You are on page 1of 19

Journal of South American Earth Sciences 29 (2010) 381–399

Contents lists available at ScienceDirect

Journal of South American Earth Sciences


journal homepage: www.elsevier.com/locate/jsames

A stratigraphic chart of the Late Carboniferous/Permian succession of the


eastern border of the Paraná Basin, Brazil, South America
Michael Holz a,*, Almério B. França b, Paulo A. Souza a, Roberto Iannuzzi a, Rosemarie Rohn c
a
UFRGS-Av. Bento Gonçalves 9500, C.Postal 15001-91501-970-Porto Alegre, RS, Brazil
b
Petróleo Brasileiro-Depex, E&P-EXP-Av. República do Chile, 65/1302-20031-912-Rio de Janeiro, RJ, Brazil
c
UNESP, C. Postal 178, 13506-900-Rio Claro, SP, Brazil

a r t i c l e i n f o a b s t r a c t

Article history: Sequence stratigraphy, lithostratigraphy and biostratigraphy of the Late Carboniferous/Permian succes-
Received 15 October 2008 sion of the eastern border of the Paraná Basin are organized in the form of a detailed chart in order to
Accepted 17 April 2009 provide a useful and updated synthesis of that stratigraphic interval. The traditional lithostratigraphic
subdivision is shown together with a third-order stratigraphic framework and a complete biostrati-
graphic scheme based upon palynomorphs, plants and invertebrate macrofossils. Based on the regional
Keywords: occurrence of features that indicate base level fall and formation of a sequence boundary (e.g., marine
Late Carboniferous
to fluvial facies shift, pebbly lags, bonebeds) seven LPTS’s (=Late Paleozoic Third-Order Sequences) were
Permian
Paraná Basin
recognized and are shown in the stratigraphic chart, enclosed as an color-printed appendix at the end
Sequence stratigraphy of the paper. The text includes a brief characterization of the main facies and depositional systems of
Biostratigraphy the lithostratigraphic units, a description of the fossil content of each lithostratigraphic unit, and a char-
acterization in terms of sequence stratigraphy, with a brief description of eight third order sequence
boundaries as depicted in the stratigraphic chart, including recent advances on radiometric dating.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction The prevalence of eustatic-tectonic cycles, which controlled


sedimentation in the Paraná Basin, has generated a stratigraphic
The Paraná Basin is a huge intracratonic basin on the South- record that is marked by numerous interruptions caused by ero-
American platform, located in southernmost Brazil and north/ sion and non-deposition. Milani et al. (1994, 1998, 2007) consid-
northwestern Uruguay, parts of Paraguay and Argentina (Fig. 1A). ered that the fill of the basin is constituted of six second order
The basin covers a surface area of about 1,700,000 km2, has a pres- depositional sequences, ranging in age from Late Ordovician to Late
ent NE–SW elongated shape, and is approximately 1750 km long Cretaceous (Fig. 1B). The stratigraphic interval studied herein cor-
and 900 km wide. The sedimentary fill of the basin was condi- responds to the third sequence of Milani et al. (2007), named as
tioned by tectonic-eustatic cycles linked to the evolution of West ‘‘Gondwana I Supersequence”, ranging from the Late Carboniferous
Gondwana during Palaeozoic and Mesozoic times. to the Late Permian.
The Basin has been emplaced over different geotectonic do- The present paper deals with the sedimentary succession of the
mains, comprising Archean and Paleoproterozoic cratonic terrains eastern border of the basin (Fig. 1A), showing a comprehensive
and Neoproterozoic mobile belts related to the Pan-African and integration of lithostratigraphy, biostratigraphy and third order se-
Brasiliano orogenies, responsible for the assemblage of West quence stratigraphy (i.e. sequences with a typical 1–11 m.y. dura-
Gondwana. Hence, the geotectonic framework of the basement is tion, e.g., Krapetz, 1996). The main goal is to offer an integrative
characterized by several cratonic blocks and intervening mobile overview of both published and new data, as well as interpreta-
belts, forming a complex framework of lineaments and crustal dis- tions on the geological evolution of these deposits. Although se-
continuities which influenced sedimentation due to differential quence stratigraphy deals rather with analytic tools than with
subsidence and uplift of the tectonic blocks (Holz et al., 2006). formal stratigraphic units (as lithostratigraphy does), we have cho-
sen to present the overview with subdivisions based on lithostra-
tigraphy, which still is a kind of ‘‘reference for communication”
between stratigraphers. Insofar, sentences like ‘‘part of the Rio Bo-
* Corresponding author. Address: Universidade Federal do Rio Grande do Sul,
nito Formation integrates sequence 2”, although conceptually
Instituto de Geociências, Av. Bento Gonçalves 9500, 91501 970, Porto Alegre, Brazil.
Tel.: +55 51 3308 6836; fax: +55 51 3308 7047. wrong (because different concepts are mixed up), will appear in or-
E-mail address: michael.holz@ufrgs.br (M. Holz). der to facilitate comprehension, because lithostratigraphical

0895-9811/$ - see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jsames.2009.04.004
382 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

Fig. 1. (A) Simplified geologic map of the eastern/southeastern margin of the Paraná Basin and the main geographic information. (B) The stratigraphic chart (see Appendix A)
dealt with herein shows an approximately northeast–southwest oriented section, covering the four states São Paulo (SP), Paraná (PR), Santa Catarina (SC) and Rio Grande do
Sul (RS). B – the stratigraphic chart of the Paraná Basin (simplified after Milani et al. (2007)). The dotted rectangle marks the target of the present paper.

names are ‘‘classics” and widely known, in contrast to the sequence nozones are not related to the lithostratigraphical unit boundaries,
stratigraphic schemes. as well as the sequence boundaries. They correspond to important
Based on these premises, the text is organized as follows: variation in the flora composition, in response to major climate
changes.
 A brief characterization of the main facies and depositional sys- Geochronology and published radiometric data are also dis-
tems of the lithostratigraphic units; cussed. We adopted the time scale of Gradstein et al. (2004) to cal-
 A descriptive list of the fossil content of each lithostratigraphic ibrate all sequence stratigraphic and biostratigraphic surfaces.
unit, stressing mainly bivalves, conchostraceans, plants and pal- Although the main figure and goal of the paper is the enclosed
ynomorphs, on which the biostratigraphic schemes and correla- stratigraphic chart, some additional figures are shown for clarifica-
tions were based; tion of particular aspects of the stratigraphy shown in the chart.
 A characterization of the succession in terms of sequence stra- The stratigraphic chart, enclosed as a fold-out page at the end of
tigraphy, with a brief description of the sequence boundaries the paper, shows the formal stratigraphic units (geochronology,
as depictured in the stratigraphic chart. lithostratigraphy, biostratigraphy) at the left, while sequence stra-
tigraphy appears at the right side, showing that the Late Carbonif-
The header of each chapter will refer to the name of the strati- erous–Permian second-order sequence is herein divided into seven
graphic unit (e.g.; ‘‘Rio Bonito Formation”) followed by the approx- third-order sequences, labeled ‘‘Late Paleozoic third order
imate time frame of deposition (e.g., ‘‘Early Permian, Sakmarian to sequences” (LPTS) 1–7.
Artinskian”), without discussing lithostratigraphic hierarchy (e.g.,
whether a particular unit is treated as a group, a subgroup or for-
mation by other authors). Extensive literature revisions are 2. Itararé Group (Late Carboniferous to Early Permian:
avoided herein, but key references to every stratigraphic unit are Bashkirian/Moscovian to Early Sakmarian)
given.
Palynozones are assigned for each lithostratigraphic unit 2.1. Main facies, depositional systems and lithostratigraphy
according to Souza and Marques-Toigo (2005) and Souza (2006)
zonal schemes, which are derived from data of several boreholes This unit comprises up to 1500 m of glacio-continental and gla-
and outcrops widespread in the Basin, integrating the intervals cio-marine facies associations, consisting mainly of clastic litholo-
proposed by Daemon and Quadros (1970). Limits between paly- gies including diamictites, sandstones, rhythmites, mudstones,
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 383

shales, conglomerates and minor coal seams. The Itararé rests uppermost Itararé situated in São Paulo and Rio Grande do Sul. It
either directly upon Precambrian basement, or on Cambrian/Ordo- is characterized by the impressions–compressions of the first
vician strata of the Camaquã and Itajaí basins (in southern Brazil), appearance of glossopterid elements and by the abundance of
or on Devonian strata of the Paraná Group. The facies association shrubby sphenophytes. Among the glossopterid elements, the
points to deposition predominantly by rain out processes followed Gangamopteris-type leaves dominate, while Glossopteris-type
by downslope resedimentation, sediment gravity flows, turbidity leaves are less common in the assemblages. The Early Permian
currents, and minor lodgment tills. age of this flora is adopted on the basis of its association with ele-
França and Potter (1988) have divided its sedimentary succes- ments of the Vittatina costabilis Zone and the fact of the first
sion into the Lagoa Azul, Campo Mourão and Taciba Formations. appearance of glossopterids is considered as a marker of the Perm-
In Rio Grande do Sul and Santa Catarina, the topmost Itararé is ian strata throughout Gondwana (Iannuzzi and Souza, 2005).
mostly marine, known as the Rio do Sul Formation, with two Mem- In few localities of the Paraná Basin, there are occurrences of
bers called Budó (a sponge-rich marine mudstone) and Suspiro (a some typical elements (bivalves) related to Australian ‘‘Eurydesma
tillite). A recent stratigraphic overview can be found in Holz et al. Fauna”, recovered from the upper part of the Itararé Group (Rocha-
(2008). Campos and Rösler, 1978). The ‘‘Eurydesma Fauna” is considered
throughout Gondwana deposits as earliest Permian, e.g., Asselian
2.2. Fossils and biostratigraphy – Early Sakmarian interval (Iannuzzi and Souza, 2005; Iannuzzi,
2007).
Most typical fossils of the glacial succession, unfortunately not
suitable for biostratigraphic zonation, are gastropods (Peruvispira
delicata), brachiopods (Atenuatella and Langella), insects (protort- 2.3. Sequence stratigraphy
hopteroids, plecopteroids, Anphiesmemenoptera; and blattoids),
besides foraminifera, sponge and paleoniscoid fish fragments, The Itararé succession is located at the base of the second-order
ostracodes, escolecodontes and ichnofossils (Petri and Souza, depositional sequence of Milani et al. (1994) and has two regional
1993; Pinto, 1995). sequence boundaries, hence containing two Late Paleozoic third
Palynomorph data indicate that the glaciogenic sedimentation order Sequences (LPTS-1 and 2).
began during the Moskovian/Bashkirian (Late Carboniferous), but LPTS-1 is marked by the basal unconformity SB-1, which is rec-
the unit reaches up to Permian times (Sakmarian) (Souza, 2006). ognizable regionally and marked by the contact between the Late
In terms of macrofloral succession, the Itararé Group is basically Carboniferous to Early Permian succession and its underlying Neo-
composed of two main kinds of floras: one Carboniferous and an- proterozoic, Cambrian-Ordovician or Devonian basement. The SB-1
other Permian. The Late Carboniferous flora was previously is also a second-order boundary, marking the base of the second-
denominated as Taphoflora A, by Rösler (1978), and later as Pre- order depositional sequence of the framework of Milani et al.
Glossopteris Flora, by Iannuzzi and Souza (2005). This flora occurs (1994).
as compressed–impressed plant remains in only few localities of Within the Itararé succession there is an important erosional
São Paulo, having been recovered from the lowermost to the mid- feature linked to excavation and deposition during a major base le-
dle Itararé Group. vel fall, the so-called Lapa Sandstone and Vila Velha Sandstone,
It is characterized by the absence of glossopterid elements and which form a continuous channel-lobe system in southern Paraná
contains the typical plant taxa found in the ‘‘Nothorhacopteris-Bot- (França et al., 1996). The Lapa Sandstone (Fig. 2A) is interpreted as
rychiopsis-Ginkgophyllum Flora” for the Late Carboniferous of a huge incised valley fill (65 km long, 2 km wide and 100 m thick)
Argentina. Some of the phytofossiliferous beds are interbedded excavated during a sea level lowstand at the end of the Carbonifer-
with glaciogenic deposits. The Late Carboniferous age of this flora ous, and filled by sands from glacial outwash during the subse-
is based on the association with palynomorphs belonging to the quent transgression (França et al., 1996). The Vila Velha
Ahrensisporites cristatus Zone and the Crucisaccites monoletus Zone Sandstone (Fig. 2B) is the result of the stacking of three widespread
(Souza, 2006). lobes about 30 m thick each and 2 km wide, deposited probably
The Permian flora of this unit was designed as Phyllotheca- during the excavation of the Lapa incised valley.
Gangamopteris Flora by Iannuzzi and Souza (2005). It corresponds Therefore, this erosional excavation feature is indicative of an
to the ‘‘Transitional” or Taphoflora ‘‘A-B” of Rösler (1978). This flora important base-level fall, marking a sequence boundary within
also occurs in few localities but is widely spread through the east- the Itararé succession, herein labeled SB-2, which divides LPTS 1
ern margin of the Basin, having been found in localities from the and 2 and is coincident with the transition between two pal-

Fig. 2. Within the Itararé succession there is an important erosional feature linked to excavation and deposition during a major base level fall: A – the Lapa Sandstone, the
infill of a huge incised valley originated during a base level fall in the Late Carboniferous, B – the Vila Velha lobes, formed by sandstones deposited probably during the
excavation of the Lapa incised valley.
384 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

ynomorph interval zones (i.e., the A. cristatus and C. monoletus Siderópolis (coastal and fluviatile sandstones). The shallow marine
Zones, labeled AcZ and CmZ, respectively). Paraguaçu member is not identified in the southernmost part of
the basin (the Rio Grande do Sul State), where the Rio Bonito For-
3. Rio Bonito Formation (Early Permian: Early Sakmarian to mation is not formally subdivided into members.
Middle Artinskian)
3.2. Fossils and biostratigraphy
3.1. Main facies, depositional systems and lithostratigraphy
This unit is characterized by the Glossopteris–Brasilodendron (G-
The Rio Bonito is the coal-bearing unit of the Paraná Basin. B) flora that corresponds to the typical ‘‘Glossopteris Flora” associ-
Some coal seams, very thin (<0.5 m) and laterally discontinuous, ated with the Gondwana coal measures (Iannuzzi and Souza,
occur in a fluvial and deltaic system, as interpreted by facies asso- 2005). It is equivalent to Taphofloras ‘‘B” and ‘‘C” of Rösler (1978).
ciation comprised by orthoconglomerates and coarse to fine In this flora, the compressed–impressed lycopsid stems, pecopterid
subarkose sandstones, with planar and trough cross bedding and fronds and codaitalean and glossopterid leaves are locally abun-
fining-upwards cycles, topped by mudstones and coal. The interflu- dant; leaves of the genus Glossopteris become dominant and replace
ves are characterized by muddy facies with argillaceous coal the previous Gangamopteris-type glossopterids. The age of this flora
seams. This fluvial-deltaic interval occurs characteristically at the is primarily given by its association with the V. costabilis Zone of the
base of the Rio Bonito Formation. late Sakmarian-Artinskian interval. In Santa Catarina, there was
Thick (up to 2.5 m) and laterally continuous (up to 40 km) coal identified a marine megafauna dominated by bivalves, named as
seams occur in lateral association with quartz-arenites and mas- ‘‘Taió Assemblage”, recorded in the middle Rio Bonito (Rocha-Cam-
sive or laminated mudstones with lenticular and wavy bedding. pos and Rösler, 1978). An Early Permian age (probable Artinskian) is
The most conspicuous sedimentary structures of the sandstones suggested for this fauna based on the affinities of the bivalves to
are tabular cross stratification and planar-parallel bedding, many Western Australian taxa (Rocha-Campos and Rösler, 1978).
levels with flaser, mud drapes and wavy bedding and very abun-
dant hummocky cross stratification, besides icnofacies Skolithos 3.3. Sequence stratigraphy
(Holz, 1999).
The features of these facies associations point to a wave and tide Detailed (third-order) sequence stratigraphy analysis of the
influenced estuarine system evolving to a barrier-lagoon deposi- interval, which corresponds to the Rio Bonito Formation (e.g., Holz,
tional system (e.g., Holz, 2003), where the mires were formed be- 1999; Holz et al., 2006), clearly shows two distinct regional uncon-
hind barrier island (Fig. 3). formities caused by important base level changes. These unconfo-
The succession of fluvial-deltaic and marine deposits is subdi- rmities are third order sequence boundaries and roughly
vided into three major packages with the hierarchy of members, la- correspond to the lithostratigraphic limits.
beled from base to top: Triunfo (coastal and fluvial sandstones), The first of these unconformities is located at the base of the Rio
Paraguaçu (marine mudstones and fine-grained sandstones) and Bonito Formation, marked as sequence boundary SB-3 in the chart.

Fig. 3. In southern Brazil (Rio Grande do Sul/Santa Catarina), the most exploited and economically important coal seams are not linked to delta systems, but to
retrogradational lagoon-barrier systems. A – Normal regression of shoreface and back-barrier peat-forming environments; B – after sea level raise and erosional shoreface
retreat, coal seams are topped by shoreface sandstones.
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 385

Fig. 4. Core expression of sequence boundary SB-3 in a representative borehole in the southern study are, where the unconformity is characterized by fluvial deposits
overlying marine sediments. Detail in A illustrates the typical rhythmic sand/mudstone couplets interpreted as deposited by distal turbidity currents derived from
subaqueous glacial outwash fans, detail in B shows marine mudstones from the topmost zone of the glacial succession, detail in C shows medium to coarse-grained
sandstones with trough cross bedding, and detail in D illustrates estuarine sandstones and mudstones with bioturbation. Modified from Holz (1997).

This is a third order sequence boundary recognized and correlated uaçu Member, which is not identified in the southernmost part
regionally, separating the glacio-marine mudstone and sandstone of the study area (the Rio Grande do Sul). The fluvial and coastal
of the Itararé from the fluvial and deltaic sandstones of the Rio Bo- sandstones of the topmost Rio Bonito Formation (labeled as the
nito Formation (Fig. 4), as recorded at several locations of the east- Siderópolis Member) are laterally equivalent to the marine depos-
ern border of the Paraná Basin (e.g., França, 1993; Alves and Ade, its of the basal portion of Palermo Formation (Fig. 5). Detailed cor-
1996; Holz, 1997; Milani et al., 1998). relation has shown that sequence boundary SB-4 in proximal
Eventually the fluvial succession scours even the transgressive setting has the typical ‘‘fluvial over marine” signature while in
systems tract of the prior sequence, indicating a considerable the distal zone a pebbly, transgressively reworked lag of shoreface
base-level fall which is recognized at the entire southern margin sandstone can be found (Fig. 6).
of the Paraná Basin, as described by several authors (Alves and Hence, the Rio Bonito succession embraces two third-order se-
Ade, 1996; Holz, 1997). quences – LPTS-3 and 4. The Siderópolis Member and its coal
The second sequence boundary of this interval is stratigraphi- seams, as occurring in Santa Catarina and Paraná, are linked to
cally located close to the top of the Rio Bonito Formation, and is LPTS-4, while the coal seams in Rio Grande do Sul belong to the
marked by fluvial to paralic facies (delta front, foreshore and shore- underlying sequence LPTS-3.
face sandstones – lithostratigraphically equivalent to the Sideróp-
olis Member) overlying a marine offshore succession
(lithostratigraphically equivalent to the Paraguaçu Member). This 4. Palermo Formation (Early Permian: Middle Artinskian)
sequence boundary, described by Holz et al. (2006), is labeled
SB-4 in the stratigraphic chart. 4.1. Main facies, depositional systems and lithostratigraphy
This limit is recognizable towards the northern region of Paraná
(Tognoli et al., 2003) and probably is present in São Paulo also The main facies are light to dark gray mudstones and fine-
(within the Tatuí Formation), as discussed ahead. grained sandstones with hummocky and wavy bedding, showing
The Rio Bonito Formation is a coal-bearing unit at the entire a characteristic yellow color in outcrops due to weathering. The
eastern margin of the basin, but the sequence stratigraphic ap- muddy facies is frequently bioturbated, with dominance of the ich-
proach, using a maximum flooding surface as a datum for correla- nogenus Teichichnus. The depositional system is interpreted as
tion, shows that the Rio Bonito has a marine intercalation in the shallow marine (shoreface, marine bayments). The average thick-
central part of the study area (Santa Catarina/Paraná), the Parag- ness is about 100 m, reaching up to 140 m. The Palermo Formation
386 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

Fig. 5. Correlation between the Early Permian succession of Rio Grande do Sul and Santa Catarina, based upon sequence stratigraphic analyses. Datum is the maximum
flooding surface of LPTS-3. Note that the marine interval called the Paraguaçu Member in Santa Catarina is equivalent to the coal-bearing shoreface-barrier system in Rio
Grande do Sul. After formation of sequence boundary SB-4 the main source area is Santa Catarina, with sandstones and coals assigned to the Siderópolis Member. This
member is time-equivalent to offshore facies in Rio Grande do Sul, labeled the Palermo Formation (modified from Holz et al., 2006). SB = sequence boundary, LPTS = Late
Paleozoic third order sequence.

is the topmost lithostratigraphic unit of the Guatá Group and cor- 5. Tatuí Formation (Early Permian: Late Sakmarian to Middle
relates northwards with the Tatuí Formation. Artinskian)

4.2. Fossils and biostratigraphy 5.1. Main facies, depositional systems and lithostratigraphy

The Palermo Formation is marked by the basal part of the The Tatuí Formation is mainly formed of a succession of mud-
Lueckisporites virkkiae interval Zone (LvZ) (Souza and Marques- stones and fine-grained sandstones with hummocky cross bedding,
Toigo, 2005) and by the transition from the Glossopteris–Brasiloden- pointing to a shallow marine origin, interbedded with minor
dron to the Polysolenoxylon–Glossopteris floras, reflecting an overall coarse-grained sandstones and conglomerates. Assine et al.
change in flora composition probably due to the strong transgres- (2003) interpreted the facies association of sandstones with sig-
sive component which controlled sedimentation of the Palermo moidal stratification and hummocky cross bedding as the result
Formation (Iannuzzi and Souza, 2005), hence inducing retrograda- of the progradation of a coastal alluvial fan system onto a marine
tion of the terrestrial facies belt accompanied by evolutionary shelf. At northern São Paulo State, the progradation of a coarse-
modification of the associated floras. grained clastic wedge (the ‘‘Ibicatu Conglomerate”) indicates tec-
Other fossils are scarce in the Palermo Formation and include tonic uplifts and reactivation of source areas during the deposi-
Dadoxylon sp. (in the form of fragments of silicified wood trunks), tional time of the Tatuí Formation (e.g., Stevaux et al., 1986).
teeth from the amphibian Loxomma, beside bivalve shells (Schnei- Paleocurrents indicate depositional dip towards the south,
der et al., 1974). paleoshoreline with an approximate east–west direction and sedi-
ment provenance from the North (Assine et al., 2003).
The Tatuí Formation correlates southwards with the Rio Bonito
4.3. Sequence stratigraphy
and Palermo formations, as already recognized by several authors
(e.g., Gordon Jr., 1946; Barbosa and Almeida, 1949; Soares, 1972).
As depicted in the chart, LPTS-4, limited at the bottom by se-
quence boundary SB-4, has a complex framework, because the
lithostratigraphic unit known as Palermo Formation is laterally 5.2. Fossils and biostratigraphy
equivalent to the Siderópolis Member of the Rio Bonito Formation
(in Santa Catarina/Paraná), and to the upper Tatuí Formation (in Besides silicified wood logs, palynological and paleobotanical
São Paulo). data are scarce in this unit. Fragments from fresh and salt-water
The Palermo succession contains the second-order maximum fishes and possible tetrapod remains, concentrated in wave-re-
flooding surface of the sequence stratigraphic framework of Milani worked transgressive lags, are described by Assine et al. (2003)
et al. (1994). and Chahud (2007).
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 387

Fig. 6. Schematic correlation section showing that the fluvial and coastal sandstones of the topmost Rio Bonito (=Siderópolis Member) are laterally equivalent to the marine
Palermo Formation. A – Base level fall and progradation of fluvial to coastal sands (=topmost Rio Bonito Formation) are coeval to marine deposits, B – resulting sequence
stratigraphy: the sequence boundary in proximal setting has the typical ‘‘fluvial over marine” signature, whereas in the distal zone a pebbly, transgressively reworked lag can
be found. C – Outcrop expression of the ‘‘fluvial over marine” signature of sequence boundary SB-4, D – distal expression of sequence boundary SB-4: offshore mudstones
with lenses of transgressively reworked sandstone, E – detail of the coarse massive sandstone lenses. Modified from Holz et al. (2006).

5.3. Sequence stratigraphy unconformity SB-3 is also recognized in the northern part of the basin,
where the contact between the Itararé Subgroup and the overlying
The base of the Tatuí unit is delimited by the same sequence bound- Tatuí Formation is marked by a prominent erosional surface inter-
ary as its lithostratigraphic equivalent, the Rio Bonito unit, because the preted as an unconformity (e.g., Soares, 1972; Assine et al., 2003).
388 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

At the top, the transition between the Tatuí and Irati formations sequence stratigraphic viewpoint, the Tatuí unit would be formed
is interpreted as a base level fall followed by a transgressive event by two depositional sequences (Fig. 7).
(Assine et al., 2003), separating a highstand systems tract of the Data on high resolution correlation are not available, but
Tatuí succession from a transgressive systems tract represented a comparison between Tatuí’s southern counterpart – the
by the Taquaral Member of the Irati Formation. The contact is Rio Bonito/Palermo – suggests that the discontinuity mapped
marked by a discontinuous but widespread centimeter-thick level within the Tatuí Formation by Soares (1972) may be correl-
of a pebbly lag, characterized by extraclasts, silex grains and bio- ative to the unconformity SB-4 mapped within the Rio Bonito
clasts (mainly fish teeth, bones and scales). This lag characterizes Formation in the southern part of the study area (e.g., Holz
a sequence boundary which is marked in the stratigraphic chart et al., 2006), as tentatively shown in the stratigraphic chart.
(SB-5), and is overlying the entire Tatuí in São Paulo, including Hence, the Tatuí Formation embraces two Late Paleozoic third-
the coarse-grained prograding wedge historically known as the order sequences (LPTS 03 and 04, Fig. 7) with boundaries
‘‘Ibicatu Conglomerate” of the Tatuí Formation. correlatable southwards, as depicted in the stratigraphic
Within the Tatuí succession, Soares (1972) recognized a discon- chart.
tinuity labeled L2, characterized by fluvial sandstones with petri- This would indicate that the tectonically controlled base level
fied wood logs. fall and the resulting paleorelief inversion and reactivation of
This surface has also been described and recognized as a source areas, as described by Holz et al. (2006), was a basin-
sequence boundary by Assine et al. (2003). Therefore, under a scale event which affected the entire eastern margin of the

Fig. 7. Correlation of representative composite profiles of the Tatuí Formation mapped in São Paulo. In terms of sequence stratigraphy, the Tatuí Formation is formed by two
depositional sequences, delimited by regional unconformities. Data are from Assine et al. (2003), incorporating information from Soares (1972).
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 389

Paraná Basin during the final Sakmarian. However, the precise The Taquaral Member was deposited in a shallow marine envi-
correlation of SB-4 through the entire basin’s margin is still a ronment (epicontinental sea) with very restricted connection to
matter of investigation. the open ocean, and with relatively good water circulation in com-
parison to the overlying Assistência Member.
The depositional system of the Assistência Member was com-
6. Irati Formation (Early Permian: Late Artinskian) posed of internal, intermediary and distal ramps tilted to south-
west, admitting a possible connection to the Pantalassa Ocean
6.1. Main facies, depositional systems and lithostratigraphy only at the southernmost region of South America. This very
large and shallow sea, according to the latitude and the influx
The Irati Formation, a well-known unit for its oil-prone rocks of continental water, could have presented some marginal re-
and famous mesosaurid fossils, is the basal part of the Permian gions dominated by fresh water (e.g., favorable for the preserva-
Passa Dois Group and extends almost throughout the entire Paraná tion of Botryococcus), whereas other marginal regions presented
Basin. It has an average thickness of 40 m, with peaks of about hypersaline conditions (Araújo, 2001).
70 m in the southern part of the basin. In the northeastern part of the basin, the basal portion of the
The contact with the underlying Palermo Formation is marked Assistência Member records the most proximal parts of the ramp,
by the appearance of dark mudstone. characterized by evaporitic conditions or exposition of carbonate
Formally, the Irati Formation is divided into the lower Taquaral muds (gipsite and almost autochtonous carbonate breccias) (Araú-
Member, comprising siltstones and gray to black mudstones, and jo, 2001). Cyclic alternation of dry and humid climates coupled
the upper Assistência Member, formed by organic-rich mud- with minor sea level changes are evidenced by rhythmites of dolo-
stones and shales interbedded with limestones (packstones, mite and betuminous shales. Relatively deeper depositional condi-
wackestones). tions are interpreted for the southern parts of the basin, where

Fig. 8. Typical core log showing facies associations of the Palermo/Irati transiton in the southern study area. (A) Shows the lithostratigraphic contact between Palermo and
Irati formations (a flooding surface under sequence stratigraphic viewpoint), (B) displays the breccias marking sequence boundary SB-5, and (C) illustrates the wave-
influenced facies of basal Irati Formation.
390 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

organic-rich shales predominate, but even there bone rich carbo- 1991; Soares, 2003). Crustaceans of the genus Clarkecaris are rela-
natic tempestites occur. tively common in the Taquaral Member and other pyg-
ocephalomorph genera occur in the Assistência Member. Insects
6.2. Fossils and biostratigraphy as a dragonfly from São Paulo and representants of Mecoptera,
Neuroptera and Homoptera of Rio Grande do Sul are significant
The unit is characterized by the L. virkkiae Interval Zone and (Rösler et al., 1981; Pinto and Ornellas, 1981).
the Polysolenoxylon–Glossopteris flora which, differently from Scattered fish scales are relatively common in the non-bitumi-
underlying floras, is mainly composed of fossil logs, including nous siltstone in the middle part of the Assistência Member, but
elements with probable glossopterid, cordaitalean and conifer abundant fragmented and abraded scales, teeth and bony parts
alliances (Iannuzzi and Souza, 2005). A Glossopteris leaf and a from several types of fishes, as well as from possible tetrapods
conifer branch were also found (Ricardi-Branco et al., 1999). are found at the base of the Taquaral Member (e.g., Castro et al.,
The plants are partially equivalent to the Taphoflora ‘‘D” as de- 1993; Assine et al., 2003; Lages, 2004; Chahud, 2007).
scribed by Rösler (1978), and include fossil logs of the genus Cyanophyta and Chlorophyta microfossils were discovered in
Barakaroxylon (Alves, 2006). Rare acritarchs and Botryococccus al- early diagenetic silex formed in the basal contorted carbonatic
gae were found in the Taquaral and Assistência members (Araú- interval of the Assistência Member in São Paulo (Calça and Fair-
jo, 2001; Lages, 2004). child, 2005; Calça, 2008). In the same interval, microbialites were
Besides palynomorphs and plants, the fossil content of the Irati recognized by Araújo (2001). Finally, other important fossils are
is characterized by the mesosaurid reptiles preserved as articu- the large stromatolites in the most northeastern outcrop region
lated skeletons and disarticulated and scattered remains, which of the Passa Dois Group (Suguio and Sousa, 1985; Ricardi-Branco
are typical in carbonatic, shaly and bone bed facies in the State et al., 2006), as well as stromatolites in the northern part of the ba-
of São Paulo, in carbonatic tempestite facies in Rio Grande do Sul sin (Fairchild et al., 1985). Stromatolites associated to mesosaur
and also in the black shales of Santa Catarina and Paraná (Lavina, fossils were also found in equivalent levels of the Huab Basin in Na-

Fig. 9. Late Paleozoic third-order sequence 5 (equivalent to the Irati Formation) as occurring in the northern part of the study area (Paraná and São Paulo). Note the thickness
compensation of the Taquaral and Assistência members in the Ponta Grossa Arch region, which evidences distinct tectonic tendencies during the deposition of each member
and their probable separation by an unconformity. Datum for correlation is the base of the Irati Formation (modified from Rohn, 2007).
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 391

mibia, which was formerly connected to the Paraná Basin (Holzför-


ster et al., 2000).

6.3. Sequence stratigraphy

The base of the Irati records a base level fall followed by a trans-
gressive event separating a highstand systems tract of the Tatuí
succession from a transgressive systems tract represented by the
Taquaral Member of the Irati Formation in São Paulo (Assine
et al., 2003; Araújo, 2001; Rohn et al., 2003a; Lages, 2004). The
base of the Taquaral Member of the Irati Formation corresponds
to an unconformity which separates the highstand systems tract
of the Tatuí succession from a transgressive systems tract (Assine
et al., 2003; Araújo, 2001; Rohn et al., 2003a; Lages, 2004).
This level, marked by a pavement composed of fish and possible
tetrapod remains, is conspicuous in the eastern margin of the ba-
sin. From northern Paraná towards northeast, this pavement
grades to a decimeter-scale pebbly sandstone with abundant
vertebrate fragments (Mezzalira, 1957; Soares, 1972; Castro
et al., 1993; Assine et al., 2003; Chahud, 2007). In Rio Grande do
Sul, the boundary is marked by a few centimeter thick breccia beds
at the base of the typical siltstones and gray to black mudstones of
the Taquaral Member (Fig. 8). The beds are an indication of an
overall sea level drop and reworking of the sedimentary material
of the exposed marginal zone of the basin (Holz, 1997).
In summary, the above described features record a regional base
level fall, labeled SB-5 in the stratigraphic chart.
A minor sequence boundary is suggested at the Taquaral–
Assistência members’ interface, according to their relative thick-
nesses or respectively distinct subsidence rates/tectonic tenden-
cies in the Ponta Grossa Arch and in the neighboring areas
(Fig. 9) (Lages, 2004; Rohn, 2007). The abrupt replacement of silt-
stones and mudstones (Taquaral Member) by limestones and or-
ganic-rich mudstones (Assistência Member) is another evidence
of total change of the depositional systems, respectively from a
siliciclastic intracratonic to a mixed carbonatic-anoxic siliciclastic
ramp.

7. Serra Alta Formation (Early Permian: Kungurian)

7.1. Main facies, depositional systems and lithostratigraphy

The 60–90 m thick Serra Alta Formation is formed of laminated


dark gray mudstones and black shales, besides carbonatic concre-
tions, decimeter-thick layers of fine grained sandstones and occa-
sional layers of micritic carbonates and thin fish bone-beds
(Meglhioratti et al., 2005). The thickness of this unit decreases very
quickly in the northeastern region of the basin and the dark shales
give place to red mudstones. The boundary with the overlying
Teresina Formation is transitional in most of the basin margin
and is characterized by gradual increase of fine-grained sandstone
interlaminations with lenticular to wavy bedding.
The Serra Alta Formation is interpreted to be formed in a marine Fig. 10. Representative well log for the Santa Catarina/Rio Grande do Sul area. Note
environment below storm wave base, under mainly dysoxic condi- the retrogradational to progradational pattern of the Serra Alta/Teresina transition,
tions (Schneider et al., 1974; Araújo, 2001). However, no fossils and the arrow pointing high gamma ray values which mark the second-order
maximum flooding surface of the Permian succession, within the Palermo
prove a true marine origin. In comparison to the Assistência Mem-
Formation.
ber, the depositional environment changed to less stagnant or less
stratified water bodies, the climate became more humid, the pri-
mary productivity declined, organic matter was more destroyed
by oxidation and the accommodation space increment turned 7.2. Fossils and biostratigraphy
more stable at a relatively high rate (Araújo, 2001). There are no
proximal lithofacies represented in the Serra Alta Formation and Although palynomorphs are very scarce in this unit, guide spe-
it may be inferred that the relatively deep sea surpassed widely cies of palynomorphs assign the Serra Alta Formation to the L. Virk-
the present eastern margin of the basin. kiae Interval Zone (Souza and Marques-Toigo, 2003). Bivalves occur
392 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

in siltstones some meters above the base of the formation and cor- tic sedimentation. In the past, it was interpreted as a shallow
respond to the Barbosaia angulata Zone (Rohn and Rösler, 2000). In marine system and, in part, associated to wide coastal areas with
micrites near to the transition to the Teresina Formation, bivalves tidal flats (e.g., Schneider et al., 1974; Sousa et al., 1991). However,
of the Pinzonella illusa Zone were identified (Meglhioratti, 2006). the adequately identified fossils do not evidence a connection be-
Fish scales and teeth, as well as coprolites, are found in bone-beds tween the epicontinental sea of the Paraná Basin and the Panta-
or loosely dispersed in the mudstone of the Serra Alta Formation. lassa Ocean (e.g., Mendes, 1984; Rohn, 1994, 2007; Simões et al.,
1998). The general low diversity of the fossils is evidence against
7.3. Sequence stratigraphy a normal marine environment. Ghilardi and Simões (2002) specu-
late that the bivalves lived in shallow water habitats, frequently af-
The change from Irati to Serra Alta units is marked by a base le- fected by storms, and punctuated by episodes of intense
vel fall which is marked not by exposure, fluvial deposition or evaporation and hypersalinity with influence of local freshwater
extensive and clear facies shifts, but by a thin fish bone-bed or inputs, denoting a situation of high environmental stress. Sum-
breccias, by the disappearance of the bituminous shales, and by ming up all sedimentological and paleontological evidence, the
tiny layers of breccia, intraformational conglomerates or sand- paleosalinity of the Teresina Formation depositional environment
stones. This level is in general separating a more progradational seemed to have been varying from one extreme (freshwater) to
pattern of sedimentation below from a more retrogradational pat- the other (hypersalinity). The lycophytes – the most abundant
tern above, and is interpreted as a transgressive lag (Etgeton, 1997; megaplants found in the Teresina Formation – probably lived par-
Holz, 1997; Menezes, 2000; Lages, 2004; Rohn, 2007). Therefore, in tially in the water and tolerated saline conditions.
the stratigraphic chart we indicate a sequence boundary separating The carbonatic rocks may correspond to the phases of high
the Irati and Serra Alta units (SB-6). water salinity, which may have taken place during relatively dry
The transition from Serra Alta to Teresina is gradational in most climatic intervals (Rohn, 1994; Meglhioratti, 2006). The wide-
of the basin margin, therefore these two lithostratigraphic units spread distribution of oosparites during apparently certain time
form the LPTS-6 sequence. That sequence begins with the Serra Alta intervals (according to correlations of cores and geophysical well
Formation, marked by a clear retrogradational pattern; with the logs) indicate that the ooid shoals were not formed only near to
gama-ray values in bore-holes reaching a maximum near the top the coast, but also in more central areas of the basin (Rohn,
of the formation in Rio Grande do Sul and Santa Catarina (Fig. 10). 1994, 2001; Rohn et al., 2003b; Meglhioratti, 2006). Mud cracks
Northwards, the gama-ray logs suggest that the maximum flooding show the same large geographic distribution. These facts indicate
surface is placed to a gradually lower position in the formation that shallow water conditions predominated in the basin, with
(Meglhioratti, 2006), and a minor sequence boundary seems to oc- occasional exposure of large areas.
cur at the basal portion of the Teresina Formation in the northern
part of the study area, as described in the next section. 8.2. Fossils and biostratigraphy

8. Teresina Formation (Early to Mid Permian: Late Kungurian- Bivalves are common in carbonatic rocks (micrites and oospa-
Rodian) rites) and sometimes in sandstones (mainly in the northeastern
part of the basin), keeping in mind that bivalve coquinas also occur.
8.1. Main facies, depositional systems and lithostratigraphy Mendes (1949, 1952) proposed the first formal biostratigraphic
zones for bivalves of São Paulo – the Pinzonella illusa–Plesiocyprinel-
The Teresina Formation is about 280–330 m thick in the central la carinata Zone and the Pinzonella neotropica–Jacquesia brasiliensis
to the eastern portion of the Paraná Basin. The main facies of this Zone. Many other authors considered these biozones valid, also
unit are gray mudstones usually interlaminated with very fine extending their geographic range further south (Rohn, 1994). The
sandstones, showing linsen and flaser bedding, hummocky cross bivalves are endemic to the Paraná Basin (Simões et al., 1998),
stratification and wavy bedding, sometimes with small mud but they are not typical freshwater mollusks (Ghilardi and Simões,
cracks. An abnormally thick sandy interval is observed in the lower 2002).
part of the formation in the northern study area. Centimeter to Ostracods are relatively frequent in the micrites, but they were
decimeter thick carbonatic rocks (mainly micrites sometimes with not thoroughly studied. Sohn and Rocha-Campos (1990) and Mar-
mudcracks and brecciated intervals, oosparites and stromatolitic anhão (1995) ascribed them to freshwater environments. Never-
biolitites), as well as bivalve coquinas and fish bone-beds, fre- theless, Almeida and Do Carmo (2005), according to a short
quently occur interbedded with siliciclastic successions. Planolites published abstract, considered some ostracods from the northern
ichnofacies occurs profusely in the muddy rocks. part of the basin as marine and cosmopolite. Some other ‘‘marine”
In the northern/northeastern part of the basin, the stratigraphic microfossils were announced, as rare foraminifers, sponge spicules
interval equivalent to the Serra Alta and Teresina formations is and even conodonts, but they were never formally presented in
usually designated as Corumbataí Formation (e.g., Schneider publications.
et al., 1974), taking into consideration that the mudstones become Fish scales, teeth, dental plates and other fish remains occur in
red, the total thickness of the units decreases, the proportion of bone beds or scattered in various lithologies. Rare amphibian teeth
sandstones, coquinas and bone-beds apparently increases, and car- were also found in bone beds. Some of these vertebrate remains
bonatic rocks reduce gradually. In subsurface, the mudstones con- indicate freshwater living environments (e.g., petalodonts,
tinue gray for a considerable interval above the Irati Formation and amphibians), but they could have been transported to more saline
there are no clear reasons for the usage of distinct stratigraphic depositional environments (Toledo, 2001).
names for the Serra Alta and Teresina formations in the north- Few palynological studies were done for the Teresina Forma-
ern/northeastern part of the basin. It is opportune to note that tion. The assemblages may be ascribed to the L. virkkiae Interval
the same stratigraphic interval, including the lower part of the Zone. Fern spores predominate in core samples from the central–
overlaying Rio do Rasto Formation, is designated as ‘‘Estrada Nova north region of the State of Paraná (Neregato, 2007; Neregato
Formation” by some authors (e.g., Maranhão and Petri, 1997). et al., in press), although lycophyte macrofossils are more abun-
The Teresina depositional system is interpreted as the result of dant. Other plants are glossopterids, very rare conifers branches
sedimentation in storm-influenced offshore to coastal plain set- (Fanton et al., 2006) and gymnospermous petrified stems. The
tings, with low carbonate contribution in comparison to siliciclas- ferns (Marattiales) are known as leaf (impressions or rare petrifica-
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 393

tions) or as petrified stems. Carophytes, normally related to conti- One is formed of meter-scale thick sandy beds in the lower part
nental lakes, are abundant in some localities. The plants of the of the Teresina Formation (Fig. 10), appearing after a short inter-
Teresina Formation correspond to the Lycopodiopsis derbyi Zone, val with the first coarsening upward successions, intercalated
which is defined as the interval of total range of L. derbyi and re- with micritic rocks (Rohn, 2001, 2007; Meglhioratti, 2006). The
lated lycopod stems (Rohn and Rösler, 2000). other minor sequence boundary is formed of carbonate breccias
Other remarkable fossils of the formation are stromatolites and is apparently coincident with the top of the P. illusa Zone/
found in several stratigraphic levels (Rohn, 1994; Rohn et al., base of the P. neotropica Zone. This faunal change may have been
2003b; Meglhioratti, 2006). caused by a large exposure of the basin in the northern part of
the study area.
8.3. Sequence stratigraphy
9. Rio do Rasto Formation (Mid to Late Permian: Wordian to
In terms of sequence stratigraphy, in the southern part of the Wuchiapingian)
study area, the Teresina succession can be considered as a high-
stand systems tract of LPTS-6, i.e., a progradational sedimentary 9.1. Main facies, depositional systems and lithostratigraphy
unit where the facies succession reflects overall shoaling upwards
conditions, which is clearly reflected by the stacking pattern of the The Rio do Rasto Formation is a succession with an overall
strata as shown in the gama-ray curve in Fig. 10. coarsening upwards trend, subdivided into the 150–250 m thick
In the northern part (São Paulo, Paraná), however, LPTS-6 Serrinha Member formed dominantly of mudstones and fine-
apparently embraces two minor sequence boundaries (Fig. 11). grained sandstones at the base, and 250–300 m thick Morro

Fig. 11. Detailed correlation in the northern part of the study area has revealed minor sequence boundaries within the Serra Alta–Teresina succession. These boundaries are
only mapable in São Paulo and northern Paraná (modified from Rohn, 2007).
394 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

Pelado Member formed mainly of fine- to medium-grained sand- tral–north Paraná (Neregato, 2007; Neregato et al., in press). The
stones. The basal part may be distinguished from the Teresina fern spore rich associations were correlated to the L. virkkiae Zone.
Formation mainly by a substantial increasing of sandstone beds, The bivalves are completely different from those of the Teresina
the reduction of the wavy and lenticular bedding and the color Formation. In the carbonate rocks and coquinas at the north of the
change of the mudstones from gray to green, violet or red (Rohn, Ponta Grossa Arch, they are recrystalized or silicified; in other re-
1994). gions and in other stratigraphic intervals, they are preserved as
Borehole data from the northeastern region of the Ponta Grossa molds. Most part of the Serrinha Member corresponds to the Lein-
Arch (center-north of the State of Paraná), show that the typical zia similis Zone (still an informal zone, defined by Rohn, 1994); the
siliciclastic and carbonatic rocks of the Teresina Formation are transitional Serrinha-Morro Pelado deposits and great part of the
abruptly superposed by 1–1.5 m thick sandstones with interca- Morro Pelado Member encompasses the Palaeomutela? platinensis
lated 0.5–1.5 m thick massive mudstones (Rohn et al., 2003b; Meg- Zone (also still an informal zone, defined by Rohn, 1994). The spe-
lhioratti, 2006). The sandstone beds are massive or show incipient cies L. similis was also found in the Gai-As Formation in Namibia;
ripple drift laminations; they may present granules at the base and volcanic ashes located few meters above the last occurrence of this
a slight tendency of fining upwards. The gamma ray values of the species in Africa indicate an absolute age of 265.5 ± 2.2 Ma, corre-
borehole decrease abruptly at the base of the Rio do Rasto sponding approximately to the Wordian–Capitanian boundary
Formation. (Stollhofen et al., 2000a).
The depositional systems of the Serrinha Member are inter- Conchostraceans are the most abundant fossils of the Rio do
preted mainly as shallow lakes, sometimes influenced by storm Rasto Formation and suggest fresh-water conditions for the depo-
waves or by fluvial incursions which produced hyperpicnal or sitional environment. Rohn and Rösler (1990) presented a bio-
rarely homopicnal flows (Rohn, 1994). The depositional system of stratigraphic scheme for the conchostraceans, which was
the Morro Pelado Member had many distinct interpretations, vary- improved in Rohn (1994) and in Ferreira-Oliveira (2007). Some
ing from meandering fluvial to mixed lacustrine, deltaic and aeo- identified genera are coherent with a Mid to Late Permian age of
lian environment (e.g., Castro and Medeiros, 1980; Lavina, 1991; the Rio do Rasto Formation. Hemicycloleaia mitchelli, only found
Rohn, 1994). Recently, a new interpretation has risen: the Morro at three localities in the upper part of the formation, is particularly
Pelado Member may have been deposited in very distal alluvial important for correlations and suggests a Wuchiapingian age (early
fan conditions with laterally extensive coalescent fluvial plains, Late Permian or early Lopingian; Ferreira-Oliveira and Rohn, 2009).
including the deposition of crevasse splay deposits, inundites and Fish scales are relatively common in the Rio do Rasto Forma-
occasional shallow fluvial channels (Rohn et al., 2005). tion, but tetrapods are especially remarkable because they can
In outcrops of the central–east region of the Paraná Basin, the be used, in part, for correlations with South African paleofaunas.
Serrinha Member is apparently transitional to the Morro Pelado In Rio Grande do Sul (Aceguá, Posto Queimado), the assemblages
Member in view of the gradual increase of lenticular, sigmoidal may be correlated to the South African Tapinocephalus Zone
or lobed sandy beds with ripple drift laminations or with large (Malabarba et al., 2003) of probable Wordian age (Langer
tangential cross stratifications and intercalated red mudstones. et al., in press). In the Ponta Grossa Arch region, a reptile and
Southwards, in Santa Catarina, the Serrinha Member becomes amphibians were found, but their real affinities are under discus-
thinner and is easier distinguished from the typical Morro Pelado sion. Their age may be a little younger, perhaps Capitanian (Lan-
Member. ger et al., in press).
In the northeastern region of the Ponta Grossa Arch, the Rio do Many deposits of the Rio do Rasto Formation show bioturbation
Rasto Formation is unconformable overlain by the Triassic aeolian (mainly the Cruziana ichnofacies).
and fluvial redbeds of the Pirambóia Formation (e.g., Soares et al.,
in press, 2008). In central and southern Santa Catarina, the Pira- 9.3. Sequence stratigraphy
mbóia Formation is missing and the Rio do Rastro Formation has
an unconformable contact with the Early Cretaceous Botucatu For- The unclear transition between the two formations as visual-
mation, not shown in the chart. In Rio Grande do Sul, the unconfor- ized in outcrops would be used as an argument that they are not
mity is marked by the contact with the Triassic Sanga do Cabral separated by an unconformity. However, according to available
Formation, a fossil-bearing unit described by Andreis et al. (1980). borehole core data of the Serrinha Member, the lithofacies change
is too abrupt to consider it as a simple lateral shift of the deposi-
tional environments (Walther’s Law). The lithostratigraphic
9.2. Fossils and biostratigraphy boundary probably corresponds to a superposition of two distinct
depositional systems and, therefore, may be interpreted as a se-
The L. derbyi Zone does not pass the Teresina-Rio do Rasto tran- quence boundary. The coarse lag and the angular unconformity
sition. In the State of Paraná and north of Santa Catarina, the Serr- at the base of the formation in Rio Grande do Sul corroborate this
inha Member shows a notable increase of megaplant occurrences interpretation (e.g., Lavina, 1991; Menezes, 2000). In borehole logs,
and the assemblages are dominated by several glossopterid spe- this stratigraphic level is characterized by a clear shift of the gam-
cies, some pecopterids, rare fertile fern leaves and abundant sphe- ma-ray pattern accompanied by a sharp color shift from medium
nophyte stems, all preserved as impressions or molds. grey to light red facies at the Teresina-Rio do Rasto transition. This
Sphenophyllum specimens are not very abundant, but their strati- third-order sequence boundary is labeled SB-7 in the stratigraphic
graphic range was used for the definition of the Sphenophyllum par- chart.
anaense Zone (Rohn and Rösler, 2000). In the transitional portion The carbonatic interval of the Serrinha Member northeast from
from the Serrinha to the Morro Pelado members, Sphenophyllum the Ponta Grossa Arch may be possibly correlated to the lithologies
gives place to another important Gondwanan sphenophyte – Schiz- with carbonatic cement at the south and probably represent a time
oneura gondwanensis – which range defines the second megaplant span of relatively dry climate, without aeolian dune deposition, re-
zone (Rohn and Rösler, 2000). The fossil occurrences become grad- lated to a lowstand system tract of the lake.
ually rarer towards the top of the formation and the diversity The Morro Pelado Member shows a coarsening upwards lithof-
decreases. acies succession from red mustones to sandstones which may be
Palynomorphs were recently recorded in the lower part of the interpreted as a gradual shallowing of a waterbody (as a flooded
Serrinha Member in cores from the mentioned borehole in cen- area or a lake), caused by water evaporation and/or sedimentary
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 395

infilling. The sandstones in the upper part are sometimes aeolian Absolute ages were also obtained from several tonsteins within
deposits, obviously indicating that the waterbody dried up. These the coal seams of the Rio Bonito Formation. However, some of
successions represent regressive intervals of relatively high order them are from a single locality, which present controversial results.
sequences (i.e., higher than 3rd order). The aeolian dune deposits As example, analysis from tonsteins of the Candiota Coalfield re-
may have been deposited during time spans of low accommoda- sulted ages of 267.1 ± 3.4 Ma (Matos et al., 2001), 290.6 ± 1.5 Ma.,
tion space, but the fact that they were preserved indicates that, 296.9 ± 1.65 Ma and 296 ± 4.2 Ma (Guerra-Sommer et al., 2008a,
after accumulation, the relative phreatic level has risen anew. 2008b), while those from Faxinal Coalfield indicated an age of
These successions were probably mainly controlled by climate. 285.4 ± 8.6 Ma (Guerra-Sommer et al., 2008c). According to these
In the upper part of the Morro Pelado Member in the Ponta last authors, the glaciogenic strata related to the Itararé Group,
Grossa Arch region, large intraclasts and some apparently erosive must be restricted to the Late Pennsylvanian. This interpretation
surfaces, including angular relationships between strata, probably is not accepted herein, based on the following data.
indicate other important sequence boundary, marked as a minor No radiometric data are available from the A. cristatus Zone and
sequence boundary in the chart. the C. monoletus Zone (Paraná Basin) and the P-L (Chacoparaná Ba-
The top of the Rio do Rasto Formation is marked by a wide- sin) palynozones. Their ages are supported by palynological corre-
spread erosional surface, labeled SB-8 in the chart. In the southern- lation with Late Carboniferous strata in Argentina (e.g., Paganzo
most part of the study area (i.e., Rio Grande do Sul), this surface Basin), Africa and Australia.
separates the Late Permian strata (labeled Pirambóia Formation) In Australia, Gondwanan palynostratigraphic schemes have
from the Early Triassic strata known as Sanga do Cabral Formation. been established and partially calibrated with marine faunas.
However, the definitive stratigraphic relationship of Sanga do Cab- There, the Pseudoreticulatispora confluens and Striatopodocarpites
ral and Pirambóia is still unclear: for some authors the base of fusus Zones were assigned to the Early Permian: Asselian/Tastubian
these units are of Late Permian age (e.g., Lavina, 1991; Milani (Archbold et al., 2004; Césari, 2007). These palynozones are equiv-
et al., 2007), while for others the Sanga do Cabral Formation is alent, in the studied basins, to the V. costabilis Zone (Paraná Basin)
exclusively Triassic (e.g., Cisneros, 2008), and Pirambóia Formation and the Cristatisporites Zone (Chacoparaná Basin), taking into ac-
has a Middle Triassic age (e.g.; Soares et al., in press, 2008). count the stratigraphic ranges of the two mentioned species. Then,
The SB-8 unconformity encloses a variable hiatus, because in the underlying biozones (A. cristatus, C. monoletus and P-L palynoz-
Rio Grande do Sul, the Rio do Rasto Formation has a Guadalupian ones) must be older than the Permian. Furthermore, absolute ages
age, based on tetrapod faunas, while northwards, in Santa Catarina, obtained from the Dwyka Group in Africa (Bangert et al., 1999), the
the age of Rio do Rasto Formation is Early Lopingian (Cisneros et al., regional equivalent of the Itararé Group, where the A. cristatus Zone
2005). Further northwards (i.e., towards São Paulo), the hiatus in- and the C. monoletus Zone were recognized, supporting a Pennsyl-
creases and Pirambóia Formation occurs in direct contact with vanian age for most of the glacial event in Western Gondwana, rep-
the Teresina Formation (e.g., Rohn, 2007, see also Fig. 11) resented by the Itararé Group in Brazil, which bears those
SB-8 is also a second-order sequence boundary, separating the palynozones.
Gondwana I and II supersequences in the stratigraphic framework The Fusacolpites fusus-Vittatina subsaccata Assemblage Biozone
of Milani et al. (2007). of western Argentina occurs within a sedimentary sequence equiv-
alent to a sedimentary one bearing interbedded basalts dated as
302 ± 6 Ma and 288 ± 7 Ma. This biozone is correlated with the V.
10. Geochronology costabilis Zone of the Paraná Basin, corroborating an Early Permian
age for both palynozones (see Césari, 2007), that includes the
Radiometric datings are scarce for the Gondwanan Paleozoic uppermost Itararé Group and the base of the Rio Bonito Formation.
strata. In addition, marine faunal assemblages with index species Geochronological analysis based on Gondwana sections in Afri-
suitable for intercontinental correlation are missing. Then, only ca reveals significant results. An absolute age of 270 Ma was ob-
the meager data on absolute ages are discussed herein, reinforcing tained from tuff beds of the Collingham Formation, in South
remarks presented by Santos et al. (2006) and Césari (2007). Africa, which overlies the Whitehill Formation (Stollhofen et al.,
Shales and tuffs of the studied interval have been target of 2000b). The latter is the regional equivalent of the Irati Formation
radiometric dating by several authors in the last decade, but until in the Paraná Basin, which bears the L. virkkiae Interval Zone. Con-
the present moment, the most reliable data were obtained for the sidering that there is no marked diachronism between these two
Irati shales by Santos et al. (2006) and Rocha-Campos et al. (2007), lithostratigraphical units, the Irati Formation should likewise be
in the Brazilian portion of the Paraná Basin, which indicated an age regarded as older than 270 Ma, reinforcing the absolute age ob-
of 279.9 ± 4.8 Ma and 278 ± 2.2 Ma (Artinskian), respectively. tained for this unit by Santos et al. (2006).
These data are used in the chronostratigrahic chart proposed here- These data corroborate intercontinental correlations estab-
in, as well as correlation of palynozones across Gondwana, as sum- lished from the Eurydesma marine fauna throughout Gondwana,
marized below. regarded as Early Permian (Asselian/Sakmarian) in age, on the

Table 1
Statistic overview on the Late Paleozoic stratigraphic units discussed in the paper. Note variable thicknesses and different sedimentation rates. LPTS = Late Paleozoic Third-order
Sequences, m = meter, m.y. = millions of years, m/m.y. = meters per million of years.

Lithostratigraphic units Aproximate maximum thickness Aproximate time span (duration) Mean sedimentation rate Sequences
Rio do Rasto 650 m 14 m.y. 46 m/m.y LPTS -7
Teresina 400 m 4.5 m.y 90 m/m.y. LPTS -6
Serra Alta 100 m 3.5 m.y. 30 m/m.y. LPTS -6
Irati 70 m 3 m.y. 23 m/m.y. LPTS -5
Palermo 280 m 3 m.y. 90 m/m.y LPTS -4
Rio Bonito 270 m 8 m.y 30 m/m.y. LPTS -3
Itararé 1500 m 22 m.y. 70 m/m.y. LPTS -1 and 2
396 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

top of the glacial sequence (Itararé Group). In the Brazilian Paraná  between LPTS-5 and 6 (the ‘‘Irati-Serra Alta transition”), marked
Basin, equivalent marine elements occur in the upper Itararé, by the appearance of bivalves and disappearance of a plant zone;
where the V. costabilis Zone occurs.  between LPTS-6 and 7 (the ‘‘Teresina-Rio do Rasto transition”),
The table below (Table 1) gives a statistic overview on the marked by bivalves and plants.
stratigraphic units discussed herein, showing ciphers on maximum
thicknesses and using the geochronological data as shown above. Actually, the only sequence boundary dissociated from any
The obtained rates of sedimentation are noteworthy variable and biostratigraphic marker is SB-5, linked to the Irati unit. Hence,
probably reflect several tectonic pulses during the Late Paleozoic from the six sequence boundaries which are exclusively of
(e.g.; Holz et al., 2006) which caused base level variations and reju- third-order hierarchy (SB-2 to SB-7), five are coincident with
venation of source rock areas and overprinted its signature to the major breaks in the biostratigraphic scheme. This seems to
trans-regressive second-order cycle represented by the Gondwana indicate that the FADs (i.e., first appearance datum) and LADs
Supersequence 1 of Milani et al. (2007). (i.e., last appearance datum) of fossils used for the biostrati-
graphic zonation have a stratigraphic rather than biotic/eco-
11. Conceptual comments and conclusions logic control.
However, there are biostratigraphic zones which fall within dif-
The issue of the paper is to show an updated and integrative ferent LPTS. This is the case of the transition from the A. cristatus
overview of both published and new data and interpretations on Zone to the C. monoletus Zone: the limit of this palynomorph zones
the geological evolution of the Late Paleozoic succession of the is coincident with a third-order maximum flooding surface. An-
Paraná Basin, and to summarize that information into an integra- other example is the appearance of the L. virkkiae Zone, which is
tive and comprehensive stratigraphic chart. not linked to a base level fall, but to a transgression. These exam-
All available biostratigraphic data, acquired by the authors and ples seem to indicate that some biostratigraphic zonation is con-
enriched by compiling published data, were updated and corre- trolled by depositional environment and ecologic conditions,
lated to the stratigraphic data. outruling in this case any stratigraphic control.
At the eastern border of the Paraná Basin, the Late Carbonifer- Why does this happen is a matter of controversy and investi-
ous/Permian succession, bounded by basinwide unconformities, gation. In the past, very few attention has been paid to strati-
records a huge trans-regressive cycle with a time span of aprox. graphic control, but since advent and popularization of
67 m.y., hence representing a second order depositional sequence, sequence stratigraphy as a key tool for understanding the gene-
as pointed out elsewhere (e.g. Milani et al., 1994, 1998). This huge sis of the sedimentary record, investigation of the stratigraphic
depositional sequence encloses several unconformities which al- control on fossil preservation has shown that the history of a
low a subdivision into several higher order sequences, which are sedimentary basin influences strongly the distribution of fossils
herein labeled Late Paleozoic third order Sequences (LPTS). Based and introduces a bias towards preservation. FADs and LADs of
on the regional occurrence of features that indicate base level fall fossils within a geologic section are strongly controlled by base
(e.g. marine to fluvial facies shift, pebbly lags, bonebeds) seven level changes (Holland and Patzkowsky, 1999; Holland, 2000;
LPTS’s were recognized and are shown in the stratigraphic chart, Holz and Simões, 2004). The detailed integration of biostrati-
the core of the present paper. Some sequences certainly need fur- graphic data and sequence stratigraphic analysis shown herein
ther subdivisions (e.g. the huge LPTS-2 has an approximately seems to corroborate this statement.
11 m.y. duration, which is the upper limit of duration of third order
sequences according to Krapetz (1996), but kind of extrapolates Acknowledgments
‘‘traditional” third order duration), while others have a relatively
short duration (e.g. LPTS-5). M. Holz, P.A. Souza and R. Iannuzzi acknowledge the Brazilian
We want to close this paper with a conceptual comment. National Research Agency (CNPq) for research support (Grants
Analyzing the data herein present under a conceptual view- 300866/2008-9, 305265/2007-5 and 309322/2007-3, respec-
point, the main conclusion is that there is a very strong correlation tively). R. Rohn acknowledges ‘‘Fundação de Amparo à Pesquisa
between biostratigraphic limits and third-order sequence bound- do Estado de São Paulo” (FAPESP -05/55027-4; 04/11215-9)
aries. This is particularly evident for the following limits: and the National Agency of Petroleum (ANP-PRH-05) for research
grants. Data from wells presented in sequence stratigraphy sec-
 between LPTS-1 and 2 (within the pro-glacial succession), tions and figures are due to CPRM-Serviço Geológico do Brasil,
marked by palynomorphs; who provided access to sample material and well cores. Many
 between LPTS-2 and 3 (the ‘‘Itararé-Rio Bonito transition”), thanks to the two anonymous reviewers who improved the
marked by plants; manuscript with many suggestions and helpful comments, and
 between LPTS-3 and 4 (within the Rio Bonito unit), marked by to Regional Editor Reinhardt Fuck for his considerable editorial
palynomorphs; effort.
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 397

Appendix A
398 M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399

References Ghilardi, R.P., Simões, M.G., 2002. Foram os Bivalves do Grupo Passa Dois (Exclusive
Formação Rio do Rasto), Neopermiano, Invertebrados Tipicamente Dulcícolas?
Pesquisas em Geociências, Porto Alegre 29 (1), 3–13.
Almeida, C.M., Do Carmo, D.A., 2005. Taxonomia e paleoecologia de ostracodes do
Gordon Jr, M., 1946. Classificação das formações gondwânicas do Paraná, Santa
Permiano da Bacia do Paraná, Estado de Goiás, Brasil. In: Congresso Brasileiro de
Catarina e Rio Grande do Sul. Boletim DNPM-DGM, Rio de Janeiro 1, 374–385.
Paleontologia, 19, Congresso Latino-Americano de Paleontologia, 6, Aracaju.
Gradstein, F.M., Ogg, J.G., Smith, A.G., Agterberg, F.P., Bleeker, W., Cooper, R.A.,
Resumos, BP. CD-Resumos.
Davydov, V., Gibbard, P., Hinnov, L., House, M.R., Lourens, L., Luterbacher, H.-P.,
Alves, L.S., 2006. A fossil wood from the Irati formation (Upper Permian), Paraná
McArthur, J., Melchin, M.J., Robb, L.J., Shergold, J., Villeneuve, M., Wardlaw, B.R.,
Basin, São Paulo State, Brazil. Geological Society of America Abstracts with
Ali, J., Brinkhuis, H., Hilgen, F.J., Hooker, J., Howarth, R.J., Knoll, A.H., Laskar, J.,
Programs 38 (7), 553.
Monechi, S., Powell, J., Plumb, K.A., Raffi, I., Röhl, U., Sanfilippo, A., Schmitz, B.,
Alves, R.G., Ade, M.V.B., 1996. Sequence stratigraphy and organic petrography
Shackleton, N.J., Shields, G.A., Strauss, H., Van Dam, J., Veizer, J., Van Kolfschoten,
applied to the study of Candiota Coalfield, RS, South Brazil. International Journal
T.H., Wilson, D., 2004. A Geologic Time Scale 2004. Cambridge University Press.
of Coal Geology 30, 231–248.
610p.
Andreis, R.R., Bossi, G.E., Montardo, D.K., 1980. O Grupo Rosário do Sul (Triássico) no
Guerra-Sommer, M., Cazzulo-Klepzig, M., Formoso, M.L.L., Menegat, R., Mendonça,
Rio Grande do Sul, Brasil. In: Congresso Brasileiro de Geologia, 31, Camboriú-SC,
J.G., 2008a. U–Pb dating of tonstein layers from a coal succession of the
1980, vol. 2. SBG, Anais, pp. 659–673.
southern Paraná Basin (Brazil): a new geochronological approach. Gondwana
Araújo, L.M., 2001. Análise da expressão estratigráfica dos parâmetros de
Research 14 (3), 474–482.
geoquímica orgância e inorgância nas seqüências Irati. Porto Alegre, UFRGS,
Guerra-Sommer, M., Cazzulo-Klepzig, M., Menegat, Rualdo, Formoso, M.L.L., Basei,
2v. 301p. (Tese de doutorado).
M.A.S., Barboza, E.G., Simas, M.W., 2008b. Geochronological data from Faxinal
Archbold, N.W., Cisterna, G.A., Simanauskas, T., 2004. The Gondwanan
coal succession in Southern Paraná Basin: a preliminary approach combining
carboniferous-permian boundary revisited: new data from Australia and
radiometric U/Pb age and palynostratigraphy. Journal of South American Earth
Argentina. Gondwana Research 7 (1), 125–133.
Sciences 25, 246–256.
Assine, M.L., Zacharias, A.A., Perinotto, J.A., 2003. Paleocorrentes, paleogeografia e
Guerra-Sommer, M., Cazzulo-Klepzig, M., Santos, J.O.S., Hartmann, L.A., Ketzer,
seqüências deposicionais da Formação Tatuí, Centro-Leste do Estado de São
J.M.M., Formoso, M.L.L., 2008c. Radiometric age determination of tonsteins and
Paulo. Revista Brasileira de Geociências 33 (1), 33–40.
stratigraphic constrains for the Lower Permian coal succession in Southern
Bangert, B., Stollhofen, H., Lorenz, V., Armstrong, R., 1999. The geochronology and
Paraná Basin, Brazil. International Journal of Coal Geology 74, 13–27.
significance of ash-fall tuffs in the glaciogenic Carboniferous-Permian Dwyka
Holland, S.M., 2000. The quality of the fossil record: a sequence stratigraphic
Group of Namibia and South Africa. Journal of African Earth Sciences 29 (1), 33–
perspective. In: Erwin, D.H., Wing, S.L. (Eds.), Deep Time – Paleobiology’s
49.
Perspective. Supplement to Paleobiology, 26, pp. 148–68.
Barbosa, O., Almeida, F.F.M., 1949. A Série Tubarão na bacia do Rio Tietê, Estado de
Holland, S.M., Patzkowsky, M.E., 1999. Models for simulating the fossil record.
São Paulo. Boletim DNPM-DGM 48, 16p.
Geology 27 (6), 491–494.
Calça, C.P., 2008. Microbiota fóssil da Formação Assistência (Subgrupo Irati,
Holz, M., 1997. Early Permian Sequence stratigraphy and paleophisiography of the
Permiano, Bacia do Paraná) no Estado de São Paulo. São Paulo, Instituto de
Paraná Basin in northeastern Rio Grande do Sul state, Brazil. Anais da Academia
Geociências, Univerisdade de São Paulo, 80p. (M.Sc. dissertation).
Brasileira de Ciências 69 (4), 521–543.
Calça, C.P., Fairchild, T.R., 2005. Uso de lâminas delgadas na paleopalinologia da
Holz, M., 1999. Early Permian sequence stratigraphy and the palaeophysiographic
Formação Assistência (Subgrupo Irati), Permiano, Estado de São Paulo. In: IXX
evolution of the Paraná Basin in southernmost Brazil. Journal of African Earth
Congresso Brasileiro de Paleontologia, 2005, Aracaju. Resumos.
Sciences 29 (1), 51–61.
Castro, J.C., Medeiros, R.A., 1980. Excursão 9: Fácies e modelos de sedimentação das
Holz, M., 2003. Sequence stratigraphy of a lagoonal estuarine system – an example
Formações Rio do Sul e Rio Bonito, leste de Santa Catarina; fácies e evolução
from the lower Permian Rio Bonito Formation, Paraná Basin, Brazil.
sedimentar do Grupo Passa Dois na BR-470-SC. In: Roteiro de Excursões 31o
Sedimentary Geology, Amsterdam 162 (3–4), 301–327.
Congresso Brasileiro de Geologia, Camboriú, SC, vol. 3, pp. 69–97.
Holz, M., Simões, M.G., 2004. Taphonomy – Overview of Main Concepts and
Castro, J,C., Maciel, U., Alves, C.F.C., Grecchi, R.C., 1993. O Grupo Guatá (Permiano
Applications to Sequence Stratigraphic Analysis. In: Koutsoukos, E.A.M. (Org),
Inferior) na margem aflorante paulista. In: Simpósio Sobre Cronoestratigrafia da
Applied Stratigraphy. Springer, pp. 251–280.
Bacia do Paraná, 1, Rio Claro. Resumos. UNESP, Rio Claro, pp. 55–56.
Holz, M., Kuchle, J., Philipp, R.P., Bischoff, A.P., Arima, N., 2006. Hierarchy of tectonic
Césari, S.N., 2007. Palynological biozones and radiometric data at the
control on stratigraphic signatures: base-level changes during the Early
Carboniferous–Permian boundary in western Gondwana. Gondwana Research
Permian in the Paraná Basin, southernmost Brazil. Journal of South American
11 (4), 529–536.
Earth Sciences 22, 185–204.
Chahud, A., 2007. Paleontologia de vertebrados da transição entre os Grupos
Holz, M., Souza, P.A., Ianuzzi, R., 2008. Sequence stratigraphy and biostratigraphy of
Tubarão e Passa Dois (Neopaleozóico) no centro-leste do Estado de São Paulo.
the Late Carboniferous to Early Permian glacial succession (Itararé Subgroup) at
USP, Instituto de Geociências. Tese de Mestrado.
the eastern-southeastern margin of the Paraná Basin, Brazil. Geological Society
Cisneros, J.C., tul=0?>Cisneros, 2008. Taxonomic status of the reptile genus
of America Special Paper 441, 115–129.
Procolophon from the Gondwanan Triassic. Paleontologia Africana 43, 7–17.
Holzförster, F., Stollhofen, H., Stanistreet, I.G., 2000. Lower Permian deposits of the
Cisneros, J.C., Abdala, F., Malabarba, M.C., 2005. Pareiasaurids from the Rio do Rasro
Huab area, NW Namibia: a continental to marine transition. Communications of
Formation, Southern Brazil: biostratigraphic implications for Permian faunas of
the Geological Survey of Namibia 12, 247–257.
the Paraná Basin. Revista Brasileira de Paleontologia 8 (1), 13–24.
Iannuzzi, R., 2007. Biostratigraphic versus Geochronologic frameworks in the Early
Daemon, R.F., Quadros, L.P., 1970. Bioestratigrafia do Neopaleozóico da Bacia do
Permian from Paraná Basin: looking forward a possible consensus. In: Iannuzzi,
Paraná. In: Congresso Brasileiro de Geologia, XXIV, Brasília, 1970. Anais, vol. 1.
R., Boardman, D.R. (Eds.), I Workshop Problems in the Western Gondwana
Sociedade Brasileira de Geologia, Brasília, pp. 359–412.
Geology. South America–Africa correlations: du Toit revisited. CIGO-UFRGS,
Etgeton, V., 1997. Aplicação de conceitos da estratigrafia de seqüências ao intervalo
Gramado, Brazil, Extended Abstracts, pp. 72–77.
Permiano-Eo-Triássico da Bacia do Paraná na região nordeste do Rio Grande do
Iannuzzi, R., Souza, P.A., 2005. Floral succession in the Lower Permian deposits of
Sul, Brasil. PPGGeo-UFRGS, Monografia de Mestrado, 137p.
the Brazilian Paraná Basin: an up-to-date overview. In: Lucas, S.G., Zigler, K.E.
Fairchild, T.R., Coimbra, A.M., Boggiani, P.C., 1985. Ocorrência de estromatólitos
(Eds.), The Nonmarine Permian: New Mexico. New Mexico Museum of Natural
silicificados na Formação Irati (Permiano) da borda setentrional da Bacia do
History and Science Bulletin, 30, 144-149.
Paraná (MT, GO). Anais da Academia Brasileira de Ciências 57 (1), 17.
Krapetz, B., 1996. Sequence stratigraphic concepts applied to the identification of
Fanton, J., Rohn, R., Ricardi-Branco, F., Rösler, O., 2006. Afloramento de Canoinhas,
basin-filling rhythms in Precambrian successions. Australian Journal of Earth
SC – Única localidade de ocorrência da conífera permiana Krauselcladus da
Science 43, 355–380.
Bacia do Paraná. In: Winge, M., Schobbenhaus, C., Berbert-Born, M., Queiroz,
Lages, L.C., 2004. A Formação Irati (Grupo Passa Dois, Permiano, Bacia do Paraná) no
E.T., Campos, D.A., Souza, C.R.G., Fernandes, A.C.S. (Eds.), Sítios Geológicos e
furo de sondagem FP-01-PR (Sapopema, PR). Rio Claro, UNESP, 117p. (M.Sc.
Paleontológicos do Brasil. Published 11/26/2006 at http://www.unb.br/ig/sigep/
Dissertation).
sitio126/sitio126.pdf
Langer, M.C., Eltink, E., Bittencour, J.S., Rohn, R., Serra do Cadeado, P.R., in press. Uma
Ferreira-Oliveira, L.G., 2007. Conchostráceos permianos da Bacia do Paraná:
janela paleobiológica para o Permiano continental sul-americano. In: Winge, M.,
taxonomia, evolução, bioestratigrafia e paleobiogeografia. Tese de Doutorado.
Schobbenhaus, C., Berbert-Born, M., Queiroz, E.T., Campos, D.A., Souza, C.R.G.,
Universidade Estadual Paulista Júlio de Mesquita Filho, UNESP- Rio Claro.
Fernandes, A.C.S. (Eds.), Sítios Geológicos e Paleontológicos do Brasil, vol. III.
Ferreira-Oliveira, L.G., Rohn, R., 2009. Leaiid conchostracans from the uppermost
<http://www.unb.br/ig/sigep/minutas/Serra_do_Cadeado_1_minuta.pdf>
Permian strata of the Paraná Basin, Brazil: Chronostratigraphic and
Lavina, E.L., 1991. Geologia sedimentar e paleogeografia do Neopermiano e
paleobiogeographic implications. Journal of South American Earth Sciences
Eotriássico (Intervalo Kazaniano-Scythiano) da Bacia do Paraná. Porto Alegre,
doi:10.1016/j.jsames.2009.03.006
Instituto de Geociências, Universidade Federal do Rio Grande do Sul, 2v., 333p.,
França, A.B., 1993. Glaciation and tectonics in an active intracratonic basin: the Late
81 figs (Doctoral Thesis).
Palaeozoic Itarare group, Parana Basin, Brazil. Sedimentology 40 (1), 1–25.
Malabarba, M.C., Abdala, F., Weiss, F., Perez, P.A., 2003. New data on the Permian
França, A.B., Potter, P.E., 1988. Estratigrafia, ambiente deposicional e análise de
fauna of Posto Queimado, Rio do Rasto Formation, Southern Brazil. Revista
reservatório do Grupo Itararé (Permo-Carbonífero), Bacia do Paraná (Parte 1).
Brasileira de Paleontologia 6, 49–54.
Boletim de Geociências da Petrobrás 2 (2/4), 147–191.
Maranhão, M.S.A.S., 1995. Fósseis das Formações Corumbataí e Estrada Nova do
França, A.B., Winter, W.R., Assine, M.L., 1996. Arenitos Lapa-Vila Velha: Um modelo
Estado de São Paulo: Subsídios ao Conhecimento Paleontológico e
de trato de sistemas subaquosos canal-lobos sob influência glacial, Grupo
Bioestratigráfico. São Paulo, IG-USP, 2 vols., 86± 361p. (Doctoral Thesis).
Itararé (C-P), Bacia do Paraná. Revista Brasileira de Geociências, Brasília 26 (1),
Maranhão, M.S.A.S., Petri, S., 1997. Novas ocorrências de fósseis nas formações
43–56.
Corumbataí e Estrada Nova de São Paulo e considerações sobre o seu significado
M. Holz et al. / Journal of South American Earth Sciences 29 (2010) 381–399 399

paleontológico e bioestratigráfico. Revista do Instituto Geológico, São Paulo 17 Rohn, R., Rösler, O., 1990. Conchostráceos da Formação Rio do Rasto (Bacia do
(1/2), 33–54. Paraná, Permiano Superior): Bioestratigrafia e implicações paleoambientais.
Matos, S.L.F., Yamamoto, J.K., Riccomini, C., Hachiro, J., Tassinari, C.C.G., 2001. Revista Brasileira de Geociências, São Paulo 9 (4), 486–493.
Absolute dating of Permian ash-fall in the Rio Bonito Formation, Paraná Basin, Rohn, R., Rösler, O., 2000. Middle to Upper Permian Phytostratigraphy of the Eastern
Brazil. Gondwana Research 4, 421–426. Paraná Basin, vol. 5. Rev. Univ. Guarulhos, Guarulhos. pp. 69–73.
Meglhioratti, T., 2006. Estratigrafia de seqüências das formações Serra Alta, Teresina Rohn, R., Lages, L.C., Penatti, J.-R.R., 2003a. Litofácies da Formação Irati no furo de
e Rio do Rasto (Permiano, Bacia do Paraná) na porção nordeste do Paraná e sondagem FP-01-PR (Permiano, borda leste da Bacia do Paraná). In: Congresso
centro-sul de São Paulo. Rio Claro, Instituto de Geociências e Ciências Exatas, Brasileiro de P&D em Petróleo & Gás, 2, Rio de Janeiro. CD ROM, 6 p., Resumos, p.
Universidade Estadual Paulista, 133 p. (M.Sc. Dissertation). 52.
Meglhioratti, T., Rohn, R., Lourenço, A.T.A., 2005. Estratigrafia do grupo Passa Dois Rohn, R., Lourenço, A.T., Meglhioratti, T., 2003b. As formações Serra Alta, Teresina e
na região de Sapopema-Congoinhas/PR (Permiano, Bacia do Paraná). Simpósio Rio do Rasto no furo de sondagem SP-23–PR (Permiano, Grupo Passa Dois,
de Geologia do Sudeste, Niterói, RJ, 2005 (Abstracts). Borda Leste da Bacia do Paraná). In: Congresso Brasileiro de P&D em Petróleo &
Mendes, J.C., 1949. Novos lamelibrânquios fósseis da Série Passa Dois, sul do Brasil. Gás, 2, Rio de Janeiro. CD ROM, 6 p., Resumos, p. 40.
Brasil. DNPM – DGM, Rio de Janeiro, Bol., 133, 40p. Rohn, R., Assine, M.L., Meglhioratti, T., 2005. A new insight on the Late Permian
Mendes, J.C., 1952. A Formação Corumbataí na região do rio Corumbataí environmental changes in the Paraná Basin, South Brazil. In: Gondwana 12,
(estratigrafia e descrição dos lamelibrânquios). Bol. Fac. Fil. Ciênc. Letr., Univ. Mendoza, 2005. Abstracts, Academia Nacional de Ciencias, p. 316.
S. Paulo, 145 (Geol., 8), 119p. Rösler, O., 1978. The Brazilian eogondwanic floral succession. Boletim IG-USP 9, 85–
Mendes, J.C., 1984. Sobre os paleoambientes deposicionais do Grupo Passa Dois. 90.
Revista do Instituto Geológico do Estado de São Paulo 5 (1/2), 15–24. Rösler, O., Rohn, R., Albamonte, L., 1981. Libélula permiana do Estado de São Paulo,
Menezes, J.R.C., 2000. Estratigrafia do Neopermiano da Bacia do Paraná no Rio Brasil (Formação Irati): Gondvanoptilon Brasiliense gen. et sp. nov. In: Congresso
Grande do Sul. In: Holz, M., De Ros, L.F. (Eds.), Geologia do Rio Grande do Sul. Latino–Americano de Paleontologia, 2, Porto Alegre. Anais, vol. 1, pp. 221–232.
Cigo-UFRGS, Porto Alegre, pp. 323–334. Santos, R.V., Souza, P.A., Alvarenga, C.J.S., Dantas, E.L., Pimentel, M.M., Oliveira, C.G.,
Mezzalira, S., 1957. Ocorrências fossilíferas novas da Série Passa Dois na região Araújo, L.M., 2006. Shrimp U–Pb Zircon dating and palinology of bentonitic
Limeira-Rio Claro-Piracicaba. Boletim da Sociedade Brasileira de Geologia 6, 37– layers from the Permian Irati Formation, Paraná Basin, Brazil. Gondwana
59. Research 9, 456–463.
Milani, E.J., França, A.B., Schneider, R.L., 1994. Bacia do Paraná. Boletim Geociências Schneider, R.L., Mühlmann, H., Tommasi, E., Medeiros, R.A., Daemon, R.F., Nogueira,
da Petrobrás 8 (2), 69–82. A.A., 1974. Revisão estratigráfica da Bacia do Paraná. In: Congresso Brasileiro de
Milani, E.J., Faccini, U.F., Scherer, C.M.S., Araújo, L.M., Cupertino, J.A., 1998. Geologia, XXVIII, 1974, Porto Alegre. Anais, Sociedade Brasileira de Geologia,
Sequences and stratigraphy hierarchy of the Paraná Basin (Ordovician to Porto Alegre, vol. 1, pp. 41–66.
Cretaceous), vol. 29. Boletim do Instituto de Geociências/USP, Southern Brazil, Simões, M.G., Rocha-Campos, A.C., Anelli, L.E., 1998. Paleoecology and evolution of
pp. 125–173. Permian pelecypod assemblages (Paraná Basin) from Brazil. In: Johnston, P.A.,
Milani, E.J., Melo, J.H.G., Souza, P.A., Fernandes, L.A., França, A.B., 2007. Bacia do Haggart, J.W. (Eds.), Bivalves – An Eon of Evolution- Paleobiological Studies
Paraná. Boletim de Geociências da Petrobrás, Rio de Janeiro. 15 (2), 265–287. Honoring Norman D. Newell. University of Calgary Press, Calgary, pp. 443–452.
Neregato, R., 2007. Estudo palinológico das formações Serra Alta, Teresina e Rio do Soares, P.C., 1972. O limite glacial-pós-glacial do Grupo Tubarão no Estado de São
Rasto nos furos de sondagem SP-23-PR e SP-58-PR, centro norte do Paraná Paulo. Anais da Academia Brasielira de Ciências, São Paulo 44, 333–341.
(Permiano, Bacia do Paraná). Programa de Pós-Graduação em Geologia Regional, Soares, M.B., 2003. A taphonomic model for the Mesosauridae assemblage of the
Instituto de Geociências e Ciências Exatas, UNESP. M.Sc. Dissertation. 107p. Irati Formation (Paraná Basin, Brazil). Geologica Acta 1 (4), 349–361.
Neregato, R., Souza, P.A., Rohn, R., in press. Registros palinológicos inéditos nas Soares, A.P., Holz, M., Soares, P.C., Bettu, D., 2008b. Variabilidade espacial no Sistema
formações Teresina e Rio do Rasto (Permiano, Grupo Passa Dois, Bacia do Aqüífero Guarani: controles estratigráficos e estruturais. Congresso Brasileiro
Paraná): implicações biocronoestratigráficas e paleoambientais. Pesquisas em de eologia, Curitiba, 2008. CD-ROM. PDF 0128.
Geociências, Porto Alegre. Soares, A.P., Soares, P.C., Holz, M., in press. Correlações conflitantes no limite permo-
Petri, S., Souza, P.A., 1993. Síntese dos conhecimentos e novas concepções sobre a triássico no sul da Bacia do Paraná: o contato entre duas superseqüências e
bioestratigrafia do Subgrupo Itararé, Bacia do Paraná, Brasil. Revista IG, São implicações na configuração espacial do Aqüífero Guarani. Pesquisas em
Paulo 14 (1), 7–18. Geociências, Porto Alegre.
Pinto, I.D., 1995. Paleobotanical and paleozoological age divergence in South Sohn, I.G., Rocha-Campos, A.C., 1990. Late Paleozoic (Gondwanan) ostracodes in the
America strata. Revista Pesquisas, Porto Alegre 22 (1-2), 46–52. Corumbataí Formation, Paraná Basin, Brazil. Journal of Paleontology 64 (1),
Pinto, I.D., Ornellas, L.P., 1981. Permian insects from Paraná Basin, South Brazil. III 116–128.
Homoptera -1-Pereboridae. In: Congresso Latino–Americano de Paleontologia, Sousa, S.H.M., Suguio, K., Castro, J.C., 1991 Sedimentary facies of the Estrada Nova
2, Porto Alegre, vol. 1. Anais, pp. 209–219. and Corumbataí Formations (Late Paleozoic of the Paraná Basin) in the State of
Ricardi-Branco, F., Bernardes-de-Oliveira, M.E., Garcia, M.J., 1999. Novos elementos São Paulo, Brazil. In: International Gondwana Symposium, 7, São Paulo,
tafoflorísticos da Formação Assistência, Subgrupo Irati, Grupo Passa Dois, Bacia Proceedings, USP, pp. 161–172.
do Paraná, provenientes de Angatuba (SP), Brasil. Revista Universidade Souza, P.A., 2006. Late Carboniferous palynostratigraphy of the Itararé Subgroup,
Guarulhos, Geociências IV(6), pp. 85–95. northeastern Paraná Basin, Brazil. Review of Palaeobotany and Palynology 138,
Ricardi-Branco, F., Caires, E.T., Silva, A.M., 2006. Campo de estromatólitos gigantes 9–29.
de Santa Rosa de Viterbo, SP – Excelente registro do litoral do mar permiano Souza, P.A., Marques-Toigo, M., 2003. An overview on the palynostratigraphy of the
Irati, Bacia do Paraná, Brasil. In: Winge, M., Schobbenhaus, C., Berbert-Born, M., Upper Paleozoic strata of the Brazilian Paraná Basin. Revista del Museo
Queiroz, E.T., Campos, D.A., Souza, C.R.G., Fernandes, A.C.S. (Eds.), Sítios Argentino de Ciencias Naturales, nueva serie 5, 205–214.
Geológicos e Paleontológicos do Brasil. Published 11/24/2006 at <http:// Souza, P.A., Marques-Toigo, M., 2005. Progress on the palynostratigraphy of the
www.unb.br/ig/sigep/sitio125/sitio125.pdf>. Permian strata in Rio Grande do Sul State, Paraná Basin, Brazil. Anais da
Rocha-Campos, A.C., Rösler, O., 1978. Late Paleozoic faunal and floral successions in Academia Brasileira de Ciências 77, 353–365.
the Paraná Basin, southeastern Brazil. Boletim IG-USP 9, 1–16. Stevaux, J.C., Souza Filho, E.E., Fulfaro, V.J., 1986. Trato deposicional da Formação
Rocha-Campos, A.C., Basei, M.A.S., Nutman, A.P., Santos, P.R., 2007. SHRIMP U–Pb Tatuí (P) na área aflorante do NE da Bacia do Paraná, Estado de São Paulo. In:
zircon ages of the late Paleozoic sedimentary sequence, Paraná Basin, Brazil. In: SBG, Congr. Bras. Geol., 34, Goiânia, Anais, vol. 1, pp. 219–229.
Simpósio Sobre Cronoestratigrafia da Bacia do Paraná, 4, 2007. Boletim de Stollhofen, H., Stanistreet, I.G., Rohn, R., Holzforster, F., Wanke, A., 2000a. The Gai-as
Resumos. Sociedade Brasileira de Paleontologia, Armação de Búzios, p. 33. lake system, Northern Namibia and Brazil. In: Gierlowski-Kordesch, E.H., Kelts,
Rohn, R., 1994. Evolução ambiental da Bacia do Paraná durante o Neopermiano no K.R. (org.), Lake Basins through Space and Time. AAPG, Tulsa, pp. 87–108.
leste de Santa Catarina e do Paraná. Programa de Pós-graduação em Geologia Stollhofen, H., Stanistreet, I.G., Bangert, B., Grill, H., 2000b. Tuffs, tectonism and
Sedimentar, Doctoral Thesis. Universidade de São Paulo, 386p. glacially related sea-level changes, Carboniferous–Permian, southern Namibia.
Rohn, R., 2001. A estratigrafia da Formação Teresina (Permiano, Bacia do Paraná) Palaeogeography, Palaeoclimatology, Palaeoecology 161, 127–150.
de acordo com furos de sondagem entre Anhembi (SP) e Ortigueira (PR). In: Suguio, K., Sousa, S.H.M., 1985. Restos de mesossaurídeos na Formação Corumbataí,
Melo, J.H.G., Terra, G.J.S. (Eds.), Correlação de Seqüências Paleozóicas Sul- Permiano da Bacia do Paraná, no Estado de São Paulo. Anais da Academia
Americanas. Ciência-Técnica-Petróleo, vol. 20. Exploração de Petróleo, Seção, Brasileira de Ciências 57 (3), 339–347.
pp. 209–218. Tognoli, F.M.W., Rohn, R., Castro, J.C., 2003. Arcabouço estratigráfico do Grupo Guatá
Rohn, R., 2007. The Passa Dois Group (Paraná Basin, Permian): investigations in no leste paranaense. 2° Congresso Brasileiro de P&D em Petróleo & Gás, Rio de
progress. In: Workshop – Problems in the Western Gondwana Geology, South Janeiro, 2003. pdf. 1157, 6p.
America–Africa Correlations: Du Toit revisited. Gramado, 2007. Extended Toledo, C.E.V., 2001. Análise paleoictiológica da Formação Corumbataí na região de
Abstracts, Porto Alegre, UFRGS, Petrobras, vol. 1, pp. 151–157. Rio Claro, Estado de São Paulo. Rio Claro, UNESP, M.Sc. Dissertation, 146p.

You might also like