditions, ecological reorganization, phases of carbonate

85

platform development and drowning episodes (Weissert et al., 1998; Immenhauser et al., 2005; Bodin

86

et al., 2006; Föllmi et al., 2006; Huck et al., 2011). However, these hypotheses have yet to be applied

87

to sections in South America, which record the response of shallow-marine carbonate depositional

88

environments of a southern hemisphere continent adjacent to the Pacific Ocean.

AC C

EP

TE D

M AN U

77

89

The western South America platform (Fig. 1) was characterised by a large oceanic basin to the

90

west and a very large uplifted continental basement area to the east. The enormous dimensions of this

91

region make this a challenging natural laboratory to understand southern hemisphere Earth System

92

behaviour under extreme climates. The aim of this paper is to document and discuss the Albian to

93

Turonian sedimentological and palaeoecological evolution of the epeiric-neritic ecosystem of the

94

western platform in central Peru (Oyon region; Jaillard, 1986; Fig. 2). The assessment of potential 4

ACCEPTED MANUSCRIPT spatial and bathymetric variability in epeiric-neritic DIC patterns is placed in context with a previously

96

published northern Peru reference composite section (Cajamarca region, Fig. 3, Navarro-Ramirez et

97

al., 2015, 2016). Specific attention is paid to environmental stressors associated to OAE1d and OAE2

98

and the manner in which they affect the main carbonate producers such as oysters, gastropods,

99

echinoids, and endemic large benthic foraminifera (Perouvianella peruviana, Steimann) that have so far only been reported from these sections (Jaillard and Arnaud-Vanneau, 1993).

101

102

2. Regional tectonic and stratigraphic setting

SC

100

RI PT

95

The South America margin was largely influenced by late stages of Gondwana breakup that

104

culminated in the separation of Africa and South America and the opening of the Equatorial Atlantic

105

Ocean during the Aptian–Albian transition (Moulin et al., 2010; Fig. 1). This allowed for the

106

connection of North and South Atlantic Ocean water masses towards the Turonian (Eagles, 2006).

107

This tectonic event caused the activation of the western South America volcanic arc in the Early

108

Cretaceous (Huarmey-Trough, Atherton and Webb, 1989; Soler and Bonhomme, 1990; Jaillard et al.,

109

1999, 2000; Winter et al., 2010) and subduction took place along the western portion of South

110

America where sedimentation was controlled by NNW–SSE-trending structures (e.g., Benavides-

111

Caceres, 1956; Jaillard, 1986, 1987; Jaillard et al., 1990; Callot et al., 2008; Robert et al., 2009).

EP

TE D

M AN U

103

Mid-Cretaceous sedimentation in what today is Peru began with the deposition of siliciclastic

113

rocks now forming the Inca Formation, which is built by iron-rich, sandy beds, assigned to an early

114

Albian age based on the ammonite Neodeshayesites nicholsoni (Robert and Bulot, 2004; Robert et al.,

115

2009). This siliciclastic unit in turn is overlain by Albian carbonates represented by the Chulec, and

116

Pariatambo formations, recording a second-order Albian transgression (Benavides-Caceres, 1956;

117

Jaillard, 1986; Robert, 2002; Robert and Bulot, 2004, 2005).

AC C

112

118

The Chulec Formation unconformably overlies the Inca Formation with a discontinuity surface

119

separating the two units (Jaillard, 1987). The Chulec Formation is characterized by marl-limestone 5

ACCEPTED MANUSCRIPT alternations with a very abundant and diversified outer shelf fauna (Jaillard, 1987). Numerous

121

ammonites have been reported and were assigned to the Knemiceras raimondii Zone (Robert and

122

Bulot, 2004; Robert et al., 2009), indicating an early Albian age. The Pariatambo Formation is

123

characterized by fossiliferous, black, bituminous, fetid marly limestones and includes finely laminated

124

and siliceous intervals. Ammonites and planktonic foraminifera are common. The presence of

125

Oxytropidoceras carbonacrium and Prolyelliceras ulrichi Zones indicate an early middle Albian age

126

(Robert et al., 2009). The Pariatambo Formation grades upsection into nodular bioclastic grey and

127

marly limestones of the Yumagual Formation in the northern Peru and by the Jumasha Formation in

128

central Peru (upper middle Albian– lower Turonian; Fig. 3).

SC

RI PT

120

At Uchucchacua (Oyon region; central Peru; Jailard, 1986) the Jumasha Formation is about 1320

130

m thick and made of very massive, thickly-bedded, light yellowish brown to brownish-grey limestones

131

(Jaillard, 1986; Fig. 3). The base of the Jumasha Formation was attributed to the upper middle Albian

132

due to ammonite associations including Lyelliceras ulrichi Knechtel and Oxytropidoceras douglasi

133

(Benavides-Caceres, 1956; Wilson, 1963; von Hillebrandt, 1970; Jaillad, 1986; Fig. 3). At section

134

metre 115 (Fig. 3) i.e., the base of the Jumasha Formation near Uchucchacua, Romani (1982) reported

135

Exogyra costagyra cf. olisoponensis and Merlingina cretacea of Cenomanian age. Further south, in

136

the upper part of the Jumasha Formation microfauna data and particularly Favusella washitensis, H.

137

delrioensis, Globigerinelloides bentonensis and Heterohelix seewashitensis, indicate a middle to late

138

Cenomanian age (von Hillebrandt, 1970; Jaillard, 1986; Fig. 3). Rotorbinella mesogeensis (Tronchetti)

139

is equally present and supports the Cenomanian age of this unit. The presence of Coilopoceras sp. at

140

the top of the Jumasha Formation suggests a Turonian age (Romani, 1982; Jaillard, 1987; Fig. 3).

AC C

EP

TE D

M AN U

129

141

It must be noted that, due to scares ammonites and planktonic foraminifera findings in the

142

Peruvian sections, the biostratigraphic scheme established by the aforementioned authors remains

143

incomplete in places. There are uncertainties about the exact position of the stage and substage

144

boundaries, although the general scheme has been confirmed by carbon isotope chemostratigraphic

145

correlation as well as Sr-isotope chronostratigraphy (Navarro-Ramirez et al., 2015, 2016). 6

ACCEPTED MANUSCRIPT 146 147

3. Methods and materials

148

3.1 Field work and thin-section microscopy Two well-exposed sections (Uchucchacua and Lauricocha localities; Fig. 2) were chosen for this

150

study to provide maximum coverage of the Albian–Cenomanian and Cenomanian–Turonian

151

transitions (Fig. 3). In total, ca. 620 m of section have been logged and studied for their

152

sedimentological, stratigraphic, and chemostratigraphic archive (Figs. 3 and 4). Outcrop observations,

153

rock samples, and thin-sections provide the fundament for petrographic and facies interpretations

154

presented here and allow us to place the chemostratigraphic results in a facies context. In analogy with

155

the nomenclature and approach presented in Navarro-Ramirez et al. (2015, 2016), non-skeletal

156

components and skeletal components were analysed semi-quantitatively with numbers indicating their

157

relative abundance: 0 = absent, 1 = present, 2 = frequent, 3 = abundant, 4 = dominant; Table 1). Facies

158

pattern and discontinuities acted as pining points for sequence stratigraphic interpretations (see also

159

Christ et al., 2015).

TE D

M AN U

SC

RI PT

149

160

3.2. Carbon isotope stratigraphy

162

3.2.1.

EP

161

Bulk micrite data (δ13CCarb)

Carbon-isotope analysis of 181 micrite samples was performed using a Thermo Finnigan MAT

164

delta-S mass spectrometer at the isotope laboratory of the Ruhr-University Bochum (see Appendix A),

165

Germany. In order to reveal the variability of carbon-isotope composition within a single rock sample,

166

several sub-samples were drilled from 20% of all the hand specimens collected. Carbonate bulk rock

167

specimens were sampled by use of tungsten drill bits. While drilling the carbonate powder, areas rich

168

in spary cement, large bioclasts and diagenetic calcite vein material were avoided. Refer to

169

Immenhauser et al. (2005) for more details of the analytical procedure. Repeated analyses of certified

170

carbonate standards (NBS 19, IAEA CO-1 and CO-8) and internal standards show an external 7

AC C

163

ACCEPTED MANUSCRIPT 171

reproducibility of ≤0.06‰ for δ13Ccarb. The δ13Ccarb values are expressed on a per mil (‰) basis relative

172

to the Vienna-Pee Dee Belemnite standard (V-PDB).

173 3.2.2.

Bulk organic matter data (δ13Corg)

RI PT

174

Carbon-isotope analyses were performed on 201 bulk organic matter samples (see Appendix B).

176

Two grams of sample powder were placed in 50mL centrifuge tubes and 6N HCl was added, this

177

procedure was repeated until no carbonate reaction was visible any more. The samples are then

178

centrifuged and the supernatant removed, the residues were rinsed with distilled water via centrifuge

179

and dried at 40 °C. The δ13Corg measurements were performed with an elemental analyser (CE 1110)

180

connected online to ThermoFinnigan Delta V Plus masspectrometer. All carbon isotope values are

181

reported in the conventional δ‐notation in permil relative to V-PDB (Vienna‐PDB). Accuracy and

182

reproducibility of the analyses was checked by replicate analyses of international or laboratory

183

standards. Organic carbon analyses were conducted at the stable isotope laboratory of the Friedrich-

184

Alexander University of Erlangen-Nuremberg, Germany.

M AN U

TE D

186

4. Data description and interpretation

EP

185

SC

175

The Uchucchacua section is here described as the representative case example for the Albian to

188

Turonian sedimentary record in the study area (Jaillard, 1986; Jaillard and Arnaud-Vanneau, 1993;

189

Figs. 3 and 4). This choice is motivated by the fact that the outcrop conditions are excellent and

190

stratigraphically near-complete (Figs. 5A–C). In addition, the Lauricocha section, which is of

191

comparably high outcrop quality, serves as a complementary section allowing to separate local from

192

regional facies changes during the late Cenomanian to early Turonian transition (Fig. 5D).

AC C

187

193

8

ACCEPTED MANUSCRIPT 194

4.1 Facies associations and depositional environments By integrating previous work (e.g., Jaillard, 1986, 1987; Jaillard and Arnaud-Vanneau, 1993;

196

Dhondt and Jaillard, 2005; Jaillard et al., 2005) with our field data, four main depositional

197

environments, each with a number of standard facies types are established for the intertidal to outer

198

ramp domain (Table 1). The documentation of these fundamental sedimentological data is important to

199

match the type of carbonate producers versus OAEs.

RI PT

195

Please note that the first depositional environment is represented by facies 1a, 1b and 1c, where

201

facies 1a is the most proximal deposits composed of sandy marls with scarce fauna (Fig. 6A).

202

Whereas, facies 1b and 1c, studied by Jaillard (1986) for the lower Cenomanian in the Uchucchacua

203

section, are generally represented by beds of dolostones with sedimentary structures like tepees,

204

birdeyes and mudcracks (facies 1b) and by algal laminites and lenticular structures (facies 1c),

205

corresponding all to a intertidal zone setting (please refer to Jaillard (1986)´s work for more details).

M AN U

SC

200

206 4.1.1

Shallow subtidal inner ramp setting

TE D

207

The Albian facies types recognized at Uchucchacua display important similarities with those

209

previously described from the Cajamarca region (Northern Peru, Navarro-Ramirez et al., 2015, 2016)

210

and are only briefly summarized here (Fig. 4 and Table 1). At Uchucchacua, facies 2a is composed of

211

grey argillaceous mudstones with scarce fauna, whereas in more distal sub-environments, facies 2b is

212

characterized by packstones and argillaceous wackestones, exhibiting a nodular fabric due to an

213

increased argillaceous content. Facies 2c is characterized by grainstone and occasionally floatstone

214

and is typically rich in mixed and fragmented shell debris, mainly derived from oysters, gastropods,

215

and echinoids (Table 1).

216 217

AC C

EP

208

The Cenomanian–Turonian facies types recorded at Uchucchacua and Lauricocha show also strong similarities as represented by the Facies 2d–h and 3b and 3c (Table 1).

9

ACCEPTED MANUSCRIPT Facies 2d — Peloid-bearing grainstones (Table 1): This facies is composed of light-grey peloidal

219

grainstones, showing cm to dm-thick lamination. The fauna yields some disperse miliolids, ostracods

220

and sparse Perouvianella peruviana (Fig. 7A) as well as Charophyte gyrogonids. In view of the

221

predominance of peloids, disperse miliolids and the low diversity of microfauna, facies 2d is

222

interpreted to reflect a restricted lagoon environment (Gräfe, 2005) with limited water circulation

223

(Masse et al., 2003) and low trophic levels (Hüneke et al., 2001).

RI PT

218

Facies 2e — Discorbids and ostracods wakestone/mudstone (Table 1): This facies is primarily

225

composed by small debris (layers) of discorbid and ostracod valvae accumulations whilst rare

226

miliolids are also present (Fig. 6B). This basic facies type has been described first by Sartoni and

227

Crescenti (1962) from the Mesozoic of Italy and was interpreted as an environment characterized by

228

restricted water circulation, medium to high trophic levels and low salinity.

M AN U

SC

224

Facies 2f — Dasycladales-bearing pack-grainstones (Table 1): This facies is built by dm- to m-

230

thick bedded, bluish-grey packstones and occasionally by grainstones. The rocks are typically

231

characterized by large rounded bioclasts composed of dasycladales remains (Figs. 7B and 8A–B). The

232

main benthic foraminifera are Perouvianella peruviana and miliolids whose tests are often abraded.

233

Occasionally peloids are present. Distribution of dasycladales may occur in shallow shoals, lagoonal,

234

semi-restricted very shallow environment and backreef environments (Flügel, 2004).

EP

TE D

229

Facies 2g — Miliolid-bearing pack-grainstones (Table 1): This facies is made of bluish-grey

236

packstones to grainstones, displaying dm-thick lamination in units of two to five metres thick (Fig.

237

6C). Bioturbation features are abundant. The fauna is dominated by miliolids, being associated to

238

Perouvianella peruviana (Fig. 7C) and numerous dasycladal algae. Peloids also occur in certain levels.

239

Associations of coated grains, skeletal elements such as miliolids, Perouvianella peruviana, and

240

dasycladales indicate deposition in the nearshore shallow zone (Flügel, 2004; Lézin et al., 2012).

241

Modern miliolids live mainly in warm, clear water lacking significant fresh-water influx and in some

242

places they are associated with reefal settings (Fang, 2003).

AC C

235

10

ACCEPTED MANUSCRIPT Facies 2h — Perouvianella peruviana-bearing grain-, rud-, and floatstones (Table 1): This facies

244

is characterized by grainstones, rudstones, and floatstones containing numerous Perouvianella

245

peruviana (Fig. 7D). Facies 2h builds thickening upwards sequences, each one being 20 to 80 m in

246

thickness. Typical biota include miliolids, Rotorbinella mesogeensis (Figs. 8C–D), agglutinated (Figs.

247

8E and G), porcelaneous foraminifera (Fig. 8F) and Perouvianella peruviana (Figs. 8H–I). According

248

to previous workers, this facies may indicate deposition in the external part of the inner platform (<20

249

m; Stein et al., 2012), located in the platform margin (Flügel, 2004) with specifically clear and well-

250

illuminated waters and low trophic levels (Lézin et al., 2012).

SC

4.1.2

Open marine middle ramp setting

M AN U

251 252

RI PT

243

In the Central Peru upper Albian interval, facies types 3a, 3b, 3c, and 4a are near-identical to those

254

identified at the Piedra Parada location (Northern Peru) as previously describe in detail in Navarro-

255

Ramirez et al. (2016; Table 1). For the sake of brevity, we only report the most important

256

characteristics here. Facies 3a is mainly characterized by oyster bioherms, interbedded by rare

257

grainstone units yielding various skeletal elements and intraclasts. Facies 3b (diverse fauna low

258

energy) in the lower Jumasha Formation is characterized by 5 to 10 cm-thick layers with variably

259

sized fragments of bryozoans, echinoids, and planktonic foraminifera. Whereas, this facies in the

260

upper Jumasha Formation, occurs as dm to m-thick bedded grey pack to grainstones and occasionally

261

by floatstones (Fig. 6D). This here includes mixed and fragmented shell debris mainly consisting of

262

Perouvianella peruviana, miliolids, gastropods, spicules and rare charophyte gyrogonites (Fig. 7E).

263

Dominance of intraclasts appear also in this facies, which are usually interpreted to be formed by

264

storm wave erosion occurring in shallow-marine environments (Flügel, 2004). Following previous

265

workers, this type of deposits indicate deposition in high-energy, shallow subtidal shoals possibly also

266

in the form of storm event beds (Elrick et al., 2009), just beneath the fair-weather wave base (Amini et

267

al., 2004; Stein et al., 2012; Lézin et al., 2012).

AC C

EP

TE D

253

11

ACCEPTED MANUSCRIPT Facies 3c (echinoid-bearing packstones to wackestones) is characterized by cm-thick laminae

269

organized in higher order units of two to five metres in thickness and is rich in echinoid debris (Figs.

270

6E and 7F). These features suggest a middle ramp environment and sediment deposition beneath the

271

fair-weather wave base (Flügel, 2004; Stein et al., 2012). This facies 3c appears also in the upper

272

Cenomanian sections at Uchucchacua and Lauricocha.

RI PT

268

Facies 4a (planktonic foraminifera-bearing mudstones) is characterized by dark grey mudstones

274

with mm to cm-thick lamination and a generally undisturbed horizontal bedding. The fauna includes

275

planktonic foraminifera and small echinoid fragments. These evidence an open marine setting below

276

the reach of the effective storm wave base (Lézin et al., 2012; Stein et al., 2012).

277 278

4.2 Sequence stratigraphic interpretation

M AN U

SC

273

A basic interpretation of the sequence stratigraphic scheme of Central Peru was presented in

280

Jaillard (1986, 1987), Jaillard and Arnaud-Vanneau (1993) and Jaillard et al. (2005) and for the

281

Northern Peru by Navarro-Ramirez et al. (2015, 2016), respectively. Following these approaches, we

282

tie the sequence stratigraphic for both regional settings as indicated in figure 3, where the Jumasha

283

Formation in Uchucchacua is divided in 5 large-scale sequences (Jumasha 1–5; Fig. 3). Here, the

284

upper Albian to lower Cenomanian is located in the first Jumasha sequence and the upper Cenomanian

285

to lower Turonian in the Jumasha 4 (Figs. 9 and 10). This is feasible as based on biostratigraphic data

286

from the northern and central carbonate ramp of Peru (Benavides-Caceres, 1956; Wilson 1963; von

287

Hillebrandt, 1970; Romani, 1982; Jaillard 1986, 1987; Jaillard and Arnaud-Vanneau, 1993; Robert

288

2002; Robert and Bulot, 2004, 2005; Dhondt and Jaillard, 2005; Jaillard et al., 2005; Robert et al.,

289

2009).

AC C

EP

TE D

279

290

The sequence stratigraphy approach used here is based on the work of Embry (2009) and Andrieu

291

et al. (2015) for shallow marine environments. Essentially, depositional sequences recognized are

292

bounded by sequences boundaries (SB), which represent either maximum regressive surfaces or 12

ACCEPTED MANUSCRIPT transgressive surfaces (van Wagoner et al., 1988). These surfaces indicate a significant shift from a

294

shallowing-upward to a deepening-upward pattern (Embry, 2009; Andrieu et al., 2015). Moreover, a

295

depositional sequence is a cycle that comprises: transgressive deposits (TD) marking a change to

296

upwards deepening facies (retrograding structures); a maximum flooding interval (MFI), a portion of

297

the section that marks the transition from deepening-upward to shallowing-upward patterns; and a

298

highstand deposit (HD), representing shallowing facies grading upwards into more proximal facies

299

(prograding structures).

RI PT

293

The lowermost medium-scale sequence at Jumasha 1 (Fig. 9) consists of thickening-upward units

301

of argillaceous, nodular limestone built by facies 3a rich in oysters. This pattern may represent a

302

highstand system tract. These units were attributed to the upper middle Albian due to ammonite

303

associations such as Lyelliceras ulrichi, Oxytropidoceras douglasi (Benavides-Caceres, 1956; Wilson,

304

1963; von Hillebrandt, 1970; Jaillad, 1986). The two overlying medium-scale sequences (Fig. 9) show,

305

in their transgressive deposits, thickening-upward transitions ranging from facies types 2c, 3a, to

306

facies 3c. These units are followed by deepening trends to outer ramp sedimentation typified by facies

307

4a with abundant planktonic foraminifera, defining the maximum flooding interval in these two

308

sequences. The highstand deposits are represented by thickening-upward successions from facies 3c to

309

3a, increasingly enriched in quartz grains (facies 1a), oysters and serpulids. At section metre 115 m,

310

measured from the base of the Jumasha Formation at Uchucchacua, Romani (1982) reported Exogyra

311

costagyra cf olisoponensis and Merlingina cretacea of Cenomanian age. Echinoderm and oyster

312

limestones characterize the last sequence (122 to 156 m; Fig. 9).

AC C

EP

TE D

M AN U

SC

300

313

The upper Cenomanian–lower Turonian portion at Uchucchacua and Lauricocha is comprised in

314

the Jumasha 4 (Fig. 10), where is showed three medium-scale sequences (MsS1–3). These are part of

315

the three depositional sequences Ro1–3 sensu Jaillard and Arnaud-Vanneau (1993). The medium-scale

316

sequences 1–2 with thickness of 60 to 140 metres (MsS1–2; Fig. 10) build the Jumasha 4. The

317

sequence boundary that limits the top of the Jumasha 4 (Fig. 6F), displays a prominent stratigraphic

13

ACCEPTED MANUSCRIPT 318

feature, defining a major change in depositional style from a regressive to a major transgression event

319

in the early Turonian (Jaillard and Arnaud-Vanneau, 1993; Fig. 10). At the base of both sections, an abrupt deepening trend from nearshore restricted (facies 2d–h) to

321

middle ramp facies (3b) is observed through medium-scale sequence 1 and may define a transgressive

322

system tract in terms of the Jumasha 4. The maximum of this deepening interval, situated at around

323

section metre 35 in both sequences, is marked by an increase in facies associations 3c. This is

324

followed by a shallowing-upward trend recorded in the medium-scale sequences 1, indicated by a

325

thickening-upward transition facies 3c upwards 2d and presence of rotaliid foraminifera Rotorbinella

326

mesogeensis, representing the regressive trend of the medium-scale sequence 1. This is followed by a

327

deepening trend represented from facies 1a, 2d to facies 3b, which form the transgressive deposits of

328

the medium-scale sequence 2. The regressive deposits of the medium-scale sequence 2 are upsection

329

indicated by the appearance of facies 2h to 2d. Medium-scale sequence 3 (MsS3, Fig. 10) is at least 40

330

m thick (but not fully exposed). This sequence includes the lower Turonian and the Jumasha 5. Facies

331

are dominated by 3c, which is well-defined in the lower part of both sections, showing a middle ramp

332

environment rich in echinoids. The transition from facies 3c into the overlaying thickly-bedded

333

limestones of facies 2d is gradual, marking the highstand system tract.

335

SC

M AN U

TE D

EP

334

RI PT

320

4.3 Carbon-isotope stratigraphy

The upper Albian to lower Cenomanian bulk micrite and organic carbon isotope stratigraphy at

337

Uchucchacua is documented in figure 9. At the lowermost part of the curve, the δ13Ccarb and δ13Corg

338

values vary between –0.3 and +2.7‰ and between –27.4 and –25.3‰. These are followed by a

339

significant positive shift from –0.2 to +3.1‰ (δ13Ccarb) and from –27.4 to –24.2‰ (δ13Corg), showing a

340

prominent positive chemostratigraphic feature. From about section metre 90 onwards (Fig. 9),

341

values decrease reaching 1.3‰ in δ13Ccarb and –26.8‰ in δ13Corg, followed by increasingly

342

enriched values reaching +2.9‰ and –24.2‰, respectively. The top of the prominent positive isotope

343

interval ends with a declining trend to a base level of +1.1‰ in δ13Ccarb and –26.7‰ in δ13Corg. 14

AC C

336

13

C-

13

C

ACCEPTED MANUSCRIPT Based on existing biostratigraphic data in Central Peru (von Hillebrandt, 1970; Romani, 1982;

345

Jaillard and Arnaud-Vanneau, 1993) and the chemostratigraphic terminology used for OAE2 interval

346

(onset, trough, plateau, and recovery, c.f., Paul et al., 1999; Tsikos et al., 2004), and datum levels A,

347

B, C and D as used in Tethyan and North Atlantic sections (Meyers et al., 2012; Du Vivier et al., 2014,

348

2015), we characterize specific sub-intervals of the Cenomanian–Turonian isotope excursion for the

349

Uchucchacua and Lauricocha sections (Fig. 10).

