You are on page 1of 28

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/337095102

Stratigraphy and ages of four Early Silurian through Late Devonian, Early and
Middle Mississippian glaciation events in the Parnaíba Basin and adjacent
areas, NE Brazil

Article  in  Earth-Science Reviews · November 2019


DOI: 10.1016/j.earscirev.2019.103002

CITATIONS READS

12 592

2 authors:

Mário Vicente Caputo R. Oliver Brasil dos Santos


Federal University of Pará 24 PUBLICATIONS   113 CITATIONS   
43 PUBLICATIONS   1,762 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Paleozoic glaciations in the Parnaíba Basin, Western Northeast Brazil View project

Juruá Orogeny- Brazil and Andean countries View project

All content following this page was uploaded by Mário Vicente Caputo on 25 June 2020.

The user has requested enhancement of the downloaded file.


Earth-Science Reviews xxx (xxxx) xxx-xxx

Contents lists available at ScienceDirect

Earth-Science Reviews
journal homepage: http://ees.elsevier.com

F
OO
Stratigraphy and ages of four Early Silurian through Late Devonian, Early and Middle
Mississippian glaciation events in the Parnaíba Basin and adjacent areas, NE Brazil
Mario Vicente Caputo Ph D Retired Professor a ,⁎ , R. Oliver Brasil dos Santos b

PR
a
Universidade Federal do Pará (UFPA), Instituto de Geociências, Faculdade de Geologia, Brazil
b
Independent Consulting, Structural Geology and Geotectonic, Brazil

ARTICLE INFO ABSTRACT

Keywords Turbidity currents, tectonism, debris flows, glaciation, and other geologic processes are considered to be respon-

D
Parnaíba Basin sible for the generation of Silurian–Mississippian diamictites in northern Brazil. Processes that best account for
Northeast Brazil the features observed in these strata are glaciation, deglaciation and sedimentary deformation. Four main glacial
Paleozoic glaciations/deglaciations ages are identified to have occurred in the Parnaíba Basin during the Paleozoic. Glacially influenced deposits are
Correlation found in the Lower Silurian, Upper Devonian, Lower Mississippian, and Middle Mississippian strata. The Parnaíba
TE
Basin is characterized by outcrops providing robust lithological evidence for three glacial ages, identified by
lithological features and dated by palynological studies. A fourth glaciation, as old as Tournaisian, was recently
recognized only in the subsurface of the basin. The glacial features were illustrated and the rocks accurately dated
with palynomorphs based on very accurate palynological studies. Additional, palynological information, derives
from outcrops and shallow boreholes drilled by a private company (Themag) and the National State Mineral Pro-
duction Department (Departamento Nacional de Produção Mineral -DNPM), and the Brazilian Geological Survey
EC

Company (Companhia de Pesquisa de Recursos Minerais-CPRM). The purpose of this reexamination, including
many new illustrations, interpretations and conclusions, is to evaluate primary sedimentological and paleonto-
logical data related to the above mentioned glaciations. Some authors remain skeptical about the record of Early
to early Late Paleozoic glaciations in western Gondwana, especially in Brazil, mainly because evidence provided
in the literature is relatively scanty. The Parnaíba Basin provides special conditions for the study of these glacia-
tions, although evidence for each of them varies considerably as discussed in the text.
RR

1. Introduction and Sergipe-Alagoas rifts of northeastern Brazil, as well as in offshore


and onshore in parts of the Accra Basin of the Republic of Ghana, Africa.
The Parnaíba Basin is an intracratonic, Paleozoic basin located in This indicates that the Parnaíba Basin originally extended far beyond the
the westerly region of northeastern Brazil (Fig. 1A), with an area of present Brazilian equatorial margin.
about~600,000 km2. It comprises a sedimentary thickness of ca. 3,000 m The basin area is bounded by the Capim or Tocantins Arch in the
CO

in addition to about 500 m of igneous rocks (diabase and basalt) in northwest, Ferrer Arch in the north, São Francisco Arch in the east and
its north side. It occupies parts of the Brazilian states of Pará (PA), southeast, and a normal fault zone in the west (Fig. 1 A).
Maranhão (MA), Piauí (PI), Ceará (CE), Tocantins (TO), and Bahia (BA) The Parnaíba Basin was also connected to the Amazon Basin (Fig.
(Fig. 1A). This basin virtually occupied almost the entire part of north- 1B) through the Marajó Rift region, where the facies of most of Serra
eastern Brazil during the Paleozoic and showed continuity with basins Grande Group formations become predominantly marine (Caputo,
of northwestern Africa (Caputo, 1984; Caputo et al., 2005). 1984).
In the Mesozoic, during and after Gondwana breakup, erosion re- After the middle Early Silurian glacial event, a long mild and
UN

duced the basin's original dimensions due to uplift and retreat of its for- non-glacial period persisted, from the latest Early Silurian to the end
mer margins. Paleozoic sedimentary remnants correlative with older for- of the Frasnian (Late Devonian). The thick, worldwide, intermittent
mations of the Parnaíba Basin are found in areas of the pre-Ordovician accumulation of transgressive organic black shales stopped in most
Brazilian Shield (Carvalho et al., 2018) and the Tucano, Jatobá parts of the world at the end-Frasnian due to eustatic sea-level fall
(Krebs, 1974). The widespread deposition of thick organic black shale
probably re


Corresponding author at: Av. Governador Magalhães Barata, 1012. Belém, CEP: 66060-281, Pará, Brasil.
E-mail addresses: mario_caputo@hotmail.com (M.V. Caputo); oliverbrasil9@gmail.com (R.OliverBrasil dos Santos)

https://doi.org/10.1016/j.earscirev.2019.103002
Received 6 November 2018; Received in revised form 30 October 2019; Accepted 31 October 2019
Available online xxx
0012-8252/© 2019.
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
Fig. 1. Parnaíba Basin (P.B.) index map illustrating the location in South America and its main boundaries. A) This basin occupies parts of the Brazilian states of Pará (PA), Maranhão
(MA), Piauí (PI), Ceará (CE), Tocantins (TO), and Bahia (BA). São Luís (SL) and Barreirinhas (Bar) are rift basins in the northern part of the Ferrer Arch. The drill locations UN-24 and
UN-4 were studied by Lobato (2007) and Playford et al. (2012). The geological section of Fig. 21 is indicated by the CO-1-CA-1 broken line. B) Location of the Brazilian sedimentary
basins and their Petrobras (Brazilian State-owned Oil Company) basin classification based on Klemme (1980). For space reasons the reader should use Google search to locate towns,
counties, villages or municipalities mentioned in the text.

sulted in a lowering of atmospheric pCO2 levels (Cox et al., 2001) and permost member of the group. Nowadays, the Itaim Formation is placed
mean global temperatures. These cold climatic conditions caused mas- at the base of the overlying Canindé Group (Fig. 2). The stratigraphic

D
sive stenothermal biota extinction (Copper, 1977), the disappearance chart shows the four glacial ages, and Late Devonian glacial diamictites
of reefs (House, 1975), and a decrease in global evaporite deposition overlying older formations (Figs. 2 and 21).
(Fairbridge, 1972). In addition, cooling caused the narrowing of car-
TE
bonate accumulation belts from 50° to 30° around the Equator, in both 3. Geological considerations
the northern and southern hemispheres (Heckel and Witzke, 1979).
An increased amount of probable cold-climate loess deposition was re- Modern palynological and geological studies indicate that the age
ported by Twenhofel (1932), in the Late Devonian Portage Shale in span of the Serra Grande Group, comprising the Ipu, Tianguá, and Jaicós
North America, and by Nahon and Trompette (1982) in North Africa, formations, is Ordovician–Early Devonian (Le Hérissé et al., 2001;
consistent with glaciation at this time. Positive δ18O excursions paired Grahn et al., 2005). Initially, Early Silurian glacially influenced beds
EC

with negative δ13C excursions (Buggish, 1991) and an abrupt change in the Parnaíba Basin were found in the subsurface (Caputo and Lima,
in the δ34S from +30‰ to +20‰, indicate a possible, sudden turnover 1984).
in the world ocean that might have been caused by cold, dense, oxy- Such strata also occur in other Brazilian basins, both in outcrops and
genated polar deep-water currents (Holser, 1977; Streel et al., 2000). boreholes (Caputo, 1984, 1998; Crowell, 1983; Caputo and Crow-
According to Bratton et al. (1999), marine anoxic conditions ended ell, 1985; Caputo et al., 2008). In turn, Early and Middle Mississip-
nearly 100 ky before the Frasnian-Famennian Boundary (FFB). At this pian diamictites were first found in the subsurface of the Solimões and
RR

time, fully oxygenated conditions returned, accompanied by a marine Amazon basins. They show less evidence of glacial depositional condi-
overturn of deep toxic water that may have disrupted the stable stratifi- tions than corresponding strata in outcrops. Therefore, Paleozoic glacia-
cation and sluggish circulation of the oceans and strong biota extinction tions in several Brazilian basins were discredited and invalidated (Lakin
(Wilde and Berry, 1984). However, the end-Frasnian glaciation seems et al., 2016) due to the lack of descriptions of glacially influenced
to have been strictly intracontinental and may not have reached the con- strata and their scarce documentation in the literature.
temporary shorelines. Therefore, no physical evidence of this glaciation Original and largely confusing regional stratigraphic succession and
could be found in marine deposits in Gondwana thus far. Despite this, nomenclature were subsequently clarified by systematic sampling and
CO

the above-mentioned geological data strongly suggest that a global cli- palynological dating (Loboziak et al., 2000; Le Hérissé et al., 2001;
mate cooling event occurred at the end-Frasnian in Gondwana (Copper, Grahn et al., 2001; Grahn et al., 2005; Melo and Loboziak, 2003).
1977). Another problem is that glacial erosion deeply truncated several
pre-Late Devonian formations dated by palynological studies. As ice
melted, Devonian glaciers unconformably deposited glacial diamictites
2. Parnaíba Basin stratigraphy
on older formations and on pre-Late Ordovician basement (Figs. 2 and
21). This caused difficulties in determining their true age and strati-
UN

The Paleozoic stratigraphic succession of the Parnaíba Basin begins


graphic position during field mapping. Devonian and younger diamic-
with the Serra Grande Series (Fig. 2), as proposed by Small (1914).
tites were noted to contain abundant, well-preserved palynomorphs re-
These strata, up to 900 m thick, consist mainly of conglomerates, peb-
cycled from variably older strata.
bly sandstones, sandstones, diamictites, and limestones. The underlying
The climatic conditions, during glacial deposition in northern Brazil,
folded limestone beds were excluded by Kegel (1953) due to the pres-
were very different from present day ones, that is, tropical and equa-
ence of an angular unconformity between the sandstone and pre-Or-
torial conditions. Chemical weathering of the outcrops is highly perva-
dovician limestone beds. Campbell (1949) renamed the section of this
sive and the identification of original strata is difficult, mainly due to
series as Ipu, but Rodrigues (1967) revalidated it to designate only
a high-level of iron enrichment and corrosion. These diamictite-bearing
the lower section of Small (1914). Later on, this member was raised
formations in some places lie on high-relief areas (plateaus), subject to
to the category of formation (Caputo and Lima, 1984). Additionally,
a high-level of weathering and laterization; therefore, they have been
there is an older, yet undated and unnamed, thin sandstone unit under-
incorrectly mapped as Cenozoic covers in many parts of the Parnaíba
lying disconformably the Ipu Formation in the Serra da Capivara Na-
Basin.
tional Park area. A hiatus also separates the Serra Grande Group from
the overlying Itaim Formation, used by and Carozzi (1980) as the up

2
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
D
TE
Fig. 2. Partial Paleozoic Stratigraphic Chart of the Parnaíba Basin modified from Vaz et al. (2007). Late Devonian diamictites bearing glacial features occur disconformably above older
formations down to pre-Ordovician basement. Stratigraphic Chart based on the Commission on the Geological Map of the World (2008).
EC

The identification of ancient glacial strata in cutting samples is usu- In the past, some geologists suggested that Brazilian Paleozoic di-
ally problematic in wells, where diamictites have not been cored. How- amictites originated from turbidity currents (Ludwig, 1964), tectonism,
ever, drill cuttings may indicate the presence of diamictites during and debris flows (Andrade, 1968, 1972; Lima, 1978), primarily be-
drilling through detection of larger irregular fragments of rocks and min- cause the occurrence of glaciations was considered unlikely in tropical
erals (around 0.2 to 0.5 cm across) within silt–clay–very fine-grained basins. This interpretation changed with the advent of the Plate Tecton-
RR

sand material. Fragments consist of quartz, quartzite, igneous or meta- ics Theory.
morphic pieces. Diamictites can also be recognized in cutting samples by For some time, the tillite interpretation was also discarded because
moulds or voids left by the removal of larger sand grains from the orig- glacial deposition was thought to be rare or non-existent during the
inal massive, fine-grained matrix during drilling. When such details are Early and early Late Paleozoic. This is the case for most of the north-
overlooked, diamictites in certain subsurface stratigraphic levels may ern hemisphere, but not for the Paleozoic Gondwanan continents of the
have passed unnoticed in lithological descriptions, provided by well re- southern hemisphere (Caputo and Crowell, 1985; Fielding et al.,
ports and papers (Cunha et al., 1994, 2007; Góes and Feijó, 1994; 2008; Díaz-Martinez et al., 1999).
CO

Vaz et al., 2007).


For instance, light-gray clay/silt and very fine-grained sandstone 4. Early Silurian glaciation
with larger fragments of rocks and minerals were observed at the top
of the Ipu Formation in well 1-PD-1-MA (Pindaré-Mirim), drilled in The oldest Paleozoic glaciogenic deposits of the Parnaíba Basin
the Maranhão State by Petrobras (Caputo, 1984; Caputo and Lima, (Early Silurian) were first identified in the subsurface by Caputo
1984). The same should be searched for other Parnaíba Basin older (1984); Caputo and Lima (1984), and Caputo and Crowell (1985).
wells, since this may indicate a diamictite of glacial origin. Indeed, Sil- They occur within the upper part of the Ipu Formation in many Petro-
UN

urian diamictites with the same stratigraphic position and age, bearing bras wells, and were shown to be as old as middle to late early Llan-
striated pebbles and boulders, have been found in other wells and out- dovery (Grahn et al., 2005), based on palynological grounds. Palynol-
crops of this basin and elsewhere in Brazil. ogy currently provides the best biostratigraphic tool to date Paleozoic
Many Mesozoic normal faults displaced several formations in the glacial deposits and to correlate globally the effects of glaciation. This
western portion of the Parnaíba Basin. This introduced some uncertainty is the only Silurian glacial episode recognized in the basin thus far. Co-
in the identification of various formations and glacial sediments in the eval glacial deposits have been recently discovered by one of the au-
area, because ice-influenced strata of different ages can be very sim- thors (M.V.C.) and professor Marivaldo dos Santos Nascimento in differ-
ilar in most respects. The correlation of glacial units based on lithol- ent outcrops, in the eastern edge of the basin (Fig. 4 and 5).
ogy, stratigraphic position, and paleontology ensured the assignment of These sediments correlate with glacial layers in other basins in
the different glacial formations to their proper chronostratigraphic posi- Brazil and Africa (Caputo and Crowell, 1985; Semtner and Kl-
tions. itzsch, 1994). The Early Silurian glacial diamictite occurs in the up

3
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

permost part of the Ipu Formation and is apparently conformable with tion, which is characterized by granules, pebbles, cobbles, and boulders
overlying gray and black shales of the Tianguá Formation (Fig. 3B). immersed in a silty–clayey fine-grained sandy matrix (Fig. 3 B, 4 A, B).
The Ipu Formation is composed of conglomerate, pebbly sandstone, The diamictite is massive, white- to cream-colored, and friable to
sandstone, and diamictite beds, up to 350 m thick (Fig. 2, 3 A, B). Con- well-cemented. It contains mainly rounded and angular, scattered quartz
glomerate beds are recurrent and predominant in this formation. The and quartzite clasts of variable size in a silty–clayey and very
lower Ipu conglomerate and sandstone layers are cross- to horizontally fine-grained sandy matrix. Some clasts show striations (Figs. 4B, 5 A).

F
bedded, fine- to coarse-grained, pinkish-brown, reddish-brown, and fri- The main clastic supply was located on the eastern margin of the basin,
able to well-lithified (Fig. 3 B). The upper Ipu conglomerate beds are whereas the marine opening was to the W/NW towards the Amazon
cross-bedded, mainly white, and kaolinitic with fine- to coarse-grained Basin (, ; Carozzi, 1980), close to the Marajó Basin (Fig. 1 B). At

OO
sandstone intercalations (Fig. 3 A, B). Diamictites with striated peb- the eastern and southeastern edges of the basin, diamictite outcrops
bles are present in the uppermost portion of the Ipu Forma with striated clasts occur on Serra Grande escarps, underneath marine
Tianguá Formation shales (Figs. 3 B, 6). The upper Ipu Formation con

PR
D
TE
Fig. 3. A) Partial section of coarse conglomerate, pebbly sandstone, and fine- to coarse-grained sandstone strata in the upper Ipu Formation at the Serra Grande escarpment in the south-
eastern part of the Parnaíba Basin in the Serra da Capivara National Park. Note red-dressed man standing for scale. B) Overall view of the Ipu Formation. Diamictites occur in the uppermost
part of the cliff and are overlain by shales of the Tianguá Formation. Lower Ipu is pinkish-red, and upper Ipu is grayish-white. (Photos: Serra da Capivara National Park).
EC
RR
CO

Fig. 4. A) New Silurian diamictite outcrop occurrence. Middle early Llandovery diamictite displays a light-grayish to white clayey–silty matrix, with dispersed sand- to boulder-sized
quartz, quartzite, and sedimentary clasts within the uppermost part of the Ipu Formation. It occurs along the new CE-257 road, close to Ipueiras County (Ceará State). B) This striated
boulder, 45 cm long, is now at the museum of the Faculty of Geology of Federal University of Pará, Belém. See YouTube: Matacão estriado na Formação Ipu. (Photos: M.V.C.).
UN

Fig. 5. A) Brownish–pink-colored diamictite exposed in the upper part of the Ipu Formation with a large gabbro boulder with a fully diagonal striated plane surface and in front of it
smaller clasts. B) Broken rounded large boulder in the Ipu diamictite. These diamictites are in the same stratigraphic position as those in Figs. 4 and 6, and also in the same area: Pictures:
Courtesy of the Professor Marivaldo dos Santos Nascimento.

4
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
Fig. 6. Overall view of the Tianguá–Ipu contact. The lowermost Tianguá dark-gray shale, when fresh, is laminated and includes several scattered sand grains. The Ipu diamictite with
striated clasts is massive and light-gray-colored. The two men dressed in red stand for scale at the contact of the uppermost Ipu Formation (diamictite) and Tianguá Formation shale. This
outcrop is located at a new road in construction (CE-257), in the upper part of the Serra Grande escarpment, close to Ipueiras town, Ceará State. Photo: M.V.C.

tact is an interbedded sandstone–shale facies, which is laterally equiva- cano Rift Basin, where there is a small appendix rift (Santa Brígida
lent to the overlying shales of the Tianguá Formation. Graben) (Figs. 1B and 7).

D
The Tianguá shale is up to 20 m thick along the eastern basin mar- Rocha-Campos et al. (2003a)b interpreted these striated surfaces
gin and ~270 m thick in wells within the basin depocenter. The lami- in 70 km apart localities 1 and 2, (Fig. 7) of the Santa Brígida Graben,
nated Tianguá shale beds at its lowermost part contain scattered sand inside the Tucano Basin, as caused by iceberg kill scouring on the sup-
TE
grains. Similar occurrences of scattered grains were observed in coeval posed Curituba Formation, during the Late Paleozoic ice age (Fig. 8
Silurian basal shales of the Pitinga Formation in the Amazon Basin (Ca- A). This formation consists of clayey fine- to medium-grained sandstone.
puto, 1998), and the Vila Maria Formation in the Paraná Basin (Grahn There is no biostratigraphic dating for it (Souza-Lima and Farias,
et al., 2000; Faria, 1982), where underlying diamictites with striated 2006), but a Late Carboniferous age has been inferred by several re-
clasts are present as well. searchers (Rocha-Campos, 1981; Rocha-Campos et al., 2003aa;
These Silurian diamictites apparently have no fossils, but the upper Souza-Lima and Farias, 2006). The Curituba (?) Formation rocks
EC

Ipu–lower Tianguá transitional beds contain chitinozoans, miospores, are locally much indurated, fractured, and barren of fossils. Possibly
and acritarchs. Thus, Silurian diamictites in the subsurface are corre- the sea and icebergs did not reach this area. It is seen in Fig. 8 A
lated with coeval outcropping beds of the upper Ipu Formation, which and is here interpreted as the sub-floor of a glacial valley, rather than
display glacial features, such as striated clasts and pavements, as well as the result of iceberg kill scour as proposed by Rocha Campos et al.
pebble pavements (Assis et al., 2019). (1983a). According to this interpretation, the valley walls were later
Shear horizontal zones are present. Shear zones have resulted from removed by erosion. In the outcrop (locality 2), several well-lithified
sandstone pseudo-beds show parallel-stacked subglacial striated surface,
RR

shearing caused by the motion of ice sheets and parts of the frozen sedi-
mentary substrate. Le Heron (2018) named this kind feature as intrafor- along many shear planes, and the supposed lateral berm is striated as
mational shear zone. The Tianguá transitional beds are fossiliferous, and well.
it has been investigated and dated by Le Hérissé et al. (2001) and Parallel-subhorizontal slickensides (Fig. 8 A) may have resulted
Grahn et al. (2005). Volcanic rocks or volcanic ash-fall deposits are from shearing caused by the motion of ice sheets (glaciotectonics) across
not known in the studied Paleozoic section, from Silurian to Middle Mis- valleys in glaciated highland areas on a frozen substrate. A true natural
sissippian, so absolute chronology was not obtained. lateral berm caused by an iceberg kill scouring is not striated (Wood-
CO

In the basin's subsurface, the continental conglomerates and worth-Lynas and Dowdeswell, 1994), but the berm in the supposed
coarse-grained Ipu clastics grade laterally to a shoreface facies in the NW Curituba Formation is striated according toRocha-Campos et al. (2003a,
portion of the basin region, consisting of white, fine- to medium-grained Fig. 8 A), that is, this pseudo-berm may be related to a bottom of glacial
sandstone beds with Skolithos. This suggests that the Early Silurian sea valley slope.
occupied much of the northwestern portion of the basin, close to pre- The supposed Curituba Formation in the area consists of very uni-
sent-day Marajó Rift area (Fig. 1 B). These sandstone beds in the north- form, fine- to medium-grained sandstone beds, without scattered clasts.
western outcrops of the basin are informally named Guamá Formation. This is an indication of absence of glaciomarine or glaciolacustrine sed-
UN

In conclusion, the Ipu diamictite is interpreted as deposited under imentation related to the melting of a hypothetical iceberg charged of
continental glacial conditions due to the presence of some striated clasts clasts in this region.
and pavements (Assis et al., 2019), and some flatiron-shaped, and The third striated surface, possibly as old as Early Silurian, occurs
rare pyramid-shaped pebbles in many occurrences, and its correlation in Igreja Nova County (Rocha-Campos, 1981; Rocha-Campos et al.,
with glacial diamictites in other South American and African Paleozoic 2003aa; Souza-Lima et al., 2014), in locality 3 (Fig. 7) in the Alagoas
basins. State. The surface of the Precambrian basement at Igreja Nova is un-
The Ipu Formation diamictite outcrops are ~400 km away from dulated; showing parallel, elongated, polished, and finely striated low
three well-known striated rock surfaces, two of them in the Bahia and bosses (Rocha-Campos, 1981). Overlying the gneissic basement occurs
Sergipe (SE) states, northeastern Brazil. These are located between the the Mulungu and Boacica members of the Batinga Formation at geo-
Paulo Afonso and Canindé of San Francisco counties, in the north Tu graphic coordinates 10° 15´S and 35° 52´W. In Sergipe, in the Mulungu
type-locality, the Mulungu Member is overlying the thick Karapotó For-
mation (Souza-Lima, 2006a,b), correlated with the Ipu Formation.

5
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
D
Fig. 7. Outside of the Parnaíba Basin, two striated surfaces are present in both sides of a small appendix rift basin named Santa Brígida Graben (Fig. 1 B). The graben is inside the Jatobá
Rift in Bahia and Sergipe states (localities 1 a 2). The Igreja Nova pavement is situated in locality 3 in the Alagoas State (AL) (Rocha Campos et al., 2003a).
TE
EC
RR

Fig. 8. A) Feature on the supposed Curituba Formation in the State of Sergipe, close to Canindé of São Francisco Town in the Santa Brígida Graben (Fig. 7, Locality 2). In the outcrop
several well-lithified sandstone pseudo-beds show parallel-stacked subglacial striated surfaces, along many shear planes and the supposed lateral berm is striated as well. Photo 8 A: M.V.C.
B) Mulungu diamictite (Fig. 7, Locality 3) with angular and irregular-shaped clasts and one thinly and fully striated rounded boulder with one possible transversal gouge structure. The
striated boulder is 18 cm across in Igreja Nova County, Alagoas State. Photo 8 B: Courtesy of Professor Wagner Souza-Lima.
CO

The Mulungu Member is made of massive, up to 30 m thick gray to ing Mulungu Member is textural, faciologically, and stratigraphically
pinkish-cream diamictite, bearing some striated, polished, faceted, and correlated with the Ipu Formation. Like in the Ipu Formation, the Kara-
flatiron-shaped clasts. It contains a large variety of rocks (granite, gab- potó Formation has a conglomeratic pinkish- red facies in the lower por-
bro, diabase, gneiss, quartzite, phyllite and schist) with pebble up to tion and a conglomeratic white-gray facies in the upper part in theirs
boulder sizes immersed in a clayey-silty matrix. Diamictite clasts are scarps. Likewise, the relative proximity between both glacial diamictites,
common and their shapes vary from rounded to predominantly angular, allows a possible correlation between them. Therefore, in this study, the
UN

and some are faceted and striated (Fig. 8 B). Mulungu Member is inferred to be as old as early Llandoverian. How-
The general trend of fluvial paleocurrents in the Karapotó and Ipu ever, a problem with this age and correlation is that the upper part of
formations is northwest (Carvalho et al., 2018; Assis et al., 2019). the Boacica Member of the Batinga Formation is as old as Westphalian,
The basal part of the upper Boacica Member of Batinga Formation based on florenites spores, and pollens (Uesugui and Santos, 1966),
is composed of diamictites with some striated clasts and rhythmites whereas the lower part of the Boacica Member made of rhythmites and
of thinly stratified shale/siltstone and few fine-grained sandstone in- diamictites is barren (Dino et al., 2002).
terbeds. The formation is up to 200 m thick in outcrops; however, in the The presence of spores and pollens, and the abundance of woody or-
subsurface this unit can be as thick as 318 m (Souza-Lima and Farias, ganic residues in the upper part of the Batinga Formation thin sections
2006). give evidence of deposition under continental conditions (Dino et al.,
The age of the Mulungu Member diamictite of the Batinga Forma- 2002). During Westphalian time, marine limestones, wadi, and aeolian
tion is unknown. However, the Karapotó Formation with the overly sandstone beds (Piauí Formation) were deposited in the Parnaí

6
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

ba Basin (Fig. 2) under dry and hot conditions, while limestone and This formation consists of diamictite-bearing faceted and striated clasts
evaporite cyclothems (Itaituba Formation) were accumulated in the in the basal Brocotó Member, and glaciolacustrine rhythmites, with
neighboring Amazon Basin (Fig. 1 B) in an arid climate as well (Ca- dropstones and invertebrate trails in the overlying Brejo do Arroz Mem-
puto, 1984). This indicates that the Westphalian climate was not suited ber. According to Campos (1992) and Campos and Dardenne (1994,
to a glacially influenced sediment deposition in the area. The Boacica 1997, 2002), there is a complete sequence of continental glaciation
Member is very thick and may include internal unconformities. It is pos- with glacial features such as: striated pavements, diamictite-bearing stri-

F
sible that the Mulungu and the lower Boacica members, overlying the ated clasts, varied compositional nature of the clasts, dropstones in
Karapotó Formation, are as old as early Llandoverian, and the upper shales, immature sandstones, and heterogeneous vertical and horizon-
Boacica Member, as old as Westphalian. Based on local paleomagnetic tal stratigraphy. The analyses suggest the occurrence of an icecap with

OO
measurements, Viviani et al. (2000) suggested the Sergipe–Alagoas rivers and lakes fed by water from melting-ice and water influenced
Basin was located at a low latitude of ~25 °S during the deposition of the by floating ice. Pavements are present in paleovalleys, in glacial and
upper Boacica Member. This paleolatitude in Westphalian time is com- pre-glacial areas (Campos and Dardenne, 2002). The processes in-
patible with warm lithic indicators (evaporites and desert sandstones) clude tractive currents for the deposition of sands by rivers, suspension
found respectively in the Amazon and Parnaíba basins. in glacial-lacustrine environments for the deposition of shales and rhyth-
The Mulungu Member was ~1,000 km farther north from glacially mites, and turbidity currents for the deposition of diamictites and con-
influenced outcrops of the latest Pennsylvanian Itararé Group of the glomerates in glacial lakes. Transport direction was deduced to be north-
Paraná Basin and 400 km eastward of the early Llandoverian upper Ipu east-southwestward by Campos (1992) and Campos and Dardenne

PR
Formation of the Parnaíba Basin. The Itararé Group is a mid-latitude (1994, 1997, 2002). Nevertheless, Rocha-Campos et al. (2003a)b,
glacial record in Brazil (Torsvik and Cocks, 2013). The Windhoek after examining the orientation of friction cracks and abrasion features
highlands in Namibia (Africa) fed ice into the Paraná Basin, remain- on pavements and shipbacks, suggested the possibility of the glacier
ing high throughout much of Pennsylvanian and Early Permian time flow towards NNE. Other evidence of these ice-sheets direction is de-
(Rocha-Campos et al., 2008). duced from a far traveled large itabirite (iron ore) boulder, as well as
Modern paleomagnetic data at 440 Ma places the Silurian mag- dropstones, with the provenance from a highland situated in the south-
netic paleopole close to the present day Amazon River mouth in south- eastern part of the Minas Gerais State (Fig. 9 A). This suggests a glacier

D
ern northwestern Africa, at the ~Aeronian-Rhuddanian time (Cohen et motion toward the Parnaiba Basin and northeastern regions of Brazil.
al., 2013). The paleomagnetic South Pole, from southern northwest- A Late Paleozoic age was inferred for the Floresta Formation (Brejo
ern Africa, in the early Ruddanian, displaced to south South America in do Arroz and Brocotó members) based on trace fossils Isopodichnus and
the earliest Devonian (Fig. 2 of Torsvik and Cocks, 2013). The Mu- Diplichnites in the upper glaciolacustrine facies composed of siltstone and
TE
lungu diamictite deposit was closer to the Silurian (Aeronian) South Pole shale beds (Campos and Dardenne, 1994, 1997, 2002 ). However,
than to the Westphalian South Pole located on Antarctica (Fielding et this age is in question, because the Diplichnites and Isopodichnus chronos-
al., 2008; Torsvik and Cocks, 2013). According to Rocha-Campos tratigraphic distribution reported in the literature ranges from the Early
et al. (2003a) a,b, the regional paleogeography of the Mulungu Mem- Cambrian to the Triassic. Thus, they are not a diagnostic age indicator
ber was puzzling. This perplexity is because they considered the Mu- (Fig. 10).
lungu diamictites as old as Pennsylvanian instead of Early Silurian and In many other places over the Precambrian Sanfranciscan craton,
EC

the glacial flow was toward the N and NE, in an opposite direction of striated pavements are associated to diamictites and dropstone bear-
the glacial Itararé Group outcrops as old as Westphalian, in the Paraná ing shales containing clasts of variable size and nature (Dardenne and
Basin. Campos, 2003). The Santa Fé strata, with glacial records, were also ob-
Based on sedimentary features, paleoclimate affinities, paleomag- served in the Minas Gerais State in São Romão (Fig. 9 C), Urucuia river
netic data (Smith et al., 1981), and worldwide paleogeography, a valley, near Unaí and Porto Feliz, and Paracatu river valley. A sugges-
correlation between the diamictites of the upper Ipu Formation of the tion is that ice-sheets covered almost all Brazilian pre-Ordovician high-
RR

Parnaíba Basin and the diamictites of the Mulungu Member of the land areas in Early Silurian, Late Devonian and other mentioned Paleo-
Sergipe-Alagoas Basin has been proposed by Caputo (1984) and Ca- zoic periods (Fig 10). Rocha-Campos (2003b) has interpreted the pres-
puto and Lima (1984). ence of icebergs in Sergipe and Bahia states, while Dardenne and Cam-
In the Minas Gerais (MG) State, in the Santa Fé de Minas County pos (2003) have considered the presence of ice-sheets in neighboring
(Fig. 9), there are additional striated pavements in the San Franciscan continental basement areas of Minas Gerais. The best explanation is the
Craton on a Precambrian arkose of the Três Marias Formation, with ge- presence of terrestrial ice-sheets in these central regions, not only in the
ographic coordinates 45° 31´ 54´´W and 16° 43´ 34´´S. The basal unit Early Silurian but also Late Devonian and Mississippian times.
CO

of the Phanerozoic is the Floresta Formation of the Santa Fé Group,


composed of three members: Brocotó, Brejo do Arroz, and Lavado.
UN

Fig. 9. A) Location of the study area in Minas Gerais and Goiás States. B) Location map in more detail. C) Schematic geomorphologic map of glacial deposits on Precambrian basement,
where evidence of glaciation/deglaciation occurs. The age of glacially influenced deposits is unknown, but an Early Lllandoverian age is suggested. In Fig. 9 C, dashed arches: glacier
margins; arrows: ice flow direction; gray areas: highlands; small rectangle: studied area. After Rocha-Campos et al., 2003.

7
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
D
TE
EC
RR
CO

Fig. 10. Schematic map of main Paleozoic basins of South America with glacially influenced sedimentation. Arrows indicate Early Silurian ice flow towards sedimentary basins. In the
blank area, basement striated pavements of various mentioned Paleozoic ice ages may occur. Modified from Grahn and Caputo, 1992.

5. Age of the Serra Grande glacial diamictites Grahn et al. (2005) also studied chitinozoans and miospores from
selected cores of 15 Petrobras deep wells. The chitinozoans indicate an
Caputo and Lima (1984) dated the lower part of the Serra Grande Early Silurian (late Aeronian–early Telychian) age for the Ipu–Tianguá
Group to be as old as the latest Ordovician–earliest Silurian, based on transitional strata, Tianguá, and lowermost Jaicós formations. The up-
UN

palynomorphs of the Tianguá Formation as well as inferences and cor- per part of the Jaicós Formation yielded chitinozoan species similar to
relation with other Brazilian and African glacial deposits. According to those of the latest Pragian (mid-Early Devonian). According to Grahn
Le Hérissé et al. (2001), the chitinozoan assemblages of the Tianguá et al. (2005), the palynoflora of the Tianguá Formation is dominated
Formation, above the Ipu diamictite, contain diagnostic components of by cryptospores, whereas only few trilete spores are noted. It is rep-
the Conochitina elongata Zone and the upper part of the Conochitina pro- resentative of the younger (post-Rhuddanian) portion of the Laevolan-
bosphera Desmochitina cf. densa Subzone, which are characteristic for the cis divellomedia miospore Subzone, dated as Aeronian–early Telychian.
Early Silurian (late Aeronian to early middle-Telychian). Acritarch as- The organic-walled microphytoplankton investigated by Le Hérissé et
semblages indicate an early to middle-Telychian age, whereas the few al. (2001) indicates an early middle Telychian age. A single grapto-
spores and cryptospores registered are consistent with an Aeronian to lite fragment found by one of the authors (M.V.C.) in core 52 of well
early Telychian age span. 1-BJ-1-PA (Badajós), within the Tianguá Formation, was later identified
as Climacograptus cf. scalaris scalaris (H. Jaeger, written communication

8
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

to J.H.G. Melo in March 1992). It confirms a middle or earliest late Llan- (San Gabán diamictites) and Bolivia (Cancañari diamictites)
dovery age for this section. The overlying Jaicós Formation may con- (Díaz-Martínez and Grahn, 2007) and Díaz-Martínez et al. (2011).
ceal one or two unconformities, probably of Silurian age Grahn et al. In North Africa, glaciogenic diamictites are present in the upper Tanez-
(2005). Le Hérissé et al. (2001) and Grahn et al. (2005) identified zuft Formation of Libya with Late Ordovician–Silurian age (Le Heron
and illustrated most of the palynomorph species observed in the ana- et al., 2013). In South Africa, the Pakhuis Formation has glacial nature,
lyzed wells; newly discovered species were also described and published but its age is regarded as latest Ordovician (Le Heron et al., 2018).

F
by them. Based on the examination of some North African graptolites and bra-
The lower and middle parts of the Ipu Formation are inferred to chiopods, Arthur James Boucot (personal letter communication to John
be of Late Ordovician to earliest Llandovery age. This is rather uncer- Chambers Crowell, 1982) stressed that the oldest part of the glaciation

OO
tain, since the Ipu Formation is over 300 m thick (Fig. 2) on the east- is recorded in the Caradoc (early Late Ordovician) of Libya, whereas its
ern flank of the Parnaíba Basin, and, thus far, no fossils have been ob- youngest part is documented in the middle Llandovery of Algeria.
served below the Tianguá–Ipu transitional contact. The Ipu Formation
is correlated with the Nhamundá and Autás-Mirim formations of the 6. Far-field middle–late Llandoverian glacial geochemistry
Amazon Basin, which collectively are dated as Late Ordovician-Early Sil-
urian. Thick sandstone and conglomerate beds of the lower Ipu Forma- Commonly, glaciations may be preceded by anoxic events with sig-
tion may be partly glaciofluvial, and thus related to the Late Ordovi- nificant accumulation of organic black shales during widespread trans-
cian glaciation in many areas of northwestern Africa. This is inferred be-

PR
gressions. This is the case of the Ipu glaciation. The massive accumula-
cause one of the authors (M.V.C.) found some clasts with faceted, flat- tion of marine organic matter in the uppermost Ordovician and lower-
iron, and pyramid-shaped (features typical of glacial origin) within the most Silurian of Africa (Semtner and Klitzsch, 1994) and elsewhere
pebbly sandstone and conglomerate beds of the lower Ipu Formation, during coeval major transgressions caused the lowering of the CO2 level
in the Serra da Capivara region. Kegel (1953) also recorded faceted in the atmosphere and hydrosphere, resulting in cooling and glaciation.
clasts in the lower sandstones and conglomerates of the Ipu Formation. Positive δ18O values are often associated with negative δ13C shifts in the
Metelo (1999, Fig. 24) shows a quartzite clast dropped, possibly in Cenozoic marine isotope record, which has been interpreted as the re-
the Late Ordovician, in a suspension deposit in the lowermost Ipu For- sult of an increased ice volume affecting Earth (Berger and Vincent,

D
mation, indicating clast rafting in a periglacial environment of sedimen- 1986). The high CO2 withdrawals from the air, due to organic matter
tation in the Capivara National Park area. Assis et al. (2019) men- accumulation in the sea, may have caused a decrease in global tem-
tion striated clasts too in the lower portion of this formation. These peratures and the onset of a new glaciation (Caputo, 1994; Gouldey
clasts might have been resulted from older melted glaciers. The up- et al., 2010) at higher latitudes of Gondwana. A negative δ13C excur-
TE
per Ipu diamictites correlate with the upper, glacially influenced beds sion (-1.3‰ to -1.4‰ ) was observed by Kaljo and Martma (2000)
of the Nhamundá Formation, in the Amazon Basin (Rodrigues et al., in Estonia and West Latvia (Baltica Paleocontinent) in the late Aeron-
1971; Caputo and Vasconcelos, 1971; ; Caputo and Sad, 1974; ian. Sea-level changes, ice volume and climate interaction schematically
Grahn and Caputo, 1992), and, with the lower Vila Maria Forma- is shown in Fig. 11. This negative δ13Corg excursion, associated with
tion glacial diamictites, in the Paraná Basin (Grahn et al., 2000).
sea-level lowstand and unconformity (Kaljo and Martma, 2000), may
Other Early Silurian diamictites occur in the Tarija Basin (Zapla diamic-
EC

be correlated with the upper Ipu glacial event in South America during
tites) of Argentina (Grahn and Gutíerrez, 2001), as well as in Peru
the latest Aeronian-earliest Telychian.
RR
CO
UN

Fig. 11. Sea level changes, ice volume and climate Interaction. This design was worked out based on Fischer and Arthur (1977) and Saltzman (1985).

9
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

This glaciation was preceded by widespread deposition of marine or- scribed outcrops with diamictite with striated clasts and rhythmite sedi-
ganic black shale (Hot shale I of Lüning et al., 2000) in North Africa. ments with lonestones over a striated sandstone-surface on the old road
Late Ordovician glaciation has been reported from North Africa, includ- from Canto do Buriti to São Raimundo Nonato towns. In this place, di-
ing many localities in Morocco, Algeria, Niger, Libya, Chad, Sudan, Jor- amictites rest on sandstone beds of the Cabeças Formation, which was
dan, and Saudi Arabia. confused with the Serra Grande Formation by Malzahn (1957).
Due to the retreat of the ice caps, a major glacioeustatic sea-level rise The Cabeças Formation displays thicknesses up to ca. 350 m on the

F
took place in the latest Ordovician and earliest Silurian (Rhuddanian), eastern border of the Parnaíba Basin, where it is best developed. Ac-
accompanied by the massive deposition of marine organic black shales cording to Ponciano and Della Fávera (2009), the widespread oc-
in the Murzuk and Ghadamis basins of Libya (Semtner and Klitzsch, currence and nature of sigmoidal clinoforms in the Cabeças Formation

OO
1994). Similar shales were deposited in Central Arabia. They constitute (i.e., several meters thick sandstones, bearing asymmptotic cross-stratifi-
the main source rock of gigantic Paleozoic oil and gas fields in parts of cation, and climbing lobes, without the distinctive features of tide-dom-
Saudi Arabia, Libya, Algeria, and other neighboring countries. This cor- inated facies, like mud drapes and current reversal) are considered to
responds to the Hot Shale I reported by Lüning et al. (2000). constitute evidence of flood-influenced deposition. During interglacial
The following regression of the Early Silurian sea was probably periods, immense meltwater discharge during deglaciation phases may
caused by a second glacial event, which is well-evidenced in many ar- have resulted in the establishment of periglacial catastrophic outwash
eas of Chad, Sudan, and Niger (Semtner and Klitzsch, 1994) and in deposition. Common flat bedding planes with parting lineation, char-
South America. The Early Silurian glaciation is considered to be respon- acteristic of the upper flow regime, was found in the lower fine- to

PR
sible for the end of the deposition of the Hot Shale I (Semtner and medium-grained sandstone strata of the Cabeças Formation in the Mira-
Klitzsch, 1994). The Tianguá Formation shale correlates with the Hot norte County, suggesting a high-energy depositional environment (Fig.
Shale II of Lüning et al. (2000, 2006), deposited in southern North 13 B).
Africa after the melting of Early Silurian ice caps. This glaciation may The glacial diamictite level occurs in the middle to upper parts of
have occurred at the Aeronian-Telychian boundary. This certainly in- the Cabeças Formation bounded by unconformities in the southwestern
dicates that the Ordovician and Silurian glaciations were two separate basin flank (Fig. 21, well 1-CL-1-MA). It is dark-gray and over 40 m
glacial events. Sea-level lowstands provide a greater mass of continen- thick. It consists of massive polygenic, unsorted clasts, some striated, im-

D
tal-sourced sediments in the sea, increasing the rate of 87Sr/86Sr values mersed in an unstratified clayey, silty, and very fine-grained sandy ma-
of limestone beds. High-resolution 87Sr/86Sr values from the Baltica pa- trix. The section (Fig. 12) shows at least two glacial diamictite intervals
leocontinent show an increased rate of 87Sr/86Sr in the Early Silurian with some striated clasts separated by sandstone beds. These sandstones
(Gouldey et al., 2010). This may be related to a higher rate of con- are fine-grained and contain layers of massive fine grained sandstone
TE
tinental sediment input at a time of lower sea-level stand during the with sedimentary structures such as balls, dishes, pillow and convolute
glaciation recorded in the upper Ipu Formation, as discussed above (Fig. folds. These sedimentary structures are typical of turbidity currents. The
11). A large volume of black shale accumulation took place in the east- basal part of the lower diamictite is marked by a shear zone (Fig. 12) in
ern Baltic epicontinental sea from the Rhuddanian to middle Aeronian. contact with fine-grained sandstone. The two diamictite bodies may in-
Organic shale deposition ceased in the late Aeronian, according to Kiipli dicate two glacial advances and the two sandstone bodies may indicate
et al. (2004). Some striated surfaces (Assis et al., 2019) observed in glacial retreats.
EC

the Early Silurian glacial drift in the Parnaíba Basin, and probably some
glaciations elsewhere, correspond to detached surfaces in subglacially
deformed beds (intraformational shear zones, according to Deynoux and
Ghienne (2004) and Le Heron et al. (2018).

7. Late Devonian glaciation


RR

The second ice age in the Parnaíba Basin is characterized by glacial,


glaciomarine, and glaciofluvial deposits. This glaciation took place near
the top of the Late Devonian succession, within the upper part of the
Cabeças Formation (; Carozzi, 1980; Crowell, 1983; Caputo, 1984,
1985; Caputo and Crowell, 1985; Caputo et al., 2008).
The term Cabeças Formation was introduced by Plummer (1948) ,
CO

in the Parnaíba Basin to designate a Devonian sandy section exposed


in the former village of Cabeças (now Dom Expedito Lopes Town), in
the State of Piauí. He divided the formation into three members, in
ascending order: Passagem, Oeiras, and Ipiranga. However, there is a
major unconformity at the lower part of the Oeiras Member, so that
the lower sub-unit (Passagem, bearing Middle Devonian marine inverte-
brates) is now incorporated into the Pimenteira Formation according to
UN

Beurlen, 1965; Loboziak et al., 1994, Loboziak et al., 1994), and


Grahn and Melo (2005). included the uppermost Ipiranga Member in
the overlying Longá Formation. The Cabeças Formation belongs to the
Canindé Group, which is composed of the Itaim, Pimenteira, Cabeças,
Longá, and Poti formations (Fig. 2), altogether encompassing a late
Early Devonian–Middle Mississippian age span (Vaz et al., 2007). The
Late Devonian glaciation is the best known glacial event in this basin.
It was first recorded by Kegel (1953), who recognized diamictite with Fig. 12. Partial stratigraphic column of glacial diamictites of the middle-upper Cabeças
striated and faceted clasts in some cores of the Carolina (1-CL-1-MA) Formation with intercalation of deformed sandstone beds, close to Pedro Afonso Town on
and Riachão (1-VG-1-MA) wells. He misattributed these diamictite sec- the road to Itacajá Town (Oliveira, 1997).
tions to the overlying Longá Formation. Malzahn (1957) de

10
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
Fig. 13. (A) Very poorly-sorted and stratified sandstone and conglomerate intercalations with dispersed angular boulders. Outcrop in Retiro, 5 km north of the town of São Francisco
do Piauí (Caputo, 1985). (B) Sandstone, with planar parting lamination characteristic of upper flow regime structures, suggests sediment accumulation under the influence of strongly
seasonal palaeoclimate that involves a pronounced seasonal peak in flood and runoff. The planar parting lamination follows the direction of the hammer handle in the Miranorte County.

During some periods, strong meltwater discharge may have resulted mites, usually laminated siltstones and shales with scattered clasts are

PR
in the establishment of significant deposition of very poorly-sorted peb- known from outcrops (Malzhan, 1957), and well cores of the upper
bly sandstone and conglomerate beds. Cabeças Formation (Carozzi, 1980; Streel et al., 2000) in the west-
The isolated irregular-shaped boulders of Fig. 13 A were interpreted ern part of the Parnaíba Basin, State of Tocantins , with sediments laid
as transported under catastrophic flood conditions. They occur in some down under glacial and periglacial conditions.
stratigraphic levels, and possibly, were deposited from floating ice, sug- Latest Famennian acritarchs are present and confirm the marine
gesting ice rafting of clasts (Fig. 13 A). In the Miranorte County, the character of the depositional environment of the diamictites and rhyth-
Cabeças Formation shows planar parting lamination typical of the upper mites, even though a large proportion (70%) of the palynological mater-
flow regime (Fig. 13 B). ial is reworked from Givetian/Frasnian strata (Streel et al., 2000). The

D
Several sandstone strata show parallel horizontal slickensides along latest Devonian diamictite includes far-traveled boulders, some of them
the stratification, named intraformational shear zone by Le Heron et over one meter across, composed of basement schist, gneiss, quartz,
al. (2018) that formed in some meters below the glacier basis. Below quartzite, conglomerate, and sedimentary rocks derived from older Pa-
TE
the level of the shear zone, there is no sign of this glaciotectonic struc- leozoic formations (Fig. 16 A).
ture, and glacier has moved together part of the substrate. These paral- Folded silty/fine-grained sandstone horseshoe-shaped boulder is in-
lel slickensides may be confused with a striated pavement, but they are serted in diamictite masses (Fig. 17 A) and massive diamictite boulders
stacked sedimentary rock body surfaces with slikensides due to shearing (Fig. 17 B) are included in another diamictite with shearing and folding
stress. structures, indicating two diamictite generations.
The upper sandstone overlying diamictites is generally white, fine- Glacial striated pavements are well-represented in the municipalities
EC

to medium-grained, with angular grains, micaceous, with kaolinitic ma- of Brejo do Piauí and Canto do Buriti (State of Piauí), particularly in
trix. It is commonly a much thinner section than that of the lower thick outcrops near the locality of Calembre. Here, the best known pavement
sandstone beds under the diamictites (see in Fig. 21, the upper sand- (geographic coordinates: 8° 15´ 02´´S, 42° 52´57´´W) is located about
stone section preserved in well 1-CA-1-MA). The main feature of the up- 26 km south of Canto do Buriti Town (Caputo and Ponciano, 2013).
per sandstone is that it is massive in some sections and displays many This remarkable extensive and well-preserved pavement is polished and
slump structures and soft sediment deformation, possibly connected to striated locally, with subparallel ridges and grooves that have irregular
glacier melting. decimetric spacing and centimetric depths and are sculptured into peb-
RR

Some slump structures were interpreted as undermelt structures. bly sandstone and cross-stratified coarse-grained sandstone beds (Fig.
They may be the result of stagnant ice covered by a new sedimentary 18). In Morro comprido area, many striated pavements are also seen
pack. In circumpolar areas, when the sandstone pore ice and the under- (Fig. 16 B).
lying ice finally melt, the internal sediment structure collapses over an The orientation of parallel striae is not coincident with that of the
irregular surface, completely destroying its original sedimentary fabric sandstone cross-bedding strike, indicating that these features do not re-
(Fig. 14 A, B). These collapsed sandstone beds are very common in the sult from the intersection between the bedding planes and the horizon-
CO

upper sandstone unit and are interpreted to have been deposited during tal ground surface. Lodged clasts embedded in the pavement surfaces
deglaciation processes. change from small-sized cobbles to large-sized pebbles and boulders.
Polished and striated clasts are shown in the Fig. 15 A, and rhyth- These may be aligned or irregularly dispersed into the sandstone, some
mite sediments with a lonestone in the Fig. 15 B. Varvelike rhyth showing notable features of glacial abrasion (Fig. 19 A, B).
UN

Fig. 14. Upper Cabeças sandstone. A) Collapsed massive lower sandstone, overlain by cross-bedded sandstone beds. B) Irregular surface underlying collapsed sandstone in Pedro Afonso
Town Port, Tocantins River. Photos: M.V.C.

11
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
Fig. 15. A) Faceted, striated, and polished cobble extracted from the Cabeças Formation diamictite exposed in the bank of the Tocantins River, north of the town of Pedro Afonso. The coin
is 2.5 cm in diameter. After Caputo and Crowell (1985). B) Rhythmically thin-bedded core. Note the small lonestone. It caused the disruption of lamination at the bottom and draping
of sediment over the upper surface of the clast (core 16, well 1-TM-1-MA, after ).

PR
D
TE
EC

Fig. 16. A) Pre-Middle Ordovician silicified and indurated conglomerate boulder (1.25 m long) is contained in the latest Devonian Cabeças Formation at Morro Comprido area. B) A stri-
ated pavement is seen close to the boulder area. Photos M.V.C.
RR
CO

Fig. 17. A) Recumbent-folded boulder made of silty/fine-grained sandstone that was folded and involved by the latest Devonian massive diamictite. B) Adjacent to Fig. 17A, two diamic-
tite generations. The massive diamictite boulder, one meter across (C), is included in another diamictite with shearing and folding structures. The voids in the upper part of the picture
indicate the former location of other diamictite boulders. Photos M.V.C.

Several undulating shearing surfaces (Fig. 20 A) in the diamictite and overlying formations. Therefore, these features are intraformational
UN

were observed in the area. These features are interpreted as a lodg- and atectonic and probably related to glacier motion.
ment diamictite formed by loading and glacier motion (glacial tecton- The stratigraphic position, origin, and nature of the Cabeças di-
ics). Disrupted and folded fine-grained sandstone bodies may be caused amictite were controversial and have been inadequately explained for
by ice shove. Several shearing surfaces are evident with different ori- many years. According to Moore (1963), the presence of a large num-
entations (Fig. 20 B). Barbosa et al. (2014) show many glaciotec- ber of diamictite outcrops also in the Pimenteira and Itaim formations
tonic features of the Cabeças Formation diamictites, such as intrafor- was problematic, since they were found at different stratigraphic lev-
mational breccias, clastic dikes, and diamictite sills, folds, thrust and els (Fig. 21). The lower Cabeças Formation sandstone is absent in
normal faults, sandstone pods, and detachment surfaces. This deforma- the town of Pedro Afonso. The diamictite in the Carolina well is ap-
tion is restricted to the glacial deposits and does not affect underlying proximately 40 m thick and directly overlies disconformably the Pi-
menteira Formation. The Cabeças sandstone under the diamictite is very

12
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

tion sections underlying the Late Devonian diamictites become progres-


sively older towards the western margin of the Parnaíba Basin, as in-
dicated by geological mapping and paleontological data (Loboziak et
al., 2000). Near Colinas do Tocantins Town, latest Famennian glacial
diamictites of the Cabeças Formation rest directly on the pre-Ordovician
basement and are overlain by Longá Formation shales of Tournaisian

F
age (Fig. 21).
This means that the latest Devonian glaciation deeply eroded all
older formations under the Cabeças diamictites down to the basement

OO
on the southwestern margin of the basin (Loboziak et al., 2000; Ca-
puto and Ponciano, 2013). Moreover, when the Late Devonian glac-
ier melted, it left glacial diamictites directly on these older stratigraphic
units (Fig. 21).
The glacial evidence in the Cabeças Formation includes slump struc-
tures, soft sediment deformation, diamictites bearing polished, striated
and flatiron clasts, and rhythmite sediments with lonestones, striated
pavements, intraformational shear zones, and exotic blocks. Paleomag-

PR
Fig. 18. Partial view of a large striated pavement of the Calembre Village in Brejo do Piauí
County, which is made of coarse sandstone with sparse pebbles, cobbles, and boulders netic studies indicate a high-paleolatitude of the diamictite deposition
stuck at the top of sandstones and conglomerates underlying the diamictites. The hammer
(Torsvik and Cocks, 2013) and stable isotope determinations help in-
in the central part of the picture is used for scale. Caputo and Ponciano (2013) furnish a
geographic location and more details of this striated pavement.
fer sea-water cold temperatures and climate.
In North America, a distinctive diamictite section was long inter-
preted as debris flow sedimentation by Sevon (1969, ). It was ini-
thick (300 m) in the central portions of the basin. By contrast, the sand-
tially regarded as a highly localized deposit (Sevon, 1969, ; Sevon
stone pinches out and disappears in the southwestern part, because deep
et al., 1997). considered the diamictite as resulting from glaciation
glacial erosion has removed it together with older formations (Fig. 2

D
in the Late Devonian. This Late Devonian diamictite extends for over
and 21).
400 km along the strike, with a folded width of nearly 40 km from north-
No lower sandstone beds intervene between the Cabeças diamictite
eastern Pennsylvania across western Maryland and into eastern–central
and the Pimenteira Formation in this region. The thick lower Cabeças
Formation sandstone was totally removed in well 1-CL-1-MA (Carolina) West Virginia. The polymictic diamictite in the Spechty Kopf and Rock-
TE
(Fig. 21) and Pedro Afonso area. The Pimenteira and Itaim forma well Formations deposits are contemporaneous with regressive facies
that reflect fluvial valley incision during the Late Devonian in the Ap
EC
RR

Fig. 19. A) Detail of faceted cobble stuck in the pebbly striated sandstone pavement shown in Fig. 18. B). Multi-faceted cobble stuck in the pavement shows a flatiron-shape. Photos
CO

M.V.C.
UN

Fig. 20. A) Diamictite disrupted with nearly parallel undulating shearing surfaces. The hammer is supported by a.shearing indurated surface rich in iron. B) Irregularly disrupted and
folded siltstone and fine-grained sandstone body with many shearing planes throughout the diamictite. Photos M.V.C.

13
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
D
TE
Fig. 21. The latest Devonian unconformity truncated older strata beneath the diamictites down to the pre-Ordovician basement, as shown in the schematic geological section from the
EC

Petrobras well 1-CA-1-MA (Caraíba) to Colinas do Tocantins Town (COL). Well 1-CA-1-MA displays a more complete stratigraphic section above and below the Cabeças Formation diamic-
tite, as compared to well 1-CL-1-MA. See location of CO—1-CA-1 section in Fig.1 A.

palachian foreland basin (Brezinski et al., 2008). Other latest Devon- ried over hundred kilometers by large icebergs (Clayton et al., 2010).
ian diamictites could possibly be found in southeastern North American Alpine glaciation is restricted to ground above sea-level, when only out-
basins. The source area may have been located on the Acadian moun- wash deposits reach the nearby sea. Tidewater glaciers do reach the sea,
tains and could be as old as Middle to Late Devonian (390 to 370 Ma). although they may originate in the ice fields, ice caps or even ice sheets,
RR

This chain was uplifted prior the Late Devonian glaciation. they cannot carry huge boulders for long lengths because they break up
Based on the observation and redescription of known outcrops, and numerous small icebergs.
identification of new exposures of this diamictite, Brezinski et al. According to Brezinski et al. (2008, 2010), the evidence for the
(2008, 2010) disclosed an association of lithofacies that is not typi- Late Devonian glaciation in North America is primarily based on glacial
cal of debris flow deposition. The diamictite does not exhibit stratifica- features in the area and glacial coeval deposits in high-latitude regions
tion, cross-bedding, or grading. The observed exposures show character- of South America (Caputo, 1985; Caputo et al., 2006).
CO

istics that are best interpreted as deposits produced by glacial processes, In Africa, reliable evidence of Devonian glacial sedimentation is
such as some clasts immersed in a massive matrix that exhibit stria- found in the Takoradi Formation of the Accra Basin in Ghana, both off-
tions, faceting, polishing, and flatiron appearance. Rhythmite deposits, shore and onshore (Junner, 1939; ; Bär and Riegel, 1974; Caputo,
sheared diamictites, and large exotic clasts are also present. In Rowan 1984), and the Teragh Formation diamictites (Valsardieu and Dars,
County, Kentucky, a huge granitic lonestone about 3 tons, named Robyn- 1971; Hambrey and Kluyver, 1981, Caputo, 1984) of the Tim
son, is embedded in the uppermost part of the Cleveland shale in the Lo- Mersoï Basin (Republic of Niger). The upper Witpoort diamictites of
gan Hollow Branch, 8 km northwest of Morehead. It is located ~500 km the upper Witteberg Group of the Karoo Basin (South Africa) have a
UN

west of the diamictite outcrops in Pennsylvania and Maryland (Clay- glacigenic origin (Almond et al., 2002). This interpretation is sup-
ton et al., 2010). Miospore assemblages from samples immediately be- ported by diamictites, occasional pyramidal clasts and striated surfaces
low, above and adjacent to this dropstone can be truly attributed to the within quartzites beneath the diamictites. This striated surface may be a
end-Famennian Retispora lepidophyta-Verrucosisporites nitidus (LN) Zone shear zone related to glacier motion. The Witpoort Formation is as old
(Clayton et al., 2010). as Late Devonian (Famennian) based in palynomorphs and fish (Streel
The continental diamictite section is both under- and overlain by and Theron, 1999).
marine strata (Brezinski et al., 2010). These glacial strata pinch out
between marine beds towards the west and were folded and eroded in 8. Age of Cabeças Formation diamictites
the east by the later Alleghenian orogeny in the Pennsylvanian to Per-
mian time (300 to 250 Ma). These authors consider that this glacia- The combined use of selected Euramerican and Western Gondwanan
tion and deglaciation was of alpine type. This interpretation may not miospore species as zonal and characteristic taxa permits the accu-
be correct because the huge rafted granitic boulder (Robynson) was car rate subdivision, dating, and correlation of Devonian palyniferous strata
in the Parnaíba and Amazon basins, in terms of equiva

14
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

lent miospore zones in Western Europe and the Old Red Sandstone Con- late Frasnian was followed by a decrease to lower values during the
tinent. Famennian. This may indicate an oceanic turnover deleterious to life,
The miospore biostratigraphy of core 31 of well 1-CL-1-MA (Car- which may have been caused by colder water circulation due to a cool-
olina) within the diamictite section (Fig. 21), reported by Kegel ing event. This worldwide Kellwasser phenomenon is considered as one
(1953) was revised by Loboziak et al., 1994, Loboziak et al., 1994). of the six big biocrisis events of the Phanerozoic (Kaiser et al., 2006,
In addition, equivalent glaciogenic sections were dated in the Solimões 2015).

F
and Amazon basins by Grahn et al. (2001), Streel et al., 2000; Melo A fast, worldwide drop in the sea-level also occurred in the latest Fa-
and Loboziak (2003), and Grahn et al. (2006). mennian due to the widespread buildup of ice sheets on Western Gond-
Nearly all Brazilian Devonian diamictite samples examined thus far wana, with consequent marine facies and climate change, biota extinc-

OO
contain Retispora lepidophyta, the miospore index species that charac- tion, and deep fluvial incision elsewhere (Isaacson et al., 1999, b).
terizes the latest Famennian (or “Strunian”) on a near-worldwide scale The main mechanism of CO2 extraction from the atmosphere and oceans
(Melo and Loboziak, 2000, 2003). The total stratigraphic range of is its net transfer to the biomass through biological activity and its sub-
R. lepidophyta in the Ardenne–Rhenish Massif (“Strunian” type area) is sequent incorporation to sedimentary rocks.
currently subdivided into three interval zones (Maziani et al., 1999). The climate in North American regions, where the glacial diamic-
Each biozone has its base defined by the first appearance of the cor- tites were found, was mainly hot and dry before the Famennian glacia-
responding eponymous species, from oldest to youngest: Knoxisporites tion (Brezinski et al., 2008, 2009, and 2010). These climatic condi-
tions were unsuitable to the spreading of rich land plant covers and coal

PR
literatus (LL) Zone, Indotriradites explanatus (LE) Zone, and Verrucosis-
porites nitidus (LN) Zone. However, the lowest zone (LL) has not been preservation, just before the Late Devonian glaciation. The increase of
identified in any of the Brazilian sections studied to date. This indi- land plants cover causes reduction of atmospheric and marine CO2, and
cates a minimum early Strunian hiatus below the diamictite in the Pa- cooling. Nevertheless, the low cover of land plants was inadequate to
leozoic basins of northern Brazil. Moreover, both verrucosisporites nitidus cool the climate in Late Devonian time. However, the high-burial of ma-
and Vallatisporites vallatus (another LN marker) may be either scarce or rine organic matter was more effective in reducing the atmospheric CO2
absent in several examined diamictite sections (Melo and Loboziak, and cooling the climate during high sea-level stands in Late Devonian
2000, 2001, 2003). (Fig. 11). In addition, all older ice ages originated without any interven-

D
An undifferentiated LE/LN zonal attribution was proposed for the to- tion of land plants. The global Hangenberg biocrisis event, long consid-
tal range of Retispora lepidophyta in those areas. It corresponds to the ered as a second-order extinction event, is now regarded to be as severe
praesulcata conodont zone in Western Europe (Fig. 3 in Streel and as the global Late Frasnian Kellwasser biocrisis, because it caused the
Loboziak, 1996). Thus, the latest Devonian (Strunian) diamictites of extinction of 45% of marine genera and wiped out entire ecosystems at
TE
Brazilian Paleozoic basins correspond to the Rle/LVa Gondwanan zonal the end of the Famennian (Kaiser et al., 2015, and references therein).
succession, which correlates with the LE/LN Euramerican zonal succes-
sion (Melo and Loboziak, 2003). The diamictites from Pedro Afonso 10. Early Mississippian glaciation
Town outcrops are coeval with those of well 1-CL-1-MA (Carolina)
(Loboziak et al., 2000). The third Paleozoic glaciation in the Parnaíba Basin is associated
Strunian diamictites are present in all Brazilian intracratonic basins with the Longá Formation, and was first recognized a few years ago,
EC

(Solimões, Amazon, Parnaíba, and Paraná), in the rifted Marajó and Ja- in the subsurface of this basin (Lobato, 2007, Lobato, 2010; Lobato
tobá basins of northern and northeastern Brazil, and in Andean basins and Borghi, 2007, 2014). The Longá Formation was first named by
(Caputo et al., 2008; Díaz-Martínez, 2004). Albuquerque and Dequech in 1946), but published in 1950, who de-
U/Pb zircon datings of three volcanic ash beds bracket the Hangen- scribed a black shale section cropping out in the Longá River on the
berg shale in Poland (Holy Cross Mountains) between 358.97 ± 0.11 Ma road from Campo Maior to Castelo villages, in Piauí State. It is charac-
and 358.89 ± 0.20 Ma, with a calculated duration of 0.05 + 0.14/- 0.05 terized by laminated or banded and partly bioturbated shaly siltstone
RR

Ma (Myrrow et al., 2013). They provide an indirect limit for the tim- strata. The formation was apparently deposited under shelf conditions
ing of the latest Devonian glaciation. The end of deposition of this black in a transgressive, restricted, storm-dominated sea environment (Vaz et
shale was succeeded by marl and marly shale, associated with a major al., 2007; Della Fávera, 1993), as suggested by thin sandstone and con-
sea-level drop and simultaneous faunal extinction (Kaiser et al., 2011). glomerate levels, about 30 cm thick, bearing hummocky structures and
The glaciation is coeval with the deposition of sandstone strata re- remobilized bivalve shell fossils (Santos and Carvalho, 2004, 2009).
lated to this sea-level drop before the end of the Devonian (Kaiser et This could be due to high-energy, stormy conditions.
The Longá Formation is up to 65 m and 220 m thick in outcrops and
CO

al., 2011). The younger U/Pb ash bed, located 38 cm above the topmost
Hangenberg shale, was dated at 358.57 ± 0.20 Ma (Myrrow et al., subsurface of the basin's depocenter, respectively. The upper part of the
2013). There is a 2.20 m thick section between the Hangenberg shale formation's diamictite is missing in the southwestern outcrops, where
and the Devonian-Carboniferous boundary (DCB). Based on cyclostratig- the Longá Formation may be subdivided into three informal parts. Its
raphy, using eccentricity cycles (De Vleeschouwer et al. (2013), the upper contact is disconformable with the Poti Formation, with a hiatus
accumulation rate between the Dasberg and Hangenberg shales was cal- of ~10–11 My (Melo and Loboziak, 2000; Grahn et al., 2001). The
culated to be 5.6 m/Ky. Based on this accumulation rate the DCB is 320 lower contact is also disconformable with the Cabeças Formation. In the
Ky younger than this ash bed (Myrrow et al., 2013). This indicates subsurface of the northern part of the Parnaíba Basin, Tournaisian di-
UN

that the duration of the glaciation in the end of the Devonian may have amictite beds occur in the upper portion of the Longá Formation (Fig.
been about ~100 Ky, which agrees with the estimate of Streel et al. 22).
(2012). Outcrops of Longá beds show marked color banding and strong
red-color in weathered expositions. However, they normally present
9. Far-field Late Devonian climate and glacial geochemistry dark-gray to black shales, with calcareous stringers within the upper
part of the unit. A thick diamictite, with shale and siltstone interbeds,
The latest Frasnian comprises an up to 5 cm thick pyrite bed in the is present in the upper Longá Formation in the subsurface of the north-
Jasper Basin near Medicine Lake (Alberta) and in the Selwyn Basin eastern region of the basin, which laterally changes into a shaly facies.
in the Yukon Territory, Canada, which shows a positive δ34S excur- In the central parts of the basin, diamictites are replaced by shale beds.
sion (+34‰) close to the Frasnian–Famennian Boundary, according In shallow boreholes drilled in the 1970´s by the Departamento
to Geldsetzer et al. (1987). The marked increase of the δ34S values Nacional da Produção Mineral (DNPM) and Companhia de Pesquisa
in pyrite in laminated anoxic argillaceous lime–mudstone during the de Recursos Minerais (CPRM) for a joint coal research project in

15
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

ing to Playford et al. (2012). Diamictite (black triangles), sandstones (dotted pattern)
and mudstone (black color).

the Parnaíba Basin, Lobato (2007, 2010) and Lobato and Borghi (2007,
2014) identified diamictite levels with a single striated clast. Studies
were conducted in the well 1UN-24, with the following geographic co-

F
ordinates 03° 54´ 37´´S and 42° 56´ 47´´W (Fig. 22). These authors car-
ried out facies studies of shallow boreholes (1UN-4 and 1UN-24), lo-

OO
cated in the northeastern part of the Parnaíba Basin (Fig. 1A). Only well
1UN-24 was deep enough to reach the glaciogenic section drilled in the
Maranhão State (MA). This well comprises parts of the upper Longá and
lower Poti formations (Fig. 22). The Longá Formation diamictite occurs
within the depth interval 338-276.5 m. However, the drilled glaciogenic
section is incomplete because the well did not fully penetrate the glacial
beds (Playford et al., 2012). The Longá Formation also belongs to the
Canindé Group (Fig. 2), which spans the late Early Devonian to Middle

PR
Mississippian (Vaz et al., 2007).
The Tournaisian glaciation was also evidenced in subsurface strata of
the upper Jaraqui and Oriximiná formations, in the Solimões and Ama-
zon basins respectively (Caputo and Crowell, 1985; Caputo et al.,
2008). Diamictite beds occur in the upper part of the Longá Formation
in the subsurface (Fig. 22) and include mainly sedimentary clasts and
also, subordinately, igneous and metamorphic rocks. The Longá Forma-

D
tion was divided into three informal members in the southwest outcrop
area of the basin. The thickness of each member varies, and the medium
member may pinch out completely northwards into the basin's depocen-
ter.
TE
The lower member comprises 20−30 m thick, laminated, green-gray-
ish and partially bioturbated dark-gray, micaceous, in part sideritic
and chloritic shales, which include thin interbeds of siltstone and
fine-grained sandstone. Thin conglomerate and conglomeratic sandstone
beds and lenses occur in some stratigraphic levels. The middle unit,
approximately 10−20 m thick, consists mainly of very fine-grained,
EC

well-sorted, white or yellow, ferruginous, sideritic, in part micaceous,


argillaceous, cross-laminated, and cross-bedded sandstone beds. The up-
per unit, 10−30 m thick, is composed of black, gray, and green shales
that are sideritic, pyritic, micaceous, laminated and partially biotur-
bated, with sandstone lenses and oligomictic sand-supported conglom-
erate beds, about 30 cm thick. Some thin, medium- to coarse- grained
RR

sandstone lenses and beds are frequent throughout the entire formation.
The upper Longá diamictite, rich in Tournaisian palynomorphs, re-
sembles the dark-gray diamictites that occur in the upper parts of
the Cabeças and Poti formations. However, in some places, the Longá
diamictite presents brown color due to weathering. Recycled paly-
nomorphs of Middle and Late Devonian age are likewise common in the
CO

Longá Formation.
Lobato and Borghi (2007, 2014) interpreted 13 sedimentary fa-
cies (10 lithofacies and 3 ichnofacies), grouped in 9 facies-successions
(FS), which allowed the identification of glaciomarine, wave-dominated
shallow marine, and wave-dominated fluvio-deltaic (delta front and
prodelta) depositional systems. The analysis of FS permitted the charac-
terization of forced regressive stratigraphic surfaces (breaks) of varied
UN

natures and orders, which in turn allowed the characterization of two


3rd order stratigraphic sequences, limited by subareal erosive surfaces,
including other sequences of smaller orders.
Small shale and sandstone fragments and rounded diamictite pel-
lets are dispersed in the clayey–silty–sandy matrix (Fig. 23 A). Angular
quartz sand grains and rock fragments show fabric alignment, possibly
Fig. 22. Core drill UN-24 with its lithologic log and sampling depths. The Poti and Longá due to internal shearing caused by glacier motion (Fig. 23 B).
formations contact (P/L) is disconformable. Blank intervals in the lithologic column in- Fig. 24. Photos A and B show drill cores of the Longá Formation
dicate the non-recovery. The diamictite thickness is ~50 m, but it is incomplete, accord with deformed balls and pellets of massive siltstone–fine-grained sand-
stone (SC), shale (ShC), and other clasts within a sandy diamictite, indi-
cating significant remobilization/loading. Ojeda and Perillo (1967),
in their Fig. 6, show a lonestone boulder about 18 cm across in a sus-
pension deposit (Fig. 24 C).

16
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
Fig. 23. Thin-sections (parallel nicols) of Longá Formation at a depth of 322.7 m of well UN-24 drilled by the DNPM/CPRM consortium. A) Diamictite is poorly sorted and has angular
clasts, many corroded, as large as fine- to coarse-grained sand and larger diamictite pellets, which are immersed in a clayey–silty–fine-grained sandy sideritic matrix. B) Poorly sorted
diamictite showing preferential clasts and matrix alignment and sand grain corrosion. Scale bar =1 mm (Lobato, 2007).

PR
D
TE
Fig. 24. Longá Formation. A) Deformed balls of massive siltstone–fine-grained sandstone (SC), shale (ShC) and other clasts within a sandy diamictite, indicating significant remobilization
EC

and loading. Core-drill UN-24 drilled by the DNPM/CPRM consortium with a core diameter of 4 cm. Depth: 292 m. B) Soft-sediment deformation within a diamictite with large angular
shale piece and very fine grained sandstone clasts. Depth: 285 m. After Playford et al. (2012). C) Quartzite boulder about 18 cm across included in argillaceous fine-grained sandstone
suspension deposit of the Longá Formation. It caused bending of underlying strata, interruption of lateral stratification and draping of sediment over its upper surface. Photo: Ojeda and
Perillo, 1967).

Longá Formation outcrops in the Campo Maior area show some fea- amictites of the Jaraqui Formation. Latest Famennian diamictites at-
tures similar to fossil triangular ice wedges formed in a periglacial per- tained widespread distribution in the Solimões Basin, and a large part
RR

mafrost environment (Fig. 25 A). Recent ice wedge (Fig. 25 B) is of these rocks may have been eroded and sporemorphs may have been
shown for comparison with Longá fossil triangular wedge-like outcrops recycled into the Tournaisian diamictites (Caputo et al., 2008).
(Fig. 26). Detailed stratigraphic analysis in the central Appalachian Foreland
Tournaisian diamictites are known in the subsurface in the Parnaíba, Basin in the United States shows coarse-grained, conglomeratic, and flu-
Amazon, and Solimões basins. In the latter basin ( Fig 26 A, B), most vial sandstones (Black Hand Sandstone) to be associated with an uncon-
often, these diamictites rest disconformably on Strunian di formity and sharp valley incision down to 60 m deep (
CO
UN

Fig. 25. A) Brownish-red Longá shale and siltstone layers with possible fossil triangular ice wedge structures, suggesting formation in a periglacial permafrost environment in the Parnaíba
Basin. Strata folded toward the soil thermal fracture. Road from Altos to Campo Maior Counties. B) Recent ice wedge formed in fractures in a permafrost environment in a periglacial area
for comparison with Longá outcrops. Beds fold downward along the thermal fractures. Photo A: Courtesy of the Hungarian geologist Dr. Peter Szatmari. Photo B: The Northfolk Project:
http:// www.northfolk.org.uk/geology/geoindex.html.

17
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
Fig. 26. A) Petrobras core of massive gray diamictite with many angular sand- and granule-sized quartz clasts, which are included in the clayey–silty matrix of the Tournaisian Upper

D
Jaraqui Formation of the Solimões Basin. B) In the Solimões Basin, latest Famennian and middle to late Tournaisian diamictites may occur in direct disconformable contact.

Matchen and Kammer, 2006). It is inferred that the sandstone-filled Canyon Range, Nevada, and Namur‒Dinant Basin, Belgium) for the
channel has resulted from forced regression during the Kinder- mid-Tournaisian and Viséan intervals. The mid‒Tournaisian event
TE
hookian–Osagean transition interval (Matchen and Kammer, 2006). shows increases in δ13Ccarb, δ13Corg, and δ15N bulk at Arrow Canyon
This unconformity marks a global sea-level fall event. The correlation of (from +0.6 to +7.0‰, ‒25.2 to ‒23.7‰, and ‒0.6 to +8.9‰, respec-
this widespread unconformity with marine and continental glaciation in tively) and Namur‒Dinant (from ‒0.3 to +4.7‰, ‒27.9 to ‒22.2‰,
South America was discussed by Matchen and Kammer (2006); Kam- and ‒5.7 to +5.2‰, respectively). The positive excursions in all three
mer and Matchen (2008), and b, 2008) , ).
isotopic records during this event are consistent with increased frac-
tional burial of organic matter and enhanced denitrification. It means
EC

11. Age of Longá Formation diamictites


an increase in marine productivity and expansion of hypoxia in the
global ocean and epicontinental basins. This event coincided with falls
The taxonomic identification of plant microfossils (miospores) in
in global sea-level and temperature related to the Gondwanan Tour-
eight diamictite cores and sixteen intercalated marine shale cores of well
naisian glaciation (Caputo et al., 2008). Liu et al. (2018) hypoth-
1UN-24 indicated an Early Mississippian age for the Longá Formation
esize that global cooling led to accelerated oceanic circulation and up-
diamictites (Playford et al., 2012).
welling on continental margins, triggering increased marine productiv-
RR

The lowermost part of the Longá Formation bears a poor benthic ma-
ity and redox changes within the affected upwelling zones.
rine invertebrate fauna of latest Devonian age (Santos and Carvalho,
Two positive δ13Ccarb isotope shifts have been recognized in the late
2009). However, this basal shale is separated from higher parts of the
Kinderhookian and early Osagean, in limestones of Idaho and Nevada.
formation by a hiatus encompassing an early Tournaisian or longer time
The oldest δ13Ccarb peak (+7‰%o) occurs in the isosticha conodont
span, depending on whether pre-PD strata are locally present or ab-
Zone and a younger, minor peak is present in the typicus conodont Zone
sent in the Longá Formation due to Tournaisian onlap along basin edges
(Saltzman, 2002). These values correlate with the curves from bra-
(Playford et al., 2012). These thin basal shale beds below the uncon-
CO

chiopod calcite from the Midcontinent and Western Europe (Saltzman,


formity could be considered as a shaly facies of the Cabeças Formation,
2002). Thus, the δ13C curve decreases from +7 ‰to +5‰ and +4‰,
to be formalized in the future as a new member.
The clasts observed in the Longá diamictites are grossly oriented and suggesting the reduction of marine anoxia and temperature.
consist of angular to rounded quartz, shale, sandstone, and igneous peb- Marine anoxia briefly returned in the typicus Zone after this episode.
bles and cobbles. It was dated to be as old as Early Mississippian (Tour- The detailed δ13C data obtained by Mii et al. (1999), on brachiopod
naisian) by Playford et al. (2012). Melo and Loboziak (2003) es- shell calcite from the Midcontinent region of North America also show
tablished three, successive, Tournaisian miospore zones in the Amazon positive excursions up to 5.4 ‰ in the late Kinderhookian. It is interest-
UN

Basin. These are, in ascending stratigraphic order: the Radiizonates arcu- ing that the δ18O record shows a 3% increase in the Kinderhookian–Os-
atus–Waltzispora lanzonii (AL), Spelaeotriletes balteatus–Neoraistrickia lo- agean transition interval, suggesting climate cooling caused by glacia-
ganii (BL), and Spelaeotriletes pretiosus–Colatisporites decorus (PD) interval tion in Gondwana, according to Caputo et al. (2008).
zones. Palynofloras of the Spelaeotriletes pretiosus–Colatisporites decorus Increased δ13C values may indicate enhanced organic 12C burial and
(PD) miospore Interval Zone indicate a late middle to early late Tour- sequestration of atmospheric CO2 that provoked global cooling, glacia-
naisian age for the Longá diamictites. The PD Zone is equivalent to tion, and sea level fall (values of δ13C began to drop very fast when the
the Spelaeotriletes pretiosus–Raistrickia clavata (PC) Zone of Western Eu- climate became cold). In the same way, low δ18O values were observed
rope (Melo and Loboziak, 2003). The representative palynomorphs during times of massive organic matter accumulation, whereas high val-
are identified and well-illustrated by Playford et al. (2012). ues were recorded when cooling and glaciation took place.

12. Far-field Tournaisian climate and glacial geochemistry

Liu et al. (2018) presented organic carbon and nitrogen isotope


profiles for two widely separated Mississippian regions (Arrow

18
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
D
TE
EC
RR

Fig. 27. Poti Formation section exposed in the southwestern portion of the Parnaíba Basin, according to Andrade (1968, 1972).

13. Middle Mississippian glaciation the underlying Poti Formation. Poti Formation diamictites were first rec-
ognized by Oliveira (1961), then working for Petrobras. He mapped
The fourth glacial deposits of the Paleozoic, recognized in the the Tocantins River and one of its tributaries, the Lontra River, along
Parnaíba Basin (Viséan, Middle Mississippian) are documented in the which he found glacial diamictites with a large variety of clast sizes
Tocantins River valley, near Itacajá Town (State of Tocantins), as well immersed in a clayey–silty–fine-sandy matrix. However, he incorrectly
CO

as in other areas where the Poti Formation is exposed. The term Poti Se- placed these glacial diamictites in the underlying Longá Formation. Bar-
ries was introduced by Lisboa (1914) to refer to sandstone beds crop- bosa et al. (1966) also found diamictites with striated clasts in the To-
ping out in the Poti River valley in the State of Piauí. Later, Paiva and cantins River outcrops, with striae aligned after one, two, three, or four
Miranda (1937) identified the Poti Series in the water drill No 125 different directions. Also , these authors misplaced the diamictites in the
near Terezina City (Piauí), within a depth interval of 219–566 m. Camp- Longá Formation.
bell (1949) restricted this unit to the depth interval of 219–423 m Several Petrobras field parties of the sixties ( Moore, 1963, Ojeda
in the same well. The Poti Formation is 170 m thick in the south- and Bembom, 1966; Ojeda and Perillo, 1967; I Perillo and Na-
UN

western outcrop belt and 320 m thick in the basin's depocenter. The hass, 1968; Andrade, 1968) mapped Devonian and Mississippian out-
lower contact is disconformable with the underlying Longá Formation. crops of the Cabeças, Longá, and Poti formations in the southwest-
Strata of latest Tournaisian–early Viséan age are apparently absent in ern part of the Parnaíba Basin. The lower Poti section (or informal
the Parnaíba Basin, which indicates a stratigraphic gap separating the member), as thick as 120 m, consists of pink–yellowish, micaceous, fri-
Longá and Poti formations. The same unconformity is also noted in able, fine- to medium-grained, occasionally coarse-grained sandstone,
equivalent sections of the Amazon and Solimões basins (Melo and often high-angle cross-bedded, with thin siltstone interbeds (Fig. 27).
Loboziak, 2000). There is a low-angle unconformity between Poti and It contains subrounded to angular, well-sorted grains and a large num-
the overlying Piauí Formation, of Westphalian age (Fig. 2). The basal ber of unoriented boulders of quartz, quartzite, gneiss, granite, and
conglomerate of the Piauí Formation, consisting of a large amount of shale scattered in the cross-bedded sandstone layers. This unit con-
quartz, quartzite, and igneous rock clasts, probably originated from tains several oligomictic, lenticular, sand-supported conglomerate beds
the recycling of diamictites, conglomerates, and pebbly sandstones of

19
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

with angular to subangular clasts of the same nature as those of the Sediment deposition by melting of ice rafts may also have taken
above-mentioned sandstone beds, and few silty/clayey interbeds. place (Fig. 28). This suggests the presence of ice-rafted lonestones in a
The upper Poti informal member, up to 54 m thick, consists of pink- periglacial environment, about 400Km apart from known glacial diamic-
ish–violet and greenish-gray, massive, blocky, micaceous, sometimes tite outcrops.
calciferous diamictite, characterized by pebble to boulder-sized clasts of Many shallow-wells were drilled in 2006 by the Themag Engineer-
quartz, mica schist, gneiss, and quartzite, dispersed within a compact, ing and Management Company for the construction of a dam in the To-

F
silty–argillaceous matrix. It typically shows a peculiar ferruginous black cantins River near Tupiratins Village, State of Tocantins (Fig. 29). The
alteration coat on the clast surfaces. cores from some wells were studied by Caputo et al., 2006,a, 2008).
It contains subrounded to angular, well-sorted grains and a large They contain sandstones, silty shales, and dark-gray to gray diamictites

OO
number of unoriented boulders of quartz, quartzite, gneiss, granite, and of late Viséan age (Streel et al., 2012). The Poti diamictites in Petro-
shale scattered in the cross-bedded sandstone layers. This unit con- bras and Themag wells are very similar to the end-Devonian diamic-
tains several oligomictic, lenticular, sand-supported conglomerate beds tites that occur at the top of the Cabeças Formation. However, the Poti
with angular to subangular clasts of the same nature as those of the diamictites and shales contain well-preserved late Viséan sporomorphs
above-mentioned sandstone beds, and few silty–clayey interbeds. (Andrade and Daemon, 1974; Streel et al., 2012).
Micaceous pink shale and siltstone-laminated beds, as thick as 10 m Several shale beds intercalated with diamictite deposits may be of
(Fig. 28), occur in the middle part of the lower Poti unit and later- glaciomarine origin because some of them contain lonestones made of
sedimentary clasts (Fig. 30 A). Dark-gray diamictite outcrops of the Poti

PR
ally may pinch out completely. This is documented in some sections and
wells around, where these shale beds are missing. Della Fávera and Formation in the Tocantins River valley contain angular to subrounded,
Uliana (1979) described a quartzite boulder on the road between Ga- sand- to cobble-sized clasts (Fig. 30 B).
turiano and Floriano towns (State of Piauí), which is associated with sus- Ojeda and Bembom (1966) recorded sandstone dykes ca. 3 cm
pension deposits of the Poti Formation (Fig. 28). A smaller boulder was thick and about 20 m long in the upper Poti member, which can be in-
also observed by one of the authors (M.V.C) on the same road, within a terpreted as either injected sandstone dykes or fossil ice wedges, filled
suspension deposit. with sand after ice melting in a periglacial environment.
The sedimentary facies and terrestrial megaflora suggest a predom-

D
inantly transitional to marine environment for parts of the upper Poti
Formation in outcrops of the southwestern portion of the basin. The va-
riety of sizes and nature of clasts suggest a high-energy environment
common to fluvial systems. On the other hand, due to the presence of
TE
rafted boulders (Fig. 28), a fluvioglacial environment is more likely
for the section below the diamictites. In this lower section (Fig. 28),
fluvial sandstone with pebble- to boulder-sized clasts and conglomer-
ates may represent outwash deposits caused by glacier outburst floods
(jökulhlaups) during the occasional release of large volumes of water
stored in mountainous source areas. Marine incursions are character-
EC

ized by the presence of storm-influenced marine facies (sandstone with


HCS) in the lowermost part of the Poti Formation (Ponciano and Della
Fávera, 2009). These beds contain sparse, poorly preserved marine
shelly faunas (Edmondia, Nucula, and Lingulidiscina) in the lowest part
of the lower member of the formation (Kegel, 1954; Mesner and
Wooldridge, 1964). The upper portion, above and within the diamic-
RR

tite of the Poti Formation, bears a particularly rich, yet carbonized land
plant megaflora of “Paraca type” (Iannuzzi, 1994 (,; Iannuzzi and Pf-
efferkorn, 2002) and thin coal films (Dolianiti, 1954). These fos-
Fig. 28. Quartzite boulder included in suspension deposits of the Poti Formation (Della
sil plants indicate climate improvement in the interglacial and post-
Fávera and Uliana, 1979). The boulder caused the bending of underlying strata (it ap-
pears to squish down into the bed), interruption of lateral stratification and draping of glacial intervals of the late Viséan. The genera identified include Archae-
sediment over its upper surface. calamites, Nothorhacopteris, Triphyllopteris, Sphenopteridium, and Diploth
CO
UN

Fig. 29. A) Poti Formation consisting of gray diamictite cored in a shallow well with irregular-shaped and angular clasts immersed in a clayey–silty–fine sandy matrix. The clasts are made
of granitic, basic rocks, quartz and sedimentary rocks. B) Massive gray diamictite core from the same well that was drilled in 2006 close to the Tocantins River near Tupiratins Town, State
of Tocantins. Samples courtesy of the Themag Engineering and Management Co.

20
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
PR
Fig. 30. A) Poti Formation. Core made of laminated siltstone and shale interbeds, with outsized sandy lonestones, as old as late Viséan. Photo: Courtesy of the Themag Engineering and
Management Company. B) Dark-gray diamictite outcrop, as old as Middle Mississippian (late Viséan) containing angular to subrounded sand- to cobble-sized clasts in the Tocantins River
Valley. Some voids indicate detached clasts. Photo: M.V.C.

D
TE
EC

Fig. 31. A) Cabeças Formation thin-section of Strunian diamictite from the Parnaiba Basin outcrops close to Pedro Afonso Town, with very angular to subangular corroded quartz grains
(q) with variable particle sizes (1.5–0.1 mm) and rock fragment of chlorite schist (c). Sand grains and sideritic nodules (dark-brown color) are rarely in contact with one another and “float”
in a silty/clayey sideritic-chloritic matrix. B) Poti Formation thin-section (plane polarized light) of a core from well SM 204, drilled by Themag company. The diamictite mainly shows
fine- to coarse-grained, angular and irregular-shaped sand grains “floating” in a silty–clayey sideritic matrix (dark-brown color). Almost all sand grains are corroded and rarely in contact
RR

with one another.

mema. They also occur in Mississippian beds of Africa, India, and Aus- genic nature (Swart and Hiller, 1982; Theron, 1994). The Kom-
tralia (Iannuzzi and Pfefferkorn, 2002). madagga Subgroup consists of the Miller, Swartwaters-poort, Soutkloof,
Andrade (1968, 1972), did not interpret the depositional environ- and Dirkskraal formations. This subgroup intervenes between the up-
ment of the pebbly mudstone, or did he recognize any glacial influence per Waaipoort Formation of the underlying Witteberg Group and the
CO

in the sedimentation of the Poti Formation, in a paper prepared by An- basal Dwyka Group diamictite. The Miller Diamictite occurs discontin-
drade and Daemon (1974). Later, Andrade (written communication, uously and has a sharp paraconformable basal contact with the upper
2006) raised the hypothesis that all diamictites found in the Pimenteira, Waaipoort Formation of the Wittenberg Group. It records the oldest Car-
Cabeças, Longá, and Poti formations could be the result of tectonism. boniferous glacial episode in the Late Paleozoic section (Theron, 1994).
However, the presence of clast striations in diamictites, intraformational The basal portion of the Kommadagga Subgroup, comprising the Miller
folding, and boulders within suspension deposits, suggests glacial and Diamictite, is of Viséan age (Theron, 1994; Evans, 1999). However,
periglacial deposition. Thin-sections of latest Devonian diamictites from the glacial origin for this diamictite is being challanged (see Grobbe-
UN

Parnaíba Basin outcrops (Fig. 31) A) and Middle Mississippian diamic- laar, 2015). The diamictite is massive and does not exhibit stratifi-
tites from a Themag well core (Fig. 31 B) display very similar textures. cation, cross-bedding, grading, or other sedimentary structures, which
The thin-sections show grain shapes characteristic of ice abrasion. There tend to exclude debris flow deposition. Thin-sections and hand samples
are no rounded quartz grains in those thin-sections.Iannuzzi and Pfef- of the Miller Diamictite (Fig. 32 A, B) present quartz composition, with
ferkon (2002) suggested a Late Viséan age for the fossil macroflora of angular fragments similar to those of glacial diamictites of Brazil. Its tex-
the Poti Formation. This flora is dominated by pteridosperms with as- ture is consistent with mechanical erosion and short-distance transporta-
sociated arboreal lycopsids and sphenopsids. In conclusion, pebble stri- tion under very cold climate. Older, dark-gray diamictite fragments im-
ations in diamictites, boulders in suspension deposits, and lonestones mersed in the Miller diamictite are also observed in outcrops (Fig. 32
in laminites suggest continental, periglacial, and glaciomarine environ- B).
ments in the Viséan Poti Formation. The Miller Formation is overlain by the Swartwaterspoort Forma-
The Miller Diamictite Formation (Karoo Basin, South Africa) of the tion sandstone with soft-deformation. Sandstone thickness, at Campher
Kommadagga Subgroup of the Witteberg Group is possibly of glacio

21
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

F
OO
Fig. 32. Miller Diamictite from South Africa, Karoo Basin. A) Thin-section of diamictite, with angular fragments “floating” in a silt-shale matrix, where they are rarely in contact with one
another. B) Several angular clasts, inside the gray massive Miller Diamictite, are mainly made of sedimentary rocks (white color), quartz, and feldspar minerals (pink), older diamictites
(dark-gray fragments) and coal streaks (black color). The diamictite is ungraded (Grobbelaar, 2015).

PR
spoort and Kommadagga Kop, is a few meters, reaching a maximum ian Sahara (Lanzoni and Magloire, 1969), were compared with those
92 m at Kruidfontein (Grobbelaar, 2015). in the Parnaíba Basin (Andrade and Daemon, 1974; Melo and
According to Swart (1982), these sandstone units are discontinuous Loboziak, 2000). The Poti Formation miospore assemblages were later
in the eastern portion of the basin where they are easily recognized by referred to a distincive Gondwanan Viséan biozone, the Reticulatisporites
their coarse grain size, clean white weathering nature, the presence of magnidictyus (Mag) Range Zone (Melo and Loboziak, 2003; Melo and
iron oxide staining, and frequent slump structures. A 0.1 cm–3 cm thick Playford, 2012).
horizontal lamination, defined by iron oxide staining, is usually seen in Coeval Viséan diamictites are also present in the Amazon and

D
the overlying Swartwaterspoort sandstone with soft-sediment deforma- Solimões basins (Caputo et al., 2008; Melo and Playford, 2012). Be-
tion structures as the only sedimentary structure commonly observed. sides, they were newly recognized in the Paraná Basin, underlying lat-
These slump structures mentioned by Swart (1982) were related est Pennsylvanian diamictites of the Itararé Group (Rosa et al., 2019).
to the beginning of the Dwyka glaciation by Bell (1981). This sub- Mid-late Viséan diamictites are present in Argentina and Bolivia (Reyes,
TE
ject was polemic, because there are two non-deformed formations un- 1972), indicating glacial waxings and wanings in South America. Viséan
derlying the Dwyka diamictites. The Dwyka beds rest unconformably diamictites were reliably dated in the upper member of the Cortaderas
on older stratigraphic units, covering different formations of the Kom- Formation, in the Río Blanco Basin of northwestern Argentina. The
madagga Subgroup. A large time span, of about 15 My, is present be- glacial deposits in this basin consist of shales, dropstone-bearing shales,
tween Dwyka and Miller diamictites deposition (Stapleton, 1977). In massive matrix-rich diamictites, massive clast-rich diamictites with stri-
our view, this soft-sediment deformation or slumping structures are re- ated clasts, and stratified diamictites, all of which were deposited in dis-
EC

lated to the end of the Miller glaciation like for the Late Devonian glacia- tal glaciomarine settings (Perez-Loinaze et al., 2010). The Cortaderas
tion in the Parnaíba Basin. These features were interpreted as resulting Formation diamictites are of the same age as those of the Poti Forma-
from tectonics by Grobbelaar (2015), but they may be attributed to the tion (i.e., late Holkerian to Asbian, further confirmed by radiometric dat-
retreat of glaciers as well. Microscopic study of the deformed sandstone ings in Argentina), and both formations share identical palynofloras (see
indicates that all the structures originated definitively prior to lithifica- Melo and Playford, 2012, and references therein).
tion (Bell, 1981). These soft-sediment structures suggest deglaciation at Late Mississippian (Serpukhovian) glaciogenic deposits, documented
RR

the end of the Miller diamictite deposition instead of later tectonism. in other South American countries like Peru, Bolivia, and Argentina,
In many other regions, debris flow deposits were reinterpreted as are yet unproven in Brazil, except possibly in the Paraná Basin, where
glacially influenced deposits and vice-versa. The age of the Miller di- pre-Itararé (pre-Pennsylvanian) diamictites were newly discovered and
amictite is the same as that of the Poti Formation (Middle Mississippian). dated palynologically as middle Viséan–early Serpukhovian (Rosa et
Therefore, more studies are necessary for working out this uncertainty al., 2019).
relative to age and nature of the South African Miller Diamictite. In the Acre Basin, close to Peru and Bolivia, the oldest, Phanerozoic
CO

section observed in outcrops and wells, the Apuí Formation (conglomer-


14. Age of Poti Formation diamictites ates, sandstones, thin shales, and diamictites), correlates with the Mis-
sissippian Ambo Group (Caputo, 2014) of Peru and Bolivia.
Melo and Loboziak (2000) studied well-preserved miospore
species with both southern Euramerican and Gondwanan affinities from
Poti Formation sections of some Petrobras wells in the Parnaíba Basin. 15. Viséan glacial geochemistry
They assigned the Poti Formation to the lower and middle parts of the
Indirect evidence of glaciation, preserved in unconformity-bounded,
UN

upper Viséan. Several age-significant miospore species were identified


in their study, which included a succinct systematic description of the low latitude ramp carbonate sequences, in the Illinois Basin, United
most important species and comments on their biostratigraphy. Based on States, suggests that the geographically extensive continental (and ma-
correlations with the Western European Carboniferous miospore zona- rine) glaciation of Gondwana actually began in the late Viséan (Smith
tion, the Poti Formation palynofloras were compared with the Perotrilites and Read, 2000). The late Viséan age of the Poti Formation diamictites
tessellatus–Schulzospora campyloptera (TC) and Raistrickia nigra–Triqui- correlates with that of incised valleys in the Illinois Basin.
trites marginatus (NM) zonal range (Melo and Loboziak, 2000). This Bruckschen et al. (1999, Fig. 12) reported the atmospheric CO2
characterizes the upper portion of the Holkerian and the whole Asbian decline and cooling in the Early and Middle Mississippian based on sta-
according to Higgs et al. (1988, Fig. 1), which are the British re- ble isotope studies of brachiopod shells in Western Europe and the for-
gional stages for the lower to middle parts of the upper Viséan. Co- mer USSR. This corresponds more precisely to the late middle to early
eval miospores illustrated from the Grand Erg Occidental, in the Alger late Tournaisian and early to middle parts of late Viséan glaciations in
the Parnaíba and other South American basins.

22
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

The 87Sr/86Sr record for the European Carboniferous presents a Venezuela (Caputo et al., 2008). The Devonian beds in this country
middle Viséan minimum and sharp increase during the late Viséan are very deep in the subsurface (Stover, 1967), possibly in part meta-
(Bruckschen et al., 1999). This may indicate a high flux of fluvial sed- morphosed in the Eastern Venezuela Basin Province, and have not been
iments rich in radiogenic Sr derived from continents to the deep-sea due investigated in sufficient detail thus far.
to Gondwanan glaciation and concomitant sea-level fall during the early According to Lakin et al. (2016), the interpretation of the deposi-
late Viséan. tional environment and age of Tournaisian diamictites in the Solimões

F
It should be noted that the late Viséan glacial episode identified Basin cannot be validated. However, figure 26 B herein shows the
physically in South America (Caputo et al., 2008; Fielding et al., stratigraphic relationship of Devonian and Tournaisian diamictites in
2008; Lakin et al., 2016; present study), restricted to the early to mid- some areas of this basin. The abundant miospore assemblages found in

OO
dle parts of late Viséan, is not equivalent to the Vice Event reported cored diamictites of the Parnaíba Basin are clearly representative of the
by Liu et al. (2018), which is early Viséan in age. Rocks of latest Tour- Spalaeotriletes pretiosus-Colasporites decorus (PD) miospore interval Zone,
naisian–early Viséan age are absent in the Parnaíba Basin (Melo and which specifically dates the Longá glacial sediments as late middle to
Loboziak, 2000), and also in the Amazon and Solimões basins (Melo early late Tournaisian (Playford et al., 2012). On the other hand, the
and Loboziak, 2003; Melo and Playford, 2012). Latest Tournaisian occurrence of Mississippian miospores in the higher diamictites of the
or early Viséan fossiliferous strata seem to occur in northwestern Ar- otherwise end-Devonian Itacua Formation of Bolivia suggests probable
gentina (Malimán and El Ratón formations: see discussion in Melo and superposition of latest Famennian and Early Mississippian diamictites,
Playford, 2012); however, no undoubted glacial evidence has been rec- as noted in wells in the Solimões Basin of northern Brazil (Fig. 27 B),

PR
ognized thus far in this particular interval. according to Streel et al. (2012). However, both glacial events are
widely separated in geological time.
16. Discussion Lakin et al. (2016) stated that “Viséan glaciation in Brazil has not
yet been supported by any published measured sections or palynolog-
Until recently, the Silurian glaciation was not known in Brazil, and ical descriptions”. Therefore, “detailed sedimentological and biostrati-
Devonian and/or Mississippian glaciations were unrecognized both in graphic data regarding Viséan diamictites in Brazilian basins are ex-
Brazil (Ludwig, 1964; Mabesoone, 1975, 1977; Lima, 1978), and tremely limited and difficult to validate”. However, Andrade and Dae-

D
elsewhere in the world (; Schermerhorn, 1971; Heckel and Witzke, mon (1974) and Streel et al. (2012) recorded palynomorphs as old
1979; Thompson and Newton, 1989; Brand, 1989; Dickins, 1993; as Viséan in the Poti Formation diamictites of the Parnaíba Basin.
Becker, 1993a, b; House, 1996). Clast striations in diamictites and large lonestone (boulders) in sus-
Thompson and Newton (1989) and Brand (1989) suggested a pension deposits suggest continental, periglacial, and glaciomarine de-
TE
worldwide warming instead of cooling at the end-Frasnian. According positional environments in the late Viséan Poti Formation. In addition,
to Dickins (1993), “the Late Devonian apparently had a warm climate. Smith and Read (2000) suggested that the late Viséan glaciation
The climate of the Early Carboniferous was warm and equable. No cold marks the beginning of the extensive Permocarboniferous continental
climate can be identified, and apparently no ice was present at the poles. and marine glaciation in Gondwana. However, in our opinion, a signifi-
Even a cold temperate climate is difficult to identify”. “In addition, some cant, though intermittent Paleozoic cooling, after the Early Silurian one,
reservation seems necessary on the data presented for the Late Devonian commenced at the Frasnian-Famennian boundary.
EC

in South America and South Africa by Caputo and Crowell (1985) and Tournaisian (Wicander et al., 2011) and mid-late Viséan (Reyes,
Caputo (1985)”. However, in those areas there is a robust evidence for 1972) glacially-influenced diamictites are also present in Argentina and
the latest Devonian glaciation in the international literature since 1979 Bolivia, indicating glacial waxings and wanings in other parts of South
(Carozzi, 1979, 1980; Crowell, 1983; Caputo, 1985; Caputo and America.
Crowell, 1985; Díaz-Martínez, 2004; Caputo et al., 2006, 2008). Based on present data, the glaciation in the Karoo Basin (South
These glacial events have been studied by many researchers ever since. Africa) may have begun in either the end-Devonian or Mississippian
RR

Moura (1938) was the first who interpreted Devonian diamictites (Plumstead, 1964). The upper Witpoort diamictites of the upper Witte-
as deposited under glacial conditions based on the rock texture observed berg Group of the Karoo Basin have a glacigenic origin (Almond et al.,
in borehole cores from the Curuá Formation of the Amazon Basin. Later, 2002). The Witpoort Formation is as old as Late Devonian (Famennian)
a Petrobras Amazon Basin study by Bouman et al. (1960) documented based in palynomorphs and fish (Streel and Theron, 1999).
striated clasts of Late Devonian age in well cores, but Ludwig (1964) Additional direct evidence of Tournaisian and Viséan glaciations/
refuted the glacial nature of these sediments by considering them as tur- deglaciations is likely to occur in Africa. Some references are available,
CO

bidites. Other authors (Caputo and Vasconcelos, 1971; Rodrigues et but the information is still poor due the lack of geological mapping and
al., 1971; Caputo and Vasconcelos, 1971; , ; Carozzi, 1979, 1980) subsurface data. Glacial sediments, at the base of sedimentary sections
reinterpreted the Late Devonian diamictites as a result of glacial activity of the Congo Basin and Zambezi Block (Chappell and Humphreys,
based on rock texture, clast striae, lonestones in rhythmites, and strati- 1970), may have been deposited in the Mississippian (Frakes and
graphic correlations. In spite of its short duration, this glacial episode Crowell, 1970). In South Africa, the Miller Formation (lowest part of
attained widespread distribution in Brazilian sedimentary basins. the Kommadagga Subgroup) may include glacially influenced strata of
The Late Devonian diamictite is very rich in clay, which was prob- Viséan age. The formation contains diamictites with clasts derived from
UN

ably incorporated from the Curuá Formation shale substratum during the underlying Cape Supergroup, and soft-sediment deformation was as-
glacier advances. The abundant presence of latest Famennian marine cribed by Streel and Theron (1999). The study of these authors con-
palynomorphs in the coeval diamictites and rhythmites (Streel et al., firms an age not older than middle Tournaisian based on the presence of
2000b) attests that the glaciation reached the Brazilian Paleozoic in- miospores like Schopfites cf. delicatus and Spelaeotriletes pretiosus in the
tracratonic marine basin margins, and was not restricted to mountain- unit underlying the Miller Diamictite, that is, the Waaipoort Formation.
ous areas. Latest Devonian glaciers are recorded on almost the entire The Tournaisian age of the Waaipoort Formation matches that of the up-
South American continent, from Venezuela to southern Brazil (Caputo per Longá Formation in the Parnaíba Basin (Playford, et al., 2012),
et al., 2008), including also Peru, Bolivia (Díaz-Martínez, 2004), but it has not found yet glacial evidence in this unit.
and possibly central Argentina. Venezuela is on the northern side of The Miller Diamictite corresponds to the Poti Formation, of early
the Guiana Shield, i.e., the main location of glacial centers that fed late Viséan age. Far-field geochemical evidence, discussed in the text for
ice to the northern margins of the Solimões and Amazon basins, and Devonian–Mississippian glaciations, is abundant in Europe and North
possibly also to basins on the northern side of the Guiana Shield in America.

23
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

17. Conclusions Declaration of Competing Interest

There is robust evidence for the deposition of various Paleozoic sedi- The authors declare that they have no known competing financial in-
mentary formations under glacially influenced conditions, based on typ- terests or personal relationships that could have appeared to influence
ical glacial depositional, erosional and deformational sedimentary fea- the work reported in this paper.
tures. They were probably deposited by ice masses fed from multiple up-

F
land glacial centers along the former borders of the Parnaíba Basin. Sed- Acknowledgements
iments generated by the oldest Paleozoic (Early Silurian) glaciation were
deposited in a chiefly terrestrial setting, whereas those of the other three We gratefully acknowledge Petrobras and Themag Engineering and

OO
(Strunian, Tournaisian, and Viséan) were accumulated in glacio-deltaic Management Company for permission to study well cores and pub-
(transitional) and/or glaciomarine environments. The Silurian continen- lish the results. We acknowledge doctors Marivaldo dos Santos Nasci-
tal glaciation in the central portion of the Parnaíba Basin may have mento, Peter Szatmari and Wagner Souza-Lima for the pictures in fig-
shifted to transitional and glaciomarine environments, because there ures where their names are mentioned. We also acknowledge Francisco
was a major marine opening on the western and northwestern sides of de Assis Mattos de Abreu and Wagner Souza-Lima for field assistance,
the Parnaíba Basin (Fig. 1 B). and anonymous journal reviewers. We did not receive any specific grant
Striated pavements in the Santa Brígida Graben and Sergipe–Alagoas from funding agencies in the public, commercial, or not-for-profit sec-
Basin are envisaged here as possibly related to Silurian glaciers. These tors.

PR
pavements were attributed to the Pennsylvanian by several previous au-
thors cited in the text, but the coeval climate was warmer according References
to lithological indicators in neighboring areas (Caputo, 1984). There
are no fossils in the Alagoas State Mulungu diamictites and below and Aguiar, G A, 1971. Revisão geológica da Bacia Paleozoica do Maranhão. In: 25o Congresso
Brasileiro de Geologia. Anais..... SBG, 3, 69–109.
above the mentioned pavements; therefore, their age is an open ques- Albuquerque, O R, Dequech, V, 1950. Contribuição para a geologia do Meio–Norte,
tion. Here, this unfossiliferous unit is inferred to be as old as early Llan- especialmente Piauí e Maranhão, Brasil. In: 2° Cong. Panamer. Eng. Minas e Geol.,
dovery and may be correlated with the upper Ipu diamictites of the Anais…, 3. Brasil, pp. 69–109.

D
Almond, J, Marshall, J, Evans, F, 2002. Latest Devonian and earliest carboniferous glacial
Parnaíba Basin. The Mulungu diamictites overlie the Karapotó Forma- events in South Africa. In: 16th International Sedimentological Congress Abstract
tion (formerly named the Tacaratu Formation), which possibly corre- Volume 2002.
lates with the Ipu Formation of the Parnaíba Basin. Silurian diamictites Anderson, M M, 1961. The Late Paleozoic glaciation of Gondwanaland in relation to West
Africa and to north South America during Carboniferous time: Geol. Mag. R. Free
were found in sedimentary basins and on cratonic areas as well.
TE
Hosp. Lond. Engl. Geological Magazine 44 (Supl.), 231–236.
Glacially influenced deposits of Devonian and Mississippian age were Andrade, S M, 1968. Geologia do Sudeste de Itacajás. Petrobras Internal Report, 293.
accumulated in close association with marine sediments, favouring their Belém, Brasil 38 p..
Andrade, S M, 1972. Geologia do Sudeste de Itacajá, Bacia do Parnaíba (Estado de Goiás)
easy burial, preservation, and dating. Sheared diamictites and deeply
[Doctorate Dissertation]. Universidade de São Paulo, Escola de Engenharia de São
striated pavements indicate that glacial deposition also took place in Carlos, São Paulo 87 p.
marginal outcrop basin areas under a thick and heavy ice cover. The Andrade, S M, Daemon, R F, 1974. Litoestratigrafia e bioestratigrafia do flanco sudeste
da Bacia do Parnaíba (Devoniano e Carbonífero). in: 28° Congr. Bras. de Geol., Porto
latest Devonian (Strunian) glaciation was certainly the most widespread
EC

Alegre, Anais…, v. 1. SBG, pp. 129–137.


of all in the four intracratonic basins discussed here, with well-known Bär, P, Riegel, W, 1974. Les microflores des séries Paleozoïques du Ghana (Africa
records in parts of three modern continents (Africa, South America, and Occidentale) et leurs relations paleofloristiques. Sci. Geol. Strasbourg Bull. 27 (1–2),
North America). 39–58.
Barbosa, R C, de, M, Nogueira, A C R, Domingos, F H G, 2015. Famennian glaciation in the
Late Mississippian (Serpukhovian) glaciogenic deposits, documented eastern side of the Parnaíba Basin, Brazil: Evidence of advance and retreat of glacier
in other South American countries like Peru, Bolivia, and Argentina, in the Cabeças Formation. Braz. J. Geol. 45 (Suppl 1), 13–27.
are yet unproven in Brazil, except possibly in the Paraná Basin, where Barbosa, O, Ramos, J R A, Gomes, F, Hembold, R, 1966. Geologia estrutural e econômica
RR

da área do projeto Araguaia. Dept. Nac. Prod. Min., Div. Geol. Miner., Monogr. 19, 94.
pre-Itararé (pre-Pennsylvanian) diamictites were newly discovered and Assis, A P, Porto, A L, Schmitt, R S, Linol, B, Medeiros, S R, Correa Martins, F, Silva, D
dated palynologically as middle Viséan–early Serpukhovian (Rosa et S, 2019. The Ordovician-Silurian tectono-stratigraphic evolution and paleogeography
al., 2019). of eastern Parnaiba Basin, NE Brazil. J. South Am. Earth Sci. 95 (November).
doi:10.1016/j.jsames.2019.102241.
Younger (latest Pennsylvanian) glaciation/deglaciation ice ages oc- Becker, R T, 1993a. Anoxia, eustatic changes, and Upper Devonian to lowermost
curred in southern Brazil (e.g., Itararé Group in the Paraná Basin) and Carboniferous global ammonoid diversity. In: House, M R (Ed.), The Ammonoidea
southern South America, whereas northern and northeastern Brazilian Environment, Ecology and Evolucionary Change, Vol. 47. Clarendon Press,
Systematics Association Special, Oxford, pp. 115–163.
CO

basins experienced warm climate conditions during most of the Penn- Becker, R T, 1993b. Analysis of ammonoid palaeogeography in relation to the global
sylvanian time. The glaciation continued elsewhere in Gondwana in re- Hangenberg (terminal Devonian) and lower Alum shale (middle Tournaisian) events.
sponse to the migration of glacial centers across Gondwana during the Ann. Soc. Belg. 115 (2), 459–473.
Bell, C M, 1981. Sediment deformation of sandstone related to Dwyka glaciation in South
Paleozoic (Caputo and Crowell, 1985; During the drift of the Gond- Africa. Sedimentology 28, 321–329.
wana Supercontinent, northwestern Gondwana rotated toward the trop- Berger, W H, Vincent, E, 1986. Deep sea carbonates: reading the carbon isotope. Geol.
ics, whereas eastern Gondwana moved toward the South Pole. Rundsch. 75 (1), 249–269.
Beurlen, K, 1965. Observações no Devoniano do Estado do Piauí. Ann. Acad. Bras. de
The glacially influenced beds and pavements of the Santa Fé Group
Ciências. Brasil 37 (1), 61–67.
of the San Francisco Craton need to be studied in greater detail in order
UN

Bouman, Q C, Mesner, J C, Padden, M, 1960. Amazon Basin Study. Petrobras Internal


to accurately determine their geological age. Report, Belém, Brazil, no 454-A, 4 v., 1102 p. Petrobras.
Brand, U, 1989. Global climatic changes during the Devonian-Mississippian: Stable isotope
Additional mapping is required to reliably identify more Paleozoic
biogeochemistry of brachiopods. Palaeogeogr. Palaeoclimatol. Palaeoecol. 75,
glaciogenic strata in other parts of Brazil and elsewhere in Gondwana. 311–329.
Bratton, J F, Berry, W B N, Morrw, J R, 1999. Anoxia pre-dates Frasnian–Famennian
boundary mass extinction horizon in the Great Basin. USA. Palaeogeography,
Palaeoclimatology, Palaeoecology 154, 275–292.
Uncited references Brezinski, D K, Cecil, C B, Skema, R, Stamm, R, 2008. Late Devonian glacial deposits from
eastern United States signal an end of the mid-Paleozoic warm period. Palaeogeogr.
,,,,,,,,,. Palaeoclimatol. Palaeoecol. 268, 143–151.
Brezinski, D K, Cecil, C B, Kertis, C A, 2009. Evidence for long-term climate change
in Upper Devonian strata of the Central Apalachians. Palaeogeogr. Palaeoclimatol.
Palaeoecol. 284, 315–325.

24
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

Brezinski, D K, Cecil, C B, Skema, V W, 2010. Late Devonian glacigenic and associated Clayton, G, Mason, C E, Ettensohn, F R, Lierman, R T, Goodhue, R, 2010. Palynological
facies from the central Appalachian Basin, eastern United States. Geol. Soc. Am. Bull. correlation of a Late Devonian dropstone in Kentucki with diamictite-bearing sections
122 (1–2), 265–281 101130/B26556.1. in Maryland and Pennsylvania. In: Denver Annual Meeting. Colorado. Abstract. Paper
Bruckschen, P, Oesmann, S, Veizer, J, 1999. Isotope stratigraphy of European 275-12. GSA.
Carboniferous: proxy signals for ocean chemistry, climate and tectonics. Chem. Geol. Cohen, K M, Finney, S C, Gibbard, P L, Fan, J X, 2013. Updated the ICS International
161 (1–3), 127–163. Chronostratigraphic Chart. Episodes 36, 199–204.
Buggish, W, 1991. The global Frasnian-Famennian Kelwasser event. Geol. Rundschau 1, Copper, P, 1977. Paleolatitudes in the Devonian of Brazil and the Frasnian-Famennian
49–72. mass extinction. Palaeogeogr. Palaeoclimatol. Palaeoecol. 21, 165–207.

F
Campbell, D F, 1949. Revised Report on the Reconnaissance Geology of the Maranhão Cox, P M, Betts, R A, Jones, C D, Spall, S A, Totterdell, I J, 2001. Modeling vegetation and
Basin. Petrobras Internal Report No 7, Pp 117. the carbon cycle as interactive elements of the climate system.. In: Pearce, R (Ed.),
Campos, J E G, 1992. A Glaciação Permo-Carbonífera nas Regiões de Canabrava e Santa Meteorology at the Millennium. Academic Press, pp. 259–279.
Fé de Minas–MG. Dissertação de Mestrado,. Universidade de Brasília, p. 104. Crow, A T, 1952. The rocks of the Sekondi Series of the Gold Coast ., no 18, 65 p.. Gold

OO
Campos, J E G, Dardenne, M A, 1994. A Glaciação neopaleozoica na porção meridional da Coast Geol. Survey Bull.
Bacia Sanfranciscana. Rev. Bras. Geosci. 24, 65–76. Crowell, J C, 1983. Ice ages recorded on Gondwanan continents. South Afr. J. Geol. 86 (3),
Campos, J E G, Dardenne, M A, 1997. Estratigrafia e sedimentação da Bacia 238–261.
Sanfranciscana: Uma revisão. Revista Brasileira de Geologia 27 (3), 269–282. Cunha, P R C, Gonzaga, F G, Coutinho, L F C, Feijó, F J, 1994. Bacia do Amazonas. Boletim
Dardenne, M A, Campos, L E G, 2003. Glacigenic facies stratigraphy of the Santa Fé de Geociências da Petrobras. Rio de Janeiro 8 (1), 47–55.
Group in the Sanfranciscan Basin. 3rd Latinamerican Congress of Sedimentology. Cunha, P R C, Melo, J H G, Silva, O B, 2007. Bacia do Amazonas. Boletim de Geociências
Belém, Pará, Brazil, Abstracts. In: Rossetti, D F., Truckenbrodt, W., Góes, A. M. Museu da Petrobras 15 (2), 227–251.
Paraense Emilio Goeldi. UFPA, pp. 118–119. Della Fávera, J C, 1990. Tempestitos da Bacia do Parnaíba [Doctorate. Dissertation].
Campos, J E G, Dardenne, M A, 2002. Pavimentos estriados do Grupo Santa Universidade Federal do Rio Grande do Sul, Porto Alegre 2 v, 400 p.
Fé–Neopaleozoico da Bacia Sanfranciscana, MG: Registro de abrasão do Della Fávera, J C, Uliana, M A, 1979. Bacia do Maranhão-facies e ambientes. Petrobras

PR
Neopaleozoico. In: Schobbenhaus, C, Campos, D A, Queiroz, E T, Winge, M, Internal Report (Brazil). Petrobras pp-112.
Berbert-Born, M (Eds.), Sítios Geológicos e Paleontológicos do Brasil, 1. Beca, São De Vleeschouwer, D, Rakocinski, M, Racki, G, Bond, D P G, Sobion, K, Klaeys, P, 2013. The
Paulo, pp. 161–164. astronomical rhythm of Late Devonian climate change (Kowala section, Holy Cross
Caputo, M V, 1984. Stratigraphy, tectonics, paleoclimatology and paleogeography of Mountain. Poland. Earth Planet. Science Letters. 365, 25–37.
northern Basins of Brazil. University of California, Santa Barbara XX + 583 p. Ph.D Díaz-Martínez, E, 2004. La glaciación del Devónico Superior in Sudamérica: Estado del
Thesis. conocimiento y perspectivas. 12° Congreso Peruano de Geología. Resúmenes
Caputo, M V, 1985. Late Devonian glaciation in South America. Palaeogeogr. Extendidos. Sociedad Geológica del Perú.. SGP, pp. 440–443.
Palaeoclimatol. Palaeoecol. 51 (1–4), 291–317. Díaz-Martinez, E, Vavrdová, M, Beck, J, Isaacson, P E, 1999. Late Devonian (Famennian)
Caputo, M V, 1994. Atmospheric CO2 depletion as glaciation and biotic extinction agent: glaciation in Western Gondwana: Evidence from central Andes. In: Feist, R, Talent, J
The Late Devonian-Early Carboniferous glacial examples. Boletim de Resumos A, Daurer, A (Eds.), North Gondwana: Mid Paleozoic terranes, stratigraphy and biota..

D
Expandidos do Simpósio de Geologia da Amazônia,, 4. SBG, pp. 194–197. Abhadlungen Der Geologischen Bundesanstalt. Abh. Geol-B.-A,, pp. 213–237.
Caputo, M V, 1998. Ordovician-Silurian glaciations and global sea-level changes. In: Díaz-Martínez, E, Grahn, Y, 2007. Early Silurian glaciation along the western margin of
Landing, E, Johnson, M E, Hall, J (Eds.), Silurian Cycles - Linkages of Dynamic Gondwana (Peru, Bolívia and northern Argentina): Paleogeographic and Geodynamic
Stratigraphy with Atmospheric, Oceanic and Tectonic Changes, 491. New York State setting. Palaeogeogr. Palaeoclimatol. Palaeoecol. 245, 62–81.
Museum Bulletin, pp. 15–25. Díaz-Martínez, E, Vavrdová, M, Isaacson, M, Grahn, Y, 2011. Early Silurian VS. Late
TE
Caputo, M V, 2014. Juruá Orogeny: Brazil and Andean countries. Braz. J. Geol. 44 (2), Ordovician glaciation in South America. In: Gutiérrez-Marco, J C, Rábano, I,
181–190. García-Bellido, D (Eds.), Ordovician of the World. Cuadernos del Museo Geominero
Caputo, M V, Sad, A R E, 1974. Geologia do baixo Rio Negro e trecho a BR-174. Petrobras 14. Instituto Geológico y Minero de España, Madrid, pp. 127–134.
Internal Report- Sistema de Informação de Exploração 130-5170. Petrobras, Belém, Dickins, J M, 1993. Climate of the Late Devonian to Triassic. Palaeogeogr. Palaeoclimatol.
Brasil, p. 28. Palaeoecol. 100, 213–237.
Caputo, M V, Streel, M, Melo, J H G de, Vaz, L F, 2006a. Glaciações neodevonianas e Dino, R, Antonioli, L, Bergamaschi, S, Rodrigues, M A, da, C, 2002. Palinologia de
eocarboníferas na América do Sul. SBG 43o Congresso Brasileiro de Geologia. Anais depósitos aflorantes da Formação Batinga, Membro Boacica, Carbonífero Superior da
EC

....Simpósio/Sessão Técnica, 2 2, pp. 103. Bacia de Sergipe. 41o Congresso Brasileiro de Geologia, João Pessoa, Paraíba. Boletim
Caputo, M V, Lima, E C, 1984. Estratigrafia, idade e correlação do Grupo Serra de Resumos. SBG p. 659.
Grande-Bacia do Parnaíba. in: SBG 23° Congresso Brasileiro de Geologia. Rio de Dolianiti, E, 1954. A flora do Carbonífero Inferior de Terezina, Piauí, Departamento
Janeiro, Anais… v. 2, pp 740-753. Nacional da Produção Mineral, Divisão de Geologia e Mineralogia. Rio de Janeiro, Bol.
Caputo, M V, Crowell, J, 1985. Migration of glacial centers across Gondwana during 148, 56.
Paleozoic Era. Geol. Soc. Am. Bull. 96 (8), 1020–1036. Evans, F J, 1999. Paleobiology of Early Carboniferous lacustrine biota of the Waaipoort
Caputo, M V, Melo, J H M de, Streel, M, Vaz, L F, 2006b. Late Devonian and Early Formation (Witteberg Group) South Africa. Paleontology Africana 35, 1–6.
Carboniferous glaciations. vol .38. Geological Society of America Abstracts with Fairbridge, R W, 1972. Climatology of a glacial cycle. Quat. Res. 2, 283–302.
Programs 266. Faria, A de, 1982. Formação Vila Maria: Nova unidade litoestratigráfica siluriana da Bacia
RR

Caputo, M V, Ponciano, L C M de O, 2013. Pavimento estriado de Calembre, Brejo do do Paraná. Ciências da Terra - Earth Sci. J. 3, 12–15.
Piauí–Registro de geleiras continentais há 360 milhões de anos no nordeste do Brasil. Fielding, C R, Frank, T D, Isbell, J L, 2008. Resolving the Late Paleozoic ice age in time
V 3, pp. 163-174. (SIGEP 052). Sítios Geológicos e Paleontológicos do Brasil. CPRM, and space. Geological Society of America Special paper 441, 354.
pp. 163–174. Fischer, A G, Arthur, M A, 1977. Secular variations in the pelagic realm. In: Cook, H E,
Caputo, M V, Iannuzzi, R, Fonseca, V M M da, 2005. Bacias Sedimentares Brasileiras. Bacia Enos, P (Eds.), Deep Water Carbonate Environments. SEPM Special Publication, pp.
do Parnaíba. 81. Fundação Paleontológica Phoenix, pp. 1–6. 19–50.
Caputo, M V, