Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná Basin, southern Brazil

Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná Basin, southern Brazil

Accepted Manuscript Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná Basin, southern Brazil Mariane Candido, Joice Ca...

3MB Sizes 0 Downloads 30 Views

Accepted Manuscript Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná Basin, southern Brazil Mariane Candido, Joice Cagliari, Francisco Manoel Wohnrath Tognoli, Ernesto Luiz Correa Lavina PII:

S0895-9811(18)30315-8

DOI:

https://doi.org/10.1016/j.jsames.2018.12.008

Reference:

SAMES 2065

To appear in:

Journal of South American Earth Sciences

Received Date: 30 July 2018 Revised Date:

11 December 2018

Accepted Date: 13 December 2018

Please cite this article as: Candido, M., Cagliari, J., Wohnrath Tognoli, F.M., Correa Lavina, E.L., Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná Basin, southern Brazil, Journal of South American Earth Sciences (2019), doi: https://doi.org/10.1016/j.jsames.2018.12.008. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

1

ACCEPTED MANUSCRIPT

Stratigraphic modeling of a transgressive barrier-lagoon in the Permian of Paraná

2

Basin, southern Brazil

3

Mariane Candido*, Joice Cagliari, Francisco Manoel Wohnrath Tognoli, Ernesto Luiz

4

Correa Lavina

5

Programa de Pós-Graduação em Geologia, Universidade do Vale do Rio dos Sinos, Av. Unisinos, 950,

6

Cristo Rei, São Leopoldo, Rio Grande do Sul CEP 93022-000, Brazil.

8

* Corresponding author.

9

E-mail address: [email protected]

M AN U

10

SC

7

RI PT

1

ABSTRACT

12

Barrier-lagoon systems occur in most coastal regions worldwide. The sedimentary

13

record of these environments is of scientific and economic relevance because important

14

coal deposits are associated with them. We show the evolution of an ancient coal-

15

bearing barrier-lagoon system in the subsurface on the southern Paraná Basin, Brazil,

16

from a stratigraphic modeling approach. The main objective is to simulate and

17

understand the origin and evolution of this system controlled by relative sea level

18

changes and sediment supply. Database comprises 75 logged boreholes that provided

19

data about the main coal bed deposited in the back-barrier, as well as 19 cored and

20

logged boreholes distributed along two stratigraphic sections, one parallel (NE-SW) and

21

another perpendicular (NW-SE) to the paleocoastline. The dataset was uploaded in the

22

Diffusive Oriented Normal and Inverse Simulation of Sedimentation (DIONISOS)

23

software to establish basic parameters and simulate different scenarios. Described facies

24

are interpreted as swamp, lagoon, sandy barrier and tide-influenced deposits, developed

25

during a relative sea level rise of 16 m. The coal bearing succession was deposited

26

during the transgressive system tract of fourth-order depositional sequence and included

AC C

EP

TE D

11

1

2

ACCEPTED MANUSCRIPT

peat deposition in a swamp in the back-barrier, barrier breakup and landward migration.

28

Results are of high relevance for understanding this paleoenvironment evolution and

29

similar systems in intracratonic basins. This environment occurs both in ancient and

30

recent coasts, and the present model supports the understanding of several barrier-

31

lagoon systems with coal formation in the world.

32

RI PT

27

Keywords: barrier-lagoon system, coal-bearing succession, Rio Bonito Formation,

34

stratigraphic modeling parameters, barrier transgression.

SC

33

35

1.

Introduction

M AN U

36

Coastal depositional environments host complex interactions between

38

continental and marine processes. Barrier islands are highly dynamic coastal systems

39

that are characteristic of these transitional environments, which require detailed studies

40

of their geology and geomorphology to understand their genesis. Barrier islands can

41

isolate lagoon bodies and are formed by the action of agents such as wind, waves and

42

currents, which transport, deposit and rework sediments along the coastline (Kjerfve,

43

1994; Boyd et al., 1992). The combination of transgressive and regressive cycles,

44

coastal physical characteristics and sedimentary supply directly influences the

45

structuring of these systems.

EP

AC C

46

TE D

37

The formation and evolution of barrier islands are mainly influenced by waves

47

and secondarily by tides, winds and river processes (Mulhern, 2017). Recent barrier

48

island systems (coast of Delaware - U.S.A., Yellow Sea - China, Wadden Sea –

49

Denmark, Adriatic Sea – Italy) were developed mainly in micro-tidal and meso-tidal

50

coasts (Hayes, 1975) where wave action and drift currents transport and deposit

51

sediments parallel to the coast, forming an extensive sandy or gravelly bar (Davis, 1994;

2

3

ACCEPTED MANUSCRIPT 52

Reinson, 1992). One or more inlets can segment the sandy bar. Nevertheless, on the

53

landward side, estuaries, lagoons, and marshes can develop protected from the marine

54

influence processes. The preservation of ancient barrier islands in the rock record depends upon the

56

amount of erosion occurring during the subsequent transgression. Higher preservation

57

rates are related to transgressive rather than regressive settings, and the best

58

preservation occurs when the system is drowned in place during rapid relative sea level

59

rise (Reinson, 1992). Moderate to higher preservation rates are also related to

60

intracratonic basins settings, since they are usually characterized by a less intense

61

tectonism activity, when compared to other basins. Thus, the Paraná Basin, an

62

intracratonic basin in southern Brazil, provides excellent record of Permian barrier

63

island systems developed along the paleocoastlines of the epicontinental sea in the south

64

of Gondwana. In this basin, most of the economic coal measures are related to peat

65

(Cagliari, 2014; Kalkreuth et al., 2010; Kern, 2008; Holz, 2003; Lopes et al., 2003b,

66

2003c; Holz et al., 2002; Alves and Ade, 1996; Lavina and Lopes, 1987).

SC

M AN U

TE D

67

RI PT

55

The central region of Rio Grande do Sul presents a great sedimentary record of these deposits, in a high density of boreholes. They can be studied due to the availability

69

of boreholes intercepting a coal bed exploited during the 1970s and 1980s. The upper

70

Iruí coal bed belongs to the Iruí coal measure and corresponds to coal formed in swamp

71

associated with barrier-lagoon context discussed in this paper. The bed adds to 1,665 ×

72

106 t of coal. The Rio Bonito Formation is the unit of the Paraná Basin that hosts the

73

most economically important Brazilian coal measures. This formation records coastal

74

and shallow marine environments bordering an epicontinental sea (Milani et al., 2007)

75

after the end of the Late Paleozoic Ice Age (LPIA) whose sediments are preserved in the

76

Itararé Group (Gordon Jr., 1947). Due to glacial melting, Rio Bonito Formation was

AC C

EP

68

3

4

ACCEPTED MANUSCRIPT 77

deposited during approximately 20 Ma (Cagliari et al., 2014) of continuous relative sea

78

level rise, the so-called Permian transgression (Lavina et al., 1985). The complex relationship between accommodation space and sediment supply

79

that controlled barrier island migration and peat preservation is elucidated from

81

stratigraphic modeling. This study explains the sedimentary dynamics of the barrier

82

island system and evolution, with the peat-forming environment in the back-barrier,

83

preserved in the Rio Bonito Formation succession. This evolution occurred during the

84

transgressive and highstand system tract, in a fourth-order depositional sequence,

85

showing the retrogradational stacking of this system, segmentation of barrier and later

86

progradation. According to the stratigraphic modeling, a relative sea level rise of 16 m

87

in 2 Ma produced similar vertical succession as recorded in the stratigraphic sections.

88

The described facies are interpreted as swamp, lagoon, sandy barrier and tide-influenced

89

deposits. The modeling contributes to the understanding of the evolution of the barrier-

90

lagoon system in the period studied, allowing the establishment of relative sea-level

91

curve and likely sediment source locations.

93 94

2.

Geological setting

EP

92

TE D

M AN U

SC

RI PT

80

The Rio Bonito Formation is the base of the Guatá Group and is part of a second-order depositional sequence, the Gondwana I Supersequence, of the Paraná

96

Basin (Milani, 1997). The Paraná Basin is an intracratonic basin filled with sedimentary

97

and volcanic rocks between the Late Ordovician and the Late Cretaceous (Milani et al.,

98

2007). The basin is partly preserved in Brazil, Paraguay, Argentina, and Uruguay and

99

covers an area of approximately 1,600,000 km2. The basin is divided into six second-

AC C

95

100

order depositional sequences, as follows: Rio Ivaí (Ordovician-Silurian), Paraná

101

(Devonian), Gondwana I (Carboniferous-Early Triassic), Gondwana II (Late Triassic),

4

5

ACCEPTED MANUSCRIPT 102 103

Gondwana III (Jurassic-Cretaceous) and Bauru (Late Cretaceous) (Milani et al., 2007). Gondwana I Supersequence records a wide change in the paleoenvironmental conditions, starting with the LPIA in the Carboniferous (Holz et al., 2008; Rocha-

105

Campos et al., 2008; França and Potter, 1988; Rocha-Campos, 1967; Gordon Jr., 1947),

106

passing through warmer and humid climate in the Permian (during Rio Bonito

107

Formation deposition) (Iannuzzi, 2013), up to arid conditions in the Triassic (Faccini,

108

2007; Faccini et al., 2003; Zerfass et al., 2003). Thus, the Rio Bonito Formation is a

109

post-glacial unit characterized by the deposition of coal, mudstone, siltstone,

110

carbonaceous siltstone, sandstone and conglomerate (Netto et al., 2012; Gandini et al.,

111

2010; Tognoli and Netto, 2003; Buatois et al., 2001; Lopes and Lavina, 2001; Castro et

112

al., 1999; Bortoluzzi et al., 1987; Lavina and Lopes, 1987; Schneider et al., 1974). The

113

interpreted depositional environment changed through the sedimentary succession from

114

fluvial, to tide-dominated estuarine and wave-dominated shoreface (Boardman, 2013;

115

Netto et al., 2012; Gandini et al., 2010; Buatois et al., 2007; Lopes et al., 2003a, 2003b,

116

2003c; Lopes and Lavina, 2001; Lavina and Lopes, 1987; Lavina et al., 1985).

TE D

M AN U

SC

RI PT

104

Rio Bonito coal beds are numerous with a thickness of up to 4 m. Coal measures

118

are distributed along the eastern border with the most important reserves in the southern

119

basin (Fig. 1A), including Candiota, Capané, Iruí, Leão, Morungava-Chico Lomã, Santa

120

Terezinha and Sul-Catarinense (Gomes et al., 1998). The peat-forming environment is

121

interpreted as delta interdistributary bays, back-barrier, and estuaries (Lopes et al.,

122

2003a). Early Permian climate conditions that favored peat accumulation and

123

preservation were warm and humid (Slonski, 2002) with precipitation probably over

124

2,400 mm/yr (Ziegler et al., 1987).

125 126

AC C

EP

117

In the Iruí coal measure (study area, item 3.1), the succession reaches 120 m in thickness and the sedimentary facies were interpreted as floodplain, overbank, fluvial

5

6

ACCEPTED MANUSCRIPT and tidal bars, lagoon, bay-head delta, swamp, tidal flat, estuarine bay, foreshore,

128

shoreface and offshore deposits (Cagliari, 2014; Gandini et al., 2010; Kern, 2008;

129

Lavina and Lopes, 1987; Lavina et al., 1985). SW-NE elongated sand bodies composed

130

of foreshore and shoreface facies, laterally correlated to the swamp and lagoonal

131

deposits, were interpreted as barrier island systems that isolated a swamp or lagoon in

132

the back-barrier (Cagliari, 2014; Gandini et al., 2010; Kern, 2008; Lavina and Lopes,

133

1987). Sandy barrier deposits are recurrent throughout the succession (Cagliari, 2014)

134

and their formation was related to longshore currents, which drifted to N-NE during

135

summer and to SW during winter according to Cacela (2008).

SC

RI PT

127

Guatá Group (Rio Bonito and Palermo Formations) is part of the lowstand and

M AN U

136

transgressive system tracts of a second-order depositional sequence (Lopes and Lavina,

138

2001; Lopes, 1995), the Gondwana I Supersequence of Milani (1997). In the southern

139

part of the basin, the Guatá Group is subdivided into four third-order depositional

140

sequence boundaries by erosional surfaces (subaerial unconformity) that developed

141

during relative sea level fall (Lopes, 1995) (Fig. 2). In all sequences defined by Lopes

142

(1995), the subaerial unconformity is replaced by a transgressive surface and the

143

sequences record predominantly the transgressive system tract. The first depositional

144

sequence (Sequence A) is constrained to the deepest valleys and the fourth sequence

145

(Sequence D) records the transition of the Rio Bonito Formation to the Palermo

146

Formation. Sequence B and C correspond to the succession in the Iruí coal measure area

147

according to Kern (2008).

AC C

EP

TE D

137

148 149

3.

Materials and Methods

150

3.1.

Study area

151

The study area is located in the southern border of the Paraná Basin, near the

6

7

ACCEPTED MANUSCRIPT Sul-Riograndense Shield, and comprehends part of the Iruí and Capané coal measures

153

(Fig. 1). In Brazil, the area is located in the center of the state of Rio Grande do Sul,

154

near the town of Cachoeira do Sul. The area is 26 × 24 km wide (624 km2) characterized

155

by a robust geological database based on cores drilled by the Brazilian Geological

156

Survey (CPRM) during the 1970s and 1980s (CPRM, 1980, 1983; Figueroa, 1983;

157

Ferreira, 1978). The database includes hundreds of logged and cored boreholes in which

158

75 boreholes were selected and 19 were described, including those previously described

159

by Gandini et al. (2010) and Kern (2008). The study focuses on one of the fourth-order

160

depositional sequences (Fig. 2) of the Rio Bonito Formation at the Iruí coal measure.

161

The depositional sequence includes the upper Iruí coal bed deposited in a back-barrier

162

environment and records the sandy barrier transgression (Cagliari, 2014; Kern, 2008).

SC

M AN U

163

RI PT

152

This study is based on borehole cores logging and gamma-ray logs interpretation (CPRM, 1980, 1983). About 11 cores were logged by Gandini et al. (2010) and Kern

165

(2008); we logged 9 additional cores with a higher resolution in the selected interval to

166

contribute to the understanding of this barrier island system formation and evolution.

167

Sedimentary facies were described, leading to the interpretation of depositional and

168

transport processes (Walker, 1992). Two stratigraphic sections were compiled, one

169

parallel (SW-NE) and another perpendicular (SE-NW) to the paleocoastline (Fig. 1B).

170

The stratigraphic datum assumed for lateral correlation between the logs is an ash fall

171

horizon (tonstein) within the upper Iruí coal bed, the same datum used by other authors

172

(Cagliari, 2014; Kern, 2008). This horizon represents one episode of the large volcanic

173

eruption in southwestern Gondwana paleocontinent (Matos et al., 2000). Stratigraphic

174

surfaces were mapped in the analyzed sections considering the Embry (1993) and

175

Embry and Johannssen (1992) approaches.

176

AC C

EP

TE D

164

All cores, sedimentary facies, and stratigraphic surfaces were imported into

7

8

ACCEPTED MANUSCRIPT Recon®, 3D software developed by Seisware International Inc. for geologic

178

interpretation. In Recon®, we modeled facies spatial distribution, the depositional

179

surface prior to the analyzed sequence that corresponds to coal bed lower boundary, and

180

the volume of sediment preserved. For these models, we used a radial algorithm (3,000

181

m of radius) to interpolate data among cores. The relation between accommodation

182

space and sediment supply that controlled the barrier island landward migration and

183

peat preservation was estimated using a forward stratigraphic modeling performed in

184

DIONISOS®.

185

187

3.2.

Forward Stratigraphic Modeling

M AN U

186

SC

RI PT

177

The DIONISOS® (Diffusive Oriented Normal and Inverse Simulation of Sedimentation) is a 3D numerical stratigraphic forward model developed by Beicip-

189

Franlab, an associate company of the Institut Français du Pétrole. The model simulates

190

erosion, transport and deposition of sediments in shallow marine and coastal

191

environments in order to obtain the geometry and facies of sedimentary units over

192

regional spatial scales. The stratigraphic model is a result of the interaction of the main

193

factors that control sediment deposition, such as sediment supply, sea level change and

194

subsidence or uplift.

EP

The diffusion equation (Eq. 1), which is based on the main principles of stream

AC C

195

TE D

188

196

flow and mass conservation, expresses sediment transport and is dependent especially

197

on the slope and efficiency of transport. In Equation 1,

198

width unit;

199

concentration and D is diffusion coefficient in m/s2; h is the height above horizontal

200

datum; and x is the horizontal distance in flux direction. Sedimentation and erosion rates

201

at each point of the basin and each time are estimated from the mass balance principles

is the modified diffusion coefficient;

is the sediment influx by =

×

, where

is sediment

8

9

ACCEPTED MANUSCRIPT 202 203

(Beicip-Franlab, 2009). = −

×

(Eq. 1)

Long-term evolution of sedimentary processes is controlled by long-term fluvial

204

and gravity transport, processes that depend on the topographic slope, diffusion

206

coefficient, and water discharge, and the short-term is induced by catastrophic rain fall,

207

slope failures, and turbidity flow, processes that are in turn depend on water velocity

208

and inertia (Beicip-Franlab, 2009). DIONISOS® was developed to simulate the long-

209

term and large-scale evolution of the sedimentary basins, although it is also considered

210

a reliable tool to model the evolution of the short-term system (Seard et al., 2013). The

211

time scale can range from tens of thousands of years to hundreds of millions of years,

212

with the shortest time step limited on tens of thousands of years. A review of the

213

DIONISOS® mathematical model and limitations can be found in the Csato et al. (2013)

214

appendix.

M AN U

SC

RI PT

205

In the same way, as in other forward stratigraphic modeling tools, DIONISOS®

TE D

215

requires a set of pre-determined parameters to perform the stratigraphic simulation

217

(Shafie and Madon, 2008). Several parameters are difficult to be extracted from ancient

218

stratigraphic succession, thus the primary uncertainty related to the stratigraphic model

219

is mostly caused by the lack of accurate and detailed input data.

221 222

AC C

220

EP

216

3.3.

Initial and boundary conditions and input data

The initial conditions correspond to relief (or basement conditions) and sea level

223

position at the beginning of simulation (when t = 0), and the boundary conditions

224

include simulation area boundary, grid spacing and interval of time considered in the

225

analysis. The basement condition, or the initial geometry of the basin, was obtained

226

from the distribution map of the coal seam basal surface (basal sequence boundary, see

9

10

ACCEPTED MANUSCRIPT Fig. 8) after applying the datum (tonstein horizon) to correlate all borehole logs. In

228

order to constrain the sea level position at the beginning of the simulation, we assumed

229

that the lagoon level was controlled by the sea level. The water depth in modern coastal

230

lagoons is typically between 1 and 3 m but can reach 5 m in inlet channels and isolated

231

relict holes or channels (Kjerfve, 1994). Thus, based on modern environment data, we

232

have constrained the lagoon water depth to 2 m in average with the deepest part

233

reaching up 5 m (Fig. 3).

RI PT

227

The simulation area was 26 × 24 km (624 km2) with the grid spacing set at 0.5 km.

235

The depositional sequence analyzed in this study was not dated but the interval of time

236

varied between 80,000 yr and 3 Ma according to Vail et al. (1991). The sequence is

237

considered third-order by Kern (2008) and fourth-order by Cagliari et al. (2014). The

238

preservation rate estimated from two radiometric datings (Cagliari et al., 2016) of the

239

Rio Bonito Formation is ~10 m/Ma. Considering this preservation rate in the thickest

240

depositional sequence, the time span simulated was 2 Ma with a time step of 200 ka.

241

We investigated current and ancient depositional rates for these environments to

242

isolate the time span of back-barrier and barrier deposition. Nichols (1989) estimated

243

sediment accumulation and relative sea level rise rates for 22 United States lagoons and

244

discovered that most lagoons have an increased rate of about 1.6 mm/yr of sediment

245

deposition, which nearly equals to the relative sea-level rise. In this study, erosion is

246

null, thus the sedimentation rate corresponds to the preservation rate. Another limitation

247

of applying this data to our study is the anthropic occupation in the surrounding areas,

248

which can impact in the sediment budget. However, all studied lagoons are shallow (~3

249

m depth) environments and share the same depositional context of relative sea-level

250

rise, providing a good analogous model.

251

AC C

EP

TE D

M AN U

SC

234

Peat accumulation rates vary with climate, type of vegetation and local drivers. A

10

11

ACCEPTED MANUSCRIPT paleolatitude estimation of the southern Paraná Basin in the Carboniferous-Permian

253

transition (~50° S; Franco et al., 2012) point out to a temperate climate, and according

254

to different studies approaches, the Rio Bonito Formation coal seams are described as

255

being formed in limno-telmatic moors, where pteridophytic arborescent and herbaceous

256

plant material was accumulated (see review of Correa da Silva, 2004). Due to the

257

absence of peat accumulation rates for the studied coal seams we have considered

258

published rates for different modern environments as an approximation: 0.1 –2.3 mm/yr

259

(lowest rate in the Arctic and the highest in the tropics, McCabe, 1984), 0.35 – 1.13

260

mm/yr (Siberian peatlands, Borren et al., 2004), and 0.16 – 0.97 mm/yr (Congo Basin

261

peatland complex, Dargie et al., 2017). However, over the trial and error simulations,

262

the best result was achieved with a peat accumulation rate of 0.03 mm/yr. The time span

263

of the peat-forming was approximated at 0.5 Ma, based on coal thickness, decompaction

264

rate of 3:1 (Ryer and Langer, 1980; Courel, 1987) and peat accumulation rates of 0.03

265

mm/yr. We used the same time span for the lagoon system because lagoonal deposits

266

are thicker than peat forming successions and the depositional rates of these

267

environments are higher. A total of 1 Ma was set for modeling barrier migration

268

landwards, once the barrier thickness is about 10 m and the preservation rate estimated

269

for the Rio Bonito Formation is ~10 m/Ma (Cagliari et al., 2016).

271 272

SC

M AN U

TE D

EP

The input data for stratigraphic simulation were extracted from the sedimentary

AC C

270

RI PT

252

record, published studies and data from depositional environments documented in modern analogous (Table 1 and Supplementary file). Multiple scenarios were simulated

273

to adjust parameters and obtain the best result fitting with real data. The quality of the

274

results was evaluated by comparing the simulated one with the vertical and lateral

275

sedimentary succession expressed in the stratigraphic sections and individual core logs.

276

11

12

ACCEPTED MANUSCRIPT 277

4.

Results

278

4.1.

Sedimentary facies The described depositional sequence ranges between 12 and 16 m in thickness.

279

Nine sedimentary facies were described and interpreted: coal, laminated siltstones,

281

structureless siltstone, heterolithic sandstone, trough cross-bedded arkosic sandstone,

282

parallel to subhorizontal lamination quartz sandstone, trough cross-bedded quartz

283

sandstone, hummocky cross-stratification quartz sandstone, and bioturbated heterolithic

284

sandstone (Tab. 2). Considering the main processes that control the deposition of the

285

sedimentary facies described above we grouped them into three facies association: the

286

wave-dominated facies, the tidal-dominated facies, and the standing-water-dominated

287

facies.

SC

M AN U

288

RI PT

280

The wave-dominated facies association includes the parallel to subhorizontal lamination quartz sandstone (F6), trough cross-bedded quartz sandstone (F7),

290

bioturbated heterolithic sandstone (F8), and hummocky cross-stratification quartz

291

sandstone (F9). Tidal-dominated facies association includes heterolithic sandstone (F4)

292

and trough cross-bedded arkosic sandstone (F5), and the standing-water-dominated

293

facies comprehends the coal (F1), laminated siltstones (F2) and structureless siltstones

294

facies (F3).

EP

The vertical succession of facies that predominates is coarsening-upward,

AC C

295

TE D

289

296

composed from base to top by, coal (F1), laminated and structureless siltstone (F2 and

297

F3), parallel to subhorizontal lamination quartz sandstone (F6), trough cross-bedding

298

quartz sandstone (F7), bioturbated heterolithic sandstone (F8) and hummocky cross-

299

stratification quartz sandstone (F9) (Fig. 5A). The transition from coal (F1) to siltstone

300

facies (F2 or F3) is gradual and from siltstone to quartz (F6-9) or arkosic sandstone (F5)

301

facies is abrupt.

12

13

ACCEPTED MANUSCRIPT F1 – Coal: Most drilled and cored coal intervals are lacking in the boxes because of

303

previous studies (the Brazilian Geological Survey (CPRM) sampled these intervals to

304

run several analyses in the 1970s and 1980s), but detailed descriptions of these missing

305

intervals by CPRM (1980, 1983), Figueroa (1983) and Ferreira (1978) are available.

306

These authors described coal in the selected cores as black frosted coal with lenses of

307

fusinite, laminae of vitrinite, pyrite nodules, cracks filled with calcium carbonate and

308

abundant plant fragments (Fig. 4A). Thickness reaches up to 3.94 m in core IC-099-RS.

309

Interpretation: Coal facies is the result of organic matter deposition in a shallow

SC

RI PT

302

water environment under anoxic conditions and with some marine influence, as

311

indicated by the presence of pyrite nodules (Frey and Basan, 1985; McCabe, 1984).

312

According to Cúneo (1996), the main vegetation in the mid early Permian was

313

glossopterids, conifers, cordaitales, early ginkgophytes, shrubby ferns and Botrychiopsis

314

(understory). The gymnosperms Rubidgea and Ginkgophyllum represent the endemic

315

forms in the Brazilian unit. The productivity of vegetation and climate contributed to

316

the formation of peat deposits, which formed the coal beds of Rio Bonito Formation in

317

southern Brazil, also yielding an assemblage dominated by Glossopteridales (Boardman

318

et al., 2012; Guerra-Sommer, 1991).

319

F2 – Laminated siltstone: Whitish-gray to dark-gray siltstone with horizontal

320

lamination, weak bioturbation, sparse plant fragments, locally carbonaceous, and bed

321

thickness ranging from 0.35 to 1.80 m (Fig. 4B).

TE D

EP

AC C

322

M AN U

310

Interpretation: Siltstone deposited by suspension settling from standing water

323

with the horizontal lamination recording the pulses of sediment supply (Collinson,

324

1996; Reinson, 1984). Weak bioturbation and sparse plant fragments facilitated the

325

preservation of the sedimentary structure.

326

F3 – Structureless siltstone: Whitish-gray to medium-gray siltstone, structureless, with

13

14

ACCEPTED MANUSCRIPT 327

root marks and plant fragments, locally with weak bioturbation (Fig. 4C). Some

328

horizons show blocky texture and slickensides. Carbonaceous siltstones occur locally

329

underlying coal facies. This facies is widespread and the thickness reaches up 7.4 m in

330

IC-050-RS. Interpretation: Siltstone deposited by suspension settling from standing water,

RI PT

331

same as F2 (Collinson, 1996; Reinson, 1984). The lack of depositional structure is

333

explained by rapid deposition or later destruction by plant roots and soil-forming

334

processes (Collinson, 2006). Blocky texture and slickensides suggest subaerial exposure

335

and soil-forming process corroborating the possibility of later destruction of primary

336

sedimentary structures. Gandini et al. (2010) described Glossopteris flora in this facies.

337

F4 – Heterolithic sandstone: Centimeter- to millimeter-thick intercalation of fine- to

338

medium-grained sandstone and dark-gray siltstone, wavy and lenticular bedding, mud

339

drapes and flaser (Fig. 4D). Siltstone is locally carbonaceous, and bioturbation range

340

from weak to moderate.

M AN U

TE D

341

SC

332

Interpretation: Alternation of bedload sand migration as ripples and suspension settling from standing water (Rahmani, 1988; Smith, 1988; Reineck and Singh, 1986).

343

Mud drapes and flaser deposits during the slack-water period were produced by the

344

variation in the tidal current over semi-diurnal to longer cyclicity (Collinson, 1996).

345

Gandini (2010) found in the same facies Helminthopsis, Palaeophycus, Planolites,

346

Thalassinoides, Thalassinoides-Palaeophycus, Thalassinoides-Palaeophycus-

347

Helminthopsis, and interpreted as estuarine bay or lower shoreface/transition offshore.

348

F5 - Trough cross-bedded arkosic sandstone: Fine- to coarse-grained arkosic

349

sandstone, locally composed by granules and pebbles; grains are poorly-sorted,

350

subangular to angular, with low sphericity, exhibiting trough cross-bedding and

351

carbonaceous mud drapes (Fig. 4E). Succession is frequently fining-upward with

AC C

EP

342

14

15

ACCEPTED MANUSCRIPT 352 353

erosive boundary contact. Interpretation: This facies represents the migration of subaqueous dunes under low flow regime (Collinson, 1996; Collinson and Thompson, 1989; Miall, 1977). Grain

355

composition and characteristics indicate transport over short distances. Mud and

356

carbonaceous mud drapes, as in heterolithic facies (F4), is evidence of variation in tidal

357

current.

358

F6 - Parallel to subhorizontal laminated quartz sandstone: Fine- to medium-grained

359

quartz sandstone, locally coarse-grained, well-sorted, subrounded to rounded grains,

360

with parallel to subhorizontal cross lamination with erosive contact at the base (Fig. 4F).

SC

Interpretation: Deposition under upper flow regime with high flow velocity and

M AN U

361

RI PT

354

shallow water depth (Reinson, 1992; McCubbin, 1982; Heward, 1981; Clifton, 1969)

363

indicated by the presence of Macaronichnus icnofabric, described by Gandini et al.

364

(2010), the classical signature of high-energy shallow marine deposits. This icnofabric

365

contains information on the bioturbation rate and horizontal excavations resulting from

366

the reworking of the substrate by detritus-eating vermiform organisms (Pemberton et

367

al., 2001).

368

F7 - Trough cross-bedded quartz sandstone: fine- to medium-grained quartz

369

sandstone, locally very coarse-grained, well-sorted, high-sphericity and rounded grains,

370

medium- and small-scale uni- and bidirectional trough cross-bedding (Fig. 4G).

371

Bioturbation is absent.

EP

AC C

372

TE D

362

Interpretation: Migration of subaqueous 3D dunes under lower flow regime due

373

to unidirectional and oscillatory currents (Reinson, 1992; McCubbin, 1982). The

374

absence of bioturbation also reflects high-energy condition and high rate of velocity

375

migration.

376

F8 - Bioturbated heterolithic sandstone: Interbeds of fine-grained quartz sandstone

15

16

ACCEPTED MANUSCRIPT 377

and dark-gray, carbonaceous in places, siltstone with wavy and lenticular bedding, with

378

moderate to high bioturbated. Locally the sedimentary structures are obliterated by

379

bioturbation (Fig. 4H). Interpretation: This facies is interpreted as the alternation of suspension settling

380

from standing water and migration of ripples (Collinson, 1996; Reinson, 1992;

382

McCubbin, 1982). The high intensity of bioturbation and the ichnological signature

383

reflects a relatively calm marine environment below the fair-weather wave base.

384

Gandini et al. (2010) studied the same area and described for this facies the occurrence

385

of Cruziana and Skolitos ichnofacies.

386

F9 - Hummocky cross-stratification quartz sandstone: Fine-grained quartz sandstone

387

with intercalated laminae of siltstone, moderately well-sorted and light-colored grains,

388

low-angle cross-stratification (hummocky cross-stratification) (Fig. 4I).

M AN U

SC

RI PT

381

Interpretation: Fallout of sediment from suspension and lateral tractive flow

390

under wave oscillation generated by storm (Dott and Bourgeois, 1982; Harms et al.,

391

1975). The presence of hummocky cross-stratification is the diagnosis of the offshore

392

transition zone, which lies between fair-weather and storm wave base.

393

395

4.2.

Stacking pattern of the studied succession The succession has a retrogradational stacking pattern, with lagoon and swamp

AC C

394

EP

TE D

389

396

underlying foreshore and shoreface deposits. Thus, the succession records a vertical

397

shift from proximal (swamp and lagoon) to more distal depositional environments

398

(lower shoreface). This succession does not represent a typical succession formed

399

during the transgression of a barrier island according to Reinson (1992). In this case,

400

the Walther’s Law of facies is not respected, due to the absence of intermediate facies,

401

originally located along an individual depositional profile. Otherwise, it is similar to the

16

17

ACCEPTED MANUSCRIPT 402

partial preservation of facies succession recording by erosional shoreface retreat (Kraft

403

and Chrzastowski, 1985; Reinson, 1984, 1992; Heward, 1981) where shoreface facies

404

overlie more landward facies across a planar erosion surface.

405

The succession described above does not occur in cores IB-182-RS, IC-029-R, and IC-045-RS (section AA’, where the sandy barrier is totally or partially replaced by

407

tidal-influenced facies association, heterolithic and trough cross-bedded arkosic

408

sandstone facies) (Fig. 5B and Fig. 7). In other cores, sandy barrier facies is

409

interdigitated with tidal-influenced facies (IB-010-RS and IB-006-RS, section BB’) or

410

absent (CA-040-RS, section BB’).

SC

Coal facies is distributed into two beds, the lower one named upper Iruí coal bed

M AN U

411

RI PT

406

and has a wide occurrence in the study area and the upper one is restricted to the west

413

portion. The upper Iruí coal bed has an elongated geometry with the long axis of

414

approximately 32 km in the SW-NE direction and the short axis of approximately 5 km

415

in the NW-SE direction (Fig. 6). It is laterally correlated to the laminated and

416

structureless siltstone facies (F2 and F3, respectively), as presented in the NE and SE of

417

sections AA’ and BB’ (Fig. 7), which are also widely distributed in the area although

418

the thicknesses change along the sections.

EP

419

TE D

412

The sandy barrier facies association occurs along the entire area although tideinfluenced facies association locally interrupts it. This feature likely suggests that tidal

421

channels segmented the sandy barrier. However, tidal deposits would not have been

422

deposited in the inlets due to the high-energy flow. The tidal sediments would be

423

probably deposited in a more protected area, as like as the flood tidal delta. Its location

424

in a partially protected environment is indicated by carbonaceous silty interlaminations

425

that occur in the middle of poorly selected sandstones, possibly close to a tidal, flat or a

426

barred lagoon for example.

AC C

420

17

18

ACCEPTED MANUSCRIPT 427 428 429

4.3.

Sequence stratigraphy and depositional model The depositional sequence is bounded below and above by subaerial

unconformities (Fig. 7). The lower sequence boundary (SB1) is at the base of the upper

431

Iruí coal bed and records an abrupt contact between the coal facies (F1) and the

432

underlying structureless/laminated siltstones (F2 and F3). Root marks and blocky

433

texture in the underlying siltstone facies likely indicate subaerial exposure. As showed

434

in Figure 2 and described by Gandini et al. (2010), the SB1 marks the contact between

435

lagoonal and swampy depositional environment.

SC

RI PT

430

The upper sequence boundary (SB2) records the contact between marine

437

deposits below (F6-F9) and coastal (F2-F5) or shallow marine above in the BB’ section

438

(Fig. 7B). The same contact is seen in some of the AA’ section cores. The deposition of

439

coastal facies overlying marine facies records a coastline regression. A sharp and

440

erosive contact between coastal (F2 and F3) deposits below and shallow marine (F6-F9)

441

above records a transgressive surface (TS), more specifically the wave ravinement

442

surface. Considering the gamma ray log shape and the stacking pattern of facies a

443

maximum flooding surface (MFS) is identified in the upper portion of the depositional

444

sequence. The change between the lower retrogradational and the upper progradational

445

stacking pattern in the sandy barrier succession marks the MFS and records the most

446

extensive marine incursion in this depositional sequence.

TE D

EP

AC C

447

M AN U

436

Facies succession between SB1 and MFS records the transgressive systems tract

448

and facies succession between MFS and SB2 the highstand systems tract. In the

449

transgressive systems tract, the stacking pattern of facies is retrogradational with facies

450

deposited in even more marine and deep conditions towards the top. In the highstand

451

systems tract, the low thicknesses preserved prevent the recognition of the stacking

18

19

ACCEPTED MANUSCRIPT 452 453

pattern. The elongated and widespread coal bed and the vertical succession of facies indicate that deposition took place in a swamps environment in a protected coastal area.

455

Nodules of pyrite indicate occasional seawater incursion, e.g. through tidal inlets or

456

washover fans formed during storms (cf. McCabe 1984) although these features were

457

not identified in the study area. Based on that, the SW-NE coal bed orientation likely

458

indicates the coastline position.

The structureless and laminated siltstone facies that are laterally correlated and

SC

459

RI PT

454

overlies the coal facies were deposited in a lagoon and records an increase in the

461

sediment influx and ventilation conditions (increase in the concentration of oxygen in

462

water) that prevents the continuous establishment of the peat-forming environment. The

463

sediment supply and oxygen increase can be explained by the presence of inlets that had

464

segmented the sandy barrier, thus changing the previous sheltered environment into a

465

wave- and tide-influenced one, or by the climatic condition that supplied more sediment

466

from the river run-off.

TE D

M AN U

460

The parallel to sub-horizontal lamination quartz sandstone, trough cross-bedded

468

quartz sandstone, bioturbated heterolithic sandstone and hummocky cross-stratification

469

quartz sandstone facies are interpreted as foreshore to shoreface/offshore-transition

470

during storm conditions deposits. This facies, as explained before, comprehends the

471

sandy barrier system, which has a thickness ranging between 3.6 m (IC-045-RS) and

472

10.7 m (IB-10-RS) and abruptly overlies the siltstone and/or coal facies. The sandy

473

barrier is locally eroded by tide-influenced channel bars in section AA’ and in the

474

northwest of section BB’, which likely indicate the barrier segmentation by tidal inlets.

475 476

AC C

EP

467

The sedimentary facies characteristics and lateral and vertical successions suggest a depositional environment consisting of a beach barrier and a protected back-

19

20

ACCEPTED MANUSCRIPT barrier lagoon, belonging to a barrier-lagoon system. The stratigraphic reconstruction

478

shows the establishment of a barrier-island system and landward migration due to

479

increasing accommodation space. The coastal area was influenced by the combination

480

of wave and tidal processes although wave processes predominated to develop extensive

481

sandy barrier strandplain.

RI PT

477

482

484

4.4.

Forward stratigraphic model

During the processes of stratigraphic simulation, a relative sea level curve was

SC

483

established for the deposition of the analyzed succession, the result of the balance

486

between the sediment supply and accommodation space (Fig. 8). The relative sea level

487

rose 16 m during the sequence deposition simulation, with rates ranging from 0 to 10

488

m/Ma. This assumption was made based on the total amount of preserved thickness in

489

the depositional sequence and the estimated interval time, therefore has a relatively low

490

degree of stratigraphic resolution. The relative sea level rise and the sediment supply

491

controlled the formation of the barrier-lagoon system and further barrier island

492

landward migration.

TE D

The coal facies was formed in a back-barrier environment widespread over an

EP

493

M AN U

485

area of 32 km in the SW-NE direction and 5 km in the NW-SE (Fig. 9A). The estimated

495

peat production rate was 30 m/Ma, the equivalent of 15 m during 500 ka (the time

496

interval simulated for peat deposition), which corresponds with the maximum depth of

497

peat generation (5 m) transformed by the adopted rate 3:1 (5 m coal = 15 m peat) (Ryer

498

and Langer, 1980; Courel, 1987). Peat was deposited during the transgressive systems

499

tract and total peat accumulation was controlled by the 5 m relative sea level rise during

500

500 ka (sea level rise rate of 1 cm/ka). The peat-bearing environment was probably

501

developed in the borders of a coastal lagoon, which is preserved only in the core CA-

AC C

494

20

21

ACCEPTED MANUSCRIPT 502

40-RS (BB’, Fig. 7), recorded in the laminated/structureless siltstones facies laterally

503

correlated with the upper Iruí coal bed. At this time, the lagoon is outside the studied

504

area and was not simulated. The lagoon environment, where the laminated/structureless siltstone facies was

505

likely deposited, was simulated until reaching the preserved thickness of siltstone in the

507

stratigraphic sections (Fig. 9B). The distribution of siltstone facies over the upper Iruí

508

coal bed suggests that sediment was supplied from the northeast. However, only this

509

source was not sufficient to distribute the sediments, being necessary to add another

510

source to the simulation. Considering that this source is continental, it was positioned in

511

the southwest, since there are tidal channels near that are filled with arkosic sandstones.

512

Furthermore, there are sandstones (or even conglomerates) directly overlying the Upper

513

Iruí coal bed in some boreholes (IB-097-RS, IB-098-RS, and IB-157-RS) which are out

514

of the stratigraphic section, but inside coal bed area, indicating a probable fluvial

515

source.

SC

M AN U

TE D

516

RI PT

506

In order to establish values for discharge (in m³/s), it was considered a system with low flow due to the existence of peat initially and after the predominance of

518

siltstones. Thus, values below 250 m³/s were tested and the values that generated the

519

best result of sedimentary distribution were used. The best result was obtained

520

considering a balance between marine (source northeast seawater discharge = 110 m3/s)

521

and a continental (fluvial discharge = 100 m3/s) source. To simulate the siltstone facies

522

deposition and the lagoon system evolution, we considered 3 m relative sea level rise

523

during 500 ka (0.6 cm/ka). Although the preserved sediment corresponds to 0.9

524

km³/Ma, the best result was achieved with 0.8 km³/Ma (corresponding to 0.4 km³ of

525

sediment supplied during the time interval simulated).

526

AC C

EP

517

Sandy barrier was simulated considering the continuous increase in the relative

21

22

ACCEPTED MANUSCRIPT sea level during the transgressive systems tract, and a relative sea level still stand,

528

recorded in the highstand systems tract (Fig. 9C, 9D). The sandy barrier thickness

529

preserved in the geologic record reached 10.7 m (IB-10-RS). The forward stratigraphic

530

model shows the barrier transgression over 1 Ma and producing a similar thickness as

531

presented in the sedimentary record. The sandy barrier is composed mainly of quartz

532

sandstone transported onshore and secondarily by arkosic sandstone (tide-influenced

533

fluvial sediments) deriving from continental sources. The 3.9 km³/Ma of sediment

534

supplied and the relative sea level rise of 8 m during 1 Ma (0.8 cm/ka) caused a barrier

535

island landward shift of 9 km.

SC

RI PT

527

The resulting sedimentary succession is shown in Figure 10. In the simulated

M AN U

536

sections, sandy sediments are not differentiated by composition. The simulated sections,

538

as well as the stratigraphic sections, first show the migration of the barrier-lagoon

539

system landwards and seawards. This landward migration, according to the model,

540

occurred during about 1.6 Ma, when the progradation of the barrier began at the

541

maximum sea level ingression, corresponding with the flooding surface. The

542

transgressive surface shows the overlap of marine sediments on to the underlying

543

lagoonal deposits.

EP

TE D

537

Erosive processes were not simulated, but the resulting forward stratigraphic

544

model shows the sediment distribution and succession similar to the stratigraphic

546

model. Especially the siltstones are shown thicker in the simulating boreholes compared

547

to real boreholes (Fig. 11). Since the absence of erosion produced thicker successions

548

than the sedimentary record, this difference can indicate, in a simplified way, the

549

erosion occurred during the relative sea level rise.

AC C

545

550 551

5.

Discussion

22

23

ACCEPTED MANUSCRIPT 552

The simulated stratigraphic sections (Fig. 10) are similar to the stratigraphic sections (Fig. 7) when considering the sediment distribution both vertically and

554

laterally. The sediment supply in the best scenario (0.8 km³/Ma for the back-barrier

555

system and 3.9 km³/Ma for the sandy barrier) corresponds with more than 80% of the

556

estimated from the sedimentary record (0.9 km³/Ma and 4.6 km³/Ma, respectively).

557

After the drowning of the lagoon, the barrier moved seawards while the relative sea

558

level was rising at increasingly lower rates and there was a positive balance of

559

sediments, corresponding with the highstand systems tract.

SC

560

RI PT

553

Continuous and less continuous thick coal beds are usually deposited during the middle transgressive systems tract when accommodation space is high (Bohac and

562

Suter, 1997). However, in the stratigraphic model, peat was likely deposited during the

563

early transgressive systems tract, first in the morphologically depressed sectors of the

564

investigated area and then over a larger area during the late relative sea level rise. This

565

suggests that even at lower rates of relative sea level rise, the amount of organic matter

566

allowed the generation of peat early in TST. The upper Iruí coal bed has high mineral

567

content indicating that during peat formation sediments were deposited in the peat-

568

forming environment. This indicates the peat-forming environment was located on the

569

border of a lagoon next to the sandy barrier, thus receiving sporadically sediments from

570

the barrier.

TE D

EP

AC C

571

M AN U

561

The estimated relative sea level curve (Fig. 8) was constructed from data

572

extracted from the facies thickness, considering that the sediment deposition in coastal

573

and marine environments are mainly controlled by relative sea level changes and,

574

subordinately, by sediment discharge from river run-off, alongshore transport and by in

575

situ sediment production. However, these additional elements have not been taking into

576

account, for simplicity. The curve was refined through modeling by trial and error

23

24

ACCEPTED MANUSCRIPT 577

during successive simulations in DIONISOS®. Despite the rate of the sea level rise

578

reduction hypothesized after the accumulation of the peat (1 cm/ka for 0.6/ka), the

579

interruption of its deposition can be explained by the increase of the sedimentary supply

580

compared to accommodation. In coastal environments with gentle slopes, sediment deposition can be

RI PT

581

controlled by many factors, but in this simulation, it was observed that the main

583

controlling factors are relative sea level change and sediment supply, once the

584

subsidence was not considered as input parameter. DIONISOS® software was originally

585

developed for deltaic and shelf environment simulation, both with significant slopes.

586

However, this software simulates long-term basin scale sediment distribution in large-

587

scale, what is crucial in this case. As seen from the simulation results, relative sea level

588

change controlled the barrier island landward migration over the lagoon deposits.

589

Therefore, this result is in accordance with the depositional history recorded in the

590

sedimentary succession. The forward stratigraphic simulation, not even considering the

591

erosion, displays a similar sediments distribution as in the stratigraphic section,

592

although the thickness is slightly different.

M AN U

TE D

Results presented here can be applied to the construction of system

EP

593

SC

582

understanding in intracratonic sedimentary basins such as the Paraná basin, especially in

595

wave-dominated paleocoastlines. A better understanding of environments helps set the

596

evolutionary history of basins, including processes and products generated during their

597

formation. Economically, depositional sequences are better understood associated with

598

deposits such as coal, greater optimization of economic viability and exploration

599

studies. These facies associations are found all over the world, being a matter of general

600

interest, because a large part of coal deposits is found associated with barrier-lagoon.

AC C

594

601

24

25

ACCEPTED MANUSCRIPT 602 603

6.

Conclusions Forward stratigraphic simulation applied to the stratigraphic succession

preserved in the subsurface of the Paraná Basin shows that a possible scenario that

605

explains the formation and preservation of a barrier-lagoon system during a relative sea

606

level rise of 16 m. The analyzed stratigraphic interval of the Rio Bonito Formation was

607

strongly influenced by waves, which controlled the longshore drift and thus the

608

sediment distribution along the coast. Both marine and fluvial flows have significant

609

control in the sediment distribution, as well as the depositional surface slope, sediment

610

supply, and relative sea level change.

SC

Stratigraphic modeling contributed to the understanding of the processes

M AN U

611

RI PT

604

controlling the Rio Bonito Formation barrier-lagoon system evolution through time,

613

focusing on the estimation of the relative sea level rise and the likely location of the

614

sediment sources. The model shows that clastic sedimentary supply derives initially

615

from continental and marine sources, which are equally important and, later, almost

616

exclusively from marine sources (e.g., alongshore transport and flood-tidal deltas

617

entering through tidal inlets breaching the barrier). Competition between marine and

618

continental environments is visualized along the sea level curve created during the

619

modeling, as time steps progress.

EP

AC C

620

TE D

612

Stratigraphic modeling is a powerful tool to reconstruct sedimentary settings and

621

estimate specific conditions that are not preserved in the geological record. Although

622

stratigraphic modeling, as such as other modeling approaches, presupposes a

623

simplification of the reality, the obtained model is satisfactory to explain the presented

624

questions. Thus, the results and interpretations addressed in this study may clarify issues

625

not only regarding the evolution of the barrier-lagoon systems of the Rio Bonito

626

Formation, in the Paraná Basin, but also in similar contexts of other basins. 25

26

ACCEPTED MANUSCRIPT 627

Acknowledgments

629

This article is the result of the master's thesis of the first author developed in the

630

Graduate Program in Geology of the Unisinos University (Universidade do Vale do Rio

631

dos Sinos) and supported by CAPES/PROSUP and FAPERGS [ARD/PPP 16/2551-

632

000274-6]. The authors thank the Brazilian Geological Survey (CPRM) for permission

633

to cores description and Léo Afraneo Hartmann for his contributions.

RI PT

628

636

References

Alves, R.G., Ade, M.V.B., 1996. Sequence stratigraphy and coal petrography applied to

M AN U

635

SC

634

637

the Candiota Coal Field, Rio Grande do Sul, Brazil: A depositional model.

638

International Journal of Coal Geology, 30, 231-248.

640 641

Beicip-Franlab. 2009. Training guide for Dionisos 3.0 Industrial Version: 3D basin modeling. Beicip-Franlab, Paris, 137 p.

TE D

639

Boardman, D., Souza, P.A., Iannuzzi, R., Mori, A.L.O., 2013. Paleobotany and palynology of the Rio Bonito Formation (Lower Permian, Paraná Basin, Brazil)

643

at the Quitéria Outcrop. Ameghiniana, 49, 451-472.

EP

642

Bohacs, K.M., Suter, J., 1997. Sequence stratigraphic distribution of coaly rocks:

645

Fundamental controls and examples. American Association of Petroleum

646

AC C

644

Geologists Bulletin, 81, 1612-1639.

647

Borren, W., Bleuten, W., Lapshina, E.D. Holocene peat and carbon accumulation rates

648

in the southern taiga of western Siberia. Quaternary Research, 61, 42-51.

649

Bortoluzzi, C.A., Awdziej, J., Zardo, S.M., 1987. Geologia da Bacia do Paraná em

650

Santa Catarina. In: Silva, L.C., Bortoluzzi, C.A. (Eds.), Textos básicos de

651

geologia e recursos minerais de Santa Catarina - Mapa geológico do Estado de

26

27

ACCEPTED MANUSCRIPT 652

Santa Catarina escala 1:50.000: texto explicativo e mapa. Departamento

653

Nacional de Produção Mineral, Brasil, 135-167.

654 655

Boyd, R., Dalrymple, R., Zaitlin, B.A., 1992. Classification of clastic coastal depositional environments. Sedimentary Geology, 80, 139–150. Buatois, L.A., Netto, R.G., Mángano, M.G., 2001. Reinterpretación paleoambiental de

657

la Formación Rio Bonito (Pérmico de la Cuenca de Paraná) en el yacimiento de

658

carbón de Iruí, Rio Grande do Sul, Brasil: integración de análisis de fácies,

659

icnología y estratigrafia secuencial de alta resolución. Geogaceta, 29, 27-30.

SC

660

RI PT

656

Buatois, L.A., Netto, R.G., Mangano, M.G., 2007. Ichnology of Permian marginal- to shallow-marine coal-bearing successions: Rio Bonito and Palermo Formations,

662

Paraná Basin, Brazil. In: MacEachern, J.A., Bann, K.L., Gingras, M.K.,

663

Pemberton, S.G. (Eds.), Applied Ichnology e SEPM Short Course Notes. SEPM,

664

Tulsa, 167-178.

M AN U

661

Cacela, A.S.M., 2008. Paleoclima e dinâmica costeira como fatores controladores da

666

distribuição de arenitos em sistemas parálicos – Um estudo para reservatórios

667

análogos no Eopermiano da Bacia do Paraná. Master thesis, UFRGS University. Cagliari, J., 2014. Contexto deposicional e modelagem estratigráfica direta dos

EP

668

TE D

665

depósitos sedimentares Permocarboníferos nas jazidas de carvão Capané e Iruí

670

Central (Formação Rio Bonito, Bacia do Paraná). Ph.D. thesis, UNISINOS

671 672

AC C

669

University.

Cagliari, J., Tognoli, F.M.W., Lavina, E.L.C., 2014. Influência dos controles externos

673

na deposição da Formação Rio Bonito na borda sudeste da Bacia do Paraná: uma

674

análise a partir da modelagem estratigráfica direta. In: Contexto deposicional e

675

modelagem estratigráfica direta dos depósitos sedimentares Permocarboníferos

676

nas jazidas de carvão Capané e Iruí Central (Formação Rio Bonito, Bacia do

27

28

ACCEPTED MANUSCRIPT 677 678

Paraná). PPG Universidade do Vale dos Sinos, Brasil, 188-216. Cagliari, J., Philipp, R.P., Buso, V.V., Netto, R.G., Hillebrand, P.K., Lopes R. da C., Basei, M.A.S., Faccini, U.F., 2016. Age constraints of the glaciation in the

680

Parana Basin: evidence from new U-Pb dates. Journal of the Geological Society,

681

173, 871-874.

RI PT

679

Castro, M.R., Perinotto, J.A.J., Castro, J.C., 1999. Fácies, análise estratigráfica e

683

evolução pós-glacial do Membro Triunfo/Formação Rio Bonito, na faixa

684

subaflorante do norte catarinense. Revista Brasileira de Geociências, 29, 533-

685

538.

687

Christofoletti, A., 1999. Modelagem de Sistemas Ambientais, Edgard Blücher, São

M AN U

686

SC

682

Paulo.

688

Clifton, H.E., 1969. Beach lamination: nature and origin. Marine Geology, 7, 553-559.

689

Collinson, J.D., 1996. Alluvial sediments. In: Reading, H.G. (Ed.), Sedimentary Environments: Process, Facies and Stratigraphy. Blackwell Science, Oxford, 37-

691

82.

694 695

and Hall, London.

EP

693

Collinson, J.D., Thompson, D.B., 1989. Sedimentary structures, second ed. Chapman

Collinson, J.D., Mountney, N., Thompson, D.B., 2006. Sedimentary Structures, third ed. Dunedin Academic Press, United Kingdom.

AC C

692

TE D

690

696

Companhia de Pesquisa de Recursos Minerais (CPRM), 1980. Projeto Iruí/Butiá, Bloco

697

Iruí, Carvão Mineral: Relatório Final de Pesquisa, Alvarás n° 2396, 2397, 2471,

698 699

3835, 3956, 3957, 3996, 3997/77 e 7957/78. DNPM/CPRM, Porto Alegre. Companhia de Pesquisa de Recursos Minerais (CPRM), 1983. Projeto Iruí/Butiá Área

700

Cordilheira: Relatório Final de Pesquisa, Alvarás nº 2573, 4616/80.

701

DNPM/CPRM, Porto Alegre, 2v.

28

29

ACCEPTED MANUSCRIPT 702 703

Corrêa da Silva, Z.C., 2004. Coal facies studies in Brazil: a short review. International Journal of Coal Geology, 58, 119-124. Courel, L. 1987. Stages in the compaction of peat; examples from the Stephanian and

705

Permian of the Massif Central, France. Journal of the Geological Society, 144,

706

489-493.

707

RI PT

704

Csato, I., Granjeon, D., Catuneanu, O., Baum, G.R. 2013. A three-dimensional

stratigraphic model for the Messianian crisis in the Pannonian Basin, eastern

709

Hungary. Basin Research, 25, 121-148.

711 712

Cúneo, N.R., 1996. Permian phytogeography in Gondwana. Palaeogeography, Palaeoclimatology, Palaeoecology, 125, 75-104.

M AN U

710

SC

708

Dargie, G.C., Lewis, S.L., Lawson, I.T., Mitchard, E.T.A., Page, S.E., Bocko, Y.E.,

713

Ifo,A.A. Age, extend and carbon storage of central Congo Basin peatland

714

complex. Nature, 542, 86-90.

716

Davis, R.A., 1994. Geology of Holocene barrier island systems, Springer-Verlag, New York.

TE D

715

Dott, R.H.Jr., Bourgeois, J., 1982. Hummocky stratification: Significance of its variable

718

bedding sequences. Geological Society of America Bulletin, 93, 663-680.

720 721 722

Embry, A.F., 1993. Transgressive–regressive (T–R) sequence analysis of the Jurassic succession of the Sverdrup basin, Canadian Arctic Archipelago. Canadian

AC C

719

EP

717

Journal of Earth Sciences, 30, 301–320.

Embry, A.F., Johannessen, E.P., 1992. T–R sequence stratigraphy, facies analysis and

723

reservoir distribution in the uppermost Triassic- Lower Jurassic succession,

724

western Sverdrup basin, Arctic Canada. In: Vorren, T.O., Bergsager, E., Dahl-

725

Stamnes, O.A., Holter, E., Johansen, B., Lie, E., Lund, T.B. (Eds.), Arctic

726

Geology and Petroleum Potential, Special Publication. Norwegian Petroleum

29

30

ACCEPTED MANUSCRIPT 727

Society, Norway, 121–146.

728

Faccini, U.F., 2007. Tectonic and climatic induced changes in depositional styles of the

729

Mesozoic sedimentary record of Southern Parana Basin, Brazil. In: Problems in

730

Western Gondwana Geology I, Gramado, 42–45. Faccini, U.F., Giardin, A., Garcia, A.J.V., Machado, J.L.F., 2003.

RI PT

731

Megaheterogeneidades e hidroestratigrafia do Aqüífero Guarani na região de

733

SantaMaria (RS). In: Paim, P.S.G., Faccini, U.F., Netto, R.G. (Eds.), Geometria,

734

arquitetura e heterogeneidade de corpos sedimentares: estudo de casos. Unisinos,

735

São Leopoldo, 78-92.

738 739 740

M AN U

737

Ferreira, J.A.F., Süfert, T., Santos, A.P., 1978. Projeto Carvão no Rio Grande do Sul: Relatório Final. DNPM/CPRM, Porto Alegre, 16v.

Figueroa, I., Dias, A. de A., 1983. Projeto Carvão na Área do Iruí: Relatório Final. DNPM/CPRM, Porto Alegre, 5v.

Franco, D. R., Ernesto, M., Ponte-Neto, C. F., Hinnov, L. A., Berquo, T. S., Fabris, J.

TE D

736

SC

732

D., Rosie, C. A., 2012. Magnetostratigraphy and mid-palaeolatitude VGP

742

dispersion during the Permo-Carboniferous Superchron: results from Parana ́

743

Basin (Southern Brazil) rhythmites. Geophysical Journal International, 191, 993-

744

1014.

746 747 748

França, A.D., Potter, P.E., 1988. Estratigrafia, ambiente deposicional e análise de

AC C

745

EP

741

reservatório do Grupo Itararé (Permocarbonífero), Bacia do Paraná (parte I). BoI. Geoc. Petrobrás, 2, 147-191.

Frey, R.W., Basan, P., 1985. Coastal Salt Marshes. In: Davis, R.A. (Ed.), Coastal

749

sedimentary environments. Spring-Verlag, Massachusetts, 101-169.

750

Gandini, R., 2010. Assinaturas icnológicas da sucessão sedimentar Rio Bonito no

751

bloco central da jazida carbonífera de Iruí, Cachoeira do Sul (RS). Master thesis,

30

31

ACCEPTED MANUSCRIPT 752 753

UNISINOS University. Gandini, R., Netto, R.G., Kern, H.P., Lavina, E.L.C., 2010. Assinaturas icnológicas da

754

sucessão sedimentar Rio Bonito no bloco central da jazida carbonífera de Iruí,

755

Cachoeira do Sul (RS). Gaea, 6, 21-43.

757 758

Gomes, A.P., Ferreira, J.A.F., Albuquerque, L.F., Süffert, T., 1998. Carvão Fóssil.

RI PT

756

Estudos Avançados, 12, 89-106.

Gordon Jr., M.J., 1947. Classificação das formações gondwânicas do Paraná, Santa Catarina e Rio Grande do Sul. Notas Preliminares e Estudos da Divisão de

760

Geologia e Mineralogia do DNPM, 38, 1-20.

Granjeon, D., 2014. 3D forward modelling of the impact of sediment transport and

M AN U

761

SC

759

762

base level cycles on continental margins and incised valleys. Int. Assoc.

763

Sedimentol. Spec. Publ., 46, 453–472.

764

Harms, J.C., Southard, J.B., Spearing, D.R., Walker, R.G., 1975. Depositional environments as interpreted from primary sedimentary structures and

766

stratification sequences, first ed. Society of Economic Paleontologists and

767

Mineralogists, Dallas.

TE D

765

Hayes, M.O., 1975. Morphology of sand accumulation in estuaries: an introduction to

769

the symposium. In: Eugene Cronin, L. (Ed.), Estuarine Research. Geology and

770

engineering. Academic Press, New York, 3–22.

772 773

AC C

771

EP

768

Heward, A.P., 1981. A review of wave-dominated clastic shoreline deposits. Earth-Sci. Rev., 17, 223-276.

Holz, M., 2003. Sequence stratigraphy of a lagoonal estuarine system - an example from

774

the lower Permian Rio Bonito Formation, Paraná Basin, Brazil. Sedimentary

775

Geology, 162, 305–331.

776

Holz, M., Kalkreuth, W., Banerjee, I., 2002. Sequence stratigraphy of paralic coal-

31

32

ACCEPTED MANUSCRIPT 777

bearing strata: an overview. International Journal of Coal Geology, Amsterdam,

778

48, 1– 33.

779

Holz, M., Souza, P.A., Ianuzzi, R., 2008. Sequence stratigraphy and biostratigraphy of the Late Carboniferous to Early Permian glacial succession (Itararé Subgroup) at

781

the eastern-southeastern margin of the Paraná Basin, Brazil. Geological Society

782

of America Special, 441, 115–129.

784 785

Huang, Z., Gradstein, F., 1990. Depth-porosity relationship from deep sea sediments. Sci. Drill., 1, 157-162.

SC

783

RI PT

780

Ianuzzi, R., 2013. The Carboniferous-Permian floral transition in the Paraná Basin. In: Lucas, S.G., DiMichele, W.A., Barrick, J.E., Schneider, J.A., Spielmann, J.A.

787

(Eds.), The Carboniferous-Permian Transition. Museum of Natural History and

788

Science, New Mexico, 132-136.

789

M AN U

786

Kalkreuth, W., Holz, M., Mexias, A., Balbinot, M., Levandowski, J., Willett, J., Finkelman, R., Burger, H., 2010. Depositional setting, petrology and chemistry

791

of Permian coals from the Paraná Basin: South Santa Catarina Coalfield, Brazil.

792

International Journal of Coal Geology, 84, 213–236. Kern, H.P., 2008. Arquitetura estratigráfica de corpos arenosos gerados por ondas e

EP

793

TE D

790

marés no Bloco Central da Mina do Iruí (Formação Rio Bonito, eopermiano da

795

Bacia do Paraná, RS). Master thesis, UNISINOS University.

AC C

794

796

Kjerfve, B., 1994. Coastal lagoons. Elsevier Oceanography Series, 60, 1–8.

797

Kraft, J.C., Chrzastowski, M.J., 1985. Coastal stratigraphic sequences. In: Davis, R.A.

798

(Ed.), Coastal sedimentary environments. Spring-Verlag, Massachusetts, 187-

799

224.

800 801

Lavina, E.L., Nowatzki, C.H., Santos, M.A.A., Leão, H.Z., 1985. Ambientes de sedimentação do Super-Grupo Tubarão na região de Cachoeira do Sul. Acta

32

33

ACCEPTED MANUSCRIPT 802 803

Geolog. Leopoldensia, 9, 5-75. Lavina, E.L.C., Lopes, R.C., 1987. A Transgressão Marinha do Permiano Inferior e a

804

Evolução Paleogeográfica do Supergrupo Tubarão no Estado do Rio Grande do

805

Sul. Paula-Coutiana, 1, 51-103. Lopes, R.C., 1995. Arcabouço aloestratigráfico para o intervalo “Rio Bonito Palermo”

RI PT

806 807

(Eopermiano da Bacia do Paraná), entre Butiá e São Sepé, Rio Grande do Sul.

808

Master thesis, UNISINOS University.

Lopes, R.C., Lavina, E.L.C., 2001. Estratigrafia de sequências nas formações Rio

SC

809

Bonito e Parlermo (Bacia do Paraná), na região carbonífera do Jacuí, Rio Grande

811

do Sul. In: Ribeiro, H.J.P.S. (Eds.), Estratigrafia de Sequências: Fundamentos e

812

Aplicações. Edunisinos, São Leopoldo, 391-419.

813

M AN U

810

Lopes, R.C., Faccini, U.F., Paim, P.S.G., Garcia, A.J.V., Lavina, E.L., 2003a. Barras de maré na formação Rio Bonito: elementos arquiteturais e geometria dos corpos

815

(Iruí e Capané - RS). In: Paim, P.S.G., Faccini, U.F., Netto, R.G. (Eds.),

816

Geometria, arquitetura e heterogeneidade de corpos sedimentares: estudo de

817

casos. Unisinos, São Leopoldo, 78 -92. Lopes, R.C., Lavina, E.L., Paim, P.S.G., Goldberg, K., 2003b. Controle estratigráfico e

EP

818

TE D

814

deposicional da gênese dos carvões da região do Rio Jacuí (RS). In: Paim,

820

P.S.G., Faccini, U.F., Netto, R.G. (Eds.), Geometria, arquitetura e

821 822 823

AC C

819

heterogeneidade de corpos sedimentares: estudo de casos. Unisinos, São Leopoldo, 187-206.

Lopes, R.C., Paim, P.S.G., Lavina, E.L., 2003c. Modelo de reservatório em arenitos

824

litorâneos: ilha de barreira Permiana na Formação Rio Bonito (Minas do Leão -

825

RS). In: Paim, P.S.G., Faccini, U.F., Netto, R.G. (Eds.), Geometria, arquitetura e

826

heterogeneidade de corpos sedimentares: estudo de casos. Unisinos, São

33

34

ACCEPTED MANUSCRIPT 827 828

Leopoldo, 60-77. Matos, S.L.F., Yamamoto, J.K., Hachiro, J., Coimbra, A.M., 2000. Tonsteins da

829

Formação Rio Bonito no depósito de carvão Candiota, RS. Revista Brasileira de

830

Geociências, 30, 679-684.

832 833

McCabe, P.J., 1984. Depositional envinronments of coal and coal-bearing strata. Spec.

RI PT

831

Publs int. Ass. Sediment., 7, 13-42.

McCubbin, D.G., 1982. Barrier-Island and Strand-Plain Facies. In: Scholle, P.A., Spearing, D. (Eds.), Sandstone Depositional Environments. American

835

Association of Petroleum Geologist, Tulsa, 247-279.

837 838

Miall, A.D., 1977. A review of the braided river depositional environment. Earth Science Reviews, 13, 1-62.

M AN U

836

SC

834

Milani, E.J., 1997. Evolução tectono-estratigráfica da Bacia do Paraná e seu relacionamento com a geodinâmica fanerozóica do Gondwana sul-ocidental.

840

Ph.D. thesis, UFRGS University.

841 842

TE D

839

Milani, E.J., Melo, J.H.G., Souza, P.A., Fernandes, L.A., França, A.B., 2007. Bacia do Paraná. Boletim de Geociências da Petrobrás, 8, 69-82. Mulhern, J.S., Johnson, C.L., Martin, J.M., 2017. Is barrier island morphology a

844

function of tidal and wave regime? Marine Geology, 387, 74-84.

846 847 848

Netto, R.G., 2001. Icnologia e Estratigrafia de Sequências. In: Ribeiro, H.J.P.S. (Ed.),

AC C

845

EP

843

Estratigrafia de Sequências: Fundamentos e Aplicações. Edunisinos, São Leopoldo, 219-259.

Netto, R.G., Tognoli, F.M.W., Gandini, R., Lima, J.H.D., Gibert, J.M., Bosetti, E.P.,

849

2012. Ichnology of the Phanerozoic Deposits of Southern Brazil: Synthetic

850

Review. In: Netto, R.G., Carmona, N.B., Tognoli, F.M.W. (Eds.), Ichnology of

851

Latin America - Selected Papers. Sociedade Brasileira de Geologia, Porto

34

35

ACCEPTED MANUSCRIPT

853 854 855 856

Alegre, 37-68. Nichols, M.M., 1989. Sediment accumulation rates and relative sea-level rise in lagoons. Marine Geology, 88, 201-219. Nirmala Khandan, N., 2002. Modeling tools for environmental engineers and scientists, CRC Press LLC, Florida.

RI PT

852

Pemberton, S.G., Spila, M., Pulham, A.J., Saunders, T., MacEachern, J.A., Robbins, D.,

858

Sinclair, I.K., 2001. Ichnology and sedimentology of shallow to marginal marine

859

systems: Ben Nevis and Avalon reservoirs, Jeanne D’Arc Basin, Geological

860

Association of Canada Short Course Notes, St John’s.

Rahmani, R.A., 1988. Estuarine tidal channel and nearshore sedimentation of a Late

M AN U

861

SC

857

862

Cretaceous epicontinental sea, Drumheller, Alberta, Canada. In: Boer, P.L., Van

863

Gelder, A., Nio, S.D. (Eds.), Tide-Influenced Sedimentary Environments and

864

Facies. D. Reidel Publishing Company, Holland, 433-471.

866 867

Reineck, H.E., Singh, I.B., 1986. Depositional sedimentary environments with reference

TE D

865

to terrigenous clastics, second ed. Springer-Verlag, Berlin. Reinson, G.E., 1984. Barrier island and associated strand-plain systems. In: Walker, R.G. (Ed.), Facies Models. Geological Association of Canada , Toronto, 119-

869

140.

871 872 873

Reinson, G.E., 1992. Transgressive barrier island and estuarine systems. In: Walker,

AC C

870

EP

868

R.G., James, N.P. (Eds.), Facies Models - Response to Sea Level Change. Geological Association of Canada Publications, Canada, 179-194.

Ryer, T. A., Langer, A. W., 1980. Thickness change involved in the peat-to-coal

874

transformation for a bituminous coal of Cretaceous age in central Utah. Journal

875

of Sedimentary Petrology, 50, 987-992.

876

Rocha-Campos, A.C., 1967. The Tubarão group in the Brazilian portion of the Paraná

35

36

ACCEPTED MANUSCRIPT 877 878 879 880

basin. Problems in Brazilian Gondwana Geology, 27-102. Rocha-Campos, A.C., Santos, P.R., Canuto, J.R., 2008. Late Paleozoic glacial deposits of Brazil: Paraná Basin. Geological Society of America Bulletin, 441, 97-114. Schneider, R.L., Mühlmann, H., Tommasi, I.E., Medeiros, R.S., Daemon, R.F., Nogueira, A.A., 1974. Revisão estratigráfica da Bacia do Paraná. In: XXVIII

882

Congresso Brasileiro de Geologia, Porto Alegre, 41-66.

883

RI PT

881

Sclater, J.G., Christie, P.A.F., 1980. Continental stretching: Na explanation of the postmid-Cretaceous subsidence of the Central North Sea Basin. Journal Geophysical

885

Research, 85, 3711-3739.

Seard, C., Borgomano, J., Granjeon, D., Camoin, G. Impact of environmental

M AN U

886

SC

884

887

parameters on coral reef development and drowning: forward modeling on the

888

last deglacial reefs from Tahiti (French Polynesia, IODP Expedition #310).

889

Sedimentology, 60, 1357-1388.

Schafie, K.R., Madon, M. 2008. A review of stratigraphic simulation techniques and

TE D

890 891

their applications in sequence stratigraphy and basin analysis. Geological

892

Society of Malaysia, 54, 81-89.

Slonski, G.T., 2002. Interpretação paleoclimática do Permiano Inferior da Bacia do

EP

893

Paraná em Santa Catarina, Brasil (Formação Rio Bonito). Master thesis, UFSC

895

University.

AC C

894

896

Smith, D.B., 1988. Bypassing of sand over sand waves and through a sand wave field in

897

the central region of the southern North Sea. In: Boer, P.L., Van Gelder, A., Nio,

898

S.D. (Eds.), Tide-Influenced Sedimentary Environments and Facies. D. Reidel

899

Publishing Company, Holland, 39-50.

900 901

Tognoli, F.M.W., Netto, R.G., 2003. Ichnological signature of paleozoic estuarine deposits from the Rio Bonito-Palermo succession, eastern Paraná Basin, Brazil.

36

37

ACCEPTED MANUSCRIPT 902 903

Asociación Paleontológica Argentina, 1, 141-155. Vail, P.R., Audemard, F., Bowman, S.A., Eisner, P.N., Perez-Cruz, G., 1991. The stratigraphic signature of tectonics, eustasy and sedimentation: An overview. In:

905

Einsele, G., Ricker, W., Seilacher, A. (Eds.), Cycles and Events in Stratigraphy.

906

Springer-Verlag, Berlin, 617-659.

907

RI PT

904

Walker, R., 1992. Facies, facies models and modern stratigraphic concepts. In: Walker, R.G., James, N.P. (Eds.), Facies Models - Response to Sea Level Change.

909

Geological Association of Canada Publications, Canadá, 1-14.

911 912

Wentworth, C.K., 1992. A Scale of Grade and Class Terms for Clastic Sediments. The Journal of Geology, 30, 377-392.

M AN U

910

SC

908

Zerfass, H., Lavina, E.L., Schultz, C.L., Garcia, A.J.V., Faccini, U.F., Chemale Jr., F., 2003. Sequence-stratigraphy of continental strata of Southernmost Brazil: a

914

contribution to Southwestern Gondwana palaeogeography and palaeoclimate.

915

Sedimentary Geology, 161, 85-105.

916

TE D

913

Ziegler, A.M., Raymond, A.L., Gierlowski, T.C., Horrell, M.A., Rowley, D.B., Lottes, A.L., 1987. Coal, climate and terrestrial productivity: the present and early

918

Cretaceous compared. Geological Society, 32, 25-49.

920 921

TABLE CAPTIONS

AC C

919

EP

917

922

Table 1 – Input data of stratigraphic simulation.

923

Table 2 – Summary of sedimentary facies of Rio Bonito Formation in analyzed

924

depositional sequence.

925 926

FIGURE CAPTIONS

37

38

ACCEPTED MANUSCRIPT 927

Figure 1 – A: Study area located in the southern Paraná Basin; B: Detailed study area

929

with cores and stratigraphic sections (white section is from Cagliari, 2014, see Fig. 2);

930

Insert figure shows the location in South America; C: Paleogeographic map of the

931

eastern border of the Paraná Basin (Lavina and Lopes, 1987).

932

Figure 2 – Capané-Iruí stratigraphic section with third-order depositional sequences

933

according to Lopes (1995) and fourth-order depositional sequence, which is the focus of

934

this study (adapted from Cagliari, 2014).

935

Figure 3 – Paleobathymetric map for the studied area prior to coal deposition. The NE-

936

SW direction is parallel to the paleocoastline.

937

Figure 4 – Core photographs of the sedimentary facies.

938

Figure 5 – Vertical succession of facies. A: Detailed profile showing sandy barrier

939

succession; B: Sandy barrier interleaved with lagoonal deposits.

940

Figure 6 – Isopach map of the upper Iruí coal seam. Isocontours are in meters.

941

Figure 7 – Stratigraphic sections. AA’ corresponds to SW-NE section, parallel to

942

paleocoastline, and BB’ corresponds to NW-SE section, perpendicular to

943

paleocoastline. Facies code: F1 - Coal; F2 - Laminated siltstone; F3 - Structureless

944

siltstone; F4 – Heterolithic sandstone; F5 - Trough cross-bedded arkosic sandstone; F6 -

945

Parallel to subhorizontal lamination quartz sandstone; F7 - Trough cross-bedded quartz

946

sandstone; F8 – Bioturbated heterolithic sandstone; F9 - Hummocky cross-stratification

947

quartz sandstone.

948

Figure 8 – Relative sea level curve and sediment supply established for the stratigraphic

949

simulation.

950

Figure 9 – Results of stratigraphic simulation in different stages, shown as sediments

951

relative distribution. A: Peat accumulation at 0.5 Ma; B, C: Lagoon deposition at 0.7

AC C

EP

TE D

M AN U

SC

RI PT

928

38

39

ACCEPTED MANUSCRIPT Ma and 1.05 Ma; D: Sandy barrier transgression over continental deposits.

953

Figure 10 – A: W-E simulated stratigraphic section, corresponding to stratigraphic

954

section AA’; B: N-S simulated stratigraphic section, corresponding to stratigraphic

955

section BB’ (complete section in Figure 7). The red arrows and values show the

956

locations and extensions (in kilometers) of the sedimentary sources.

957

Figure 11 – Real (left) and predicted (right) lithologies. The real cores are from Figure 7

958

and predicted ones are the result of the forward stratigraphic simulation. A: Similar

959

thickness obtained for IB-174-RS; B-C: Predicted deposits are shifted upward due to the

960

absence of the erosion processes in the simulation (note coal in B and siltstone in C); D:

961

Sandstone succession in core IB-012-RS is thicker than predicted one.

M AN U

SC

RI PT

952

AC C

EP

TE D

962

39

ACCEPTED MANUSCRIPTReference Value

Parameter

Sand: ≤ 2 mm Silt: ≤ 0.06 mm Coal: ≤ 0.004 mm

Wentworth (1922)

Amplitude of the relative sea level rise

16 m

Estimated from the sedimentary record considering the preserved thickness equal to the creation of the accommodation space

Rate of the relative sea level rise

0-10 m/Ma

Obtained from the stratigraphic simulation

Compaction parameters

Sand: ∅ = 0.49; c = 270 Silt: ∅ = 0.75; c = 580 Coal: ∅ = 0.48; c = 900

Sclater and Christie (1980) Huang and Gradstein (1990) Beicip-Franlab (2009)

Total buried depth

3,000 m

Cagliari (2014)

Erosion

Not considered

-

Subsidence

Not configured

Volume of compacted sediment

Sand: 4.6 km³ Silt: 0.45 km³ Coal: 0.59 km³

Total sediment supply

2.38 km³/Ma

Obtained from stratigraphic simulation

Peat accumulation

30 m/Ma

Obtained from boreholes and simulation, according to paleobathymetry

Precipitation

2,400 mm/y

Annual averages necessary for peat accumulation according to Ziegler et al. (1987)

SC

TE D

Supply

Obtained from stratigraphic simulation

Average depositional slope

Continental: 0.2 m/km Marine: 200 m/km

Beicip-Franlab (2009) Beicip-Franlab (2009)

Gravity-driven (sand + silt)

Continental: 26.52 km²/ka Marine: 0.018335 km²/ka

Obtained from stratigraphic simulation

Water-driven (sand + silt)

Continental: 26.52 km²/ka Marine: 0.018335 km²/ka

Obtained from the stratigraphic simulation

Wave influence (sand + silt)

53.14 km²/ka

Obtained from the stratigraphic simulation

Wave incidence

Summer waves (constructive): NE Winter waves (destructive): SW

Rio Bonito Formation waves incidence estimated by Cacela (2008)

AC C Transport

Estimated from the volume of sediment preserved in the studied depositional sequence

Position: wide - flow rate East: 0.05 km - 22 m3/s South: 2 km - 5 m3/s North: 0.05 km - 37.5 m3/s South: 0.05 km - 12.5 m3/s East: 1.8 km - 2.5 m3/s North: 1 km - 3.75 m3/s West: 0.8 km - 37.5 m³/s

EP

Fluvial discharge (average)

-

M AN U

Basin

RI PT

Grain size Lithology

∅ : depositional porosity; c: exponential coefficient.

ACCEPTED MANUSCRIPT Code

Sedimentary Facies

F1

Coal

Facies Association

Interpretation Swamp

Heterolithic sandstone

F5

Trough cross-bedding arkosic sandstone

F6

Parallel to subhorizontal lamination quartz sandstone

F7

Trough cross-bedding quartz sandstone Bioturbated heterolithic sandstone

F9

Hummocky crossstratification quartz sandstone

EP AC C

Channel-fill bars

Foreshore

Tide-influenced deposits

Upper shoreface Wave-dominated facies association

TE D

F8

Tidal-dominated facies association

Estuary/open lagoon

RI PT

F4

Back-barrier

Lagoon

SC

Laminated siltstones and massive siltstone

M AN U

F2, F3

Standing-waterdominated facies

Lower shoreface

Lower shoreface under storm conditions

Sandy barrier

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

ACCEPTED MANUSCRIPT - Facies interpretation: swamp, lagoon, sandy barrier and tide-influenced deposits; - The coal bearing succession was deposited during the transgressive system tract; - Peat deposition, barrier breakup and migration landward occurred in sequence; - Relative sea level curve and sedimentary sources location were obtained from modeling;

AC C

EP

TE D

M AN U

SC

RI PT

- Relative sea level rose 16 meters during simulation period;