RI PT

344

In terms of their δ13C amplitudes and trends, the Uchucchacua and Lauricocha isotope curves are

351

directly comparable. In both sections, the lower part of the prominent carbon isotope excursion related

352

to OAE2 was not covered due to a gap in exposure. However, the OAE2 interval at Uchucchacua

353

displays a significant rise from mean values of +3.7‰ to a first peak of +4.6‰ in δ13Ccarb and a high

354

ratio –22.6‰ in δ13Corg, whereas at Lauricocha the δ13Ccarb curve shows

355

specific segment is here interpreted to represent the upper part of the onset of OAE2. Upsection at

356

both locations, both bulk micrite and organic carbon ratios are characterized by decreasing 13C-values

357

of 3.3‰ in δ13Ccarb and –25‰ in δ13Corg, followed by a rapid trend to

358

+4.2‰ and –22.1‰, respectively. This pattern corresponds with a `trough´ according to the Tethyan

359

scheme (Du Vivier et al., 2014; Fig. 10). Thereafter, a plateau interval of δ13Ccarb and δ13Corg values

360

ranging between +1.78 to +3.3‰ and –23 to –20.9‰ is recorded. The recovery interval is

361

characterized by a rapid decline to a base level of +2.5‰ in δ13Ccarb and –25.9‰ in δ13Corg.

13

13

C depleted values. This

C-enriched values reaching

EP

TE D

M AN U

SC

350

AC C

362

363

5

Discussion

364

5.1 Environmental evolution of the Western Platform from the Albian towards the Turonian

365

Jaillard et al. (1990) suggested that the opening of the North Atlantic (Western Tethys) in the

366

Early Cretaceous lead to the formation of a Tethyan oceanic seaway that extended through the Palaeo-

367

Pacific realm. This interoceanic communication may explain the Tethyan affinities of some Pacific

368

provinces during the Aptian to Turonian (e.g., Iba and Sano, 2007; Robert, 2002; Robert and Bulot, 15

ACCEPTED MANUSCRIPT 2004; Dhondt and Jaillard, 2005). According to Robert (2002) and Robert and Bulot (2004), the

370

earliest Albian was dominated by migration of Tethyan Engonoceratidae specimens to the Western

371

Platform due to an Albian transgression (Robert, 2002). Cretaceous Tethys taxa have also been

372

recorded in the Berriasian to early Albian of the north-western Pacific, perhaps evidencing a Tethyan

373

biotic realm in the Pacific in that time (Iba and Sano, 2007; Robert and Bulot, 2004). In addition to

374

this palaeontological observation, bivalve taxa of the Albian-Cenomanian, recorded at Pongo de

375

Rentema in Peru, are characterized by taxa known from Tethyan faunas, North Africa, southwestern

376

Europe, and Texas (Dhondt and Jaillard, 2005). It is thus conceivable that water masses from the

377

North Atlantic (Western Tethys) reached the Western Platform during the Albian-Cenomanian

378

(Robert, 2002; Jaillard et al., 1990; Dhondt and Jaillard, 2005).

M AN U

SC

RI PT

369

Moreover, in the earliest middle Albian, the South America margin witnessed the onset of the

380

Huarmey-Trough volcanism, leading the depositions of large volumes of basaltic pillow lavas

381

(Atherton and Aguirre, 1992; Winter et al., 2010) . This volcanic activity has been temporally

382

associated to an increased convergence rate of the rifting of South America and Africa (Soler and

383

Bonhomme, 1990; Moulin et al., 2010; Winter et al., 2010). The uplift of the Huarmey-Trough may

384

have also isolated the Western Patform from vigorous exchange and mixing with water masses of the

385

Pacific Ocean in the middle Albian (Fig. 11A). The rifting of South America and Africa continued

386

through the Cenomanian, forming continuous deep-water connection between the North and South

387

Atlantic Ocean (Wagner and Pletsch, 1999; Poulsen et al., 2003). This tectonic event triggered a major

388

reorganization of the global shallow and deep oceanic circulations and regional climate changes

389

(Poulsen et al., 2003). Another feature related to this pattern is the presence of new species of

390

ammonites, bivalves, and benthic foraminifers, which prospered in shallow-marine environments in

391

the Pacific (e.g., Chinzei, 1986; Toshimitsu et al., 1990; Jaillard and Arnaud-Vanneau, 1993; Robert,

392

2002; Robert and Bulot, 2004; Dhondt and Jaillard, 2005; Iba and Sano, 2007).

AC C

EP

TE D

379

393

These new Peruvian species are exemplified by Glottoceras moorei and Engonoceras gr. stolleyi-

394

hilli Böhm from the middle Albian (Robert and Bulot, 2004). Bivalve communities also prospered in 16

ACCEPTED MANUSCRIPT the Western Platform in the late middle Albian to late Cenomanian (Dhondt and Jaillard, 2005). The

396

lack of fossil remains of reef constructers (Jaillard, 1987) suggests the establishment of a heterozoan

397

ramp devoid of any morphological features such as a rimmed platform morphology (Fig. 11A). The

398

absence of a rim or barrier may have caused rather uniform facies belts across wide portions of the

399

Western Platform, where sediment and organic material were transported and re-distributed by storm

400

action and currents (Navarro-Ramirez et al., 2015). The presence of argillaceous material in facies 1a

401

and 2a-b points to continent-derived sediments probably transported by rivers. According to Jaillard

402

(1987), during the late Albian to late Cenomanian, a diachronic progradation of terrigenous supratidal

403

facies was accentuated, resulting in a NE-oriented deltaic system fed by runoff from the Brazilian

404

continental basement (Fig. 11A). This may indicate the installation of high trophic levels and more

405

humid climate with all corresponding changes in sediment runoff and type on the Western Platform

406

(Fig. 11A).

M AN U

SC

RI PT

395

In the late Cenomanian towards the early Turonian, a key feature of Jumasha 4 recorded at

408

Uchucchacua and Lauricocha (central Peru) is the ubiquitous presence of Perouvianella peruviana-

409

dominated carbonate facies (Fig. 10). During this period, central Peru was isolated from the Pacific

410

and from eastern deltaic influx of the Brazilian continental basement due to the uplift of the Marañon

411

Massif during the early Cenomanian (Jaillard, 1986; Fig. 11B). Evidences of this tectonic event is

412

given by a shallowing maximum, associated with tectonic breccias and synsedinemtary normal faults,

413

occurring from the lower to the middle Cenomanian (Jaillard, 1986). Furthermore, the presence of a

414

morphological barrier is given by the lack of an argillaceous material in these sections (Jaillard, 1986,

415

1987), probably implying the installation of an oligotrophic environment (Fig. 11B). This later

416

palaeoceanographic feature might have triggered – or at least favoured – the mass occurrence of

417

Perouvianella peruviana and associated foraminifera, being only known from these Central Peruvian

418

rocks (Jaillard, 1986 and references therein). In contrast to northern Peru (Pongo de Rentema), benthic

419

taxa are represented by inoceramids such as Mytiloides mytiloides, M. opalensis, and Sergipia sp.

420

(Dhondt and Jaillard, 2005). The genus Sergipia has a wide distribution, being reported in Brazil,

AC C

EP

TE D

407

17

ACCEPTED MANUSCRIPT 421

Nigeria, Mexico, Japan, and France (Hessel, 1988; Jolet et al., 1997; Dhondt and Jaillard, 2005), and

422

indicates the installation of an open marine setting in northern Peru (Fig. 11B). Perouvianella peruviana is a porcelaneous benthic foraminifera with a complex shell architecture

424

and flattened shape. The divisions of chamber lumen below the lateral wall may indicate the loci for

425

endosymbionts. In recent oceans, the large discoidal porcelaneous Archaias hosts clorophycean algae

426

symbionts in their shells which restrict their habitat to the first 20 m of the water column. Besides,

427

Marginopora, Amphisorus and Sorites possess dinophycean algae as symbionts allowing them to

428

reach a life depth of about 50 m. (Leutenegger, 1984). On the level of a working hypothesis, similar

429

environmental conditions are inferred for Perouvianella peruviana. Concerning the reproduction cycle

430

of Perouvianella peruviana, in the fourth Jumasha sequence, two A-forms have been found (see Fig.

431

8I) but no B-forms are present. This may be compared with the population of Amphisorus hemprichii

432

from the Red Sea (Zohary et al., 1980). Dasylcadacean and Codiacean algae (Figs. 8A–B) in

433

combination with miliolids, hyaline and agglutinated foraminifera (Figs. 8E–G) associated with

434

Perouvianella peruviana support a local oligotrophic environment within the photic zone in which

435

these foraminifera thrived. A foraminifer that is perhaps similar to Perouvianella peruviana has been

436

found the Cenomanian of Mexico (Jaliscella sigali, Fourcade et al., 1990). The presence of

437

Perouvianella and Jaliscella on the American continent, and the lack of alveolinoideans (the genus

438

Sellialveolina indicated by Jaillard and Arnaud-Vanneau (1993) is not found in our samples) may

439

confirm an endemic Central Peruvian palaeobioprovince that is in contrast with the Tethyan biotic

440

realm for the late Cenomanian–early Turonian time interval.

442

SC

M AN U

TE D

EP

AC C

441

RI PT

423

5.2 Evidence for a disturbed carbon cycle

443

Shallow water carbonates have been shown to record, under favourable conditions, first order

444

patterns of the global marine carbon isotope signature (Weissert et al., 1998; Immenhauser et al.,

445

2005; Föllmi et al., 2006; Huck et al., 2010, 2011, 2012, 2013, 2014; Andrieu et al., 2015). The global

446

seawater δ13CDIC preserved in the rock record are related to changes in the ratio of marine carbonate 18

ACCEPTED MANUSCRIPT 447

carbon and organic carbon burial (Scholle and Arthur, 1980; Weissert, 1989; Immenhauser et al.,

448

2008). With reference to the Peruvian sections documented here, we argue that the mid-Cretaceous low-

450

amplitude sea-level fall was insufficient to expose significant portions of the carbonate ramp studied.

451

Reasons for that might include rapid basement subsidence, a feature that is supported by the

452

stratigraphically thick successions found for the OAE2 interval (Ucchucchacua = 210 m and

453

Cajamarca composite section = 52 m; Fig. 12). Further evidence for rapid basement subsidence comes

454

from the remarkable lack of karst-related features, bleaching, isotopic shifts or other patterns assigned

455

to meteoric diagenesis (Allan and Matthews, 1977; Christ et al., 2012; Huck et al., 2014) beneath

456

discontinuity surfaces recorded in the sections studied. This may imply only weak or no meteoric

457

diagenesis and suggest that the carbonates studied here experienced mainly marine and subsequent

458

burial diagenesis. Judging from the direct comparability of chemostratigraphic sections in Peru with

459

such in many other locations worldwide (Sageman et al., 2006; Jarvis et al., 2011), we tentatively

460

suggest that most of these carbonates stabilized in the presence of marine pore waters. This is turn

461

implies that at least the chemostratigraphic patterns, but perhaps not absolute isotope values, represent

462

palaeoceanographic features as opposed to diagenetic patterns. The question to which degree these

463

patterns reflect global features versus regional water mass properties merits discussion.

TE D

M AN U

SC

RI PT

449

In the Tethyan Ocean, OAE1d is associated to carbon isotope shifts with amplitudes between 0.5

465

and 1.5‰ (e.g., Bornemann et al., 2005; Gambacorta et al., 2015 and references therein; Fig. 9).

466

Despite the limitations of the biostratigraphic control used here, the δ13Ccarb and δ13Corg records at

467

Uchucchacua indicated shifts of 3‰ during the late Albian being more pronounced than those

468

recorded in the Cajamarca composite section (1.5‰ δ13Ccarb and 2.5‰ for δ13Corg; Fig. 12). The

469

carbonates facies, within the limitations of facies characterization, are near-identical in both areas (Fig.

470

11A). This implies that facies-control on chemostratigraphy is weak to absent. It is likely that the

471

regional offset in δ13Ccarb ratio amplitude may have been induced by differences in the water mass

AC C

EP

464

19

ACCEPTED MANUSCRIPT 472

residence time or differential

473

Immenhauser et al., 2008).

13

C sources (Lloyd, 1964; Holmden et al., 1998; Saltzman, 2003;

For OAE2, the contrast-comparison of δ13Ccarb and δ13Corg records from the Uchucchacua and

475

Lauricocha sections (situated 36 km apart) agrees with the notion of mainly palaeoceanographic

476

patterns and a weak diagenetic overprint (Fig. 10). However, discrepancies exist in the amplitudes of

477

the δ13Ccarb and δ13Corg chemostratigraphic curves when these are compared with the reference section

478

of the Cajamarca region situated approximately 500 km to the north (Fig. 12; Navarro-Ramirez et al.,

479

2016). The differential amplitudes of δ13Ccarb and δ13Corg records are more pronounced in Cajamarca

480

location than in sections at the Uchucchacua and Lauricocha sites. Probably, differences in

481

palaeogeographic settings significantly affected or overprinted global patterns at Uchucchacua and

482

Lauricocha as indicated by mesotrophic and oligotrophic environments in northern and central Peru,

483

respectively (Fig. 11B).

M AN U

SC

RI PT

474

484 5.2.1

Western Platform patterns during the mid-Cretaceous OAEs

TE D

485

In the Western Platform domain, the uppermost Aptian Jacob Level (1st black-shale level of the

487

OAE1b set), which is associated to relative sea-level fall near the Aptian–Albian boundary, caused

488

erosional surfaces, disconformities, basal conglomerates, and ferruginous deposits as reported at the

489

base of the Inca Formation (Jaillard, 1987; Fig. 12). This relative sea-level fall had an amplitude of at

490

least 50m (Maurer et al., 2013) and is probably linked to global cooling and ice sheet dynamics at

491

high-latitudes (Frakes and Francis, 1988; Weissert and Lini, 1991; Frakes et al., 1995; Bodin et al.,

492

2015) responding to global scale and prolonged episodes of organic carbon (Menegatti et al., 1998;

493

Gröcke, 1998; Bralower et al., 1999; Herrle et al., 2004; Westermann et al., 2010; McAnena et al.,

494

2013; Bodin et al., 2015). For this time interval, decreased rates of continental weathering as

495

suggested by low 87Sr/86Sr values and the coeval positive shift of δ18O have been proposed (Bodin et

496

al., 2015).

AC C

EP

486

20

ACCEPTED MANUSCRIPT The Kilian and Paquier levels (2nd and 3rd black-shale levels of the OAE1b set) led to transition

498

stages from a siliciclastic-dominated sedimentation to a heterozoan-neritic carbonate factory, followed

499

by incipient phases of platform demise of the Tethyan biotic realm in the Pacific (Robert and Bulot,

500

2004; Iba and Sano, 2007). The earliest Albian corresponds to a change from low to high sea-level,

501

marked by a transition between coldhouse to greenhouse climate mode (Bodin et al., 2015). This

502

climatic pattern is perhaps due to the activity of the South Kerguelen plateau around 112–110 Ma

503

(Coffin et al., 2002; Duncan, 2002; Frey et al., 2003; Bryan and Ferrari, 2013). Early Albian global

504

warming led to a transgressive event that transported water masses from the North Atlantic (Western

505

Tethys) onto the neritic domain of the Western Platform. Moreover, global warming also shifted the

506

Inter-Tropical Convergence Zone (ITCZ) towards the equator, installing more humid climate in the

507

Western Platform domain (Hay and Flögel, 2012; Hasegawa et al., 2012). The enhanced hydrological

508

cycle may have increased continental weathering of the Brazilian continental shield, delivering

509

nutrients and clay-sized material to the ramp, promoting high productivity of benthic organisms

510

(oysters, gastropods, and echinoids). High productivity and an abrupt change in lithology and biota

511

have been associated to the OAE1b set around the globe, indicating the significance of this event

512

during the Cretaceous (Erbacher et al., 1996, 1998, 1999; 2001; Leckie et al., 2002).

TE D

M AN U

SC

RI PT

497

The Leenhardt level (i.e., the last black-shale level of the OAE1b set) is contemporaneous to a

514

major demise of the Cretaceous Tethyan taxa in the Pacific and neritic carbonate production. In the

515

field in Peru, this feature is marked by a regional discontinuity and the disappearance of all lower

516

Albian ammonite species with the exception Glottoceras crassinodosum Sommermeier (Robert, and

517

Bulot, 2004). High atmospheric CO2 output by the South Kerguelen plateau reduced carbonate

518

availability and seawater pH, thereby causing demise episodes of the neritic carbonate production, as

519

well as the loss of calcareous plankton during the broad interval of the OAE1b set (Leckie et al.,

520

2002).

AC C

EP

513

521

In the middle Albian, the increased flux of isotopically light CO2 at times of the activation of the

522

western South America volcanic arc, shifted the δ13Ccarb towards more negative values as perhaps 21

ACCEPTED MANUSCRIPT 523

documented in the Cajamarca composite section (Navarro-Ramirez et al., 2015, 2016; Fig. 12). This

524

feature is evidenced by reduced carbonate production in the Western Platform domain. The Albian global warming continued until the early Turonian, reaching a mid-Cretaceous climate

526

maximum (Pagani et al., 2014). This led to a rapid diversification and a pronounced increase in the

527

degree of calcification of the planktic foraminifera, increased upper water column stratification,

528

initiation of widespread chalk deposition and oligotrophic conditions in (hemi-) pelagic settings in the

529

Atlantic and Tethys (Leckie, 1989; Premoli Silva and Sliter, 1999; Leckie et al., 2002). In contrast,

530

global warming resulted in an enhanced hydrodynamic circle with erosion delivering nutrients and

531

promoting diversification of benthic communities dominated by oyster bioherms and large endemic

532

benthonic foraminifera in the Western Platform (Fig. 11A). On the other hand, in the Tethyan realm

533

and elsewhere, OAE1d, MCEI, and OAE2 have been associated to high turnover rates affecting the

534

planktic foraminifera and radiolarians, anoxia, organic-rich deposition, and demise of the carbonate

535

platform (e.g., Erbacher et al., 1996, 2001; Gale et al., 2000; Wilson and Norris, 2001; Leckie et al.,

536

2002; Bornemann et al., 2005; Gertsch et al., 2010; Schröder-Adams et al., 2012; Scott et al., 2013;

537

Owens et al., 2013; Giraud et al. 2013; Melinte-Dobrinescu et al., 2015; Gambacorta et al., 2015;

538

Andrieu et al., 2015). These patterns are in clear contrast to phases of enhanced carbonate deposition

539

documented for the sections in Peru discussed here. Persistent deposition of benthic communities in

540

these sections supports the notion of well-oxygenated water masses throughout the middle Albian–

541

early Turonian interval.

AC C

EP

TE D

M AN U

SC

RI PT

525

542

In the Tethyan realm, several groups of foraminifera such as the alveolinids, rhapydionids,

543

pseudorahpydioninids, and nezzazatids characterize the Cenomanian (see Schroeder and Neumann,

544

1985; Calonge et al., 2002; Vicedo et al., 2011; Consorti et al., 2015 and the work cited therein). All of

545

these foraminifera disappear at the Cenomanian–Turonian boundary. The finding of Perouvianella

546

peruviana in Jumasha 4 (Fig. 10) in the central Peru sections is here considered evidence for the

547

continuation of a stable, albeit endemic, marine ecosystem during and after OAE2 contrasting the

548

Tethyan biotic patterns (Parente et al., 2008; Fig. 12). In combination with similar findings in neritic 22

ACCEPTED MANUSCRIPT 549

carbonates in the Pacific - lacking evidence of anoxic conditions during OAE2 (Elrick et al., 2009;

550

Takashima et al., 2011) - this suggests that anoxic conditions did not establish in major portions of the

551

Pacific Ocean during OAE2 (Takashima et al., 2011). In the Cajamarca composite section, a prominent negative shift is observed at the onset of OAE2

553

(Fig. 12). This negative δ13C excursion has also been identified in other sections in the Pacific realm

554

(California, USA; Hokkaido, Japan; Takashima et al., 2011). According to Takashima et al. (2011),

555

this negative δ13C excursion is linked to an increase in pCO2 induced by the formation of the

556

Caribbean large igneous province (LIP; ~95.1–92.2 Ma, Snow et al., 2005). Arguments for the claim

557

that the Caribbean massive volcanism triggered OAE2 come from a negative 34Ssulphate shift ca. 500 kyr

558

before OAE2 (Adams et al., 2010) as well as from a negative

559

2014) and increased pCO2 levels (e.g., Barclay et al., 2010).

SC

RI PT

552

Os/188Os shift (e.g., Du Vivier et al.,

M AN U

187

This significant increase in pCO2 drove the climate to a hothouse mode (Kidder and Worsley

561

2010), perhaps leading to an acidified ocean (Du Vivier et al., 2015) and a perturbed hydrological

562

cycle, that enhanced weathering rates at the beginning of the onset of OAE2 (Frijia and Parente, 2008;

563

Blätter et al., 2011; van Helmond et al., 2013; Pogge von Strandmann et al., 2013). The beginning of

564

the onset of OAE2 in the Western Platform is coeval to an increased influx of argillaceous material

565

and reduced carbonate production in northern Peru (Fig. 11B). However, continuous deposition of

566

floatstones and grainstones containing Perouvianella peruviana is observed in the central Peru

567

sections (Fig. 10) possibly due to the restriction of continental nutrient influx (Fig. 11B). Consumption

568

of significant amount of pCO2 by the enhanced global organic carbon burial and by the continental

569

weathering after the onset of the Caribbean LIP (Sinninghe Damsté et al., 2010; Jarvis et al., 2011; van

570

Helmond et al., 2013; Pogge von Strandmann et al., 2013) conditioned the oceans to gradually be

571

more oxic (Pogge von Strandmann et al., 2013) and might have caused the reduction of benthic

572

communities in the Western Platform of Peru after OAE2 (Figs. 10 and 12).

AC C

EP

TE D

560

573

23

ACCEPTED MANUSCRIPT 574

6

Conclusions The mid-Cretaceous sections of the Western Platform of Peru document a predominance of

576

oyster-rich, mixed siliciclastic-carbonate facies in the middle Albian–late Cenomanian. Regional

577

evidence suggests that this rather uniform facies belt expanded across wide portions of the spatially

578

very extensive carbonate ramp in northern and central Peru. In the late Cenomanian and the early

579

Turonian, the deposition of an endemic large benthic foraminifera facies (Perouvianella peruviana) is

580

a key facies element that continues throughout the OAE2 interval without discernible adaptation to

581

environmental change. During this time, central Peru was isolated from the Pacific water masses and

582

from deltaic influx from the emerged shield of Brazil in the east. This specific palaeoceanographic

583

setting might have promoted the mass occurrence of Perouvianella peruviana. Probably due to the

584

complex setting of the Western Platform in the mid-Cretaceous, clear evidence for a carbonate crisis

585

associated with OAE 1d and OAE 2 in the sections of the central Peru is lacking.

M AN U

SC

RI PT

575

Global warming during the Albian–early Turonian resulted in a more humid climate, high trophic

587

levels, and a diversification of benthic communities in Western Platform carbonates, influenced by

588

palaeogeographic modifications. Even though the benthic communities thrived during extremely warm

589

conditions in the late middle Albian and towards the early Turonian, they were exposed to high

590

nutrient levels caused by a perturbed hydrological cycle at the beginning of the onset OAE2. The

591

intensified chemical weathering on the Brazilian shield during the mid-Cretaceous may have

592

contributed to trigger the OAE in the North Atlantic nutrient trap. The work documented here is a

593

clear case example for the superposition of regional and global patterns in palaeoceanography and

594

complements scarce existing data from South America for this critical interval in Earth’s history.

596

EP

AC C

595

TE D

586

Acknowledgements

597

24

ACCEPTED MANUSCRIPT This project was supported by the Deutsche Forschungsgemeinschaft (DFG, project n° BO-365/2-1)

599

and by the Deutscher Akademischer Austauschdienst (DAAD) through a scholarship to J.P.N. (PKZ:

600

91540654). We thank the Geological survey of Peru (INGEMMET) for important logistical support.

601

Esmeralda Caus (Barcelona) is gratefully acknowledged for a critical reading of the text in its early

602

version. Analytical work was performed in the isotope laboratories at Bochum and Erlangen-

603

Nuremberg. We greatly acknowledge the editorial advice of E. Koutsoukos and the comments of two

604

anonymous journal reviewers.

RI PT

598

SC

605

References

607

Adams, D.D., Hurtgen, M.T., Sageman B.B., 2010. Volcanic triggering of a biogeochemical cascade

608

during oceanic anoxic event 2. Nature Geosciences 3, 201–204.

609

Allan, J.R., Matthews, R.K., 1977. Carbon and oxygen isotopes as diagenetic and stratigraphic tools:

610

surface and subsurface data, Barbados, West Indies. Geology 5, 16–20.

611

Amini, Z., Adabi, M.H., Burrett, C.F., Quilty, P.G., 2004. Bryozoan distribution and growth form

612

associations as a tool in environmental interpretation, Tasmania, Australia. Sedimentary Geology 167,

613

1–15.

614

Andrieu, S., Brigaud, B., Rabourg, T., Noret, A., 2015. The Mid-Cenomanian Event in shallow marine

615

environments: Influence on carbonate producers and depositional sequences (northern Aquitaine

616

Basin, France). Cretaceous Research 56, 587–607.

617

Arthur, M.A., Dean, W.E., Schlanger, S.O., 1985. Variations in the global carbon cycle during the

618

Cretaceous related to climate, volcanism and changes in atmospheric CO2. In: Sundquist, E.T.,

619

Broecker, W.S. (Eds.), the carbon cycle and atmospheric CO2: Natural variations Archean to Present.

620

AGU Geophysical Monograph 32, pp. 504–529.

AC C

EP

TE D

M AN U

606

25

ACCEPTED MANUSCRIPT Atherton, M.P., Aguirre, L., 1992. Thermal and geotectonic setting of Cretaceous volcanic rocks near

622

Ica, Peru, in relation to Andean crustal thinning. Journal of South American Earth Sciences 5, 47–69.

623

Atherton, M.P., Webb, S., 1989. Volcanic facies, structure and geochemistry of the marginal basin

624

rocks of Central Peru. Journal of South American Earth Sciences 2, 241–261.

625

Barclay, R.S., McElwain, J.C., Sageman, B.B., 2010. Carbon sequestration activated by a volcanic

626

CO2 pulse during Ocean Anoxic Event 2. Nature Geoscience 3, 205–208.

627

Blakey,

628

http://cpgeosystems.com/index.html.

629

Blättler, C.L., Jenkyns, H.C., Reynard, L.M., Hendxerson, G.M., 2011. Significant increase in global

630

weathering during Oceanic Anoxic Events 1a and 2 indicated by calcium isotopes. Earth Planetary

631

Science Letters 309, 77–88.

632

Barron, E.J., Washington, W.M., 1985. Warm Cretaceous climates: high atmospheric CO2 as a

633

plausible mechanism. In: Sunquist, E.T., Broecker, W.S., (Eds), the carbon cycle and atmospheric

634

CO2: Natural variations from Archean to Present. AGU Geophysical Monograph 32, pp. 546–553.

635

Benavides-Caceres, V.E., 1956. Cretaceous system in Northern Peru. American Museum of Natural

636

History Bulletin 108, 353–494.

637

Berner, R.A., Lasaga, A.C., Garrels, R.M., 1983. The carbonate silicate geochemical cycle and its

638

effect on atmospheric carbon dioxide over the past 100 million years. American Journal of Science

639

283, 641–683.

640

Bodin, S., Godet, A., Vermeulen, J., Linder, P., Föllmi, K.B., 2006. Biostratigraphy, sedimentology

641

and sequence stratigraphy of the latest Hauterivian - early Barremian drowning episode of the

642

Northern Tethyan margin (Altmann Member, Helvetic nappes, Switzerland).Eclogae geologicae

643

Helvetiae 99, 157–174.

Paleogeographic

map

of

the

Middle

Cretaceous,

Website:

SC

Global

M AN U

2011.

AC C

EP

TE D

R.,

RI PT

621

26

ACCEPTED MANUSCRIPT Bodin, S., Meissner, P., Janssen, N.M.M., Steuber, T., Mutterlose, J., 2015. Large igneous provinces

645

and organic carbon burial: Controls on global temperature and continental weathering during the Early

646

Cretaceous. Global and Planetary Change 133, 238–253.

647

Bornemann, A., Pross, J., Reichelt, K., Herrle, J. O., Hemleben, C., Mutterlose, J., 2005.

648

Reconstruction of short term palaeoceanographic changes during the formation of the Late Albian

649

ʻNiveau Breistrofferʼ black shales (Oceanic Anoxic Event 1d, SE France). Journal of the Geological

650

Society, London 162, 623–639.

651

Bralower, T.J., CoBabe, E., Clement, B., Sliter, W.V., Osburn, C.L., Longoria, J., 1999. The record of

652

global change in mid-Cretaceous (Barremian-Albian) sections fromthe Sierra Madre, northeastern

653

Mexico. Journal of Foraminiferal Research 29, 418–137.

654

Bryan, S.E., Ferrari, L., 2013. Large igneous provinces and silicic large igneous provinces: Progress in

655

our understanding over the last 25 years. Geological Society of America Bulletin 125, 1053–1078.

656

Callot, P., Sempere, T., Odonne, F., Robert, E., 2008. Giant submarine collapse of a carbonate

657

platform at the Turonian1–Coniacian transition: The Ayabacas Formation, southern Peru. Basin

658

Research 20, 333–357.

659

Calonge, A., Caus, E., Bernaus, J. M., Aguilar, M., 2002. Praealveolina (foraminifera): a tool to date

660

Cenomanian platform sediments. Micropaleontology 48, 53–66.

661

Chinzei, K., 1986. Shell structure, growth, and functional morphology of an elongate cretaceous

662

oyster. Palaeontology 29, 139–154

663

Christ, N., Immenhauser, A., Amour, F., Mutti, M., Preston, R., Whitaker, F., Peterhänsel, A.,

664

Egenhoff, A.O., Dunn, P.A., Agar, S.M., 2012. Triassic Latemar cycle tops — subaerial exposure of

665

platform carbonates under tropical arid climate. Sedimentary Geology 265–266, 1–29.

AC C

EP

TE D

M AN U

SC

RI PT

644

27

ACCEPTED MANUSCRIPT Christ, N., Immenhauser, A., Wood, R.A., Darwich, K. and Niedermayr, A., 2015. Petrography and

667

environmental controls on the formation of Phanerozoic marine hardgrounds. Earth-Science Reviews,

668

151, 176-226.

669

Coffin, M.F., Pringle, M.S., Duncan, R.A., Gladczenko, T.P., Storey, M., Müller, R.D., Gahagan,

670

L.A., 2002. Kerguelen Hotspot Magma Output since 130 Ma. Journal of Petrology 43, 1121–1137.

671

Consorti, L., Caus, E., Frijia, G., Yazdi-Moghadam, M., 2015. Praetaberina new genus (type species:

672

Taberina bingistani Henson, 1948): a stratigraphic marker for the Late Cenomanian. Journal of

673

Foraminiferal Research 45, 370‒389.

674

Dhondt, A.V., Jaillard, E., 2005. Cretaceous bivalves from Ecuador and northern Peru. Journal of

675

South American Earth Sciences 19(3), 325–342.

676

Du Vivier, A.D.C., Jacobson, A.D., Lehn, G.O., Selby, D., Hurtgen, M.T., Sageman, B.B., 2015. Ca

677

isotope stratigraphy across the Cenomanian–Turonian OAE 2: Links between volcanism, seawater

678

geochemistry, and the carbonate fractionation factor. Earth and Planetary Science Letters 416, 121–

679

131.

680

Du Vivier, A.D.C., Selby, D., Sageman, B.B., Jarvis, I., Gröcke, D.R., Voigt, S., 2014. Marine

681

187

682

Oceanic Anoxic Event 2. Earth and Planetary Science Letters 389, 23–33.

683

Duncan, R.A., 2002. A Time Frame for Construction of the Kerguelen Plateau and Broken Ridge.

684

Journal of Petrology 43, 1109–1119.

685

Eagles, G., 2006.New angles on South Atlantic Opening. Geophysical Journal International 7 (28).

686

doi:10.1111/j.1365-246X.2006.03206.x.

687

Elrick, M., Molina-Garza, R., Duncan, R., Snow, L., 2009. C-isotope stratigraphy and

688

paleoenvironmental changes across OAE2 (mid-Cretaceous) from shallow-water platform carbonates

689

of southern Mexico. Earth and Planetary Science Letters 277, 295–306.

TE D

M AN U

SC

RI PT

666

AC C

EP

Os/188Os isotope stratigraphy reveals the interaction of volcanism and ocean circulation during

28

ACCEPTED MANUSCRIPT Embry, A.F., 2009. Practical Sequence Stratigraphy. Canadian Society of Petroleum Geologists, 79

691

pp.

692

Erba, E., 1994. Nannofossils and superplumes: the early Aptian “nannoconid crisis”.

693

Paleoceanography 9, 483–501.

694

Erbacher, J., Huber, B.T., Norris, R.D., Markey, M., 2001.Increased thermohaline stratification as a

695

possible cause for an oceanic anoxic event in the Cretaceous period. Nature 409, 325–327.

696

Erbacher, J., Hemleben, C., Huber, B.T., Markey, M., 1999. Correlating environmental changes during

697

early Albian oceanic anoxic event 1B using benthic foraminiferal paleoecology, Marine

698

Micropaleontology 38, 7 –28.

699

Erbacher, J., Gerth, W., Schmiedl, G., Hemleben, C., 1998. Benthic foraminiferal assemblages of late

700

Aptian-early Albian black shale intervals in the Vocontian Basin, SE France. Cretaceous Research 19,

701

805–826.

702

Erbacher, J., Thurow, J., Littke, R., 1996. Evolution patterns of radiolaria and organic matter

703

variations: A new approach to identify sea-level changes in mid-Cretaceous pelagic environments.

704

Geology, 24: 499–502.

705

Fang, Q., 2003. Quantitative Analysis of Biofacies and Genetic Sequence Stratigraphy of the Yegua

706

Formation, Houston Salt Embayment, Northern Gulf of Mexico, U.S.A. In Olson, H.C., Leckie R.M.,

707

(Eds) Micropaleontologic Proxies for Sea-Level Change and Stratigraphic Discontinuities, SEPM

708

Special Publication Number 75, pp. 201–228.

709

Fourcade, E., Fleury, J.J., Michaud, F., Bourgois, J. 1990. Jaliscella sigali n.gen, n.sp., nouveau

710

foraminifère Soritacea du Cénomanien du Mexique. Palaeontographica Abteilung A Band A214

711

Lieferung 1-2, 13‒29.

712

Flügel, E., 2004. Microfacies of Carbonate Rocks. Springer, Berlin, 996 pp.

AC C

EP

TE D

M AN U

SC

RI PT

690

29

ACCEPTED MANUSCRIPT Föllmi, K.B., Godet, A., Bodin, S., Linder, P., 2006. Interactions between environmental change and

714

shallow-water carbonate build-up along the northern Tethyan margin and their impact on the early

715

Cretaceous carbon-isotope record. Paleoceanography 21, PA4211.

716

Föllmi, K.B., Weissert, H., Bisping, M., Funk, H., 1994. Phosphogenesis, carbon isotope stratigraphy,

717

and carbonate-platform evolution along the Lower Cretaceous northern Tethyan margin. Geological

718

Society of America, Bulletin 106, 729–746.

719

Frakes, L.A., Alley, N.F., Deynoux, M., 1995. Early Cretaceous ice rafting and climate zonation in

720

Australia. International Geology Review 37, 567-583.

721

Frakes, L.A., Francis, J.E., 1988. A guide to Phanerozoic cold polar climates from high latitude ice-

722

rafting in the Cretaceous. Nature 333, 547–549.

723

Frey, F.A., Coffin, M.F., Wallace, P.J., Weis, D., 2003. Leg 183 synthesis: Kerguelen Plateau–Broken

724

Ridge — a large igneous province, in: Frey, F.A., Coffin, M.F., Wallace, P.J., Quilty, P.G., (Eds.),

725

Proceedings of the Ocean Drilling Program, Scientific Results 183, pp. 1-48.

726

Frijia, G., Parente, M., 2008. Strontium isotope stratigraphy in the upper Cenomanian shallow-water

727

carbonates of the southern Apennines: Short-term perturbations of marine 87Sr/86 Sr during the Oceanic

728

Anoxic Event 2. Palaeogeography, Palaeoclimatolology, Palaeoecolology 261, 15–29.

729

Gale, A.S., Smith, A.B., Monks, N.E.A., Young, J.A., Howard, A., Wray, D.S., Huggett, J.M., 2000.

730

Marine biodiversity through the Late Cenomanian–Early Turonian: palaeoceanographic controls and

731

sequence stratigraphic biases. Journal of the Geological Society, London, 157, 745–757.

732

Gambacorta, G., Jenkyns, H.C., Russo, F., Tsikos, H., Wilson, P.A., Faucher, G., Erba, E., 2015.

733

Carbon- and oxygen-isotope records of mid-Cretaceous Tethyan pelagic sequences from the Umbria–

734

Marche and Belluno Basins (Italy). Newsletters on Stratigraphy 48/3, 299–323.

AC C

EP

TE D

M AN U

SC

RI PT

713

30

ACCEPTED MANUSCRIPT Gertsch, B., Adatte, T., Keller, G., Tantawy, A.A.A.M., Berner, Z., Mort, H.P., Fleitmann, D., 2010.

736

Middle to late Cenomanian oceanic anoxic event in shallow and deeper shelf environments of western

737

Morocco. Sedimentology 57, 1430–1462.

738

Giraud, F., Olivero, D., Baudin, F., Reboulet, S., Pittet, B., Proux, O., 2003. Minor changes in surface-

739

water fertility across the oceanic anoxic event 1d (latest Albian, SE France) evidenced by calcareous

740

nannofossils. International Journal of Earth Sciences, 92, 267–284.

741

Glikson, A., 2010. Climate, Fire and Human Evolution. Approaches in Solid Earth Sciences 10, DOI

742

10.1007/978-3-319-22512-8_1.

743

Gräfe, K.U., 2005. Late Cretaceous benthic foraminifers from the Basque-Cantabrian Basin, Northern

744

Spain. Journal of Iberian Geology 3, 277–298.

745

Gröcke, D., 1998. Carbon-isotope analyses of fossil plants as a chemostratigraphic and

746

palaeoenvironmental tool. Lethaia 31, 1-13.

747

Hasegawa, H., Tada1, R., Jiang, X., Suganuma, Y., Imsamut, S., Charusiri, P., Ichinnorov, N., Khand,

748

Y., 2012. Drastic shrinking of the Hadley circulation during the mid-Cretaceous Supergreenhouse.

749

Climate of the Past 8, 1323–1337.

750

Hay, W.W., 1998. Detrital sediment fluxes from continents to oceans. Chemical Geology 145, 287-

751

323.

752

Hay, W.W., Floegel, S., 2012. New thoughts about the Cretaceous climate and oceans. Earth-Science

753

Reviews 115, 262–272.

754

Herrle, J., O., Schröder-Adams, C.J., Davis, W., Pugh, A.T., Galloway, J.M., Fath, J., 2015. Mid-

755

Cretaceous High Arctic stratigraphy, climate, and Oceanic Anoxic Events. Geology doi:

756

10.1130/G36439.1.

757

Herrle, J.O., Kössler, P., Friedrich, O., Erlenkeuser, H., Hemleben, C., 2004. High-resolution carbon

758

isotope records of the Aptian to Lower Albian from SE France and the Mazagan Plateau (DSDP Site 31

AC C

EP

TE D

M AN U

SC

RI PT

735

ACCEPTED MANUSCRIPT 545): a stratigraphic tool for paleoceanographic and paleobiologic reconstruction. Earth and Planetary

760

Science Letters 218, 149-161.

761

Hessel, M.H.R., 1988. Lower Turonian inoceramids from Sergipe, Brazil: systematics, stratigraphy

762

and palaeoecology. Fossils Strata 22–49.

763

Holmden, C., Creaser, R.A., Muehlenbachs, K., Leslie, S.A., Bergström, S.M., 1998. Isotopic

764

evidence for geochemical decoupling between ancient epeiric seas and bordering oceans: Implications

765

for secular curves. Geology 26, 567–570.

766

Huck, S., Rameil, N., Korbar, T., Heimhofer, U., Wieczorek, T.D., Immenhauser, A., 2010.

767

Latitudinally different responses of Tethyan shoal-water carbonate systems to the Early Aptian

768

oceanic anoxic event (OAE 1a). Sedimentology 57, 1585–1614.

769

Huck, S., Heimhofer, U., Rameil, N., Bodin, S., Immenhauser, A., 2011. Strontium and carbon-isotope

770

chronostratigraphy of Barremian–Aptian shoal-water carbonates: Northern Tethyan platform drowning

771

predates OAE 1a. Earth and Planetary Science Letters 304 (3–4), 547–558.

772

Huck, S., Heimhofer, U., Immenhauser, A., 2012. Early Aptian algal bloom in a neritic proto-North

773

Atlantic setting: Harbinger of global change related to OAE1a? GSA Bulletin 124 (11–12), 1810–

774

1825.

775

Huck, S., Heimhofer, U., Immenhauser, A., Weissert, H., 2013. Carbon-isotope stratigraphy of Early

776

Cretaceous (Urgonian) shoal-water deposits: Diachronous changes in carbonate-platform production

777

in the north-western Tethys. Sedimentary Geology 290, 157–174.

778

Huck, S., Stein, M., Immenhauser, A., Skelton, P.W., Christ, N., Föllmi, K.B., Heimhofer, U., 2014.

779

Response of proto-North Atlantic carbonate-platform ecosystems to OAE1a-related stressors.

780

Sedimentary Geology 313, 15–31.

AC C

EP

TE D

M AN U

SC

RI PT

759

32

ACCEPTED MANUSCRIPT Hüneke, H., Joachimski, M., Buggisch, W., Lützner, H., 2001. Marine Carbonate Facies in Response

782

to Climate and Nutrient Level: The Upper Carboniferous and Permian of Central Spitsbergen

783

(Svalbard). Facies 45, 93–136.

784

Iba, Y., Sano, S., 2007. Mid-Cretaceous step-wise demise of the carbonate platform biota in the

785

Northwest Pacific and establishment of the North Pacific biotic province. Palaeogeography,

786

Palaeoclimatology, Palaeoecology 245, 462–482.

787

Immenhauser, A., Hillgärtner, H., van Bentum, E., 2005. Microbial–foraminiferal episodes in the

788

Early Aptian of the southern Tethyan margin: ecological significance and possible relation to oceanic

789

anoxic event 1a. Sedimentology 52, 77–99.

790

Immenhauser, A., Holmden, C., Patterson, W.P., 2008. Interpreting the carbon-isotope record of

791

ancient shallow epeiric seas: lessons from the recent. In: Pratt, B.R., Holmden, C. (Eds.), Dynamics of

792

Epeiric Seas, Geological Association of Canada, Toronto, Ontario, Special Publication vol. 48, pp.

793

135–174.

794

INGEMMET, 1980. Carta Geológica Nacional. San Borja, Lima.

795

Jaillard, E., Bengtson, P., Dhondt, A.V., 2005. Late Cretaceous marine transgressions in Ecuador and

796

northern Peru: A refined stratigraphic framework. Journal of South American Earth Sciences 19, 307–

797

323.

798

Jaillard, E., Hérail, G., Monfret, T., Diaz-Martinez, E., Baby, P., Lavenu, A., Dumont, J.F., 2000.

799

Tectonic Evolution of the Andes of Ecuador, Peru, Bolivia and Northernmost Chile.In : Tectonic

800

Evolution of South America, 31st International Congress, Geology and sustainable Development :

801

481-559.

802

Jaillard, E., Laubacher, G., Bengtson, P., Dhondt, A., Bulot, L.G., 1999. Stratigraphy and evolution of

803

the Cretaceous forearc Celica- Lancones basin of southwestern Ecuador. Journal of South American

804

Earth Sciences 12, 51–68.

AC C

EP

TE D

M AN U

SC

RI PT

781

33

ACCEPTED MANUSCRIPT Jaillard, E., Arnaud-Vanneau, A., 1993.The Cenomanian–Turonian transition on the Peruvian margin.

806

Cretaceous Research 14, 585–605.

807

Jaillard, E., Soler, P., Carlier, G., Mourier, T., 1990. Geodynamic evolution of the northern and central

808

Andes during early to middle Mesozoic times: A Tethyan model. Journal of the Geological Society,

809

London, 147, 1009–1022.

810

Jaillard, E., 1987. Sedimentary evolution of an active margin during middle and upper Cretaceous

811

times: the North Peruvian margin from Late Aptian up to Senonian. Geologische Rundschau 76, 677–

812

697.

813

Jaillard, E., 1986. La sédimentation crétacée dans les Andes du Pérou central: exemple de la

814

Formation Jumasha (Albien moyen-supérieur à Turonien supérieur) dans la région d’Oyón

815

(département de Lima). Géodynamique 1 (2), 97–108.

816

Jarvis, I., Gale, A.S., Jenkyns, H.C., Pearce, M.A., 2006. Secular variation in Late Cretaceous carbon

817

isotopes and sea-level change: evidence from a new δ13C carbonate reference curve for the

818

Cenomanian–Campanian, 99.6–70.6 Ma. Geological Magazine 143, 561–608.

819

Jarvis, I., Lignum, J.S., Gröcke, D.R., Jenkyns, H.C., Pearce, M.A., 2011. Black shale deposition,

820

atmospheric CO2 drawdown, and cooling during the Cenomanian–Turonian Oceanic Anoxic Event.

821

Paleoceanography 26, PA3201.

822

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

823

11, 1–30.

824

Jolet, P., Philip, J., Thomel, G., Lopez, G., Tronchetti, G., 1997. Comptes Rendus de l’Académie des

825

Sciences, Paris 325 (2), 703–709.

826

Larson, R.L., 1991. Geological consequences of superplumes. Geology 19, 963–966.

AC C

EP

TE D

M AN U

SC

RI PT

805

34

ACCEPTED MANUSCRIPT Larson, R.L., Erba, E., 1999. Onset of the mid-Cretaceous greenhouse in the Barremian-Aptian:

828

Igneous events and the biological, sedimentary and geochemical responses. Paleoceanography 14,

829

663-678.

830

Lechler, M., Pogge von Strandmann, P.A.E., Jenkyns, H.C., Prosser, G., Parente, M., 2015. Lithium-

831

isotope evidence for enhanced silicate weathering during OAE1a (Early Aptian Selli event). Earth and

832

Planetary Science Letters 432, 210–222.

833

Leckie, R.M., Bralower, T.J., Cashman, R., 2002. Oceanic anoxic events and plankton evolution:

834

Biotic response to tectonic forcing during the mid-Cretaceous. Paleoceanography 17(3), 1–29.

835

Leckie, R. M., 1989. An oceanographic model for the early evolutionary history of planktonic

836

foraminifera, Palaeogeography, Palaeoclimatology, Palaeoecology 73, 107– 138.

837

Leutenegger, S., 1984. Symbiosis in benthic foraminifera, specificity and host adaptation. The Journal

838

of Foraminiferal Research 14, 16‒35.

839

Lézin, C., Andreu, B., Ettachfini, El M., Wallez, M.-J., Lebedel, V., Meister, C., 2012. The upper

840

Cenomanian–lower Turonian of the Preafrican Trough, Morocco. Sedimentary Geology 245–246, 1–

841

16.

842

Lloyd, M.R., 1964. Variations in the oxygen isotope ratios of Florida Bay molluscs and their

843

environmental significance: Journal of Geology 72, 84–111.

844

Kidder, D.L., Worsley, T.R., 2010. Phanerozoic Large Igneous Provinces (LIPs), HEATT (Haline

845

Euxinic Acidic Thermal Transgression) episodes, and mass extinctions. Palaeogeography,

846

Palaeoclimatology, Palaeoecology 295, 162–191.

847

McAnena, A., Flögel, S., Hofmann, P., Herrle, J.O., Griesand, A., Pross, J., Talbot, H.M., Rethemeyer,

848

J., Wallmann, K., Wagner, T., 2013. Atlantic cooling associated with a marine biotic crisis during the

849

mid-Cretaceous period. Nature Geoscience 6, 558–561.

AC C

EP

TE D

M AN U

SC

RI PT

827

35

ACCEPTED MANUSCRIPT Manabe, S., Bryan, K., 1985. CO2-induced change in a coupled ocean-atmosphere model and its

851

paleoclimatic implications. Journal of Geophysical Research, DOI: 10.1029/JC090iC06p11689.

852

Masse, J.-P., Fenerci, M., Pernarcic, E., 2003. Palaeobathymetric reconstruction of peritidal

853

carbonates. The Late Barremian, Urgonian, sequences of Provence (SE France). Palaeogeography,

854

Palaeoclimatology, Palaeoecology 200, 65–81.

855

Maurer, F., van Buchem, F.S.P., Eberli, G.P., Pierson, B.J., Raven, M.J., Larsen, P.-H., Al-Husseini,

856

M.I., Vincent, B., 2013. Late Aptian long-lived glacio-eustatic lowstand recorded on the Arabian

857

Plate. Terra Nova 25, 87-94.

858

Méhay, S., Keller, C.E., Bernasconi, S.M., Weissert, H., Erba, E., Bottini, C., Hochuli, P.A., 2009. A

859

volcanic CO2 pulse triggered the Cretaceous Oceanic Anoxic Event 1a and a biocalcification crisis.

860

Geology 37, 819–822.

861

Melinte-Dobrinescu, M. C., Roban, R.-D., Stoica, M., 2015. Palaeoenvironmental changes across the

862

Albian–Cenomanian boundary interval of the Eastern Carpathians. Cretaceous Research 54, 68–85.

863

Menegatti, A.P., Weissert, H., Brown, R.S., Tyson, R.V., Farrimond, P., Strasser, A., Caron, M., 1998.

864

High resolution δ13C stratigraphy through the early Aptian “Livello Selli” of the Alpine Tethys.

865

Paleoceanography 13, 530-545.

866

Meyers, S.R., Siewert, S.E., Singer, B.S., Sageman, B.B., Condon, D.J., Obradovich, J.D., Jicha, B.R.,

867

Sawyer, D.A., 2012. Intercalibration of radioisotopic and astrochronologic time scales for the

868

Cenomanian–Turonian boundary interval, Western Interior Basin, USA. Geology 40, 7–10.

869

Moulin, M., Aslanian, D., Unternher, P., 2010. A new starting point for the South and Equatorial

870

Atlantic Ocean. Earth-Science Reviews 98, 1–37.

871

Navarro-Ramirez, J.P., Bodin, S., Heimhofer, U., Immenhauser, A., 2015. Record of Albian to early

872

Cenomanian environmental perturbation in the eastern sub-equatorial Pacific. Palaeogeography,

873

Palaeoclimatology, Palaeoecology 423, 122–137.

AC C

EP

TE D

M AN U

SC

RI PT

850

36

ACCEPTED MANUSCRIPT Navarro-Ramirez, J.P., Bodin, S., Immenhauser, A., 2016. Ongoing Cenomanian – Turonian

875

heterozoan carbonate production in the neritic settings of Peru. Sedimentary Geology 331, 78–93.

876

Ogg, J.G., Hinnov, L.A., 2012. Chapter 27, Cretaceous. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.D.,

877

Ogg, G.M. (Eds.). The Geological Time Scale. Elsevier, Amsterdam, pp. 793–853.

878

Owens, J.D., Gill, B.C., Jenkyns, H.C., Batesa, S.M., Severmann, S., Kuypers, M.M.M., Woodfine,

879

R.G., Lyons, T.W., 2013. Proceedings of the National Academy of Sciences of the United States of

880

America 110, 18407–18412.

881

Pagani, M., Huber, M., Sageman, B., 2014. Greenhouse climates. In: Holland, H., Turekian, K. (Eds.),

882

Treatise on Geochemistry 6, 281–304.

883

Paul, C.R.C., Lamolda, M.A., Mitchell, S.F., Vaziri, M.R., Gorostidi, A., Marshall, J.D., 1999. The

884

Cenomanian–Turonian boundary at Eastbourne (Sussex, UK): a proposed European reference section.

885

Palaeogeography, Palaeoclimatology, Palaeoecology 150, 83–122.

886

Parente, M., Frijia, G., Di Lucia, M., Jenkyns, H. C., Woodfine, R. G., Baroncini, F., 2008. Stepwise

887

extinction of larger foraminifera at the Cenomanian- Turonian boudary: a shallow-water perspective

888

on nutrient fluctuation during Oceanic Anoxic Event 2 (Bonarelli Event). Geology 36, 715‒718.

889

Pogge von Strandmann, P.A.E., Jenkyns, H.C., Woodfine, R.G., 2013. Lithium iso-tope evidence for

890

enhanced weathering during Oceanic Anoxic Event 2. Nature Geoscience 6, 668–672.

891

http://dx.doi.org/10.1038/ngeo1875.

892

Poulsen, C.J., Gendaszek, A.S., Jacob, R.L., 2003. Did the rifting of the Atlantic Ocean cause the

893

Cretaceous thermal maximum? Geology 31, 115–118.

894

Premoli Silva, I., Sliter, W.V., 1999. Cretaceous paleoceanography: Evidence from planktonic

895

foraminiferal evolution, in Evolution of the Cretaceous Ocean-Climate System, edited by E. Barrera

896

and C. C. Johnson, Geological Society of America Special Papers 332, 301– 328.

AC C

EP

TE D

M AN U

SC

RI PT

874

37

ACCEPTED MANUSCRIPT Robert, E., Latil, J.L., Bulot, L.G., 2009. Albian ammonite faunas from South America: the genus

898

Tegoceras Hyatt, 1903. Revue de Paléobiologie 28, 43–51.

899

Robert, E., Bulot, L.G., 2005. Albian ammonite faunas from Peru, the genus Neodeshayesites Casey,

900

1964. J. Paleont., 79, 611–618.

901

Robert, E., Bulot, L.G., 2004. Origin, phylogeny, faunal composition, and stratigraphical significance

902

of the Albian Engonoceratidae (Pulchelliaceae, Ammonitina) of Peru. Journal of South American

903

Earth Sciences 17, 11–23.

904

Robert, E., 2002. La transgression albienne dans le Bassin Andin (Pérou, Équateur). Paléontologie

905

(ammonites), biostratigraphie et séquences de dépôts. Strata, 38, 380 p.

906

Romani, M., 1982. Géologie de la région minière Uchucchatua-Hacienda Otuto. Pérou.Thèse 3e cycle,

907

Grenoble 176 p.

908

Sageman, B.B., Meyers, S.R., Arthur, M.A., 2006. Orbital time scale and new C-isotope record for

909

Cenomanian–Turonian boundary stratotype. Geology 34, 125–128.

910

Saltzman, M.R., 2003. The Late Paleozoic Ice Age: pCO2 or oceanic gateway? Geology 31, 151–154.

911

Sartoni, S, Crescenti, U., 1962. Ricerche biostratigrafiche nel Mesozoico dell'Appennino meridionale.

912

Giornale di Geologia 29, 161‒388.

913

Sch, P.A., Arthur, M.A., 1980. Carbon isotope fluctuations in Cretaceous pelagic limestones —

914

potential stratigraphic and petroleum-exploration tool. AAPG Bulletin- American Association of

915

Petroleum Geologists 64 (1), 67–87.

916

Schroeder, R., Neumann, M. (eds.), 1985. Les grands Foraminifères du Crétacé moyen de la région

917

méditerranéenne. Geobios, Memoire Special 7, 161 p.

918

Scott, R.W., Formolo, M., Rush, N., Owens, J.D., Oboh-Ikuenobe, F., 2013. Upper Albian OAE1d

919

event in the Chihuahua Trough, New Mexico, U.S.A. Cretaceous Research 46,136–150.

AC C

EP

TE D

M AN U

SC

RI PT

897

38

ACCEPTED MANUSCRIPT Schröder-Adams C.J., Herrle, J.O., Tu Q., 2012. Albian to Santonian carbon isotope excursions and

921

faunal extinctions in the Canadian Western Interior Sea: Recognition of eustatic sea-level controls on a

922

fore bulge setting. Sedimentary Geology 281, 50–58.

923

Sinninghe Damsté, J. S., van Bentum, E.C., Reichart, G.J., Pross, J., Schouten, S., 2010. A CO2

924

decrease‐driven cooling and increased latitudinal temperature gradient during the mid‐Cretaceous

925

Oceanic

926

doi:10.1016/j.epsl.2010.02.027.

927

Snow, L.J., Duncan, R.A., Bralower, T.J., 2005, Trace element abundances in the Rock Canyon

928

anticline, Pueblo, Colorado, marine sedimentary section and their relationship to Caribbean plateau

929

construction

930

doi:10.1029/2004PA001093.

931

Soler, P., Bonhomme, M.G., 1990. Relations of magmatic activity to plate dynamics in Central Peru

932

from late Cretaceous to present. Geological Society of America Special Paper 241, 173−192.

933

Stein, M., Arnaud-Vanneau, A., Adatte, T., Fleitmann, D., Spangenberg, J.E., Follmi, K.B., 2012.

934

Palaeoenvironmental and palaeoecological change on the northern Tethyan carbonate platform during

935

the Late Barremian to earliest Aptian. Sedimentology 59, 939–963.

936

Takashima, R., Nishi, H., Yamanaka, T., Tomosugi, T., Fernando, A.G., Tanabe, K., Moriya, K.,

937

Kawabe, F., Hayashi, K., 2011. Prevailing oxic environments in the Pacific Ocean during the mid-

938

Cretaceous oceanic anoxic event 2, Nature Communications 2, doi:10.1038/ncomms1233.

939

Tarduno, J.A., Sliter, W.V., Kroenke, L., Leckie, M., Mayer, H., Mahoney, J.J., Mus-grave, R.,

940

Storey, M., Winterer, E.L., 1991. Rapid formation of Ontong Java Plateau by Aptian Mantle Plume

941

Volcanism. Science 254, 399–403.

oxygen

Earth

anoxic

and

Planetary

event

2:

Science

Letters

SC

2.

Paleoceanography,

v.

293,

20,

97–103,

PA3005,

AC C

EP

TE D

and

Event

M AN U

Anoxic

RI PT

920

39

ACCEPTED MANUSCRIPT Tejada, M.L.G., Suzuki, K., Kuroda, J., Coccioni, R., Mahoney, J.J., Ohkouchi, N., Sakamoto, T.,

943

Tatsumi, Y., 2009. Ontong Java Plateau eruption as a trigger for the early Aptian oceanic anoxic

944

event. Geology 37, 855-858.

945

Toshimitsu, S., Kano, M., Tashiro, M., 1990. Oyster reefs from the Upper Himenoura Subgroup

946

(Upper Cretaceous), Kyushu, Japan. Kaseki (Fossils) 49, 1–12 (in Japanese with English Abstr.).

947

Tsikos, H., Jenkyns, H.C., Walsworth-Bell, B., Petrizzo, M.R., Forster, A., Kolonic, S., Erba, E.,

948

Premoli Silva, I., Baas, M., Wagner, T., Sinninghe Damsté, J.S., 2004. Carbon-isotope stratigraphy

949

recorded by the Cenomanian–Turonian oceanic anoxic event: correlation and implications based on

950

three key localities. Geological Society, London, Special Publication, 161, pp. 711–719.

951

Turgeon, S., Brumsack, H.J., 2006. Anoxic vs dysoxic events reflected in sediment geochemistry

952

during the Cenomanian–Turonian Boundary Event (Cretaceous) in the Umbria–Marche Basin of

953

central Italy. Chemical Geology 234, 321–339.

954

Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., Loutit, T.S.,

955

Hardenbol, J., 1988. An overview of the fundamentals of sequence stratigraphy and key definitions.

956

In: Wilgus, C., Hastings, B.S., Kendall, C.G., Posamentier, H.W., Ross, C.A., Van Wagoner, J.C.

957

(Eds.), Sea Level Changes: an Integrated Approach, SEPM Special Publication 42, pp. 39–46.

958

Vicedo, V., Calonge, A., Caus E., 2011. Cenomanian rhapydioninids (Foraminiferida): architecture of

959

the shell and stratigraphy. Journal of Foraminiferal Research 41, 41–52.

960

Van Helmond, N.A.G.M., Sluijs, A., Reichart, G.J., Sinninghe-Damsté, J.S., Slomp, C.P., Brinkhuis,

961

H., 2013. A perturbed hydrological cycle during Oceanic Anoxic Event 2. Geology 42, 123–126.

962

Von Hillebrandt, A., 1970. Die Kreide in der Zentral-Kordillere östlich von Lima (Peru, Südamerika).

963

Geo/Rundsch 59/3, 1180–1203.

964

Wagner, T., Pletsch, T., 1999, Tectono-sedimentary controls on Cretaceous black shale deposition

965

along the opening of the Equatorial Atlantic Gateway (ODP Leg 159), in Cameron, N.R., et al., eds.,

AC C

EP

TE D

M AN U

SC

RI PT

942

40

ACCEPTED MANUSCRIPT The oil and gas habitats of the South Atlantic: Geological Society [London] Special Publication 153,

967

p. 241–265.

968

Weissert, H., Lini, A., Föllmi, K.B., Kuhn, O., 1998. Correlation of Early Cretaceous carbon isotope

969

stratigraphy and platform drowning events: a possible link? Palaeogeography, Palaeoclimatology,

970

Palaeoecology, 137, 189–203.

971

Weissert, H., 1990. Siliciclastics in the Early Cretaceous Tethys and North Atlantic oceans:

972

Documents of periodic greenhouse climate conditions. Memorie - Società Geologica Italiana 44, 59–

973

69.

974

Weissert, H., Lini, A., 1991. Ice age interludes during the time of Cretaceous greenhouse climate, in:

975

Müller, D.W., McKenzie, J.A., Weissert, H., (Eds.), Controversies in modern geology, Academic

976

Press, London, pp. 173–191.

977

Weissert, H., 1989. C-isotope stratigraphy, a monitor of paleoenvironmental change: a case study from

978

the early Cretaceous. Surveys in Geophysics 10, 1–61.

979

Westermann, S., Föllmi, K.B., Adatte, T., Matera, V., Schnyder, J., Fleitmann, D., Fiet, N., Ploch, I.,

980

Duchamp-Alphonse, S., 2010. The Valanginian δ13C excursion may not be an expression of a global

981

oceanic anoxic event. Earth and Planetary Science Letters 290, 118-131.

982

Wilson, J.J., 1963. Cretaceous stratigraphy of Central Andes of Peru. - American Association of

983

Petroleum Geologists Bulletin 47, 1-34.

984

Wilson, P.A., Norris, R.D., 2001. Warm tropical ocean surface and global anoxia during the mid-

985

Cretaceous period. Nature, 412, 425–429.

986

Winter, L.S., Tosdal, R.M., Mortensen, J.K., Franklin, J.M., 2010. Volcanic stratigraphy and

987

geochronology of the Cretaceous Lancones Basin, Northwestern Peru: position and timing of giant

988

VMS deposits. Economic Geology 105, 713–743.

AC C

EP

TE D

M AN U

SC

RI PT

966

41

ACCEPTED MANUSCRIPT Wortmann, U.G., Herrle, J.O., Weissert, H. 2004. Altered carbon cycling and coupled changes in

990

Early Cretaceous weathering pattern: evidence from integrated Tethyan isotope and sandstone records.

991

Earth and Planetary Science Letters, 220, 69–82.

992

Zohary, T., Reiss, Z., Hottinger, L. 1980. Population dynamics of Amphisorus hemprichii

993

(Foraminifera) in the Gulf of Elat (Aqaba), Red Sea. Eclogae geologicae Helvetiae 73, 1071‒1094.

RI PT

989

994 995

Figure Captions

998 999 1000

Figure 1: Palaeogeographic reconstruction for the mid-Cretaceous (modified after Blakey, 2011)

M AN U

997

SC

996

indicating the position of what today is the Western Platform of South America (yellow arrow). Figure 2: Geological map of the Oyon region (modified after INGEMMET, 1980) showing the location of the Uchucchacua and Lauricocha sections documented in this study. Figure 3: Regional correlation of the sequence stratigraphic interpretation between the Cajamarca

1002

composite section (northern Peru; Navarro-Ramirez et al., 2016) and the Uchucchacua section (central

1003

Peru; this study). The Uchucchacua section is juxtaposed with our sections (Jumasha 1 and 4) and by

1004

the Jaillard (1986)´s work. Main biostratigraphic data are indicated (Benavides-Caceres, 1956; Wilson

1005

1963; von Hillebrandt, 1970; Romani, 1982; Jaillard 1986, 1987; Jaillard and Arnaud-Vanneau, 1993;

1006

Robert 2002; Robert and Bulot, 2004, 2005; Dhondt and Jaillard, 2005; Jaillard et al., 2005; Robert et

1007

al., 2009).

AC C

EP

TE D

1001

1008

Figure 4: Legend for figs. 3, 9 and 10, denoting colour codes for different facies types.

1009

Figure 5: Landscape images showing A): The Jumasha Formation of the upper Albian–lower

1010

Turonian at Uchucchacua, Oyon region, Central Peru, built by massive, thickly-bedded, light-grey

1011

limestones. B) Lower part of the Jumasha Formation at Uchucchacua, including the upper Albian and

1012

the lower Cenomanian rocks. C) The upper part of the Jumasha Formation at Uchucchacua, including 42

ACCEPTED MANUSCRIPT 1013

the middle Cenomanian to lower Turonian rocks. D) The upper part of the Jumasha Formation at

1014

Lauricocha, including the late Cenomanian to lower Turonian rocks. Figure 6: Field images taken at the Uchucchacua section. A) Thin lamination of facies 1a,

1016

represented by grey and argillaceous limestones intercalations. B) Nodular limestones of the facies 2e.

1017

C) Thickly-bedded limestones of facies 2g. D) Thickly-bedded limestones of medium-scale sequence

1018

10 (facies 2b). C) Floatstones of facies 3d. F) Bioturbation at the level of regressive deposits in the top

1019

of Jumasha 4.

RI PT

1015

Figure 7A): Thin section images documenting grainstones of facies 2d, represented by peloids,

1021

miliolids, and isolated specimen of Perouvianella peruviana. B) Dasycladales-bearing grainstones of

1022

facies 2f. C) Miliolids and Perouvianella peruviana of facies 2g. D) Grain-rudstones dominated by

1023

Peruvianella peruviana in facies 2h. E) Diverse fauna floatstones of facies 3b, characterized by a

1024

chaotic fabric. F) Packstone rich in echinoids of facies 3c.

M AN U

SC

1020

Figure 8: Upper Cenomanian‒lower Turonian microfossil from shallow-water deposits of the

1026

Jumasha Formation in central Peru. Scale bar in images C and D is 0.5 mm, in all other images the

1027

scale bar is 1 mm. A) and B) show dasycladales, closely related to the genus Trinocladus. A.

1028

Transversal section. B. Longitudinal section (central part) and transversal section (upper right part). C)

1029

and D) Rotorbinella mesogeensis; C. axial section, note the presence of central plug. D. equatorial

1030

section. E) Agglutinated foraminifera in equatorial section. F) Porcelaneous foraminifera in equatorial

1031

section. G) Agglutinated foraminifera with spiral outline closely related to the genus Merlingina

1032

cretacea Hamaoui and Saint Marc, 1965. H) Perouvianella peroviana in perfect equatorial section. I)

1033

Perouvianella peroviana proloculi of two specimens, difference in diameter suggest the existence of

1034

multiple cycles of reproduction within the population.

AC C

EP

TE D

1025

1035

Figure 9: Sequence stratigraphic interpretation and carbon-isotope stratigraphy of the

1036

Uchucchacua section for the upper Albian (Jumasha 1). Correlation with the Tethyan composite curve

1037

of Herrle et al. (2015) is shown. Biostratigraphic data for the late Albian have been obtained from

1038

Jaillard (1986, 1987 and references therein). 43

ACCEPTED MANUSCRIPT Figure 10: Regional sequence stratigraphic interpretation and carbon-isotope stratigraphy of the

1040

Uchucchacua and Lauricocha sections for the late Cenomanian (Jumasha 4). Data points shaded grey

1041

correspond to OAE2 based on bio- and chemostratigraphic data. Biostratigraphic data of the

1042

Uchucchacua section has been taken from Jaillard (1986) and Jaillard and Arnaud-Vanneau (1993).

1043

The chemostratigraphic markers of OAE2 (A‒B) include Navarro-Ramirez et al. (2016), Eastbourne

1044

(Jarvis et al., 2011), and the Portland # 1 core (Sageman et al., 2006).

RI PT

1039

Figure 11A): Palaeogeographic interpretation of the western South America platform during the

1046

late Albian, characterized by oyster-rich mixed siliciclastic-carbonate deposition. Note regional oyster

1047

facies deposition, diachronic progradation of the NE deltaic system fed from the Brazilian continental

1048

basement, as well as the Huarmey Trough isolating the western Platform from the Pacific Ocean. In

1049

contrast to the central Peru, northern Peru was localized more proximal to the southern margin of the

1050

proto-Atlantic, resulting in a potentially stratified water column setting. B) During the late

1051

Cenomanian, the Western Platform was characterized by two distinctive palaeogeographic settings: (i)

1052

northern Peru characterized by continental deltaic influx from the emerged shield of Brazil in the east;

1053

(ii) central Peru representing an oligotrophic environment, leading to the mass occurrence of

1054

Perouvianella peruviana and associated miliolids. The four main depositional environments, each

1055

with a number of standard facies types (1a through 4a), are established for the inner to outer ramp.

1056

Interpretations are based on Benavides-Caceres (1956), Jaillard (1986, 1987), Atherton and Webb

1057

(1989), Soler and Bonhomme (1990), Jaillard et al. (2000), Callot et al. (2008), Robert et al. (2009)

1058

and Winter et al. (2010).

AC C

EP

TE D

M AN U

SC

1045

1059

Figure 12: Regional correlation of the carbon-isotope stratigraphy between the Cajamarca

1060

composite section (northern Peru; Navarro-Ramirez et al., 2016) and by the Uchucchacua section

1061

(central Peru), exhibiting the Jumasha 1 and 4 carbon-isotope data. Main OAEs and biostratigraphic

1062

data are indicated (Benavides-Caceres, 1956; Jaillard, 1986, 1987; Jaillard and Arnaud-Vanneau,

1063

1993; Dhondt and Jaillard, 2005; Robert et al., 2009). The chemostratigraphic terminology for OAE2

1064

interval (onset, trough, plateau, and recovery, c.f., Paul et al., 1999; Tsikos et al., 2004), and datum 44

ACCEPTED MANUSCRIPT 1065

levels A, B, C and D as is used in Tethyan and North Atlantic sections (Meyers et al., 2012; Du Vivier

1066

et al., 2014, 2015). The red data points in the Cajamarca composite section form part of this study. Table 1: Overview of facies classification and interpretation. Numbers indicate the relative

1068

abundance of non-skeletal and skeletal components: 0 = absent, 1 = present, 2 = frequent, 3 =

1069

abundant, 4 = dominant.

RI PT

1067

AC C

EP

TE D

M AN U

SC

1070

45

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT

New Albian–Cenomanian and Cenomanian–Turonian carbon isotope curves based on epeiric-neritic successions from Central Peru Evidence for the impact of transient carbon cycle perturbations (OAE1d and OAE2) is provided

AC C

EP

TE D

M AN U

SC

RI PT

Clear evidence for a carbonate neritic crisis associated with OAE 1d and OAE 2 is lacking