Deep-burial hydrothermal alteration of the Pre-Salt carbonate reservoirs from northern Campos Basin, offshore Brazil: Evidence from petrography, fluid inclusions, Sr, C and O isotopes

Deep-burial hydrothermal alteration of the Pre-Salt carbonate reservoirs from northern Campos Basin, offshore Brazil: Evidence from petrography, fluid inclusions, Sr, C and O isotopes

Journal Pre-proof Deep-burial hydrothermal alteration of the Pre-Salt carbonate reservoirs from northern Campos Basin, offshore Brazil: Evidence from ...

56MB Sizes 0 Downloads 44 Views

Journal Pre-proof Deep-burial hydrothermal alteration of the Pre-Salt carbonate reservoirs from northern Campos Basin, offshore Brazil: Evidence from petrography, fluid inclusions, Sr, C and O isotopes Bruno Eustáquio Moreira Lima, Leonardo Ribeiro Tedeschi, André Luiz Silva Pestilho, Roberto Ventura Santos, Joselito Cabral Vazquez, Jarbas Vicente Poley Guzzo, Luiz Fernando De Ros PII:

S0264-8172(19)30595-1

DOI:

https://doi.org/10.1016/j.marpetgeo.2019.104143

Reference:

JMPG 104143

To appear in:

Marine and Petroleum Geology

Received Date: 3 July 2019 Revised Date:

18 November 2019

Accepted Date: 18 November 2019

Please cite this article as: Lima, Bruno.Eustá.Moreira., Tedeschi, L.R., Pestilho, André.Luiz.Silva., Santos, R.V., Vazquez, J.C., Guzzo, J.V.P., De Ros, L.F., Deep-burial hydrothermal alteration of the Pre-Salt carbonate reservoirs from northern Campos Basin, offshore Brazil: Evidence from petrography, fluid inclusions, Sr, C and O isotopes, Marine and Petroleum Geology (2019), doi: https:// doi.org/10.1016/j.marpetgeo.2019.104143. This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. 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. © 2019 Published by Elsevier Ltd.

1

1

Deep-burial hydrothermal alteration of the Pre-Salt carbonate

2

reservoirs from northern Campos Basin, offshore Brazil: Evidence

3

from petrography, fluid inclusions, Sr, C and O isotopes

4 a, b

, Leonardo Ribeiro Tedeschi c, André Luiz Silva

5

Bruno Eustáquio Moreira Lima

6

Pestilho c, Roberto Ventura Santos d, Joselito Cabral Vazquez c, Jarbas Vicente Poley

7

Guzzo c, Luiz Fernando De Ros b

8 9

a

Petrobras S.A., Avenida República do Chile, 330, East Tower, Rio de Janeiro, RJ,

10

20031-170, Brazil; [email protected]

11

b

12

Gonçalves, 9500, Porto Alegre, RS, 91509-900, Brazil; [email protected]

13

c

14

Universitária, Rio de Janeiro, RJ, 21941-915, Brazil;

15

[email protected]; [email protected];

16

[email protected]; [email protected]

17

d

18

Central, Brasília, DF, 70910-900, Brazil; [email protected]

19 20 21

Graduate Program in Geosciences, Rio Grande do Sul Federal University, Av. Bento

Petrobras, Research Center (CENPES), Avenida Horácio Macedo, 950, Cidade

Geosciences Institute, University of Brasília, Campus Universitário Darcy Ribeiro, Ala

2 22

ABSTRACT

23

Petrographic, mineralogical, elemental, isotopic and fluid inclusion analyses were

24

integrated to unravel the diagenetic evolution of Brazilian Pre-Salt lacustrine carbonate

25

reservoirs of northern Campos Basin, southeast Brazilian margin. Detailed thin section

26

and cathodoluminescence petrography, scanning electron microscopy and electron

27

microprobe analyses established a paragenetic evolution of diagenetic processes and

28

products, comprising extensive dolomitization, silicification, and dissolution. A

29

paragenesis including saddle dolomite, macrocrystalline calcite, mega-quartz, Sr-barite,

30

celestine, fluorite, dickite, sphalerite, galena, and other metallic sulfides filling fractures

31

and dissolution porosity, and aqueous fluid inclusions with homogenization

32

temperatures of 92 to 152 °C and salinities between 13 to 26 wt. % eq. NaCl

33

characterized a hydrothermal system with some analogy to carbonate-hosted Pb-Zn

34

Mississippi Valley (MVT) and Irish-type deposits. Petroleum inclusions and solid

35

bitumen testify atypical oil generation and migration, associated with the hydrothermal

36

flow. The host Pre-Salt spherulitic and fascicular carbonates present highly radiogenic

37

87

38

Hydrothermal phases show δ18O values more negative than syngenetic and diagenetic

39

carbonates. The δ13C values are interpreted as result of interaction between the

40

hydrothermal fluids and the host rocks. The combined data set provides clear evidence

41

of intense hydrothermal alteration of northern Campos Basin Pre-Salt reservoirs at

42

deep-burial conditions (> 2 km), possibly related to Late Cretaceous or more probably

43

Paleogene magmatic activity. Mixed-sourced fluids bearing a basinal signature fed the

44

hydrothermal system and promoted dissolution of the host rocks. The hydrothermal

45

alterations had strong impact on the porosity, permeability, and heterogeneity,

Sr/86Sr ratios, indicating strong interaction with continental crust materials.

3 46

contributing, together with the associated fracturing, to the excellent production

47

performance of the Pre-Salt reservoirs.

48

Keywords Pre-Salt; lacustrine carbonates; Campos Basin; deep-burial; hydrothermal

49

alteration; magmatic activity

50

1. Introduction

51

Carbonate reservoirs usually have complex and heterogeneous diagenetic

52

evolution as result of their reactive character. Common association with faults and

53

fracture systems often results in intense dolomitization, silicification and/or dissolution

54

(e.g., Gregg and Sibley, 1987; Machel and Mountjoy, 1987; Packard et al., 2001;

55

Corbella et al., 2006; Wilson et al., 2007; Callot et al., 2010; Fontana et al., 2010).

56

Hydrothermal alteration of carbonate rocks has been described in diverse basins, and its

57

characterization is important, both to differentiate from the “conventional” diagenetic

58

evolution of carbonates, and because hydrothermal processes substantially impact,

59

positively or negatively, the quality and performance of carbonate reservoirs (e.g.,

60

Machel and Lonnee, 2002; Davies, 2002, 2004; Machel, 2004; Davies and Smith, 2006;

61

Lonnee and Machel, 2006; Biehl et al., 2016).

62

The hydrothermal

alteration of carbonate deposits involves complex

63

physicochemical processes of interaction with hot fluids, promoting the precipitation of

64

mineral assemblages commonly including saddle dolomite, fluorite, barite, anhydrite,

65

dickite, sphalerite, and pyrite (Neilson and Oxtoby, 2008). These processes indicate the

66

existence of geothermal anomalies, as well as permeable pathways, such as deep fault

67

systems, to allow the percolation and interaction of fluids (Davies and Smith, 2006; Xu

68

et al., 2015; Mansurbeg et al., 2016).

4 69

The Lower Cretaceous Pre-Salt carbonate deposits of the Coqueiros and Macabu

70

formations of northern Campos Basin host important hydrocarbon reservoirs (Dias et al.,

71

1988; Winter et al., 2007). In Pre-Salt deposits there are significant variations in their

72

quality (porosity and permeability) due to the complex diagenesis (Herlinger Jr. et al.,

73

2017; Lima and De Ros, 2019). Recently, evidence on hydrothermal alteration of Pre-

74

Salt deposits has been presented for the Campos Basin (Alvarenga et al., 2016; Vieira

75

de Luca et al., 2017; Lepley et al., 2017; Herlinger Jr. et al., 2017; Lima and De Ros,

76

2019) and for Kwanza Basin, offshore Angola, its western African counterpart (Poros et

77

al., 2017; Girard and San Miguel, 2017; Teboul et al., 2017, 2019). Alvarenga et al.

78

(2016) identified hydrothermal vents in seismic sections of central Campos Basin.

79

Vieira de Luca et al. (2017), Poros et al. (2017), Girard and San Miguel (2017) and

80

Teboul et al. (2019) suggested that hydrothermal fluids ascribed pervasive silicification

81

of Pre-Salt carbonates based on petrographic, isotopic data and fluid inclusions.

82

Herlinger Jr. et al. (2017) and Lima and De Ros (2019) have shown evidence of

83

hydrothermal alteration in Pre-Salt reservoirs from northern Campos Basin, based on

84

the description of hydrothermal mineral phases in thin sections, such as coarse quartz,

85

saddle dolomite, sulfides and sulfates.

86

Although evidence of hydrothermal action in Campos Basin have been discussed

87

by Vieira de Luca et al. (2017), Herlinger Jr. et al. (2017) and Lima and De Ros (2019),

88

a detailed study of hydrothermal mineral phases is needed to improve the

89

characterization of the hydrothermal processes in this and other Pre-Salt areas, and has

90

not been published yet. Particularly, it is important to examine the fracture-related

91

dolomitization, silicification, dissolution, and the precipitation of typical hydrothermal

92

mineral phases, such as saddle dolomite, macrocrystalline calcite, mega-quartz, sulfides

93

and sulfates filling secondary porosity. With this aim, an integrated study was

5 94

performed on four cored wells, comprising detailed petrographic characterization

95

performed through microprobe, electron and optical microscopy, temperature and

96

salinity data from fluid inclusion analyses in diagenetic carbonates, quartz and Sr-barite,

97

and C, O and Sr isotopic analyses in carbonates and Sr-barite. Understanding the

98

processes and patterns of hydrothermal alteration of Pre-Salt deposits is of key

99

importance for reservoir quality assessment during exploration and for the production

100

101

optimization of these extraordinary reservoirs.

2. Geological setting

102

The Campos Basin is situated in southeastern Brazilian margin and covers an

103

area of approximately 100,000 km2, with only 5,800 km2 onshore. The study area is

104

located in the offshore part of the northern Campos Basin (Fig. 1). The stratigraphy of

105

the study area is based on the stratigraphic chart from Winter et al. (2007) (Fig. 2). The

106

reservoirs being addressed here are part of the Coqueiros (rift stage) and Macabu (sag

107

phase) formations from of the Pre-Salt interval of the basin.

108

The rift sequence lies on a basement of the Neoproterozoic Ribeira Fold Belt,

109

which includes a variety of metamorphic and magmatic rocks, such as tonalities,

110

granodiorites, granites and gabbros (Winter et al., 2007; Tupinambá et al., 2012). The

111

initial rift evolution of the basin is represented by Hauterivian volcanics from the

112

Cabiúnas Formation and by Barremian to Lower Aptian sediments of the Itabapoana,

113

Atafona, and Coqueiros formations (basal interval of the Lagoa Feia Group; Fig. 2). The

114

Itabapoana and Atafona formations consist of conglomerates, sandstones, siltstones,

115

arenites of stevensitic ooids and shales (Armelenti et al., 2016). The Coqueiros

116

Formation is composed by bioclastic rudstones and grainstones intercalated with

6 117

dolostones and organic-rich shales (Baumgarten et al., 1988; Dias et al., 1988; Castro,

118

2006; Thompson et al., 2015).

119

The post-rift is defined by the upper interval of the Lagoa Feia Group, during

120

which evolved the so-called sag stage of the basin. In addition to Itabapoana Formation,

121

this sequence comprises the Gargaú, Macabu, and Retiro formations that were deposited

122

during the Middle/Upper Aptian (Fig. 2). The Macabu Formation comprises

123

argillaceous and carbonate laminites as well as spherulitic and fibrous carbonate crusts.

124

These later strata are interpreted as chemical precipitates deposited in alkaline lacustrine

125

environments under arid climate conditions (Wright, 2011, 2012; Tosca and Wright,

126

2014; Wright and Barnett, 2014, 2015; Wright and Tosca, 2016; Herlinger Jr. et al.,

127

2017; Lima and De Ros, 2019).

128

The Retiro Formation (Fig. 2), a thick accumulation of evaporites, was deposited

129

during the late Aptian (Late Alagoas local stage) marine incursion, under arid climate

130

conditions (Leyden et al., 1976; Winter et al., 2007), being composed essentially of

131

anhydrite, halite, and also by bittern salts, such as sylvite and carnallite (Rodriguez et al.,

132

2018). These evaporites provide the stratigraphic seal for the large Pre-Salt hydrocarbon

133

accumulations, as well as for the migration of hydrothermal fluids through deep faults

134

(Lima and De Ros, 2019). Above the evaporites, a carbonate/clastic sequence was

135

deposited from Albian to Recent, which are out of the scope of the present study.

136

Magmatic activity has been recorded in different stages of Campos Basin

137

evolution. Winter et al. (2007) described important extrusive (basalts and hyaloclastites)

138

and intrusive (diabase and gabbro) magmatic events with Ar/Ar ages from 81.5 to 83.2

139

Ma. In the Paleogene, volcanic activity is represented by basaltic lava flows, dated 65.5

140

Ma (close to the Cretaceous-Paleogene limit), 62 Ma (Danian, Paleocene), 53 Ma

7 141

(Ypresian, Eocene), and alkaline basalts, diabase and volcanic tuffs at approximately 43

142

Ma (Lutetian, Eocene).

143

3. Analytical methods

144

More than 300 core plugs and sidewall core samples from four wells were taken

145

from pervasively dolomitized and silicified intervals of the Campos Basin Pre-Salt

146

rocks (Fig. 1). Oil and salt were removed from all samples, using toluene and methanol

147

in a Soxhlet apparatus. Standard-thickness thin sections were prepared from blue epoxy

148

resin-impregnated samples for transmitted light microscopy. All thin sections were

149

examined in petrographic microscopes under uncrossed (//P) and crossed polarizers

150

(XP).

151

The paragenetic evolution of diagenetic mineral phases was examined in

152

approximately forty polished thin sections using a combination of cathodoluminescence,

153

scanning electron microscopy and microprobe analyses. Cathodoluminescence (CL)

154

was performed with a CITL Mk5-2 equipment on a Zeiss Axiocam MRC microscope.

155

Scanning electron microscopy images with secondary electrons (SEM) and

156

backscattered electrons (BSE) were obtained with a JSM-IT300 JEOL electron

157

microscope. Mineral chemical composition was verified with the support of

158

wavelength-dispersive X-ray spectroscopy (WDS) in a JEOL JXA-8230 electron

159

microprobe, and energy-dispersive spectrometry (EDS) in an OXFORD X-MaxN

160

equipment. Mineralogical mapping was also performed by QEMSCAN 650 (FEI)

161

equipment. These analyses were performed at the Petrobras Research Center (CENPES)

162

and at the Regional Center for Technological Development and Innovation (CRTI) of

163

Goiás Federal University (UFG).

8 164

Fluid inclusion (FI) petrography and microthermometry was performed on eight

165

double-polished thin sections (120 µm-thick) prepared from sidewall core samples

166

following standard procedures (Goldstein and Reynolds, 1994). These sections were

167

selected to represent the lithologies and mineral phases of the rift and sag sections (red

168

squares in Fig. 3). FIs were analyzed in macrocrystalline quartz (MQ), Sr-barite (SB),

169

saddle dolomite (SD), blocky calcite in recrystallized bioclasts (MBCB), and

170

macrocrystalline (MC), fascicular (FC) and spherulitic (CS) calcite as described in the

171

results section. A total of 203 determinations of homogenization temperature (Th) and

172

ice melting temperature (Tm ICE) of fluid inclusions hosted in syngenetic, diagenetic and

173

hydrothermal mineral phases were obtained using a Fluid Inc. USGS gas flow and a

174

Linkam THMSG600 heating and freezing stages (Table 1), under plane polarized light

175

and ultraviolet (UV; 380 nm) illumination.

176

FIs where described using the approach of Roedder (1984), where FIs where

177

classified according to its genetic type (primary, secondary, and pseudo-secondary) and

178

tied-in to the paragenetic sequence. FI types were characterized according to their

179

location, relationship to the host mineral, and consistency of visual parameters (e.g.,

180

apparent liquid/vapor ratio). Stage calibration was carried out using H2O-CO2 and pure

181

H2O-critical density synthetic inclusions. Homogenization and melting temperatures

182

were measured with a precision of ±0.1 °C. Minimum salinities were estimated using

183

the ice melting temperatures computed from NaCl-H2O system (Bodnar and Vityk,

184

1994). Salinities of inclusions displaying CaCl2-NaCl-H2O eutectic temperatures (-

185

52 °C) were estimated by combining of hydrohalite and ice melting temperatures

186

(Steele-MacInnis et al., 2011). For petroleum inclusions, API gravity was estimated

187

with a microthermometric technique, based on the oil fluorescence results (Goldstein

188

and Reynolds, 1994).

9 189

A total of 69 carbon (δ13C) and oxygen (δ18O) isotopes analyses were performed

190

on 51 samples derived from core plugs and sidewall cores of three wells (A, B and C)

191

(Table

192

microcrystalline/blocky calcite from bioclasts (MBCB) and laminites (MBCL),

193

microcrystalline/blocky dolomite from dolostones (MBDD), saddle dolomite (SD), and

194

macrocrystalline (MC), fascicular (FC) and spherulitic (CS) calcite. Forty-eight bulk

195

rock analyses were obtained from samples with more than 90% of a single carbonate

196

phase, using an IRMS Thermo Scientific Delta XL Advantage coupled to a Gasbench II

197

device in the University of São Paulo (USP) Lab. In addition, 21 samples (~100 mg) of

198

selected carbonate phases were obtained for carbon and oxygen isotopes from eight

199

polished thin sections (120 µm-thick), using a computer-monitored New Wave

200

Research MicroMill TM equipment, and a Thermo Scientific Delta V Isotope Ratio

201

Mass Spectrometer (IRMS), coupled to a Gasbench II unit in CENPES. All data were

202

reported relative to the Vienna Pee Dee Belemnite (V-PDB) international standard in

203

per mil units (‰). External precision and accuracy were checked in relation to the

204

reference material NBS 19 (TS-Limestone; Coplen et al., 2006), yielding results better

205

than ±0.09‰ for δ13C and ±0.10‰ for δ18O for both laboratories.

2;

blue

circles

The analyses of

206

87

in

Fig.

3).

The

analyzed

mineral

phases

were

Sr/86Sr ratios of thirty samples were obtained at the

207

Geochronology Lab of the University of Brasília (UnB). Roughly 1 mg of carbonate

208

powder was placed into SavillexTM (PFA tubes) and digested using an 0.5 mol/L acetic

209

acid solution in order to avoid leaching of non-carbonate phases. After centrifugation

210

the supernatant was dried in a hot plate and dissolved in nitric acid (2.9 mol/L). Sr was

211

separated from Rb in Teflon chromatographic columns filled with SrSpec (100-150 µm)

212

resin.

213

mass spectrometer in static multicollector (TIMS). The Sr ratios were normalized

87

Sr/86Sr ratios were determined using a Thermo Triton Plus thermal ionization

10 214

assuming a 88Sr/86Sr ratio of 0.1194. External reproducibility of 87Sr/86Sr ratios from the

215

NBS-987 standard yielded 0.710271 ± 0.000017 (n=26). Sr isotope analyses were

216

normalized to NBS-987 value of 0.710250.

217

4. Results

218

4.1. Petrography

219

Bioclastic grainstones and rudstones, stevensitic ooidal arenites (stevensite is a

220

smectite group clay mineral), and dolostones are the main lithologic types in the rift

221

interval (Herlinger Jr. et al., 2017; Lima and De Ros, 2019; Fig. 4A and B). Bivalves,

222

ostracods, and gastropods are the main bioclasts in the grainstones and rudstones.

223

Common diagenetic processes in these rocks include bioclast micritization,

224

neomorphism, recrystallization, replacement and dissolution. Most of the bivalves occur

225

recrystallized, and/or dissolved and filled by diagenetic cements. Cementation of inter-

226

and intraparticle, intercrystalline, moldic, vugular, channel and fracture porosity types

227

occurs mainly by drusy/rim pore-lining calcite, blocky and saddle dolomite, fibrous

228

chalcedony, microcrystalline, prismatic to coarse mosaic quartz, radiated and prismatic

229

celestine and Sr-barite, pyrite, svanbergite and bitumen. The bioclasts and the

230

stevensitic ooids are heterogeneously replaced by dolomite, calcite, microcrystalline

231

quartz, fibrous chalcedony, sulfides and svanbergite.

232

Lithologies described in the sag section include fascicular calcite crusts,

233

stevensitic claystones with calcite spherulites, intraclastic rudstones and grainstones,

234

laminites, dolostones, and cherts (Herlinger Jr. et al., 2017; Lima and De Ros, 2019; Fig.

235

4C to 4F). As for the rift bioclasts, most spherulitic and fascicular calcites are

236

recrystallized and/or replaced by silica or dolomite. The cherts consist of micro- and

237

macrocrystalline quartz, and spheroidal, fibrous chalcedony replacing syngenetic and

11 238

diagenetic constituents and filling pores. Rift lithologic types present a diagenetic

239

sequence similar to that of sag rocks. Therefore, we represent all diagenetic sequence by

240

using Figure 5, which shows the paragenetic sequence interpreted for the sag fascicular

241

calcite crusts, including 28 syngenetic, eodiagenetic, mesodiagenetic and hydrothermal

242

processes and products, schematically represented in Figure 6 (for more details in Lima

243

and De Ros, 2019).

244

Deposits of stevensite and other Mg silicates (event 1 in Fig. 5) occur throughout

245

the whole Aptian Pre-Salt succession, forming the substrate to the precipitation of

246

fascicular calcite crusts and spherulites (events 2 and 3 in Fig. 5) in a hyper-alkaline sag

247

lacustrine environment (Fig. 6A). Fascicular calcite crusts occur intercalated with

248

discontinuous levels of clay laminations, clay peloids and ooids of and siliciclastic

249

grains, such as muscovite and quartz, which were partially to totally replaced by

250

dolomite, calcite, microquartz, chalcedony and pyrite. Calcite spherulites were formed

251

as concretions, replacing and/or displacing these fine-grained deposits. Transition forms

252

between spherulitic and fascicular aggregates are common, including asymmetrical

253

spherulites, vertically elongated spherulites with lobate borders, hemi-spherulites, and

254

single or multiple fascicular aggregates nucleated on spherulites.

255

Sag lithologic types often occur partially to fully recrystallized, dolomitized,

256

silicified, cemented and/or dissolved. The Mg clay deposits occur frequently dissolved,

257

and/or replaced by silica, calcite, dolomite, pyrite or magnesite (events 4–7 in Fig. 5;

258

Fig. 6B). Lamellar magnesite replaced pseudomorphically the clay laminations in

259

subordinate amounts (event 6 in Fig. 5). Blocky dolomite, microcrystalline silica, quartz

260

and chalcedony commonly replaced the Mg clay matrix, calcite spherulites and

261

fascicular aggregates, and fill primary and secondary pores (events 8–10 in Fig. 5; Fig.

262

6B, C). Recrystallization, fracturing and pressure dissolution of fascicular and

12 263

spherulitic calcite was heterogeneous (events 11–15 in Fig. 5; Fig. 6C, D). Stylolites

264

occur

265

concentrations.

frequently associated

with

dolomitized

intervals

and

residual

pyrite

266

In some dolomitized or silicified samples, recognizing the original carbonate

267

fabrics modified by these processes is difficult, if not impossible. Dolomitization and

268

silicification are pervasive as disseminated, replacive, and filling cavities and fractures

269

associated with hydrothermal features and minerals (Figs. 7, 8). The common sag

270

replacive and pore-filling phases are cryptocrystalline to microcrystalline silica (events

271

10, 17 and 21 in Fig. 5; Fig. 7C, E); fibrous, spherulitic chalcedony (events 10 and 17 in

272

Fig. 5; Fig. 7D, E); mosaic, bladed and prismatic rims to drusiform calcite (events 12,

273

16 and 23 in Fig. 5; Figs. 7B, 8A, B, C, E, F); blocky, rims, mosaic, and saddle

274

dolomite (events 17 and 20 in Fig. 5; Figs. 7A, B, C, 8B, F); and macrocrystalline,

275

flamboyant and prismatic rims to drusiform quartz (events 21 and 22 in Fig. 5; Fig. 7D,

276

E, F).

277

Partial to total diagenetic recrystallization, dolomitization and/or silicification of

278

fascicular and spherulitic calcite are common (events 16–17 in Fig. 5; Fig. 6D, 8F). Late,

279

coarse, subhedral saddle or baroque dolomite, mega-quartz, and macrocrystalline calcite

280

(events 20, 22 and 23 in Fig. 5; Figs. 7, 8) is widely observed filling vugs and veins.

281

Saddle dolomite exhibits a characteristic xenotopic-C texture with curved crystal faces

282

and sweeping extinction (Gregg and Sibley, 1984; Davies and Smith, 2006; Fig. 7A).

283

The pervasive hydrothermal silicification (events 17, 21 and 22 in Fig 5; Fig. 6D, E, F)

284

is texturally distinct from the eodiagenetic silica crusts (event 10; Fig. 6C), which are

285

concordant

286

dolomitization. A complex evolution in pore fluid chemistry can be interpreted for the

287

hydrothermal alterations of Pre-Salt carbonate rocks.

with

depositional

bedding,

and

clearly occurred

subsequent

to

13 288

Svanbergite (SrAl3(PO4)(SO4)(OH)6) 1 to 100 µm crystals (event 18 in Fig. 5)

289

occur disseminated in microcrystalline silica and/or in chalcedony spherulites,

290

dominantly where silicification affected clay-rich protoliths. Radiated, prismatic and

291

oriented Sr-barite and celestine (event 24 in Fig. 5; Figs. 7B, F, 8A, C, D); fibro-

292

radiated zeolite (event 25 in Fig. 5; Fig. 8E); micro- to macrocrystalline, blocky and

293

prismatic sulfides (pyrite, sphalerite, chalcopyrite and galena; event 26 in Fig. 5; Fig.

294

7F); microcrystalline rutile associated with traces of native copper and zinc (event 27 in

295

Fig. 5); and massive bitumen with shrinkage cracks (event 28 in Fig. 5; Fig. 8F) are less

296

common. Sulfates, sulfides, zeolite and rutile occur filling vugular and fracture porosity,

297

often replacing dolomite, calcite and quartz. Sulfides are dominantly pyrite with lesser

298

sphalerite and minor chalcopyrite and galena. Fluorite and dickite were also described

299

in other wells of the area (Herlinger Jr. et al., 2017), and dawsonite was identified in the

300

XRD analyses (Lima and De Ros, 2019).

301

The pore types that occur in the sag lithologic types are growth-framework,

302

inter-crystalline aggregates, from Mg-clay dissolution, intercrystalline and intra-

303

crystalline aggregates, vugular, and fracture (Lima and De Ros, 2019). Important

304

fracturing and dissolution events occurred, which largely modified the pre-existing pore

305

systems (events 7, 13, 14, 19 and 21 in Fig. 5; Fig. 6). The hydrothermal alterations

306

were related to intense fracturing and dissolution processes (events 14, 19 and 21; Fig.

307

6D, E and F). The late hydrothermal saddle dolomite, macrocrystalline quartz and

308

calcite, sulfates, sulfides, zeolite, svanbergite, rutile, and bitumen filled growth

309

framework, and mostly fracture and dissolution pores (events 18–28 in Fig 5; Fig. 6D, E,

310

F). A later overprint of calcite-quartz-sulfides veins and breccias is also observed.

311

In general, microcrystalline (MD), blocky (BD) and saddle dolomite (SD)

312

present higher concentrations of Na2O, Al2O3, SiO2, MnO and FeO compared to

14 313

fascicular (FC), spherulitic (CS) and macrocrystalline calcite (MC) (WDS analysis; see

314

Supplementary material). SD shows higher values of SiO2 and lower MnO and FeO

315

than other dolomite types. Similarly, hydrothermal MC also has lower concentrations of

316

SiO2, SrO and BaO than FC and CS. In addition, hydrothermal sphalerite (Sp) shows

317

high Fe and Pb and low As, Co, Cu and Ni contents in WDS analyses (Supplementary

318

material). Hydrothermal euhedral pyrite (Py) filling secondary porosity also occurs with

319

high Pb and low As, Co, Cu, Zn and Ni concentrations (WDS). Hydrothermal prismatic

320

Sr-barite (SB) presents SrO content ranging from 2% to 16%, with an average of 6.7%,

321

high TiO2 and low CaO and FeO concentrations (WDS).

322

4.2. Fluid inclusion analyses

323

Aqueous and petroleum fluid inclusions (FIs) hosted in syngenetic and

324

diagenetic/hydrothermal constituents (Fig. 9) were analyzed in the main rift and sag

325

lithologic types (Figs. 4, 7, 8), such as fascicular calcite crusts, stevensitic claystones

326

with calcite spherulites, intraclastic grainstones, bioclastic rudstones and cherts. The

327

phases analyzed were recrystallized bioclasts, fascicular and spherulitic calcite, and

328

diagenetic/hydrothermal macrocrystalline calcite, saddle dolomite, mega-quartz, and

329

prismatic Sr-barite, replacing pre-existing constituents, filling fractures and dissolution

330

porosity. Fluid inclusions (both aqueous and petroleum inclusions) are trapped in

331

recrystallized fascicular and spherulitic calcite displaying randomly distribution in areas

332

with clear mineral substitution. On the other hand, the preserved carbonate elements do

333

not host petroleum inclusions. Therefore, these inclusions in the recrystallized

334

carbonates

335

microthermometric results are displayed in Table 1.

were

interpreted

as

primary.

Petrographic

observations

and

15 336

All aqueous inclusions are two-phase (liquid plus vapor) at room temperature

337

without large variation in the vapor fraction, with exception of some primary inclusions

338

in saddle dolomite and macrocrystalline calcite displaying necking-down features,

339

which were disregarded from subsequent interpretations. Six high Th results from

340

saddle dolomite (140, 144, 148 and 150 °C) and Sr-barite (144 and 152 °C) may

341

represent stretched inclusions or unrecognized necking after a phase change (Goldstein,

342

2001).

343

Analyzed aqueous FIs were hosted in the following mineral phases:

344

recrystallized fascicular calcite (n=2), mega-quartz (n=34), prismatic Sr-barite (n=20),

345

macrocrystalline calcite (n=36), and saddle dolomite (n=34). The majority of the FIs

346

observed were primary and pseudo-secondary aqueous and petroleum types (78%), with

347

subordinate secondary inclusions (22%; Table 1). Aqueous secondary FIs are more

348

significant in macrocrystalline quartz (38%). The salinities of aqueous inclusions varied

349

between 12 and 26 wt. % eq. NaCl (Table 1).

350

Petroleum inclusion from sag samples correspond predominantly to white-blue-

351

fluorescent inclusions (n=82) and less yellow-fluorescent inclusions (n=16). Most are

352

hosted in the recrystallized fascicular and spherulitic calcite, and in the macrocrystalline

353

quartz, and less in macrocrystalline calcite, Sr-barite, and saddle dolomite (Table 1).

354

Petroleum FIs in rift samples consist of mostly white-blue-fluorescent (n=20) in

355

recrystallized calcite bioclasts, and less of yellow-fluorescent (n=3) in saddle dolomite,

356

macrocrystalline quartz and Sr-barite. Estimated API gravities range from 32 to 35° for

357

the yellow-fluorescent petroleum inclusions, and between 40 and 50° for the white-

358

blue-fluorescent inclusions. The characteristics of the FIs in specific analyzed

359

constituents can be summarized as follows:

16 360

Recrystallized calcite bioclasts (CB): 20 analyses of FIs in blocky calcite in

361

recrystallized calcite bioclasts (CB) from a rift rudstone in well B (Table 1; Figs. 3,

362

10B). Only primary petroleum FIs with white fluorescence (~45 to 50° API) were

363

observed. The Th values vary from 80 to 108.6 °C, with mean and median of 94.8 and

364

96.9 °C.

365

Recrystallized fascicular calcite (FC): 22 FI analyses in a sag recrystallized

366

syngenetic fascicular calcite (FC) sample from well B (Table 1; Figs. 3, 10C). Two

367

aqueous FIs present Th of 92 and 93 °C, and salinity of 23.9 and 24.3 wt. % eq. NaCl

368

(Fig. 11). White to blue fluorescent petroleum FIs (~40 to 50° API) have Th ranging

369

from 83 to 111 °C.

370

Recrystallized spherulitic calcite (CS): 38 FIs were analyzed in recrystallized

371

eodiagenetic calcite spherulite (CS) in a sag stevensitic claystone from well A (Table 1;

372

Figs. 3, 10D, E). Both primary and secondary petroleum inclusions with white

373

fluorescence color (~45 to 50° API) were hosted in the CS. Th of 18 primary petroleum

374

inclusions varies from 63 to 112.2 °C, with mean and mode of 83.5 °C (Fig. 10D). Th of

375

20 secondary petroleum inclusions ranges from 63 to 97.4 °C, with mean of 75.7 and

376

median of 72.1 °C.

377

Macrocrystalline calcite (MC): 36 primary and secondary aqueous and

378

petroleum FIs were measured in macrocrystalline calcite (MC) filling dissolution

379

porosity of three sag samples from well C (Table 1; Figs. 3, 12G, H). The 22 primary

380

and pseudo-secondary aqueous FIs present Th varying from 101 to 130 °C, with mean

381

of 114.3 and mode of 111.5 °C. Salinity of primary FIs ranges from 15.1 to 22.9 wt. %

382

eq. NaCl (Fig. 11A). Th of 5 secondary aqueous FIs varies from 99 to 126 °C, with

383

mean of 107 and mode of 101 °C (Fig. 12H). Salinity of secondary aqueous FIs ranges

384

from 18.8 to 22.9 wt. % eq. NaCl (Fig. 11B). Yellow-fluorescent (~30 to 35 °API)

17 385

primary inclusions have Th value of 100 °C. Th in 5 secondary petroleum FIs range

386

from 69 to 124 °C, with mean of 93.8 and median of 81 °C.

387

Macrocrystalline quartz (MQ): 58 FIs were observed in pore-filling mega-quartz

388

(MQ) in a silicified intraclastic grainstone, a FC crust, a chert and a stevensitic

389

claystone with CS from the sag phase in wells A and C (Table 1; Figs. 3, 12E, F). 16

390

primary/pseudo-secondary aqueous FIs display Th from 105 to 139 °C, with mean of

391

122.7 and mode of 132.5 °C (Fig. 12E). Salinity of these inclusions ranges from 12.9 to

392

23.7 wt. % eq. NaCl (Fig. 11A). Th of secondary aqueous FIs varies from 105 to 134 °C,

393

with mean of 116.6 and mode of 118.5 °C (Fig. 12F). Salinity of secondary aqueous

394

inclusions ranges from 19.9 to 23.9 wt. % eq. NaCl (Fig. 11B). Th of 26 white to yellow

395

fluorescent petroleum inclusions (~32 to 50° API) varies from 65 to 100.9 °C, with

396

mean of 76.3 and mode of 73.1 °C (Fig. 12E).

397

Prismatic Sr-barite (SB): 20 FIs analyses in prismatic Sr-barite (SB) cement

398

filling secondary porosity in a sag silicified intraclastic grainstone in well C (Table 1;

399

Figs. 3, 12C, D). Th of 11 primary/pseudo-secondary aqueous FIs ranges from 122 to

400

152 °C, with mean of 131.1 and mode of 127.5 °C (Fig. 12C). Th of 3 secondary

401

aqueous FIs varies from 122 to 124 °C (Fig. 12D). Salinity ranges from 16.7 to 22.1

402

wt. % eq. NaCl in 10 primary aqueous (Fig. 11A), and 20.2 to 21.8 wt. % eq. NaCl in 2

403

secondary aqueous FIs (Fig. 11B). Both primary and secondary petroleum FIs show

404

yellow fluorescence colors (~32 to 34° API). Only one primary yellow-fluorescent

405

petroleum inclusion yielded a Th value of 83 °C (Fig. 12C). Secondary petroleum FIs

406

have Th ranging from 100 to 119 °C with mean of 107.8 and mode of 106 °C (Fig. 12B).

407

Saddle dolomite (SD): 34 FIs in pore-filling saddle dolomite (SD) in sag FC

408

crusts and rift bioclastic rudstones from well C (Table 1; Figs. 3, 12I). Th measured in

409

20 primary aqueous FIs in zoned saddle dolomite varies from 105 to 150 °C with mean

18 410

of 132.8 and mode of 134.5 °C (Fig. 12I). Salinity of 29 primary aqueous FIs in SD

411

ranges from 16.9 to 26.1 wt. % eq. NaCl (Fig. 11A). Th obtained in 3 yellow-

412

fluorescent primary petroleum inclusions varies from 79 to 89 °C (Fig. 12I).

413

The salinities of inclusions hosted in recrystallized FC are close to the upper

414

limit for pore-filling hydrothermal minerals (Fig. 11A), with 97% higher than 15 wt. %

415

eq. NaCl. In the rift interval, salinities range from 17.3 to 22.7 wt. % eq. NaCl. The final

416

melting temperature of aqueous FIs hosted in SD, MC, MQ and SB ranges of -9 to -

417

25.3 °C, indicating NaCl-dominant compositions comparable to CaCl2 (e.g., Roedder

418

and Bodnar, 1997; Wilkinson, 2010).

419

Th of white-blue-fluorescent primary petroleum inclusions hosted in the

420

recrystallized CB, CS and FC ranges from 63 to 112.2 °C (Table 1; Fig. 10). Th of

421

yellow-fluorescent inclusions hosted in MC, MQ, SB and SD varies between 69 and

422

124 °C. Some yellow-fluorescent oil inclusions are interpreted to have been emplaced at

423

105 to 108 °C in MC and at 122 to 130 °C in SB in the presence of hypersaline aqueous

424

fluids. Mixed, co-genetic oil-aqueous FIs with Th ranging from 79 to 124 °C were

425

observed in fascicular calcite crusts, silicified intraclastic grainstone and bioclastic

426

rudstone (Table 1), indicating a common origin from a single fluid. Aqueous inclusions

427

coexist with yellow-fluorescent petroleum inclusions in MC, SB and SD.

428

4.3. Carbon and oxygen isotopes

429

The range of δ13C values for all analyzed carbonates is quite narrow, varying

430

between -1.88 and 2.43‰ (Table 2; Fig. 13A, B). The range of δ13C values for the

431

seven different carbonate phases are: 1) 0.45 to 1.43‰ for microcrystalline/blocky

432

calcite

433

microcrystalline/blocky calcite from laminites (MBCL); 3) 1.16 to 1.44‰ for

from

bioclasts

(MBCB);

2)

0.93

to

1.52‰

for

eodiagenetic

19 434

mesodiagenetic microcrystalline/blocky dolomite from dolostones (MBDD); 4) 0.31 to

435

2.43‰ for recrystallized eodiagenetic calcite spherulites (CS); 5) 0.05 to 1.98‰ for

436

recrystallized syngenetic fascicular calcite (FC); 6) -0.27 to 1.28‰ for diagenetic

437

macrocrystalline calcite (MC); and, 7) -1.88 to 1.03‰ for saddle dolomite (SD).

438

In contrast to the δ13C values, the δ18O values from the analyzed carbonates

439

show a wider variation range (from -9.74 to 1.74‰) (Table 2; Fig. 13A, B). The range

440

of δ18O values from the seven different carbonate phases are: 1) -1.18 to 1.04‰ for

441

MBCB; 2) -0.99 to 0.63‰ for MBCL; 3) -1.51 to 0.35‰ for MBDD; 4) -2.11 to 1.74‰

442

for CS; 5) -5.70 to -1.93‰ for FC; 6) -9.74 to -4.92‰ for MC; and, 7) -9.49 to -6.35‰

443

for SD. Samples analyzed in this study may be arranged into two main groups based on

444

the δ18O values: samples (MC and SD) with more depleted oxygen isotope values (<-

445

4.92‰); and samples (MBCB, MBCL, MBDD and CS) with more enriched oxygen

446

isotope values (>-2.11‰) (Fig. 13A). The δ18O values of FC samples (-5.70 to -1.93‰)

447

fall in between these two groups (Fig. 13A). The comparison of Fig. 13A and 13B with

448

13C and 13D show that both δ13C and δ18O values from different mineral phases are not

449

biased by different wells. Likewise, the distribution of samples against stratigraphy (Fig.

450

3) shows that sampling for isotope analyses are not biased against stratigraphy.

451 452

4.4. Strontium isotopes The analyzed samples show a wide range of

87

Sr/86Sr values (0.71100 to

453

0.71394; Table 3; Fig. 13). They may be separated into two main groups: those with Sr

454

isotope ratios above 0.71299 (MBDD, CS, FC, MC) and those below 0.71226 (MC, SD,

455

SB). As shown in Table 3, the Sr isotope values have a narrow range within each group

456

of samples. For instance, SD values range between 0.71100 and 0.71124, while CS

457

samples range between 0.71299 and 0.71394. As in the δ13C and δ18O results, 87Sr/86Sr

20 458

ratios from different mineral phases are not biased against stratigraphy (Fig. 3) or by

459

different wells (Fig. 13).

460

5. Discussion

461

5.1. Fluid temperature and salinity

462

There is no correlation between the salinity and temperature of fluid inclusions

463

(Fig. 14). Except for few FIs in macrocrystalline quartz and calcite, all other phases

464

exhibit similar salinity range. In contrast, the average Th of saddle dolomite and Sr-

465

barite inclusions are higher than those measured in the mega-quartz and calcite. Based

466

on these observations, the following points may be considered: i) the salinity variations

467

do not indicate mixing of fluids, but are related to cooling and/or neutralization by fluid-

468

rock interaction; ii) the formation of mega-quartz, saddle dolomite, macrocrystalline

469

calcite and Sr-barite, which are associated with higher temperature fluids, postdated the

470

formation of the other phases. This further indicates that heat was not pervasive in the

471

system, but rather concentrated along structures such as fractures that acted as conduits

472

for the hot fluids.

473

The variability of the Th values in aqueous inclusions may be related to post-

474

entrapment modification of FIs at variable high temperatures (King and Goldstein,

475

2018). For instance, the necking-down process occurs most rapidly at higher

476

temperatures, often producing Th values without any relation to the original

477

temperatures (Roedder, 1984; Bodnar et al., 1985; Goldstein, 2001; Bodnar, 2003). The

478

similar Th values in macrocrystalline calcite and quartz indicate that there were no

479

significant post-entrapment changes in FIs. Another relevant indication of preservation

480

of the integrity of the FIs is the robust petrographic control evidenced by the relatively

481

narrow range of Th values of individual fluid inclusions, generally much narrower than

21 482

within the entire sample, or between different samples (e.g., Wilkinson, 2010).

483

Therefore, the wider dispersion in Th values observed in hydrothermal saddle dolomite

484

is probably representing different generations of FIs trapped in the zoned crystals (Fig.

485

9D).

486

The Th values of aqueous FIs hosted in minerals precipitated or recrystallized by

487

hydrothermal fluids show a relatively narrow range, varying from to 92 and 152 °C

488

(Table 1; Fig. 14), in which 95% of the results range from 100 to 140 °C. In general,

489

there is a remarkable Th overlapping among all analyzed phases, with slightly higher

490

mean values for Sr-barite and saddle dolomite. The FI data indicate that hydrothermal

491

saddle dolomite, Sr-barite, macrocrystalline calcite, and mega-quartz precipitated over a

492

similar temperature range. The only two analyses performed in aqueous FIs of

493

recrystallized fascicular calcite yielded Th near the lower limits of the hydrothermal

494

phases (Table 1; Fig. 14). Poros et al. (2017) verified a Th variation between 95 to

495

115 °C in inclusions hosted in dolomite cement, 95 to 125 °C in silica (chert and

496

chalcedony) replacing carbonate, and 105 to 125 °C in pore-filling mega-quartz in the

497

Pre-Salt of Kwanza Basin, Angola. Although those mineral phases were interpreted to

498

have been formed by a hydrothermal system in a deep-burial condition, the Th values in

499

Kwanza Basin Pre-Salt reservoirs are somewhat lower than those obtained in Campos

500

Basin.

501

All analyzed phases in the study area contain abundant oil and/or co-genetic oil-

502

aqueous inclusions, which is indicative of atypical petroleum generation and migration

503

(sensu Magoon and Dow, 1994) associated with percolation of hydrothermal fluids

504

responsible for the recrystallization and precipitation of the host mineral phases. When a

505

sedimentary sequence is percolated by hot fluids, the organic matter of host rocks can

506

be thermally altered, generating petroleum through hydrothermal or “forced” maturation

22 507

(Ilchik et al., 1986; Anderson, 1991; Davies and Smith, 2006). The liquid petroleum

508

droplets produced by this mechanism are transported by the convecting fluids

509

(Schoenherr et al., 2007), and can be either trapped as petroleum FIs in hydrothermal

510

minerals, or devolatilized in the pore space as bitumen. This type of natural hydrous

511

pyrolysis is not efficient enough to fill a reservoir or an oil field (Davies and Smith,

512

2006) but may, however, generate oil or co-genetic oil-aqueous FIs and solid bitumen

513

(Simoneit, 1990; Simoneit, 2018, and references therein) comparable to those described

514

here, which may be associated to hydrothermal dolomitization and ore deposits

515

described in diverse geotectonic settings (Disnar, 1996; Wilson and Zentilli, 1999;

516

Parnell et al., 2003; Bertrand et al., 2003; Neilson and Oxtoby, 2008; Suchý et al., 2010;

517

Greenwood et al., 2013 and references therein; Ostendorf et al., 2015).

518

With progressive burial, ‘normal’ background mesodiagenetic processes

519

proceeded in Campos Basin Pre-Salt independently of hydrothermal events (Fig. 5).

520

Thus, it should be noted that part of the petroleum inclusions hosted in the recrystallized

521

calcite bioclasts and spherulites (Th = 80–111 °C), in which no aqueous inclusions were

522

observed, may have been trapped during the migration of oil generated by the

523

conventional petroleum system. According to Mello et al. (1994), conventional burial

524

petroleum generation, migration and accumulation started in Campos Basin only in the

525

Miocene, being thus subsequent to the atypical hydrothermal petroleum migration. This

526

conventional petroleum system was responsible for the large volume of oil and gas that

527

fills the Pre-Salt reservoirs.

528

Salinities are consistently high in aqueous inclusions in the studied hydrothermal

529

phases (12 to 26 wt. % eq. NaCl; Table 1), suggesting that they were formed from

530

hypersaline brines. In this context, significant Th variation within a narrow salinity

531

range indicates temperature variation within the hydrothermal system during mineral

23 532

precipitation (Fig. 14). Nevertheless, no correlation was observed between Th and

533

salinity data. This, together with homogeneous high salinity values, indicates that there

534

was no fluid mixing in the studied hydrothermal system, and that the minerals

535

precipitated from a single hydrothermal system. Rare salinities below 15 wt. % eq.

536

NaCl from aqueous FIs hosted in macrocrystalline calcite and quartz (Table 1; Fig. 14)

537

could indicate some incipient isothermal mixing between the predominant hydrothermal

538

fluid (approximately 20–21 wt. % eq. NaCl) with a moderate-salinity fluid

539

(approximately 13 wt. % eq. NaCl). However, since 97% of the salinity results are

540

above 15 wt. % eq. NaCl, this relatively lower values may be related to salinity

541

variation in recharge area (Khaska et al., 2013).

542

The Th data in the range 92 to 152 °C with salinities of 13 to 26 wt. % eq. NaCl

543

of Campos Basin Pre-Salt are very similar to the majority of hydrothermal systems

544

observed in the hydrothermal dolomites (HTD), carbonate-hosted Pb-Zn Mississippi

545

Valley (MVT) and Irish-type deposits (Davies and Smith, 2006; Leach et al., 2001;

546

Wilkinson, 2001; Paradis et al., 2007; Wilkinson, 2010).

547

5.2. Carbon, oxygen and strontium isotopes

548

The δ13C values of host rock mineral phases (MBCB, MBCL, MBDD, FC and

549

CS) and hydrothermal mineral phases (SD and MC) present a very narrow variation

550

(between -1.88 and 2.43‰; Fig. 13, Table 2). These values are similar to those

551

previously reported from Pre-Salt carbonates from Campos, Santos and Kwanza Basins.

552

δ13C bulk-rock data from the cuttings of well CP-5 in Campos Basin vary between -0.38

553

and 2.71‰ for the Macabu Formation, and between -0.88 and 1.35‰ for the Coqueiros

554

Formation (Dias, 1998; Rodrigues, 2005; Muniz and Bosence, 2015). The δ13C bulk-

555

rock data (-1.52 to 3.20‰) of spherulite, shrub, laminate, calcarenite and calcirudite

24 556

lithofacies from core samples of the sag Barra Velha Formation from Santos Basin

557

(Farias et al., 2019) are similar to reported values of lithofacies from the Macabu

558

Formation (Herlinger Jr. et al., 2017; Lima and De Ros, 2019).

559

Syn-rift and sag intervals from Kwanza Basin have lithofacies and mineral

560

phases similar to those described for the rift (Coqueiros Formation) and sag (Macabu

561

Formation) intervals of the Campos Basin (Fig. 2). The δ13C values of spherulitic and

562

shrubby calcite and early dolomite from the sag interval of Kwanza Basin vary between

563

-4.3 and 3‰, whereas neomorphic calcite and calcite cement filling mollusk molds and

564

interparticle pores from syn-rift interval of Kwanza Basin usually display δ13C between

565

-3.0 and 2.0‰ (Saller et al., 2016; Sabato Ceraldi and Green, 2016; Fig. 13). Although

566

Sabato Ceraldi and Green (2016) present δ13C and δ18O values from rift and sag Pre-Salt

567

sections from Kwanza Basin, it is not possible to identify the carbonate analyzed. The

568

lithofacies, mineralogy and δ13C data of Campos, Santos and Kwanza basins suggest

569

that the depositional and diagenetic processes were quite similar within the vast Aptian

570

lacustrine system, indicating either a single water body or hydrogeological connected

571

lakes.

572

Data from the present study show that δ18O values of host rock mineral phases

573

(MBCB, MBCL, MBDD and CS) vary between -2.11 and 1.74‰, except for FC that

574

has values raging between -5.70 and -1.93‰ (Fig. 13, Table 2). Similar pattern was

575

obtained by Sabato Ceraldi and Green (2016). In contrast, hydrothermal mineral phases

576

(SD and MC) present much lower isotope values (-9.74 to -4.92‰; Fig. 13, Table 2),

577

suggesting either differences in the temperature of the fluid and/or in their oxygen

578

isotopic composition (δ18Ow). In addition, δ18O values of FC samples suggest

579

recrystallization with temperature and/or δ18Ow different from both other host rock

580

mineral phases and hydrothermal mineral phases. The comparison between the δ18O of

25 581

carbonates from the Pre-Salt and others formed elsewhere is not straightforward,

582

because their isotopic composition depends on the temperature, carbonate phase and

583

fluid δ18Ow (Sharp, 2007). Equations (1) and (2) show the relationship between δ18Ocarb,

584

δ18Ow and temperature for calcite (O’Neil et al., 1969) and dolomite (Matthews and

585

Katz, 1977), respectively:

586

10  ∝ = 2.78 ×

587

10  ∝ = 3.06 ×

   

− 2.89 (1) − 3.24

(2)

588

Where ( 10  ∝ ) is approximately equal to δ18Ocarb - δ18Ow, and T is the

589

temperature in Kelvin. These equations were used to calculate the δ18Ow of fluid in

590

equilibrium with each mineral phase (FC, CS, SD and MC) based on Th obtained from

591

aqueous and petroleum FIs, similar to the approach by King and Goldstein (2018). The

592

δ18Ow VPDB (Vienna Pee Dee Belemnite) was then converted to δ18Ow VSMOW

593

(Vienna Standard Mean Ocean Water) by using the equation 3 (O’Neil et al., 1969):

594

δ# O% &'()* = 1.03086 × δ# O% &+,- + 30.86 (3)

595

Calculated δ18Ow for FC samples vary between 5.4 and 12.4‰ (Table 4) based

596

on the maximum and minimum measured δ18Ocarb (-5.7 and -1.93‰) and Th derived

597

from petroleum FIs (83–111 °C). A variation of 6.5 to 10.4‰ was calculated for FC

598

δ18Ow using just two measurements of Th from aqueous FIs, both of 92.2 °C. Calculated

599

δ18Ow of CS samples vary between 6.3 and 16.3‰ based on the maximum and

600

minimum measured δ18Ocarb (-2.11 and 1.74‰) and Th derived from petroleum FIs (63–

601

112.2 °C). Those values calculated from Tho suggest that δ18Ow was the oxygen isotope

602

composition of pore water during the recrystallization of both mineral phases (FC and

603

CS). These high δ18Ow values are similar to those calculated for the isotopic

604

composition of fluids in equilibrium with shrub and spherulite lithofacies from Well

26 605

SB-2 of the Barra Velha Formation, Santos Basin (Farias et al., 2019). Based on oxygen

606

isotope compositions and clumped temperatures, the recalculated δ18Ow in equilibrium

607

with carbonates from these lithofacies vary between 5.0 to 10.7‰. These highly

608

enriched δ18Ow are comparable to the isotopic composition of residual evaporative

609

brines (Epstein and Mayeda, 1953) and/or of fluids resulted from extensive rock-water

610

interaction with basinal materials (King and Goldstein, 2018). Therefore, we suggest

611

that those values represent δ18Ow modified from residual evaporative brines during

612

burial, in which δ18Ow is the product of extensive rock-water interaction and

613

recrystallization. Considering this scenario, the FIs in spherulites and shrubs were

614

trapped during the recrystallization of these constituents during advanced burial. Some

615

FIs in these phases may, however, have been inherited from earlier stages of

616

neomorphism/recrystallization.

617

The oxygen isotopic composition of aqueous fluids (δ18Ow) in equilibrium with

618

MC are calculated between 3.3 and 11.1‰ (Table 4), based on the maximum and

619

minimum measured δ18Ocarb (-9.74 and -4.92‰) and on Th derived from aqueous FIs

620

(101−130 °C). These values are similar to those calculated for the aqueous fluids in

621

equilibrium with SD (δ18Ow between 2.3 and 10.0‰), which are based on aqueous FIs

622

in the range of 105 to 150 °C and maximum and minimum measured δ18Ocarb between -

623

9.49 and -6.35‰. These results indicate that MC and SD were formed from the same

624

hydrothermal fluid.

625

The

87

Sr/86Sr ratios from of host rock mineral phases (FC, CS and MBDD –

626

from 0.71299 to 0.71394) are significantly higher than those from hydrothermal mineral

627

phases (SB, SD and MC – from 0.71100 to 0.71226) (Fig. 13, Table 3). Reported

628

87

629

2017), which is the same well labeled as well A in this study, were obtained by leaching

Sr/86Sr values for the Macabu Formation (0.712991−0.713365) in well C (Tedeschi,

27 630

carbonate phases with weak acid, without differentiation of carbonate phases or micro-

631

sampling. The results presented here are also similar to previous data from other

632

Brazilian Pre-Salt wells at Campos and Santos basins (Dias, 1998; Pietzsch et al., 2018;

633

Farias et al., 2019; Fig. 13F).

634

The

87

Sr/86Sr ratios of carbonate samples from the rift Coqueiros and sag

635

Macabu formations range from 0.7117 to 0.7118 (n=2) and 0.7128 to 0.7137 (n=2)

636

(Dias, 1998). The

637

spherulite, shrub, laminate, calcarenite and calcirudite, from Santos Basin SB-2 well

638

vary from 0.7130 to 0.7138 (Farias et al., 2019). In addition, the 87Sr/86Sr ratios of 194

639

sidewall samples from well S10 in Santos Basin vary between 0.7105 and 0.7140

640

(Pietzsch et al., 2018). The isotope ratios from the rift Itapema Formation (Jiquiá local

641

stage) were usually lower than 0.7115, while the ratios from the sag Barra Velha

642

Formation (Alagoas local stage) were roughly between 0.7120 and 0.7140. These values

643

agree with five diagenetically modified ostracod samples analyzed by Tedeschi (2017),

644

three of which are from Hourcqia sp. valves (Itapema Formation;

645

0.711553 to 0.712012) and two are from Pattersoncypris ssp. valves (Barra Velha

646

Formation;

647

isotopes and 87Sr/86Sr data from different studies, depositional and diagenetic processes

648

were similar over a regional scale in Campos and Santos Basins, suggesting a single

649

vast lake or lakes that were hydrogeologically connected. Additionally, these data

650

establish constraints on

651

comparison with the hydrothermal mineral phases.

652

The

87

Sr/86Sr ratios from 43 samples from the sag phase, including

87

Sr/86Sr ratios from

87

Sr/86Sr ratios from 0.712756 to 0.713427). As indicated by the stable

87

87

Sr/86Sr ratios at a regional scale that can be used for

Sr/86Sr ratio of Pre-Salt carbonates indicate a strong interaction with 87

Sr/86Sr ratios,

653

continental crust materials, being significantly higher than marine

654

specifically than seawater values from Early Cretaceous (from 0.707074 to 0.707541;

28 655

Jenkyns et al., 1995; McArthur et al., 2012; Bodin et al., 2015; Yamamoto et al., 2013;

656

Ando, 2015) as well as that of Phanerozoic seawater (<0.7095; McArthur et al., 2012;

657

Fig. 13F). Based on Sr isotopes, Pietzsch et al. (2018) presented a geological model to

658

discriminate the Sr source of lacustrine rift and sag sections in the Santos Basin.

659

According to this model, Sr isotopes of carbonate rocks from sag Barra Velha

660

Formation were controlled by groundwater with long residence time of interaction with

661

felsic basement rocks. Although one may argue that leaching of waters through felsic

662

basement was not enough to deliver ions (Mg2+, Ca2+ and others) to develop the Barra

663

Velha Formation, a mass balance using waters from the basement has not been

664

performed. Therefore, although this process seems to be feasible at local scale, as

665

pointed out by the geochemical models of Teboul et al. (2016), the isotopic and

666

elemental aspects of the huge Pre-Salt volumes still need to be explained in basinal

667

scale.

668

Otherwise, coquinas from the Itapema Formation would have been deposited in

669

a rift lacustrine environment with higher fluvial recharge, and more variable catchment

670

area and subsurface recharge. This hydrological model for the Itapema Formation would

671

allow mixing of groundwater recharge with long residence time with waters derived

672

from Cabiúnas volcanic and volcaniclastic rocks and Piçarras Formation siliciclastic

673

sediments. Considering the 87Sr/86Sr ratios obtained from the Campos Basin, we would

674

consider the same hydrological patterns to explain the differences between

675

ratios from the rift Coqueiros and the sag Macabu formations.

676

The

87

Sr/86Sr

87

Sr/86Sr values of SD, MC and SB overlap each other (Fig. 13, Table 3).

677

Although the petrographic sequence shows minor dissolution of SD followed by

678

formation of MC and SB, there is not a definitive petrographic evidence that these

679

mineral phases are derived from different fluids (Figs. 5, 8). The same 87Sr/86Sr values

29 680

and the absence of petrographic evidence suggest that SD, MC and SB were formed

681

from the same fluid. In addition, strontium isotope composition of carbonate

682

hydrothermal phases (SD and MC) shows respectively positive and negative correlation

683

with δ13C and δ18O values, which also suggest a same trend of evolution from the same

684

fluid (Fig. 13). The relationship between δ13C and

685

observed in hydrothermal baroque dolomite from Cambrian-Ordovician Arbuckle

686

Group in USA midcontinent, whereas the relationship between δ18O and 87Sr/86Sr ratios

687

is the same (King and Goldstein, 2018). These observations suggest the same strontium

688

source for all these three mineral phases. The negative correlation between δ18O and

689

87

690

different rocks or that hydrothermal fluids at higher temperature and/or with lower

691

δ18Ow would promote stronger host rock dissolution, becoming more radiogenic.

692

Likewise, the positive correlation between

693

increased host rock dissolution by the hydrothermal fluids. According to this

694

interpretation, strong dissolution of basement and rift rocks precede the formation of

695

these hydrothermal mineral phases (SD, MC and SB).

696

87

Sr/86Sr is the opposite of those

Sr/86Sr (Fig. 13) suggests either that different set of fluids were in equilibrium with

87

Sr/86Sr and δ13C would be related to

5.3. Hydrothermal alteration of carbonate rocks

697

The term hydrothermal alteration was originally applied to situations with the

698

presence of hot waters and ore deposits associated with magmatic activities (e.g.,

699

Gilbert, 1875; Morey and Niggli, 1913; Holmes, 1928; Stearns et al., 1935). More

700

recently, hydrothermal alteration was defined as a replacement of the original minerals

701

by hydrothermal fluids that deliver reactants and remove aqueous reaction products

702

(Utada, 1980; Henley and Ellis, 1983; Inoue, 1995; Reed, 1997). Hydrothermal fluids

703

represent aqueous solutions of different compositions and origins (magmatic,

30 704

metamorphic, groundwater, or meteoric waters), which are hotter than the wall-rock

705

(Skinner, 1997).

706

Hydrothermal alterations represent the evidence of a geothermal anomaly,

707

requiring both a mechanism (heat source) and conduits for the flow of fluids (e.g. deep

708

fault systems). Hydrothermal systems are characterized by focusing of fluids through

709

fault systems at very high flow rates, in a transient or episodic regime (Davies, 2004).

710

Such focalization of fluids and their interaction with the host rocks are linked to the

711

origin of carbonate-hosted base metal deposits (e.g., the Pb-Zn Mississippi Valley type

712

deposits; Leach, 1994) and may also modify significantly the original properties of

713

some hydrocarbon reservoirs (e.g., Smith, 2006; Zhu et al., 2015). The so-called

714

hydrothermal petroleum is generated in volcanic and geothermal marine and continental

715

environments, and sampled from submarine and sublacustrine oil seeps (vent fluids),

716

mud volcanoes, hot springs, chimneys, fumaroles and other surface manifestations

717

(Ventura et al., 2012; Simoneit, 2018).

718

Despite several studies on hydrothermal ore deposits, the characterization and

719

definition of hydrothermal processes in sedimentary successions are still a matter of

720

heated discussion (e.g., Machel and Lonnee, 2002; Davies, 2004; Davies and Smith,

721

2006). White (1957) defined hydrothermal processes solely based on a temperature

722

contrast of 5 °C or more between the inflowing aqueous solutions and the wall-rocks,

723

disregarding any temperature limits or source of the fluids. Similarly, Machel and

724

Lonnee (2002) suggested that minerals formed in sedimentary rocks should be

725

considered hydrothermal minerals only if precipitated from fluids with temperatures at

726

least 5 to 10 °C hotter than the host rock temperature, disregarding the fluid sources and

727

the pathways. More loosely, Davies and Smith (2006) defined hydrothermal fluids as

728

those arising to the surface at temperatures higher than those of the depositional

31 729

environment, or that are introduced into the host rocks at a higher temperature than the

730

host. Following such line of definition, sedimentary ore-forming hydrothermal systems

731

may be either depositional (syngenetic) or diagenetic (epigenetic), depending on the

732

time, depth, geometry and mechanism of the flow of fluids. In still other lines of

733

definition, the minerals are hydrothermal if considered “unusual”, or “exotic” in relation

734

to the common carbonate rock mineralogy, including saddle dolomite, fluorite, barite,

735

anhydrite, dickite, sphalerite, and pyrite (Neilson and Oxtoby, 2008).

736

As reliable determination of paleotemperature contrasts in the order of only 5 to

737

10 °C is impossible in most cases, and as minerals that may be considered “exotic” by

738

some authors are widespread in sedimentary successions, or simply do not occur in

739

several recognized hydrothermal systems, we opted for a more practical and operational

740

definition for hydrothermal processes and products. This definition is based on the

741

focusing of the flow of relatively heated fluids coming from below through faults,

742

fractures, unconformities and similar conduits, and on the consequent concentration of

743

mineral precipitation and/or dissolution in the proximity of such conduits. Therefore, in

744

this sense, temperature and/or compositional requirements are not mandatory for the

745

definition of hydrothermal alterations.

746

5.3.1. Hydrothermal dolomitization

747

Hydrothermal dolomitization was defined by Davies and Smith (2006) as

748

occurring under subsurface conditions, commonly at shallow burial depths (< 1 km),

749

and developed along structural lineaments by hypersaline fluids presenting higher

750

temperatures and pressures than those of the host formation, which are usually

751

carbonate rocks. According to these authors, reservoir facies formed by structurally-

752

controlled hydrothermal dolomitization (HTD) are the largest hydrocarbon producers in

753

North America and have great potential in other basins. King and Goldstein (2018)

32 754

interpreted basin-derived hydrothermal system as responsible for baroque dolomite

755

precipitation with elevated Th and salinities in the Cambrian-Ordovician Arbuckle

756

Group. Hydrothermal calcite cement filling secondary porosity in hydrothermal

757

dolomites has also been described by some authors (e.g., Coveney Jr. et al. 2000,

758

Lonnee and Machel, 2006, Biehl et al., 2016, Mansurbeg et al., 2016).

759

Hydrothermal dolomitization has been reported in the Ordovician of the United

760

States, Canada, and China (e.g., Al-Aasm, 2003; Davies and Smith, 2006; Luczaj, 2006;

761

Luczaj et al., 2006; Conliffe et al., 2010; Xu et al., 2015), the Devonian and

762

Mississippian basins of Canada (e.g., Boreen and Colquhoun, 2001; Boreen and Davies,

763

2004), the Carboniferous of the United States and Spain (e.g., Gasparrini et al., 2006;

764

Hiemstra and Goldstein, 2015), the Permian of Germany (e.g., Biehl et al., 2016), the

765

Mesozoic of the Atlantic passive margins (e.g., Wierzbicki et al., 2006), the Jurassic-

766

Cretaceous of Spain, the Persian Gulf in Iraq (e.g., Mansurbeg et al., 2016), and Saudi

767

Arabia (e.g., Cantrell et al., 2004). Based on petrography, isotopic, fluid inclusions

768

characteristic, and on the operational definition previously proposed, the analyzed

769

saddle or baroque dolomites (Figs. 5, 6, 7, 8, 9, 11, 12, 13, 14) are defined as

770

hydrothermal, as also the extensive dolomitization observed in part of the associated

771

Pre-Salt rocks.

772

5.3.2.

Hydrothermal silicification

773

Silicification is a common diagenetic process, affecting a wide variety of

774

originally non-siliceous sediments (e.g., Namy, 1974; Meyers, 1977; Meyers and James,

775

1978; Hesse, 1989). Hydrothermal silicification can affect a wide variety of rock types

776

in multiple environments and geological contexts. However, few studies specifically

777

addressing the hydrothermal silicification of carbonate deposits have been published

778

(e.g., Bellanca et al., 1984; You et al., 2018). In most studies on the hydrothermal

33 779

alteration of carbonate sequences, silicification is considered a subordinate process in

780

comparison to dolomitization (e.g., Packard et al., 2001). Nevertheless, hydrothermal

781

silicification of Pre-Salt carbonates has been previously described in Campos Basin

782

(Vieira de Luca et al., 2017; Lima and De Ros, 2019), as well as of the African basins

783

(Poros et al., 2017; Teboul et al., 2017).

784

An example of the association of hydrothermal silicification and HTD

785

corresponds to the Upper Devonian (Famenian) Wabamun Group gas-condensate

786

reservoirs of the Parkland Field, Canada, which were generated by hydrothermal

787

dolomitization, silicification and dissolution (Packard et al., 2001). In those reservoirs,

788

hydrothermal silica occurs as lenses or as diffuse and discontinuous masses of

789

microquartz associated or not with dolomitization, as well as microquartz cement filling

790

fractures after saddle dolomite precipitation. The dolomitization and silicification of

791

those carbonate reservoirs occurred in relatively rapid succession and at shallow burial

792

depths. The pervasive precipitation of microcrystalline silica, macrocrystalline quartz

793

and spherulitic chalcedony observed in the study area (Figs. 5, 6, 7, 9, 11, 12, 14) was

794

defined as hydrothermal, based on the petrographic and fluid inclusions characteristics,

795

and on the operational definition previously presented.

796

5.4. Timing and source of the hydrothermal system

797

The hydrothermal processes recorded in northern Campos Basin Pre-Salt

798

lacustrine carbonates are imprinted in rocks of the sag section, and therefore cannot be

799

related to the rift magmatism. The scarcity of hydraulic breccias, the distribution of the

800

hydrothermal phases in the host carbonates, and the high homogenization temperature

801

(Th) of FIs indicate a hydrothermal system was active during effective burial, and not

802

related to the Aptian magmatism either. This indicates that the hydrothermal fluids

34 803

responsible for dissolution and for precipitation of mineral phases filling fractures that

804

cross-cut the studied carbonates are possibly related to Late Cretaceous and/or

805

Paleogene magmatic events.

806

The timing of the hydrothermal processes affecting northern Campos Basin Pre-

807

Salt deposits can be better evaluated within the framework of the burial and thermal

808

history of the area (Fig. 15). Based on this, and on petrographic and FIs evidence, the

809

observed hydrothermal alterations can be more probably related to the Late Cretaceous

810

(Santonian/Campanian), Paleocene and/or Eocene magmatic events (Fig. 15). The burial

811

and thermal history of the area indicates that during the advent of these magmatic

812

activities, the Pre-Salt reservoirs were at 2 to 4 kilometers below the seafloor and at 76

813

to 98 °C (Fig. 15). This temperature range is substantially lower (> 5–10 °C) than those

814

recorded in aqueous FIs hosted in the hydrothermal phases, thus indicating another heat

815

source to the system besides burial also had existed.

816

The δ13C data of hydrothermal SD and MC are similar (between -1.88 and

817

+1.28‰; Fig. 13, Table 2), suggesting both mineral phases were precipitated by the

818

same hydrothermal fluids, and from the same carbon source. Except for a few samples,

819

these hydrothermal phases are within the range of δ13C values of host rock mineral

820

phases (MBCB, MBCL, MBDD, FC and CS), which vary between 0.05‰ and 2.43‰.

821

Therefore, the carbon composition of the hydrothermal fluids was near to isotope

822

equilibrium with the host rock mineral phases. The slightly wider range of δ13C values

823

observed in SD samples could be explained by their wide temperature formation range

824

indicated by the FI data (105 to 150 °C, Table 1). For instance, the carbon isotopic

825

fractionation between calcite and CO2 vary from 1.0‰ at 150 °C to 3.4‰ at 100 °C

826

(Bottinga, 1969). This implies that temperature decrease alone could explain the carbon

35 827

isotopic range observed in SD, and that the isotopic composition of hydrothermal fluids

828

was buffered by their interaction with the host rocks.

829

Besides temperature, the observed variation in carbon isotopes could also be

830

related to mixture of fluids from different sources. The lowest δ13C value observed in

831

SD (-1.88‰) could suggest an original carbon isotopic for the hydrothermal fluids

832

before fluid-rock interaction, which is lower than 0.05‰. One possibility is that the

833

fluid had a δ13C value of -1.88‰. However, it is likely that this value is a result of fluid

834

with lower δ13C value and a higher δ13C value from the host-rock (mineral phases

835

usually display δ13C > 0.05‰). This interaction occurs by dissolving partially the

836

carbonates present in the host-rock. The dissolution of carbonate minerals from host

837

rock could be due to the low pH and/or low partial pressure of CO2 of the original fluid

838

before interaction with the host rock. As dissolution increases, higher is the CO2 from

839

host rock in modified hydrothermal fluid. This process is sensitive to temperature as the

840

system involves the interaction of H2O and CO2.

841

A possible original CO2 source for the hydrothermal fluids would be the

842

serpentinization of upper mantle (Fig. 16), occurring below areas with strongly thinned

843

continental crust or even through direct mantle exhumation (e.g., Boillot et al., 1987;

844

Manatschal and Bernoulli, 1999; Whitmarsh et al., 2001; Kusznir and Karner, 2007).

845

Mantle exhumation can liberate large volumes of Si, Ca, Mg and CO2 to the interacting

846

fluids (Frost and Beard, 2007; Pinto et al., 2015; 2017). Additionally, the exothermic

847

chemical reactions related to serpentinization (Moody, 1976; Proskurowski et al., 2006)

848

may generate hydrothermal fluids with high content of hydrogen and light hydrocarbons,

849

mainly CH4 (Allen and Seyfried Jr., 2004). Mantle exhumation would have occurred in

850

deep offshore southern and southeastern Brazilian basins, including Campos, during the

851

initial Atlantic Ocean opening, as suggested by seismic and gravimetric evidence

36 852

(Unternehr et al., 2010; Gomes et al., 2011; Zalán et al., 2011; Kumar et al., 2013;

853

Peron-Pinvidic et al., 2013; Kukla et al., 2018). Despite the lowest δ13C value observed

854

in SD (-1.88‰) is significantly higher than mantle isotopic composition, which shows a

855

peak carbon isotope signature of -5‰ (Deines, 2002), we cannot rule out this hypothesis.

856

Nevertheless, the 87Sr/86Sr ratios of both the syngenetic and the diagenetic/hydrothermal

857

Pre-Salt carbonates are much higher than the values expected for the derivation of Sr

858

(and of Ca, which has similar geochemical behavior; Banner, 1995) from mantle

859

exhumation and serpentinization.

860

More probable source for these δ13C values involves the contribution of light

861

carbon derived from organic matter alteration by the hydrothermal fluids, or their

862

interaction with sediments with lighter isotopic composition, such as the rift carbonates

863

from the Coqueiros Formation (e.g., well CP-5 of Dias, 1998; Fig. 16), and/or laminite

864

and spherulite lithofacies similar to those from Santos Basin Barra Velha Formation

865

(e.g., well SB-2 of Farias et al., 2019). An additional mechanism for the lowest δ13C

866

values could include the alteration of organic matter by thermal sulfate reduction (TSR;

867

Machel, 2001). The presence of sulfides and sulfates derived presumably from the same

868

hydrothermal fluid could suggest a possible role of TSR during hydrothermal fluid-rock

869

interaction. However, the carbon isotope values are not as negative as expected for this

870

process, such as in the hydrothermal calcite from Cambrian-Ordovician Arbulckle

871

Group (King and Goldstein, 2018).

872

Calculated δ18Ow from hydrothermal mineral phases partially overlap the values

873

calculated for fluids in equilibrium with the host rock mineral phases (FC and CS),

874

suggesting that part of the oxygen isotope variations can be explained by processes

875

involving dissolution of host rock during hydrothermal flow, and/or mixing between the

876

host pore waters and the hydrothermal fluids. The exceptions are the lowest calculated

37 877

δ18Ow values (roughly between 3 and 5‰), which may be related to fluids of a different

878

source or with quite distinct temperature. If the fluids had different sources, those values

879

enriched in δ18Ow would suggest extensive water-rock interaction along the fluid

880

migration pathways. Considering the geology of the basin, the Pre-Cambrian basement

881

(mainly granitic-gneissic felsic rocks), volcanic rocks from the Cabiúnas Formation

882

(mainly basaltic) and sediments from the Atafona and Coqueiros formations could have

883

interacted with the hydrothermal fluids (Fig. 16). The definition of the main rocks that

884

interacted with the hydrothermal fluids is however difficult to define based exclusively

885

on oxygen isotopes.

886

As hydrothermal fluids are usually a result of mixing between their ‘original’

887

composition and the host-rock fluids and minerals, it is likely that such fluid-rock

888

interaction would influence the

889

However, the

890

significantly higher than those from hydrothermal mineral phases (SD, MC and SB) and

891

do not overlap them. In addition, SD and SB mineral phases display high total strontium

892

content (Supplementary material). Thus, it is likely that strontium concentration of the

893

hydrothermal fluids were relatively higher than the host rock before the dissolution of

894

the latter and that SD 87Sr/86Sr ratios (~0.711) represent the ratio of hydrothermal fluids

895

before fluid-rock interaction. The

896

radiogenic than those from Cretaceous seawater, than fluids associated to hydrothermal

897

alteration of predominantly mafic igneous rocks worldwide (~0.7073; e.g., Burke et al.,

898

1982; Allègre et al., 2010), and than most of rift volcanic and volcaniclastic rocks from

899

Campos Basin Cabiúnas Formation (~0.708; Mizusaki et al., 1992; Tedeschi, 2017).

900 901

87

87

Sr/86Sr ratios of hydrothermal phases as well.

Sr/86Sr ratios of host rock mineral phases (FC, CS and BD) are

Assuming similar

87

87

Sr/86Sr ratios of SD are substantially more

Sr/86Sr ratios for the rift Coqueiros Formation to those of

Itapema Formation from Santos Basin (0.7105 to 0.7120; Tedeschi, 2017; Pietzsch et al.,

38 902

2018), such radiogenic values are likely to be related to the dissolution of the Coqueiros

903

Formation carbonates (Fig. 16). However, as the rift Itapema Formation has

904

ratios lower than those from the sag Barra Velha Formation (Pietzsch et al., 2018),

905

another source of fluids would be needed to explain such

906

syngenetic sag carbonates. The higher

907

diagenetic sag carbonates could represent a product of leaching of the Precambrian

908

felsic basement (Tupinambá et al., 2012; Teboul et al., 2016; Pietzsch et al., 2018). The

909

question is that the granitic-gneissic basement is not a suitable source for the huge

910

amounts of Mg and Ca precipitated in the sag rocks. Nevertheless, the lower

911

ratios observed in the analyzed hydrothermal phases in relation to the syngenetic and

912

early diagenetic sag carbonates could represent mixing with original fluids derived from

913

mantle serpentinization (Fig. 16), and/or a product of interaction with mafic volcanic

914

and volcanoclastic rocks from Cabiúnas Formation (Bertani and Carozzi, 1985a, b;

915

Misuzaki et al., 1992). Such mixing would be necessary to lower the 87Sr/86Sr ratios of

916

the hydrothermal mineral phases in comparison to the host rocks.

87

87

87

Sr/86Sr

Sr enrichment in the

Sr/86Sr ratios of the syngenetic and early

87

Sr/86Sr

917

According to the burial-thermal history (Fig. 15), the estimated maximum

918

temperature for the Pre-Salt reservoirs was approximately 110 °C during the occurrence

919

of the Eocene magmatic episode in the Campos Basin. This temperature is significantly

920

lower than the mean and maximum Th values, respectively 123.4 and 152 °C, obtained

921

in the aqueous FIs. The homogenization temperatures obtained in the hydrothermal

922

phases SD, MC, MQ and SB are even higher than the maximum burial temperatures

923

interpreted for the study area, including the present reservoir temperature. Therefore, the

924

Th data from SD, MC, MQ and SB represent further evidence of the hydrothermal

925

origin of these mineral phases, regardless their relation to specific magmatic periods.

926

Hydrocarbon FIs in these late-stage mineral phases suggest an atypical oil migration

39 927

(sensu Magoon and Dow, 1994) occurred in association with the hydrothermal fluid

928

flow. The absence of coexisting vapor- and liquid-rich FIs, diagnostic of phase-

929

separation effects (e.g., Goldstein and Reynolds, 1994; Jones et al., 1996; Moore et al.,

930

2001), indicates the lack of boiling processes within the studied Pre-Salt hydrothermal

931

system.

932

6. Conclusions

933

The overall geological and geochemical framework, combined with a specific

934

mineral paragenetic assemblage (saddle dolomite, macrocrystalline quartz, calcite, Sr-

935

barite, celestine, fluorite, dickite, sphalerite, galena, other metallic sulfides, and

936

bitumen) and FIs with corresponding range of salinities and homogenization

937

temperatures allowed us to recognize that the Pre-Salt carbonate reservoirs of the

938

northern Campos Basin were percolated by hydrothermal fluids chemically

939

(compositionally) comparable to those that formed in Mississippi Valley Type and

940

similar deposits.

941

The elevated homogenization temperatures (up to 152 °C) measured in

942

macrocrystalline calcite, mega-quartz, saddle dolomite and Sr-barite indicate

943

entrapment temperatures higher than the maximum values interpreted from the burial

944

history, confirming hydrothermal conditions during the late-stage alteration of the

945

studied reservoirs. The hydrothermal system characterized in northern Campos Basin

946

Pre-Salt reservoirs presents temperatures and salinities similar to Mississippi Valley and

947

Irish hydrothermal systems.

948

Saddle dolomite and macrocrystalline calcite precipitated from hydrothermal

949

fluids show lower δ18O values than the syngenetic and diagenetic carbonates. The

950

fascicular calcite aggregates, characteristic of Pre-Salt reservoirs, experienced strong

40 951

recrystallization under the action of the hydrothermal fluids and, therefore, present δ18O

952

values intermediate between the hydrothermal carbonates and the early diagenetic

953

spherulitic and microcrystalline calcite phases. All carbonates analyzed in the Pre-Salt

954

reservoirs of the northern Campos Basin present high 87Sr/86Sr, with lower ratios for the

955

hydrothermal saddle dolomite and macrocrystalline calcite than for the syngenetic,

956

diagenetic and recrystallized carbonates.

957

The high temperatures and salinities measured from the fluid inclusions, and the

958

isotopic data obtained from saddle dolomite, Sr-barite, macrocrystalline calcite and

959

mega-quartz indicate that the fault-focused hydrothermal system affecting the northern

960

Campos Basin Pre-Salt reservoirs probably involved mixing of fluids derived from

961

several sources, such as the interaction with the granitic-gneissic basement, the rift

962

sedimentary succession, the Late Cretaceous and Paleogene magmatism, and possibly

963

the rising and exhumation of the asthenosphere. The final hydrothermal fluid is a blend

964

in which end-members are not quantifiable.

965

The fluid inclusions, isotopic and petrographic data, such as the scarcity of

966

hydraulic breccia, integrated to the burial and thermal history of the study area are

967

strong evidence that the hydrothermal alteration of northern Campos Basin Pre-Salt

968

reservoirs occurred under relatively deep burial situation of more than 2 kilometers. The

969

studied hydrothermal alterations had strong impact on the porosity, permeability, and

970

heterogeneity, contributing, together with the associated fracturing, to the production

971

performance of the Pre-Salt reservoirs.

972

Acknowledgements

973

The results and interpretations of this paper are part of the PhD research project

974

of BEML, funded by Petróleo Brasileiro S.A. - Petrobras. The authors wish to thank

41 975

Petrobras, for supporting this study and for the opportunity to publish this paper. In

976

particular, we are extremely grateful to Gustavo Garcia, who provided useful insight

977

and discussions about the burial-thermal history. We are grateful for the analytical and

978

technical assistance provided by Petrobras Research Center (CENPES), the Federal

979

University of Goiás (UFG) Regional Center for Technological Development and

980

Innovation (CRTI), the University of Brasília (UnB) Geochronology Laboratory, and

981

the University of São Paulo (USP). We would also like to thank the support of the

982

Graduate Geosciences Program of Rio Grande do Sul Federal University (UFRGS). The

983

authors wish also to thank the reviewers for their constructive suggestions that helped

984

improve the manuscript.

985

References

986

Al-Aasm, 2003. Origin and characterization of hydrothermal dolomite in the Western

987

Canada Sedimentary Basin. Journal of Geochemical Exploration 78–79, 9–15.

988

Allègre, C.J., Louvat, P., Gaillardet, J., Meynadier, L., Rad, S., Capmas, F., 2010. The

989

fundamental role of island arc weathering in the oceanic Sr isotope budget. Earth and

990

Planetary Science Letters 292 (1–2), 51–56.

991

Allen, D.G., Seyfried Jr., W.E, 2004. Serpentinization and heat generation: constraints

992

from Lost City and Rainbow hydrothermal systems. Geochimica et Cosmochimica

993

Acta 68 (6), 1347–1354.

994

Alvarenga, R.S., Iacopini, D., Kuchle, J., Scherer, C.M.S., Goldberg, K., 2016. Seismic

995

characteristics and distribution of hydrothermal vent complexes in the Cretaceous

996

offshore rift section of the Campos Basin, offshore Brazil. Marine and Petroleum

997

Geology 74, 12–25.

42 998 999

Anderson, G.M., 1991. Organic maturation and ore precipitation in Southeast Missouri. Economic Geology 86 (5), 909–926.

1000

Ando, A., 2015. Intersite discrepancy in the amplitude of marine negative δ13C

1001

excursion at the onset of early Aptian oceanic anoxic event 1a: Reconciliation

1002

through Sr isotopic screening of peculiar diagenetic overprint on the Pacific

1003

reference section (Deep Sea Drilling Project Site 463). In: Neal, C.R., Sager, W.W.,

1004

Sano, T., Erba, E. (Eds.), Environmental consequences of Ontong Java Plateau and

1005

Kerguelen Plateau volcanism. Geological Society of America Special Paper 511. The

1006

Geological Society of America, Boulder, USA, pp. 329–339.

1007

Armelenti, G., Goldberg, K., Kuchle J., De Ros, L.F., 2016. Deposition, diagenesis and

1008

reservoir potential of non-carbonate sedimentary rocks from the rift section of

1009

Campos Basin, Brazil. Petroleum Geoscience 22 (3), 223–239.

1010 1011

Banner, J.L., 1995. Application of the trace element and isotope geochemistry of strontium to studies of carbonate diagenesis. Sedimentology 42 (5), 805–824.

1012

Baumgarten, C.S., Dutra, A.J.C., Scuta, M.S., Figueiredo, M.V.L., Sequeira, M.F.P.B.,

1013

1988. Coquinas da Formação Lagoa Feia, Bacia de Campos: evolução da geologia de

1014

desenvolvimento. Boletim de Geociências da Petrobras 2 (1), 27–36.

1015

Bellanca, A., Censi, P., Di Salvo, P., Neri, R., 1984. Textural, Chemical and Isotopic

1016

Variations Induced by Hydrothermal Fluids on Mesozoic Limestones in

1017

Northwestern Sicily. Mineralium Deposita 19, 78–85.

1018

Bertani, R.T., Carozzi, A.V., 1985a. Lagoa Feia Formation (Lower Cretaceous),

1019

Campos Basin, offshore Brazil: rift valley stage lacustrine carbonate reservoirs, I.

1020

Journal of Petroleum Geology 8, 37–58.

43 1021

Bertani, R.T., Carozzi, A.V., 1985b. Lagoa Feia Formation (Lower Cretaceous) Campos

1022

Basin, offshore Brazil: rift valley type lacustrine carbonate reservoirs, II. Journal of

1023

Petroleum Geology 8, 199–220.

1024

Bertrand, R., Chagnon, A., Malo, M., Duchaine, Y., Lavoie, D., Savard, M.M., 2003.

1025

Sedimentologic, diagenetic and tectonic evolution of the Saint-Flavien gas reservoir

1026

at the structural front of the Quebec Appalachians. Bulletin of Canadian Petroleum

1027

Geology 51 (2), 126–154.

1028

Biehl, B.C., Reuning, L., Schoenherr, J., Lüders, V., Kukla, P.A., 2016. Impacts of

1029

hydrothermal dolomitization and thermochemical sulfate reduction on secondary

1030

porosity creation in deeply buried carbonates: a case study from the Lower Saxony

1031

Basin, northwest Germany. American Association of Petroleum Geologists Bulletin

1032

100 (4), 597–621.

1033

Bodin, S., Meissner, P., Janssen, N.M.M., Steuber, T., Mutterlose, J., 2015. Large

1034

igneous provinces and organic carbon burial: Controls on global temperature and

1035

continental weathering during the Early Cretaceous. Global and Planetary Change

1036

133, 238–253.

1037 1038

Bodnar, R.J., 2003. Reequilibration of fluid inclusions. Fluid inclusions: Analysis and Interpretation, pp. 213–230.

1039

Bodnar, R.J., Reynolds, T.J., Kuehn, C.A., 1985. Fluid-inclusion systematics in

1040

epithermal systems. In: Berger, B., Bethke, R. (Eds.), Geology and Geochemistry of

1041

Epithermal Systems. Society of Economic Geologists, Reviews in Economic

1042

Geology 2, pp. 73–97.

1043

Bodnar, R.J., Vityk, M.O., 1994. Interpretation of microthermometric data for H2O-

1044

NaCl fluid inclusions. In: De Vivo, B., Frezzotti, M.L. (Eds.), Fluid Inclusions in

1045

Minerals, Methods and Applications. Virginia Tech Blackburg, pp. 117–130.

44 1046

Boillot, G., Recq, M., Winterer, E.L., Meyer, A.W., Applegate, J., Baltuck, M., Bergen,

1047

J.A., Comas, M.C., Davies, T.A., Dunham, K., Evans, C.A., Girardeau, J., Goldberg,

1048

G., Haggerty, J., Jansa, L.F., Johnson, J.A., Kasahara, J., Loreau, J.P., Luna-Sierra,

1049

E., Moullade, M., Ogg, J., Sarti, M., Thurow, J.,Williamson, M., 1987. Tectonic

1050

denudation of the upper mantle along passive margins: a model based on drilling

1051

results (ODP leg 103, western Galicia margin, Spain). Tectonophysics 132, 335–342.

1052

Boreen, T., Colquhoun, K., 2001. Ladyfern, NEBC: major gas discovery in the

1053

Devonian Slave Point Formation. Canadian Society of Petroleum Geologists, Annual

1054

Convention, Abstracts, 112–115.

1055

Boreen, T., Davies, G.R., 2004. Hydrothermal dolomite and leached limestones in a

1056

TCF gas play: the Ladyfern Slave Point reservoir, NEBC. In: McAuley, R. (Ed.),

1057

Dolomites - the spectrum: mechanisms, models, reservoir development. Canadian

1058

Society of Petroleum Geologists, Seminar and Core Conference, June 13–15,

1059

Calgary, Alberta, p. 17.

1060

Bottinga, Y., 1969. Calculated fractionation factors for carbon and hydrogen isotope

1061

exchange in the system calcite-carbon dioxide-graphite-methane-hydrogen-water

1062

vapor. Geochimica et Cosmochimica Acta 33, 49–64.

1063

Burke, W.H., Denison, R.E., Hetherington, E.A., Koepnick, R.B., Nelson, H.F., Otto,

1064

J.B., 1982. Variation of seawater 87Sr/86Sr throughout Phanerozoic time. Geology 10

1065

(10), 516–519.

1066

Callot, J.P., Breesch, L., Guilhaumou, N., Roure, F., Swennen, R., Vilasi, N., 2010.

1067

Paleofluids characterization and fluid flow modelling along a regional transect in

1068

Northern United Arab Emirates (UAE). Arabian Journal of Geosciences 3, 413–437.

1069 1070

Cantrell, D.L., Swart, P.K., Hagerty, R.M., 2004. Genesis and characterization of dolomite, Arab-D reservoir, Ghawar field, Saudi Arabia. GeoArabia 9 (2), 1–26.

45 1071

Castro, J.C., 2006. Evolução dos conhecimentos sobre as coquinhas-reservatório da

1072

Formação Lagoa Feia no trend Badejo-Linguado-Pampo, Bacia de Campos. Boletim

1073

de Geociências da Petrobras 25 (2), 175–186.

1074

Conliffe, J., Azmy, K., Gleeson, S.A., Lavoie, D., 2010. Fluids associated with

1075

hydrothermal dolomitization in St. George Group, western Newfoundland, Canada.

1076

Geofluids 10, 422–437.

1077

Coplen, T.B., Brand, W.A., Gehre, M., Gröning, M., Meijer, H.A.J., Toman, B.,

1078

Verkouteren, R.M., 2006. New guidelines for δ13C measurements. Analytical

1079

Chemistry 78 (7), 2439–2441.

1080

Corbella, M., Ayora, C., Cardellach, E., Soler, A., 2006. Reactive transport modeling

1081

and hydrothermal karst genesis: the example of the Rocabruna barite deposit (Eastern

1082

Pyrenees). Chemical Geology 233, 113–125.

1083

Coveney Jr., R.M., Ragan, V.M., Brannon, J.C., 2000. Temporal benchmarks for

1084

modeling Phanerozoic flow of basinal brines and hydrocarbons in the southern

1085

Midcontinent based on radiometrically dated calcite. Geology 28 (9), 795–798.

1086

Davies, G.R., 2002. Thermobaric dolomitization: Transient pressure driven processes

1087

and the role of boiling-effervescence in fabric/reservoir development (abs.).

1088

Canadian Society of Petroleum Geologists, 75th Anniversary Convention Abstracts

1089

Vol., p. 105.

1090

Davies, G.R., 2004. Hydrothermal (thermobaric) dolomitization: rock fabrics and

1091

organic petrology. In: McAuley, R. (Ed.), Dolomites - the spectrum: mechanisms,

1092

models, reservoir development. Canadian Society of Petroleum Geologists, Seminar

1093

and Core Conference, June 13–15, Calgary, Alberta, p. 17.

46 1094

Davies, G.R., Smith, L.B.J., 2006. Structurally controlled hydrothermal dolomite

1095

reservoir facies: an overview. American Association of Petroleum Geologists

1096

Bulletin 90 (11), 1641–1690.

1097 1098

Deines, P. 2002. The carbono isotope geochemistry of mantle xenoliths. Earth-Science Reviews 58 (1-4), 247–278.

1099

Dias, J.L., 1998. Análise sedimentológica e estratigráfica do Andar Aptiano em parte da

1100

Margem Leste do Brasil e no Platô das Malvinas: considerações sobre as primeiras

1101

incursões e ingressões marinhas do Oceano Atlântico Sul Meridional. (PhD Thesis).

1102

Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil, p. 208 (in

1103

Portuguese).

1104

Dias, J.L., Oliveira, J., Vieira, J., 1988. Sedimentological and stratigraphic analysis of

1105

the Lagoa Feia Formation, rift phase of Campos Basin, offshore Brazil. Revista

1106

Brasileira de Geociências 18, 252–260.

1107

Disnar, J.R., 1996. A comparison of mineralization histories for two MVT deposits,

1108

Treves and Les Malines (Causses Basin, France), based on the geochemistry of

1109

associated organic matter. Ore Geology Reviews 11 (1-3) 133–156.

1110 1111

Epstein, S., Mayeda, T., 1953. Variation of 18O content of waters from natural sources. Geochimica et Cosmochimica Acta, 4, 213–224.

1112

Farias, F., Szatmari, P., Bahniuk, A., França, A.B., 2019. Evaporitic carbonates in the

1113

pre-salt of Santos Basin – Genesis and tectonic implications. Marine and Petroleum

1114

Geology 105, 251–272

1115

Fontana, S., Nader, F.H., Morad, S., Ceriani, A., Al-Aasm, I.S., 2010. Diagenesis of the

1116

Khuff Formation (Permian-Triassic), northern United Arab Emirates. Arabian

1117

Journal of Geosciences 3 (4), 351–368.

47 1118 1119

Frost, B.R., Beard, J.S., 2007. On silica activity and serpentinization. Journal of Petrology, 48 (7), 1351–1368.

1120

Gasparrini, M., Bechstaedt, T., Boni, M., 2006. Massive hydrothermal dolomitization in

1121

the Southwestern Cantabrian Zone (Spain) and its relation to the late Variscan

1122

evolution. Marine Petroleum Geology 23, 543–568.

1123

Gilbert, G.K., 1875. Report on the geology of portions of Nevada, Utah, California, and

1124

Arizona. United States Geographic and Geological Surveys W. 100th Meridian 3,

1125

17–187.

1126

Girard, J.P., San Miguel, G., 2017. Evidence of High Temperature Hydrothermal

1127

Regimes in the Pre-Salt Series, Kwanza Basin, Offshore Angola. American

1128

Association of Petroleum Geologists Annual Convention and Exhibition. Houston,

1129

Texas, USA, Abstracts.

1130 1131

Goldstein, R.H., 2001. Fluid inclusions in sedimentary and diagenetic systems. Lithos 55, 159–193.

1132

Goldstein, R.H., Reynolds, T.J., 1994. Fluid Inclusion Microthermometry. In: Goldstein,

1133

R.H., Reynolds, T.J. (Eds.), Systematics of Fluid Inclusions in Diagenetic Minerals.

1134

Society for Sedimentary Geology, SEPM Short Courses 31, Tulsa, OK, USA, pp.

1135

87–121.

1136

Gomes, P.O., Kilsdonk, W., Grow, T., Minken, J., Barragan, R., 2011. Tectonic

1137

evolution of the Outer High of Santos Basin, southern São Paulo Plateau, Brazil, and

1138

implications for hydrocarbon exploration. In: Gao, D. (Ed.), Tectonics and

1139

sedimentation: Implications for petroleum systems. American Association of

1140

Petroleum Geologists, USA, Memoir 100, pp. 1–14.

48 1141

Greenwood, P.F., Brocks, J.J., Grice, K., Schwark, L., Jaraula, C.M.B., Dick, J.M.,

1142

Evans, K.A., 2013. Organic geochemistry and mineralogy. I. Characterization of

1143

organic matter associated with metal deposits. Ore Geology Reviews 50, 1–27.

1144 1145 1146 1147 1148 1149

Gregg, J.M., Sibley, D.F., 1984. Epigenetic dolomitization and the origin of xenotopic dolomite texture. Journal of Sedimentary Petrology 54, 908–931. Gregg, J.M., Sibley, D.F., 1987. Classification of dolomite rock textures. Journal of Sedimentary Petrology 57 (6), 967–975. Henley, R.W., Ellis, A.J., 1983. Geothermal systems ancient and modern: a geochemical review. Earth-Science Reviews 19, 1–50.

1150

Herlinger Jr., R., Zambonato, E.E., De Ros, L.F., 2017. Influence of diagenesis on the

1151

quality of lower cretaceous Pre-Salt lacustrine carbonate reservoirs from northern

1152

Campos Basin, offshore Brazil. Journal of Sedimentary Research 87, 1285–1313.

1153

Hesse, R., 1989. Silica diagenesis: origin of inorganic and replacement cherts. Earth-

1154

Science Reviews 26, 253–284.

1155

Hiemstra, E.J., Goldstein, R.H., 2015. Repeated injection of hydrothermal fluids into

1156

downdip carbonates: a diagenetic and stratigraphic mechanism for localization of

1157

reservoir porosity, Indian Basin Field, New Mexico, USA. In: Agar, S.M., Geiger, S.

1158

(Eds.), Fundamental controls on fluid flow in carbonates: current workflows to

1159

emerging technologies. Geological Society of London, Special Publications 406, pp.

1160

141–177.

1161

Holmes, A., 1928. The Nomenclature of Petrology. Ed. Van Nostrand-Reinhold, 284 p.

1162

Ilchik R.P., Brimhall G.H., Schull H.W., 1986. Hydrothermal maturation of indigenous

1163

organic matter at the Alligator Ridge Gold Deposits, Nevada. Economic Geology 81

1164

(1), 113–130.

49 1165 1166

Inoue, A., 1995. Formation of clay minerals in hydrothermal environments. In: Velde, B. (Ed.), Origin and Mineralogy of Clays. Springer-Verlag, Berlin, pp. 268−330.

1167

Jenkyns, H.C., Paull, C.K., Cummins, D.I., Fullagar, P.D., 1995. Strontium isotope

1168

stratigraphy of Lower Cretaceous atoll carbonates in the Mid-Pacific Mountains. In:

1169

Winterer, E.L., Sager, W.W., Firth, J.V., Sinton, J.M. (Eds.), Proceedings of the

1170

Ocean Drilling Program, Scientific Results 143. Texas A & M University, Ocean

1171

Drilling Program, College Station, USA, pp. 89–97.

1172

Jones, B., Renaut, R.W., Rosen, M.R., 1996. High-temperature (>90°C) calcite

1173

precipitation at Waikite Hot Springs, North Island, New Zealand. Journal of the

1174

Geological Society 153, 481–496.

1175

Khaska, M., Le Gal La Salle, C., Lancelot, J., team, ASTER, Mohamad, A., Verdoux,

1176

P., Noret, A., Simler, R., 2013. Origin of groundwater salinity (current seawater vs.

1177

saline deep water) in a coastal karst aquifer based on Sr and Cl isotopes. Case study

1178

of the La Clape massif (southern France). Applied Geochemistry 37, 212–227.

1179

King, B.D., Goldstein, R.H., 2018. History of hydrothermal fluid flow in the

1180

midcontinent,

1181

unconformities and porosity distribution. In: Armitage, P.J., Butcher, A.R., Churchill,

1182

J.M., Csoma, A.E., Hollis, C., Lander, R.H., Omma, J.E., Worden, R.H. (Eds.),

1183

Reservoir Quality of Clastic and Carbonate Rocks: Analysis, Modelling and

1184

Prediction. Geological Society of London, Special Publications 435, pp. 283–320.

1185

Kukla, P.A., Strozyk, F., Mohriak, W.U., 2018. South Atlantic salt basins – Witnesses

1186

USA:

the

relationship

between

inverted

thermal

structure,

of complex passive margin evolution. Gondwana Research 53, 41−57.

1187

Kumar, N., Danforth, A., Nuttall, P., Helwig, J., Bird, D.E., Venkatraman, S., 2013.

1188

From oceanic crust to exhumed mantle: a 40 year (1970–2010) perspective on the

1189

nature of crust under the Santos Basin, SE Brazil. In: Mohriak, W.U., Danforth, A.,

50 1190

Post, P.J., Brown, D.E., Tari, G.C., Nemcok, M., Sinha, S.T. (Eds.), Conjugate

1191

Divergent Margins. Geological Society of London, Special Publications 369, pp.

1192

147–165.

1193

Kusznir, N.J., Karner, G.D., 2007. Continental lithospheric thinning and breakup in

1194

response to upwelling divergent mantle flow: application to the Woodlark,

1195

Newfoundland and Iberia margins. In: Karner, G.D., Manatschal, G., Pinheiro, L.M.

1196

(Eds.), Imaging, Mapping and Modelling Continental Lithosphere Extension and

1197

Breakup. Geological Society of London, Special Publications 282, pp. 389–419.

1198

Leach, D.L., 1994. Genesis of the Ozark Mississippi Valley-type metallogenic province.

1199

In: Fontboté, L., Boni, M., (Eds.), Sediment Hosted Zn-Pb Ores. Springer-Verlag, pp.

1200

104–138.

1201

Leach, D.L., Bradley, D., Lewchuk, M.T., Symons, D.T.A., De Marsily, G., Brannon, J.,

1202

2001. Mississippi Valley-type lead-zinc deposits through geological time:

1203

Implications from recent age-dating research. Mineralium Deposita 36, 711–740.

1204

Lepley, S., Piccoli, L., Chitale, V., Kelley, I., Quest, M., 2017. The Importance of

1205

Understanding Diagenesis for the Development of Pre-Salt Lacustrine Carbonates.

1206

American Association of Petroleum Geologists Annual Convention and Exhibition.

1207

Houston, Texas, USA, Abstracts.

1208 1209

Leyden, R., Asmus, H., Zembruscki, S., Bryan, G., 1976. South Atlantic diapiric structures. American Association of Petroleum Geologists Bulletin 60 (2), 196–212.

1210

Lima, B.E.M., De Ros, L.F., 2019. Deposition, diagenetic and hydrothermal processes

1211

in Aptian Pre-Salt lacustrine carbonate reservoirs of the northern Campos Basin,

1212

offshore Brazil. Sedimentary Geology 383, 55–81.

1213

Lonnee, J., Machel, H.G., 2006. Pervasive dolomitization with subsequent hydrothermal

1214

alteration in the Clarke Lake gas field, Middle Devonian Slave Point Formation,

51 1215

British Columbia, Canada. American Association of Petroleum Geologists Bulletin

1216

90, 1739–1761.

1217

Luczaj, J.A., 2006. Evidence against the Dorag (mixing-zone) model for dolomitization

1218

along the Wisconsin arch – A case for hydrothermal diagenesis. American

1219

Association of Petroleum Geologists Bulletin 90 (11), 1719–1738.

1220

Luczaj, J.A., Harrison, W.B., Williams, N.S., 2006. Fractured hydrothermal dolomite

1221

reservoirs in the Devonian Dundee Formation of the central Michigan Basin.

1222

American Association of Petroleum Geologists Bulletin 90 (11), 1787–1801.

1223 1224

Machel, H.G., 2001. Bacterial and thermochemical sulfate reduction in diagenetic settings - old and new insights. Sedimentary Geology 140, 143–175.

1225

Machel, H.G., 2004. Concepts and models of dolomitization: a critical reappraisal. In:

1226

Braithwaite, C.J.R., Rizzi, G., Darke, G. (Eds.), The geometry and petrogenesis of

1227

dolomite hydrocarbon reservoirs, vol. 235, pp. 7–63. Geological Society of London,

1228

Special Publications.

1229 1230

Machel, H.G., Lonnee, J., 2002. Hydrothermal dolomite: a product of poor definition and imagination. Sedimentary Geology 152, 163–171.

1231

Machel, H.G., Mountjoy, E.W., 1987. General constraints on extensive pervasive

1232

dolomitization - and their application to the Devonian carbonates of western Canada.

1233

Bulletin of Canadian Petroleum Geology 35 (2), 143–158.

1234

Magoon, L.B., Dow, W.G., 1994. The Petroleum System. In: Magoon, L.B, Dow, W.G.

1235

(Eds.), Chapter 1: The petroleum system - from source to trap. American Association

1236

of Petroleum Geologists, USA, Memoir 60, pp. 3–24.

1237 1238

Manatschal, G., Bernoulli, D., 1999. Architecture and tectonic evolution of nonvolcanic margins: present day Galicia and ancient Adria. Tectonics 18, 1099–1119.

52 1239

Mansurbeg, H., Morad, D., Othmana, R., Morad, S., Ceriani, A., Al-aasm, I., Kolo, K.,

1240

Spirov, P., Proust, J.N., Preat, A., Koyi, H., 2016. Hydrothermal dolomitization of

1241

the Bekhme formation (Upper Cretaceous), Zagros Basin, Kurdistan Region of Iraq:

1242

Record of oil migration and degradation. Sedimentary Geology 341, 147–162.

1243 1244

Mattews, A.K., Katz, A., 1977. Oxygen isotope fractionation during the dolomitization of calcium carbonate. Geochimica et Cosmochimica Acta 41, 1431–1438.

1245

McArthur, J.M., Howarth, R.J., Shields, G.A., 2012. Chapter 7 – Strontium Isotope

1246

Stratigraphy. In: Gradstein, F.M., Ogg, J.G., Schmitz, M.D., Ogg, G.M. (Eds.), The

1247

Geologic Time Scale. Elsevier, Boston, USA, pp. 127–144.

1248

Mello, M.R., Mohriak, W.U., Koutsoukos, E.A.M., Bacoccoli, G., 1994. Selected

1249

petroleum systems in Brazil. In: Magoon, L.B., Dow, W.G. (Eds.), The petroleum

1250

system - from source to trap. American Association of Petroleum Geologists, USA,

1251

Memoir 60, pp. 499–512.

1252 1253 1254

Meyers, W.J., 1977. Chertification in the Mississippian Lake Valley Formation, Sacramento Mountains, New Mexico. Sedimentology 24 (1), 75–105. Meyers, W.J., James, A.T., 1978. Stable isotopes of cherts and carbonate cements in the

1255

Lake

Valley

Formation

1256

Sedimentology 25, 105–124.

(Mississipian),

Sacramento

Mts.,

New

Mexico.

1257

Mizusaki, A.M.P., Petrini, R., Bellieni, P., Comin-Chiaramonti, P., Dias, J.L., De Min,

1258

A., Piccirillo, E.M., 1992. Basalt magmatism along the passive continental margin of

1259

SE Brazil (Campos Basin). Contributions to Mineralogy and Petrology 111 (2), 143–

1260

160.

1261

Moody, J.B., 1976. Serpentinization: a review. Lithos 9 (2), 125–138.

53 1262

Moore, J.N., Norman, D.I., Kennedy, B.M., 2001. Fluid inclusion gas compositions

1263

from an active magmatic–hydrothermal system: a case study of The Geysers

1264

geothermal field, USA. Chemical Geology 173, 3–30.

1265 1266

Morey, G.W., Niggli, P., 1913. The hydrothermal formation of silicates, a review. Journal of the American Chemical Society 35, 1086–1130.

1267

Muniz, M.C., Bosence, D.W.J., 2015. Pre-Salt microbialites from the Campos Basin

1268

(offshore Brazil): image log facies, facies model and cyclicity in lacustrine

1269

carbonates. In: Bosence, D.W.J., Gibbons, K.A., LeHeron, D.P., Morgan, W.A.,

1270

Pritchard, T., Vining, B.A. (Eds.), Microbial carbonates in space and time:

1271

implications for global exploration and production. Geological Society, London,

1272

Special Publications 418, pp. 221–242.

1273 1274

Namy, J.N., 1974. Early diagenetic cherts in the Marble Falls Group (Pennsylvanian) of central Texas. Journal of Sedimentary Petrology 44, 1262–1268.

1275

Neilson, J.E., Oxtoby, N.H., 2008. The relationship between petroleum, exotic cements

1276

and reservoir quality in carbonates - a review. Marine Petroleum Geology 25, 778–

1277

790.

1278 1279

O’Neil, J.R., Clayton, R.N., Mayeda, T.K., 1969. Oxygen isotope fractionation in divalent metal carbonates. Journal of Chemical Physics 51, 5547–5558.

1280

Ostendorf, J., Henjes-Kunst, F., Mondillo, N., Boni, M., Schneider, J., Gutzmer, J.,

1281

2015. Formation of Mississippi Valley–type deposits linked to hydrocarbon

1282

generation in extensional tectonic settings: Evidence from the Jabali Zn-Pb-(Ag)

1283

deposit (Yemen). Geology 43 (12), 1055–1058.

1284

Packard, J.J., Al-Aasm, I.S., Samson, I., Berger, Z., Davies, J., 2001. A Devonian

1285

hydrothermal chert reservoir: the 225 bcf Parkland field, British Columbia, Canada.

1286

American Association of Petroleum Geologists Bulletin 85 (1), 51–84.

54 1287

Paradis, S., Hannigan, P., Dewing, K., 2007. Mississippi Valley-type lead-zinc deposits.

1288

In: Goodfellow, W.D. (Ed.), Mineral Deposits of Canada: A Synthesis of Major

1289

Deposit-Types, District Metallogeny, the Evolution of Geological Provinces, and

1290

Exploration Methods. Geological Association of Canada, Mineral Deposits Division,

1291

Special Publication 5, pp. 185–203.

1292

Parnell, J., Baron, M., Mann, P., Carey, P., 2003. Oil migration and bitumen formation

1293

in a hydrothermal system, Cuba. Journal of Geochemical Exploration 78-79, 409–

1294

415.

1295

Peron-Pinvidic, G., Manatschal, G., Osmundsen, P.T., 2013. Structural comparison of

1296

archetypal Atlantic rifted margins: A review of observations and concepts. Marine

1297

and Petroleum Geology 43, 21–47.

1298

Pietzsch, R., Oliveira, D.M., Tedeschi, L.R., Queiroz Neto, J.V., Figueiredo, M.F.,

1299

Vazquez, J.C., Souza, R.S., 2018. Palaeohydrology of the Lower Cretaceous pre-salt

1300

lacustrine system, from rift to post-rift phase, Santos Basin, Brazil. Palaeogeography,

1301

Palaeoclimatology, Palaecology 507, 60–80.

1302

Pinto, V.H.G., Manatschal, G., Karpoff, A.M., Viana, A., 2015. Tracing mantle-reacted

1303

fluids in magma-poor rifted margins: The example of Alpine Tethyan rifted margins.

1304

Geochemistry Geophysics Geosystems 16, 3271–3308.

1305

Pinto, V.H.G., Manatschal, G., Karpoff, A.M., Ulrich, M., Viana, A.R., 2017. Seawater

1306

storage and element transfer associated with mantle serpentinization in magma-poor

1307

rifted margins: a quantitative approach. Earth and Planetary Science Letters 459,

1308

227–237.

1309

Poros, Z., Jagniecki, E., Luczaj, J., Kenter, J., Gal, B., Correa, T.S., Ferreira, E.,

1310

McFadden, K.A., Elifritz, A., Heumann, M., Johnston, M., Matt, V., 2017. Origin of

1311

silica in Pre-Salt carbonates, Kwanza Basin, Angola. AAPG Annual Convention and

55 1312

Exhibition, Houston, Texas, April 2–5. USA: AAPG Datapages/Search and

1313

Discovery, article #90291.

1314

Proskurowski, G., Lilley, M.D., Kelley, D.S., 2006. Low temperature volatile

1315

production at the Lost City Hydrothermal Field, evidence from a hydrogen stable

1316

isotope geothermometer. Chemical Geology, 229 (4), 331−343.

1317

Reed, M.H., 1997. Hydrothermal alteration and its relationship to ore fluid composition

1318

In: Barnes, H.L. (Ed.), Geochemistry of Hydrothermal Ore Deposits, 3rd edition.

1319

John Wiley and Sons, pp. 303–365.

1320 1321

Rodrigues, R., 2005. Chemostratigraphy. In: Koutsoukos, E.A.M. (Ed.), Applied Stratigraphy. Springer, Dordrecht, Netherlands, pp. 165–178.

1322

Rodriguez, C.R., Jackson, C.A.L., Rotevatn, A., Bell, R.E., Francis, M., 2018. Dual

1323

tectonic-climatic controls on salt giant deposition in the Santos Basin, offshore Brazil.

1324

Geosphere 14, 215–242.

1325 1326

Roedder, E., 1984. Fluid inclusions. Mineralogical Society of America, Reviews in Mineralogy 12, 646 p.

1327

Roedder, E., Bodnar, R.J., 1997. Fluid inclusion studies of hydrothermal ore deposits.

1328

In: Barnes, H.L. (Ed.), Geochemistry of Hydrothermal Ore Deposits. Wiley, New

1329

York, pp. 657–697.

1330

Sabato Ceraldi, T., Green, D., 2016. Evolution of the South Atlantic lacustrine deposits

1331

in response to Early Cretaceous rifting, subsidence and lake hydrology. In: Sabato

1332

Ceraldi, R.A., Hodgkinson, T., Backe, G. (Eds.), Petroleum Geoscience of the West

1333

Africa Margin. Geological Society, London, Special Publication 438, pp. 77–98.

1334

Saller, A., Rushton, S., Buambua, L., Inman, K., Mcneil, R., Dickson, J.A.D., 2016.

1335

Pre-Salt stratigraphy and depositional systems in the Kwanza Basin, offshore Angola.

1336

American Association of Petroleum Geologists Bulletin 100, 1135–1164.

56 1337

Schoenherr, J., Littke, R., Urai, J.L., Kukla, P.A., Rawahi, Z., 2007. Polyphase thermal

1338

evolution in the Infra-Cambrian Ara Group (South Oman Salt Basin) as deduced by

1339

maturity of solid reservoir bitumen. Organic Geochemistry 38 (8), 1293–1318.

1340 1341

Sharp, Z., 2007. Principles of Stable Isotope Geochemistry. Pearson Prentice Hall, Upper Saddle River, New Jersey, 344 p.

1342

Sibson, R.H., 1990. Faulting and fluid flow. In: Nesbitt, B.E. (Ed.), Fluids in

1343

tectonically active regimes of the continental crust. Mineralogical Association of

1344

Canada, Short Course Handbook 18, chapter 4, pp. 93–132.

1345 1346

Simoneit, B.R.T., 1990. Petroleum generation, an easy and widespread process in hydrothermal systems: an overview. Applied Geochemistry 5 (1-2), 3–15.

1347

Simoneit, B.R.T., 2018. Hydrothermal Petroleum. In: Wilkes, H. (Ed.), Hydrocarbons,

1348

Oils and Lipids: Diversity, Origin, Chemistry and Fate. Handbook of Hydrocarbon

1349

and Lipid Microbiology. Springer, p. 800.

1350

Skinner, B.J., 1997. Hydrothermal mineral deposits: what we do and don't know. In:

1351

Barnes, H.L. (Ed.), Geochemistry of hydrothermal ore deposits, 3rd edition. New

1352

York: Wiley and Sons, pp. 1–29.

1353

Smith, L.B.J., 2006. Origin and reservoir characteristics of Upper Ordovician Trenton–

1354

Black River hydrothermal dolomite reservoirs in New York. American Association

1355

of Petroleum Geologists Bulletin 90 (11), 1691–1718.

1356 1357

Stearns, N.D., Stearns, H.T., Waring, G.A., 1935. Thermal springs in the United States. United States Geological Survey, Water Supply Paper 679-B, 59-191.

1358

Steele-MacInnis, M., Bodnar, R.J., Naden, J., 2011. Numerical model to determine the

1359

composition of H2O-NaCl-CaCl2 fluid inclusions based on microthermometric and

1360

microanalytical data. Geochimica et Cosmochimica Acta 75, 21–40.

57 1361

Suchý, V., Dobes, P., Sýkorová, I., Machovic, V., Stejskal, M., Kroufek, J., Chudoba, J.,

1362

Matejovský, L., Havelcová, M., Maysová, P., 2010. Oil-bearing inclusions in vein

1363

quartz and calcite and, bitumens in veins: Testament to multiple phases of

1364

hydrocarbon migration in the Barrandian basin (lower Palaeozoic), Czech Republic.

1365

Marine and Petroleum Geology 27 (1), 285–297.

1366

Teboul, P.A., Durlet, C., Gaucher, E.C., Virgone, A., Girard, J.P., Curie, J., Lopez, B.,

1367

Camoin, G.F., 2016. Origins of elements building travertine and tufa: New

1368

perspectives provided by isotopic and geochemical tracers. Sedimentary Geology

1369

334, 97–114.

1370

Teboul, P.A., Durlet, C., Girard, J.P., Dubois, L., Miguel, G.S., Virgone, A., Gaucher,

1371

E.C., Camoin, G., 2019. Diversity and origin of quartz cements in continental

1372

carbonates: example from the Lower Cretaceous rift deposits of the South Atlantic

1373

margin. Applied Geochemistry 100, 22–41.

1374

Teboul, P.A., Kluska, J.M., Marty, N.C.M., Debure, M., Durlet, C., Virgone, A.,

1375

Gaucher, E.C., 2017. Volcanic rock alterations of the Kwanza Basin, offshore

1376

Angola - Insights from an integrated petrological, geochemical and numerical

1377

approach. Marine and Petroleum Geology 80, 394–411.

1378

Tedeschi, L.R., 2017. Lower Cretaceous climate records and the correlation between

1379

marine and lacustrine settings (Europe and South America). (PhD Thesis). University

1380

of Oxford, Oxford, UK, p. 343.

1381

Thompson, D.L., Stilwell, J., Hall, M., 2015. Lacustrine carbonate reservoirs from Early

1382

Cretaceous rift lakes of Western Gondwana: Pre-Salt coquinas of Brazil and West

1383

Africa. Gondwana Research 28, 26–51.

58 1384

Tosca, N.J., Wright, V.P., 2014. The formation and diagenesis of Mg-clay minerals in

1385

lacustrine carbonate reservoirs. AAPG Annual Convention and Exhibition, Houston,

1386

Texas, April 6-9. USA: AAPG Datapages/Search and Discovery, article #51002.

1387

Tupinambá, M., Heilbron, M., Valeriano, C., Júnior, R.P., Dios, F.B., Machado, N.,

1388

Silva, L.G.D.E., Almeida, J.C.H., 2012. Juvenile contribution of the Neoproterozoic

1389

Rio Negro Magmatic Arc (Ribeira Belt, Brazil): Implications for Western Gondwana

1390

amalgamation. Gondwana Research 21 (2–3), 422–438.

1391

Unternehr, P., Péron-Pinvidic, G., Manatschal, G., Sutra, E., 2010. Hyper-extended

1392

crust in the South Atlantic: in search of a model. Petroleum Geoscience 16, 207–215.

1393

Utada, M, 1980. Hydrothermal alterations related to igneous activity in Cretaceous and

1394

Neogene formations of Japan. Mining Geology, Special Issue 8, 67–83.

1395

Ventura, G.T., Simoneit, B.R.T., Nelson, R.K., Reddy, C.M., 2012. The composition,

1396

origin and fate of complex mixtures in the maltene fractions of hydrothermal

1397

petroleum assessed by comprehensive two-dimensional gas chromatography.

1398

Organic Geochemistry 45, 48–65.

1399

Vieira de Luca, P.H., Matias, H., Carballo, J., Sineva, D., Pimentel, G.A., Tritlla, J.,

1400

Esteban, M., Loma, R., Alonso, J.L.A., Jiménez, R.P., Pontet, M., Martinez, P.B.,

1401

Vega, V., 2017. Breaking barriers and paradigms in presalt exploration: The Pão de

1402

Açúcar discovery (offshore Brazil). In: Merrill, R.K., Sternbach C.A. (Eds.), Giant

1403

fields of the decade 2000–2010. American Association of Petroleum Geologists,

1404

USA, Memoir 113, pp. 177–194.

1405 1406 1407 1408

White, D.E., 1957. Thermal waters of volcanic origin. Geological Society of America Bulletin 68, 1637–1658. Whitmarsh, R.B., Manatschal, G., Minshull, T.A., 2001. Evolution of magma-poor continental margins from rifting to sea-floor spreading. Nature 413, 150–153.

59 1409

Wierzbicki, R., Dravis, J.J., Al-Aasm, I., Harland, N., 2006. Burial dolomitization and

1410

dissolution of Upper Jurassic Abenaki platform carbonates, Deep Panuke reservoir,

1411

Nova Scotia, Canada. American Association of Petroleum Geologists Bulletin 90,

1412

1843–1861.

1413 1414

Wilkinson, J.J., 2001. Fluid inclusions in hydrothermal ore deposits. Lithos 55, 229−272.

1415

Wilkinson, J.J., 2010. A Review of Fluid Inclusion Constraints on Mineralization in the

1416

Irish Ore Field and Implications for the Genesis of Sediment-Hosted Zn-Pb Deposits.

1417

Economic Geology 105, 417–442.

1418

Wilson, M.I.J., Evans, M.J., Oxtoby, N.H., Nas, D.S., Donnelly, T., Thirwall, M., 2007.

1419

Reservoir quality, textural evolution, and origin of fault-associated dolomites.

1420

American Association of Petroleum Geologists Bulletin 91, 1247–1272.

1421

Wilson, N.S.F., Zentilli, M., 1999. The role of organic matter in the genesis of the El

1422

Soldado volcanic-hosted manto-type Cu deposit, Chile. Economic Geology 94 (7),

1423

1115–1136.

1424 1425 1426 1427

Winter, W.R., Jahnert, R.J., França, A.B., 2007. Bacia de Campos. Boletim de Geociências da Petrobras 15, 511–529. Wright, V.P., 2011. Reservoir architectures in non-marine carbonates. AAPG Annual Convention and Exhibition, Proceedings, Houston, p. 75–89.

1428

Wright, V.P., 2012. Lacustrine carbonates in rift settings: the interaction of volcanic and

1429

microbial processes on carbonate deposition. In: Garland, J., Neilson, J.E., Laubach,

1430

S., Whidden, K.J. (Eds.), Advances in carbonate exploration and reservoir analysis.

1431

Geological Society of London, Special Publications 370, pp. 39–47.

60 1432

Wright, V.P., Barnett, A.J., 2014. Cyclicity and Carbonate-Silicate Gel Interactions in

1433

Cretaceous Alkaline Lakes. AAPG Annual Convention and Exhibition, Houston,

1434

Texas. Search and Discovery Article #51011.

1435

Wright, V.P., Barnett, A.J., 2015. An abiotic model for the development of textures in

1436

some South Atlantic Early Cretaceous lacustrine carbonates. In: Grotzinger, J.P.,

1437

James, N. (Eds.), Microbial carbonates in space and time: implications for global

1438

exploration and production. Geological Society of London, Special Publications 418,

1439

pp. 209–219.

1440

Wright, V.P., Tosca, N.J., 2016. Geochemical Model for the Formation of the Pre-Salt

1441

Reservoirs, Santos Basin, Brazil: Implications for Understanding Reservoir

1442

Distribution. American Association of Petroleum Geologists Annual Convention and

1443

Exhibition. Calgary, Alberta, Canada.

1444

Xu, K., Yu, B., Gong, H., Ruan, Z., Pan, Y., Ren, Y., 2015. Carbonate reservoirs

1445

modified by magmatic intrusions in the Bachu area, Tarim Basin, NW China.

1446

Geoscience Frontiers 6, 779–790.

1447

Yamamoto, K., Ishibashi, M., Takayanagi, H., Asahara, Y., Sato, T., Nishi, H., Iryu, Y.,

1448

2013. Early Aptian paleoenvironmental evolution of the Bab Basin at the southern

1449

Neo-Tethys margin: Response to global carbon-cycle perturbations across Ocean

1450

Anoxic Event 1a. Geochemistry, Geophysics, Geosystems 14 (4), 1104–1130.

1451

You, D., Han, J., Hu, W., Qian, Y., Chen, Q., Xi, B., Ma, H., 2018. Characteristics and

1452

formation mechanisms of silicified carbonate reservoirs in well SN4 of the Tarim

1453

Basin. Energy Exploration & Exploitation 36 (4), 820–849.

1454

Zalán, P.V., Severino, M.C.G., Rigoti, C.A., Magnavita, L.P., Bach, J.A., 2011. An

1455

Entirely New 3D-View of the Crustal and Mantle Structure of a South Atlantic

61 1456

Passive Margin - Santos, Campos and Espírito Santo Basins, Brazil. AAPG Annual

1457

Convention and Exhibition, Proceedings, Houston, 12p.

1458

Zhu, D., Meng, Q., Jin, Z., Liu, Q., Hu, W., 2015. Formation mechanism of deep

1459

Cambrian dolomite reservoirs in the Tarim basin, northwestern China. Marine and

1460

Petroleum Geology 59, 232–144.

1461 1462

62 1463

Figures and Tables

1464 1465

Figure 1. Location map of the study area in the northern Campos Basin, offshore

1466

southeast Brazil (modified from Dias et al., 1988; Lima and De Ros, 2019). The straight

1467

dashed lines define the limits of the Campos Basin with the Espírito Santo Basin to the

1468

North (Vitória High) and with the Santos Basin to the South (Cabo Frio High). The

1469

detail indicates the location of the four Petrobras wells used in the study.

1470 1471

Figure 2. Summarized Stratigraphic chart of the Lagoa Feia Group in Campos Basin,

1472

offshore Brazil (modified from Winter et al., 2007; Lima and De Ros, 2019). Wavy

1473

lines represent unconformities.

1474 1475

Figure 3. Detailed stratigraphic logs of the Coqueiros (rift section) and Macabu (sag

1476

phase) formations at Campos Basin, offshore Brazil, in three of the four wells used in

1477

the study. The stratigraphic datum is the top of the sag reservoir (base of evaporites

1478

from Ariri Formation, which is not shown). Dolomitization and silicification intensities

1479

are indicated by the size of the bars increasing to the left. GR: gamma ray curve; δ13C

1480

and δ18O: samples analyzed for oxygen and carbon isotopes; 87Sr/86Sr: samples analyzed

1481

for strontium isotopes; FI: samples analyzed for fluid inclusions. Red wavy lines

1482

represent unconformities.

1483 1484

Figure 4. Photomicrographs highlighting main features in rift and sag lithologic types,

1485

Campos Basin, Brazil. A) Rudstone of fragmented, partially dissolved bivalve bioclasts,

1486

with well-preserved interparticle porosity (blue) (plane-polarized light with uncrossed

1487

polarizers; //P). B) Stevensitic arenite with stevensite ooids and peloids strongly

63 1488

dissolved and silicified (white) (//P). C) Fascicular calcite crusts (CFC) intercalated

1489

with granular deposits replaced by microcrystalline dolomite (MD) (crossed polarizers;

1490

XP). D) Calcite spherulites (CS) in Mg-clay matrix replaced by microcrystalline

1491

dolomite (MD) (XP). E) Planar to wavy laminite with millimetric levels defined by

1492

predominance of clay or microcrystalline dolomite (//P). F) Intraclastic grainstone with

1493

interparticle porosity partially cemented by calcite and dolomite (//P).

1494 1495

Figure 5. Paragenetic sequences interpreted for the sag fascicular calcite crusts from

1496

northern Campos Basin, Brazil. The paragenetic sequence is divided into syngenetic

1497

(Syn), eodiagenetic, mesodiagenetic and hydrothermal phases (modified from Lima and

1498

De Ros, 2019). The thickness of the lines corresponds to the intensity or abundance of

1499

the processes and products. The red lines represent porosity loss, and the blue lines

1500

porosity increase. Tha and Tho show the mean homogenization temperatures of primary

1501

aqueous and oil fluid inclusions, respectively (minimum and maximum limits in

1502

brackets).

1503 1504

Figure 6. Schematic representation of the main syngenetic, diagenetic and hydrothermal

1505

processes of formation and alteration of the studied Aptian Pre-Salt deposits, Campos

1506

Basin, offshore Brazil (modified from Lima and De Ros, 2019). (A) Syngenetic crust of

1507

coalesced fascicular calcite aggregates covering syngenetic Mg-clay matrix partially

1508

replaced and displaced by calcite spherulites. B) Microcrystalline and blocky dolomite

1509

partially replacing the Mg-clay matrix, calcite spherulites, and fascicular aggregates.

1510

Dolomite rhombohedra partially filling inter-aggregate pores. Dissolution of the

1511

syngenetic Mg-clay matrix and part of the dolomitized calcite aggregates. C)

1512

Microcrystalline quartz and chalcedony spherulites partially to fully replacing

64 1513

preexisting constituents, and filling primary and secondary porosity. D) Silicification of

1514

some dolomitized fascicular and spherulitic aggregates. Svanbergite as pseudocubic

1515

crystals disseminated in microcrystalline silica or within chalcedony spherulites.

1516

Recrystallization of fascicular calcite and initiation of fracturing process. E) Intense

1517

silicification, fracturing and dissolution. Saddle dolomite partially filling fracture and

1518

vugular pores. F) Macrocrystalline calcite, radiated Sr-barite and celestine, fibro-

1519

radiated zeolite, euhedral to subhedral pyrite, chalcopyrite, galena and sphalerite, and

1520

massive bitumen filling the secondary pores. The numbering is in accordance with the

1521

paragenetic sequence in Figure 5.

1522 1523

Figure 7. Photomicrographs showing aspects of sag lithologic types affected by

1524

dolomitization and silicification. A) Incipiently-dissolved saddle dolomite (SD) filling

1525

vugular pore (XP). B) Macrocrystalline calcite (MC) and prismatic, elongated and

1526

oriented celestine (Clt) filling fractures in brecciated dolostone (Dol) (XP). C)

1527

Cathodoluminescence (CL) image of zoned saddle dolomite crystals (SD) partially

1528

dissolved and replaced by microcrystalline silica (MS). D) Zoned saddle dolomite

1529

intensely silicified (SSD), macrocrystalline quartz (MQ), and spherulitic chalcedony

1530

(ChS) completely filling vugular pore (XP). E) Drusiform quartz (DQ) and spherulitic

1531

chalcedony (ChS) cementing vugular porosity in partially dolomitized and silicified

1532

protolith (XP). F) Radial prismatic quartz (PQ), prismatic Sr-barite (SB),

1533

macrocrystalline calcite (MC) and euhedral pyrite (blue arrows) filling vugular porosity

1534

(XP).

1535 1536

Figure 8. Photomicrographs from petrographic features of some hydrothermal mineral

1537

phases identified in this study. A) Prismatic Sr-barite (SB), saddle dolomite (SD) and

65 1538

macrocrystalline calcite (MC) (stained red) filling vugular pore (XP). B) CL image of

1539

zoned macrocrystalline calcite (MC) and saddle dolomite (SD) filling vugular porosity.

1540

C) Elongated prismatic Sr-barite (SB) and euhedral macrocrystalline calcite (MC)

1541

partially filling vugular porosity (backscattered electrons image; BSE). D) Sr-barite

1542

(SB) filling vugular porosity in dolostone (Dol) (BSE). E) Fibro-radiated zeolite (Ze)

1543

and euhedral macrocrystalline calcite (MC) filling partially vugular pore in chert

1544

composed by microcrystalline silica (MS) (BSE). F) Blocky dolomite (BD),

1545

macrocrystalline calcite (MC) and bitumen (Bi) cementing vugular porosity in rock with

1546

totally recrystallized and partially dolomitized calcite spherulites (CS) (//P).

1547 1548

Figure 9. Photomicrographs showing examples of fluid inclusions (FIs) in double-

1549

polished sections (120 µm) showing primary aqueous and petroleum fluid inclusions

1550

(blue and red arrows, respectively) from Campos Basin Pre-Salt, Brazil. Liquid-rich

1551

aqueous inclusions hosted in: A) Macrocrystalline calcite, B) Quartz, and C) Sr-barite

1552

(//P). D) Photomicrograph shows a saddle dolomite growth zones bearing aligned

1553

primary petroleum inclusions (//P). E) Photomicrograph in (//P) and F) ultraviolet

1554

fluorescence image of saddle dolomite hosting primary aqueous and pseudo-secondary

1555

petroleum fluid inclusions.

1556 1557

Figure 10. Homogenization temperature (°C) histograms of aqueous and petroleum

1558

(hatched) fluid inclusions hosted in syngenetic and diagenetic constituents from Pre-Salt

1559

sag and rift intervals, Campos Basin, offshore Brazil. (A) Primary fluid inclusions

1560

hosted in all analyzed constituents. (B) Primary fluid inclusions hosted in recrystallized

1561

calcite bioclasts. (C) Primary fluid inclusions hosted in recrystallized fascicular calcite.

1562

(D) Primary fluid inclusions hosted in recrystallized calcite spherulites. (E) Secondary

66 1563

fluid inclusions hosted in recrystallized calcite spherulites. The average formation

1564

temperature currently measured in the depth interval where the samples for fluid

1565

inclusions were obtained is 113.7 °C.

1566 1567

Figure 11. Salinities estimated for minerals phases from Pre-Salt rift and sag intervals,

1568

Campos Basin, Brazil. Histograms of the salinities (wt. % of equivalent in NaCl),

1569

computed from NaCl-H2O system, for primary and pseudo-secondary (A) and

1570

secondary (B) fluid inclusions hosted in saddle dolomite (orange), Sr-barite (black), and

1571

macrocrystalline quartz (brown) and calcite (red) hydrothermal phases. Note that the

1572

highest salinity values were obtained in the secondary fluid inclusions. The formation

1573

water salinity sampled in the rift reservoir is 27.2 wt. % NaCl eq. (*) Salinities of

1574

aqueous fluid inclusions hosted in recrystallized fascicular calcite calculated in wt. %

1575

NaCl-CaCl2 eq.

1576 1577

Figure 12. Homogenization temperature (°C) histograms of aqueous and petroleum

1578

(hatched) fluid inclusions hosted in hydrothermal mineral phases from Pre-Salt sag and

1579

rift intervals, Campos Basin, Brazil. In the left, primary and pseudo-secondary and in

1580

the right side secondary fluid inclusions obtained in all hydrothermal phases (A and B),

1581

Sr-barite (black) (C and D), macrocrystalline quartz (brown) (E and F), macrocrystalline

1582

calcite (red) (G and H), and saddle dolomite (orange) (I). Note that the homogenization

1583

temperatures obtained from oil fluid inclusions are consistently lower than those of the

1584

aqueous fluid inclusions in all mineral phases.

1585 1586

Figure 13. Carbonate isotope composition, Campos Basin Pre-Salt, offshore Brazil.

1587

Cross-plots of: A) and B) δ18O values versus δ13C values (both in ‰ VPDB) by

67 1588

lithofacies and wells. Note the gray polygons delimiting the rift (r) and sag (s) data from

1589

Campos (Dias, 1998; Rodrigues, 2005; Muniz and Bosence, 2015), Santos (Farias et al.,

1590

2019), and Kwanza Basins (Saller et al., 2016; Sabato Ceraldi and Green, 2016). C) and

1591

D) 87Sr/86Sr ratio versus δ18O values (in ‰ VPDB). E) 87Sr/86Sr ratio versus δ13C values

1592

(in ‰ VPDB). F)

1593

rocks compared to Early Cretaceous and Phanerozoic seawater (Jenkyns et al., 1995;

1594

McArthur et al., 2012; Yamamoto et al., 2013; Bodin et al., 2015; Ando, 2015) and Pre-

1595

Salt carbonate (Dias, 1998; Tedeschi, 2017; Pietzsch et al., 2018; Farias et al., 2019).

1596

Samples of microcrystalline and blocky calcite in recrystallized bioclasts (dark green

1597

triangles), recrystallized fascicular calcite (purple squares), microcrystalline and blocky

1598

calcite in laminite (light green diamonds), recrystallized calcite spherulites (blue

1599

squares), microcrystalline and blocky dolomite (pink circles), and hydrothermal

1600

carbonates filling secondary porosity and fractures (saddle dolomite - orange circles;

1601

macrocrystalline calcite - red rhombs). Micromill and bulk samples are indicated with

1602

and without external lines, respectively. Note that the δ18O values are closer to zero for

1603

primary, syngenetic and eodiagenetic constituents, more negative for the hydrothermal

1604

phases, and intermediate for the recrystallized fascicular calcite. In the 87Sr/86Sr vs. δ18O

1605

and

1606

syngenetic and eo/mesodiagenetic constituents (blue line) and another for the

1607

hydrothermal phases (red line). In the hydrothermal paragenesis, the more radiogenic

1608

values (higher

1609

well as the correlation factors (R) are indicated.

87

Sr/86Sr values obtained in the hydrothermal phases and sag host

87

Sr/86Sr vs. δ13C cross-plots there are two well-defined trends, one for the

87

Sr/86Sr ratio) present more negative values of δ18O. The equations as

1610 1611

Figure 14. Homogenization temperature (°C) versus salinity (wt. % NaCl eq.) cross plot

1612

for aqueous fluid inclusions from Pre-Salt sag and rift samples, Campos Basin, offshore

68 1613

Brazil. Primary/pseudo-secondary (upper part) and secondary (lower part) fluid

1614

inclusions hosted in hydrothermal phases Sr-barite (black), macrocrystalline calcite

1615

(red) and quartz (brown), and saddle dolomite (orange). Salinity in primary/pseudo-

1616

secondary fluid inclusions varies in the range of 12 to 26 wt. % eq. NaCl. However,

1617

97% of the data obtained from aqueous fluid inclusions show salinities higher than 15

1618

wt. % eq. NaCl. (*) Salinity of aqueous fluid inclusions hosted in recrystallized

1619

fascicular calcite calculated in wt. % NaCl-CaCl2 eq.

1620 1621

Figure 15. Burial-thermal history diagram for northern Campos Basin Pre-Salt

1622

lacustrine reservoirs, highlighting the burial depths (black line) of the sag reservoir (S)

1623

and of the 100 °C isotherm (dashed red line). The surface temperatures estimated for rift

1624

and sag deposits are 26.9 and 27.8 °C, respectively. Color scale and red values within

1625

the white rectangles represent the temperature variation. The blue arrows in the lower

1626

right corner indicate the depths of the samples analyzed for fluid inclusions. In the left,

1627

a histogram of homogenization temperatures (Th) of primary/pseudo-secondary aqueous

1628

fluid inclusions hosted in fascicular calcite, saddle dolomite, macrocrystalline calcite,

1629

mega-quartz, and Sr-barite, indicating precipitation temperatures significantly higher

1630

than the burial temperatures. Note that subsidence was essentially continuous since the

1631

deposition of the Pre-Salt sag sequence. Magmatic activity (M) occurring in the Campos

1632

Basin during Cretaceous and Paleogene (Winter et al., 2007) is shown by purple

1633

rectangles and black dashed lines. The dashed green line indicates the peak of

1634

conventional oil generation and migration (OGM) in the Miocene of the Campos Basin,

1635

according to Mello et al. (1994).

1636

69 1637

Figure 16. Schematic representation of the deep-burial hydrothermal system affecting

1638

the northern Campos Basin Pre-Salt reservoirs, offshore Brazil. The fault-focused

1639

hydrothermal system probably involved mixing of fluids derived from several sources:

1640

(B) Pre-Cambrian basement (mainly granitic-gneissic felsic rocks), (C) volcanic rocks

1641

from the Cabiúnas Formation (mainly basaltic), (R) rift deposits from the Atafona and

1642

Coqueiros formations, and (S) serpentinization of the upper mantle. Note the intrusive

1643

magmatic (mafic) rocks (M) in the Coqueiros (rift interval) and Macabu (sag section)

1644

formations; hydrothermal alteration (H) comprising extensive dolomitization,

1645

silicification, and dissolution, with the paragenesis including saddle dolomite,

1646

macrocrystalline calcite, mega-quartz, Sr-barite, celestine, fluorite, dickite, sphalerite,

1647

galena, and other metallic sulfides filling fractures and dissolution porosity; and

1648

anhydrite thick layer (A) at the top of the sag reservoir affected by the hydrothermal

1649

alteration (base of evaporites).

1650 1651

Table 1. Statistical summary of the results obtained in aqueous and petroleum

1652

inclusions from Campos Basin Pre-Salt, offshore Brazil. Fluid inclusions are hosted in

1653

the recrystallized bioclasts and syngenetic fascicular calcite, diagenetic recrystallized

1654

calcite spherulites, saddle dolomite, macrocrystalline calcite, mega-quartz, and radiated

1655

Sr-barite. Lithologic types - br: bioclastic rudstone; fcc: fascicular calcite crust; sc:

1656

stevensitic claystones with calcite spherulites; ig: intraclastic grainstone; ct: chert.

1657

Comp.: composition; FC: fluorescence color with Nikon UV-2A filter; API hc:

1658

estimated API gravity of petroleum inclusions; Th: liquid-vapor homogenization

1659

temperature of aqueous and petroleum fluid inclusions; Tm ICE: ice melting temperature

1660

of aqueous fluid inclusions; Sal (wt% NaCl eq.): equivalent salinities estimated from

1661

final ice melting temperatures in weight percent of NaCl; n: number of inclusions

70 1662

measured; pr: primary fluid inclusions; sec: secondary fluid inclusions; psec: pseudo-

1663

secondary fluid inclusions; aq: aqueous fluid inclusions; hy: hydrocarbon fluid

1664

inclusions; wt: white fluorescence; bl: blue fluorescence; yl: yellow fluorescence; pyl:

1665

pale yellow fluorescence; mx: mixed inclusions containing aqueous fluid and

1666

petroleum; N/A: could not be determined. (*) Salinities of aqueous fluid inclusions

1667

hosted in recrystallized fascicular calcite calculated in wt. % NaCl-CaCl2 eq.

1668 1669

Table 2. δ13C and δ18O results of isotopic analyses obtained in bioclasts recrystallized

1670

to microcrystalline and blocky calcite, recrystallized fascicular calcite, laminite with

1671

microcrystalline and blocky calcite, recrystallized calcite spherulites, microcrystalline

1672

and blocky dolomite replacing pre-existing constituents, and hydrothermal saddle

1673

dolomite and macrocrystalline calcite filling dissolution porosity and fractures.

1674 87

Sr/86Sr ratio obtained from recrystallized fascicular calcite and

1675

Table 3. Results of

1676

calcite spherulites, microcrystalline and blocky dolomite, and hydrothermal saddle

1677

dolomite, macrocrystalline calcite and Sr-barite filling dissolution porosity and fractures.

1678 1679

Table 4. δ18Ow in VSMOW calculated based on Th values from fluid inclusions and

1680

measured δ18Ocarb measured in carbonates from Campos Basin Pre-Salt, Brazil. The

1681

equations of O’Neil et al. (1969) and Mathews and Katz (1977) were used to calculate

1682

δ18Ow in VPDB, which was converted to δ18Ow in VSMOW by using the conversion of

1683

O’Neil et al. (1969). Details are given in item 5.2.

1684

71 1685

Supplementary Material

1686 1687

Supplementary material 1. WDS (Wavelength-Dispersive X-Ray Spectroscopy)

1688

chemical composition of carbonates from Pre-Salt reservoir in the Campos Basin, Brazil.

1689

The box-plots show the contents of (A) Na2O, (B) Al2O3, (C) SiO2, (D) MnO, (E) FeO,

1690

(F) SrO, (G) BaO and (H) Ce2O3 in mass %, obtained from fascicular (FC; purple;

1691

n=72), spherulitic (CS; blue; n=66) and macrocrystalline (MC; red; n=140) calcite, and

1692

microcrystalline (MD; light pink; n=32), blocky (BD; dark pink; n=40) and saddle

1693

dolomite (SD; orange; n=88). Thick lines and rhombs within the boxes, and circles

1694

correspond to median, mean and outliers, respectively.

1695 1696

Supplementary material 2. Scanning electron microscopy (SEM) analyses in selected

1697

hydrothermal minerals from Campos Basin Pre-Salt, Brazil. Backscattered electrons

1698

images (BSE) and As, S, Co, Cu, Zn, Ni, Fe and Pb contents (mass%) from WDS

1699

(Wavelength-Dispersive X-Ray Spectroscopy) analyses performed on sphalerite (Sp)

1700

and euhedral pyrite (Py), and Na2O, and P2O5, SO3, CaO, TiO2, FeO, SrO and BaO

1701

(mass%) analyses performed on prismatic Sr-barite (SB). A) and B) Macrocrystalline

1702

sphalerite (Sp) engulfing and replacing microcrystalline (MD) and blocky (BD)

1703

dolomite and partially filling vugular porosity (BSE); C) and D) Euhedral pyrite (Py),

1704

Sr-barite (SB), macrocrystalline calcite (MC) and prismatic macroquartz (MQ) filling

1705

vugular and fracture porosity in chert; E) and F) Vugular porosity completely cemented

1706

by prismatic Sr-barite (SB), macrocrystalline quartz (MQ), and macrocrystalline calcite

1707

(MC).

1708

Lithologic Types

Depth (m)

Comp.

Timing

FC

API hc (o)

Th (oC) range

Th (oC) average (n)

Recrystallized Calcite B Bioclast (CB)

br

XD26.0

hy

pr

wt

45-50

Recrystallized Fascicular Calcite (FC)

fcc fcc

XB93.5 XB93.5

aq hy

pr pr

wt

45-50

wt/bl wt

40-50 45-50

pr

wt

45-50

sec

wt

45-50

80-82 87-89 93-94 95-99 101-104 108.6-108.6 92-93 83-84 91-94 95-100 100-105 105.8-105.8 110-111 63-69 76-80 82-85 85-89 95-95 112.2-112.2 63-70 71-74 77.5-77.5 81-84 86-89 91.9-91.9 97.4-97.4 101-101 105-108 106-108 109-114 117-118 119-124 125-130 112-115 120-121

81.1 (2) 88.1 (5) 93.5 (2) 96.9 (5) 102.5 (5) 108.6 (1) 92.2 (2) 83.4 (2) 92.6 (4) 97.4 (8) 102.5 (3) 105.8 (1) 110.8 (2) 66 (2) 77.7 (3) 83.5 (6) 86 (5) 95 (1) 112.2 (1) 65.8 (8) 72.1 (3) 77.5 (1) 82.4 (4) 87.1 (2) 91.9 (1) 97.4 (1) 101 (1) 106.5 (2) 107 (3) 111.5 (5) 117.5 (2) 121.5 (3) 127.5 (2) 113.5 (2) 120.5 (2)

Mineral Host

Well

B B

Recrystallized Calcite A Spherulite (CS)

Macrocrystalline Calcite (MC)

C

sc

XC46.0

sig

XB71.5

sig/fcc

XB71.5/ XB98.0

fcc fcc

XB98.5 XB98.0

fcc

XB98.5

hy

aq

pr

pr/psec

Tm ICE (oC) range

Sal (wt% NaCl eq.)

-26.8

23.9-24.3 (*)

-20.9 -16.6 to -17.0 -20.1 to -20.4 -15.4 to -19.4 -18.1 to -20.8 -14.9 to -16.6 -14.8 to -17.2 -11.1 to -13.3 -16.7 to -18.4

23.0 19.9-20.2 20.5-22.4 19.0-22.0 21.0-22.9 18.6-19.9 18.5-20.4 15.1-17.2 20.0-21.3

fcc fcc

XB98.0 XB98.5

sig fcc

XB71.5 XB98.0

fcc sig

XB98.5 XB71.5

fcc

XB98.5

ct

XB91.0

fcc ct sig ct/fcc ct

XB98.5 XB91.0 XB71.5 XB91.0/ XB98.5 XB91.0

A

sc

XC46.0

C A

sig sc

C A C

sig sc sig

C

Macrocrystalline Quartz (MQ)

Prismatic Sr-barite (SB)

C

sec

hy

pr sec

aq

pr

yl yl

N/A N/A mx mx

pr/psec

sec

wt

45-50

XB71.5 XC46.0

pyl wt

33-35 45-50

XB71.5 XC46.0 XB71.5

yl wt

32-34 45-50

hy

aq

pr

pr

pr/psec

N/A 99-103 106-106 126-126 100-100 69-73 81-81 122-124 105-105 109-109 111-111 114-115 119-119 137-139 114-114 120-123 127-127 130-135 133-133 105-106 106-109 115-122 119-121 129-134 65-69.4 70-74.4 75-75 83-83 87.3-87.3 90-94 99-99 100.9-100.9 122-127 125-130 144-144 152-152 N/A 125-130 135-135

N/A (1) 101 (3) 106 (1) 126 (1) 100 (1) 71 (2) 81 (1) 123 (2) 105 (1) 109 (1) 111 (1) 114.5 (2) 119 (1) 138 (2) 114 (1) 121.5 (2) 127 (1) 132.5 (3) 133 (1) 105.5 (3) 107.5 (3) 118.5 (5) 120 (2) 131.5 (3) 68.2 (7) 73.1 (12) 75 (1) 83 (1) 87.3 (1) 92.1 (2) 99 (1) 100.9 (1) 124.5 (3) 127.5 (2) 144 (1) 152 (1) N/A (1) 127.5 (3) 135 (1)

-20.7 -15.2 to -16.1 -15.9 -16.0

22.9 18.8-19.5 19.4 19.5

-9.5 -9.0 -17.6 -15.4 to -15.9 -18.6 -17.9 to -18.0 -19.0 -19.9 to -21.2 -16.2 -16.2 to -17.8 -22.1 -16.6 to -18.2 -19.1 to -21.9 -19.5 to -21.3 -21.7 to -22.3 -19.3 to -20.5

13.4 12.9 20.7 19.0-19.4 21.4 20.9-21.0 21.7 22.3-23.1 19.6 19.6-20.8 23.7 19.9-21.1 21.8-23.6 22.0-23.2 23.5-23.9 21.9-22.7

-12.8 to -18.2 N/A -15.3 -16.6 -19.1 -18.1 to -19.6 -18.9

16.7-21.1 N/A 18.9 19.9 21.8 21.0-22.1 21.6

sec C

Saddle Dolomite (SD) C

C

sig

XB71.5

hy

pr/psec sec

br fcc br fcc/br

aq

pr

br

XF39.8 XB98.0 XF39.8 XB98.0/ XF39.8 XF39.8

br

XF39.8

hy

pr pr/psec

yl yl

yl yl

32-34 mx

N/A mx

N/A 122-124 83-83 100-100 105-107 119-119 N/A N/A 105-105 121-124 126-131 132-137 137-137 140-143 144-148 150-150 85-85 79-79 89-89

N/A (1) 123 (2) 83 (1) 100 (1) 106 (2) 119 (1) N/A (2) N/A (7) 105 (1) 122.5 (3) 128.5 (4) 134.5 (6) 137 (1) 141.5 (2) 146 (2) 150 (1) 85 (1) 79 (1) 89 (1)

-18.2 -16.9 to -19.2

21.1 20.2-21.8

-16.6 to -17.3 -19.3 to -25.3 -16.1 -18.4 to -21.0 -13.5 to -20.7 -15.0 to -17.8 -16.7 -14.5 to -16.8 -16.9 to -17.4 -20.4

19.9-20.5 21.9-26.1 19.5 21.3-23.0 17.3-22.9 18.6-20.8 20.0 18.2-20.1 20.2-20.5 22.7

Mineral Phase

Sampling

Well Depth (m) δ13C

Microcrystalline/Blocky Calcite (MBCB) (Bioclastic Grainstone and Rudstone)

bulk

B

Microcrystalline/Blocky Calcite (MBCL) (Laminite)

bulk

B

Fascicular Calcite (FC)

bulk

C B

C

micromill

C

XD06.0 XD08.0 XD10.0 XD73.0 XE38.3 XC95.4 XD00.0 XD01.0 XD43.0 XD44.0 XA15.0 XA37.0 XC85.0 XC88.2 XC89.0 XC93.0 XB57.0 XA06.0 XA95.0 XC02.3 XC04.0 XB98.8 XC01.0 XC08.5 XB83.0 XB87.0 XB88.5

XB98.8 XC08.5 Calcite Spherulite (CS)

bulk

A

XA66.5 XA84.0

1.10 0.99 1.13 0.83 0.45 1.14 0.70 1.42 1.37 1.43 1.41 0.93 1.30 1.28 1.52 1.42 1.07 1.98 1.43 1.60 1.62 1.42 1.17 1.26 0.05 1.15 0.88 0.71 0.34 0.19 0.16 0.59 0.72 0.63 0.98 0.52

δ13C error (1s) 0.04 0.06 0.05 0.06 0.05 0.03 0.05 0.04 0.04 0.05 0.05 0.05 0.07 0.05 0.05 0.02 0.05 0.02 0.09 0.05 0.04 0.05 0.05 0.05 0.06 0.04 0.03 0.04 0.05 0.03 0.06 0.04 0.02 0.04 0.05 0.05

δ18O -0.95 1.04 0.55 -1.06 -0.80 -0.61 -0.54 0.46 -1.18 -0.90 -0.59 -0.19 -0.99 -0.97 -0.23 0.63 0.54 -2.94 -1.93 -3.94 -3.34 -4.97 -4.36 -2.34 -5.70 -5.25 -5.10 -4.95 -5.37 -5.24 -5.12 -5.19 -3.85 -2.88 1.23 1.28

δ18O error (1s) 0.06 0.05 0.08 0.10 0.09 0.06 0.08 0.05 0.04 0.08 0.11 0.08 0.08 0.06 0.06 0.04 0.07 0.08 0.07 0.08 0.08 0.07 0.07 0.07 0.06 0.09 0.02 0.05 0.02 0.03 0.06 0.03 0.03 0.04 0.07 0.07

Sampling Observation Bulk analyses in the bioclastic grainstone composed of blocky and microcrystalline calcite in recrystallized bioclasts (>90%) and microcrystalline dolomite in partially dolomitized portions (<10%).

Bulk analyses in the bioclastic rudstone composed of blocky and microcrystalline calcite in recrystallized bioclasts (100%).

Bulk analyses in the laminite lithofacies constituted by recrystallized blocky and microcrystalline calcite (>90%) partially replaced by microcrystalline dolomite (<10%).

Bulk analyses in sample composed of syngenetic fascicular calcite crusts (>90%), eodiagenetic calcite spherulite (<10%) and eo/mesodiagenetic blocky dolomite (<10%).

Punctual analyses performed in the syngenetic fascicular calcite and/or fascicular aggregates.

Bulk analyses in sample constituted by eodiagenetic calcite spherulite (>90%), syngenetic fascicular calcite (<10%) and eo/mesodiagenetic blocky dolomite (<10%).

B

C

bulk

B

micromill

C

Macrocrystalline Calcite bulk (MC) micromill

C

Microcrystalline/Blocky Dolomite (MBDD) (Dolostone)

C

XB67.0 XA03.0 XA41.2 XA58.5 XA69.8 XB59.0 XC01.0 XC91.0 XA85.0 XB53.0 XD51.0 XC87.4 XC94.0 XC08.5 XB83.0 XB84.0 XB71.5 XB83.0 XB87.0 XB88.5 XC08.5

Saddle Dolomite (SD)

bulk

A

B C

XC88.5 XC96.6 XD19.0 XD19.5 XE62.0 XF17.0 XD90.0

2.43 1.25 0.75 1.79 1.49 2.14 1.54 1.91 0.31 1.36 1.67 1.25 1.16 1.44 1.37 1.28 1.18 -0.27 0.75 0.90 0.97 0.35 0.83 0.01 0.70 0.88 -1.88 -0.57 -0.04 0.22 1.03 0.34 -0.92

0.05 0.04 0.06 0.07 0.04 0.08 0.05 0.02 0.05 0.05 0.05 0.03 0.05 0.03 0.01 0.05 0.05 0.07 0.03 0.03 0.05 0.07 0.04 0.04 0.02 0.02 0.05 0.05 0.05 0.05 0.07 0.03 0.05

1.47 -2.04 0.76 -0.10 0.07 -0.56 -1.05 1.01 -2.11 -1.54 1.74 0.35 -1.12 -1.51 -1.45 -5.90 -4.92 -9.74 -9.51 -7.85 -6.19 -8.11 -7.67 -8.22 -7.82 -7.21 -6.87 -7.39 -9.49 -7.87 -7.33 -6.97 -6.35

0.07 0.07 0.10 0.08 0.07 0.12 0.09 0.08 0.07 0.07 0.07 0.10 0.07 0.04 0.04 0.07 0.07 0.07 0.03 0.04 0.05 0.07 0.04 0.05 0.02 0.02 0.07 0.07 0.07 0.07 0.07 0.11 0.07

dolomite (<10%).

Bulk analyses in the dolostone with blocky and microcrystalline dolomite (100%). Punctual analyses performed in the eo/mesodiagenetic blocky dolomite filling secondary porosity. Bulk analyses in sample composed of hydrothermal macrocrystalline calcite (>90%) and syngenetic fascicular calcite (<10%). Punctual analyses performed in the hydrothermal macrocrystalline calcite filling vugular and fracture porosity.

Bulk analyses in the dolostone and/or chert constituted by hydrothermal saddle dolomite (>90%) and macrocrystalline calcite (<10%) filling secondary porosity.

Mineral Phase

Sampling

Well Depth (m)

87

Fascicular Calcite (FC)

micromill

C

XB83.0 XB88.5 XB98.8 XC08.5

Calcite Spherulite (CS)

bulk

A

micromill

C

XA66.5 XA84.0 XB67.0 XB89.5

XB91.0

Sr/86Sr

± 2SE

Sampling Observation

0.71317 0.71329 0.71332 0.71360

0.00002 0.00002 0.00002 0.00001

Punctual analyses performed in the syngenetic fascicular calcite and/or fascicular aggregates.

0.71317 0.71338 0.71394 0.71317 0.71367 0.71365 0.71299 0.71360 0.71315

0.00001 0.00001 0.00001 0.00001 0.00001 0.00001 0.00002 0.00001 0.00001

Bulk analyses in sample constituted by eodiagenetic calcite spherulite (>90%), syngenetic fascicular calcite (<10%) and eo/mesodiagenetic blocky dolomite (<10%). Punctual analyses performed in the eodiagenetic calcite spherulite and spherulitic aggregates .

Microcrystalline/Blocky Dolomite (MBDD)

micromill

C

XC08.5

0.71351 0.71348

0.00001 0.00002

Punctual analyses performed in the eo/mesodiagenetic blocky dolomite filling secondary porosity.

Macrocrystalline Calcite (MC)

micromill

C

XB71.5

D

XD22.3

0.71136 0.71138 0.71199 0.71225 0.71215 0.71210

0.00005 0.00001 0.00001 0.00001 0.00001 0.00001

Punctual analyses performed in the hydrothermal macrocrystalline calcite (>90%) and Sr-barite (<10%) filling vugular and fracture porosity.

XD35.7 XD38.7 Saddle Dolomite (SD)

bulk

A

XC88.5 XC96.6 XD19.0 XD19.5

0.71109 0.71100 0.71123 0.71124

0.00001 0.00001 0.00001 0.00003

Bulk analyses in the dolostone and/or chert constituted by hydrothermal saddle dolomite (>90%) and macrocrystalline calcite (<10%) filling secondary porosity.

Sr-barite (SB)

micromill

D

XD35.7 XD41.6 XD42.1

0.71226 0.71209 0.71203 0.71202 0.71204

0.00001 0.00001 0.00001 0.00001 0.00001

Punctual analyses performed in the hydrothermal Sr-barite (>90%) and macrocrystalline calcite (<10%) filling vugular and fracture porosity.

Mineral Phase

δ18Ow Min

Max

δ18Ocarb Min

Fascicular Calcite (FC)

6.5

10.4

Fascicular Calcite (FC)

5.4

Calcite Spherulite (CS)

Max

Fluid Inclusion

Th (º C) Min

Max

-5.7

-1.9

Aqueous

92.2

92.2

12.4

-5.7

-1.9

Oil

83.0

111.0

6.3

16.3

-2.1

1.7

Oil

63.0

112.2

Saddle Dolomite (SD)

2.3

5.6

-9.5

-6.4

Aqueous

105.0

150.0

Macrocrystalline Calcite (MC)

3.3

11.1

-9.7

-4.6

Aqueous

101.0

130.0

Highlights A deep-burial hydrothermal system is recognized in the Campos Basin Pre-Salt carbonates. Hydrothermal phases show low δ18O values, and high Th and salinity fluid inclusions. The geochemical and fluid inclusion signatures are similar to MVT deposits. Hydrothermal fluids are the result of mixing between different sources. Atypical oil generation and migration occurred associated with the hydrothermal action.

Declaration of interests ☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: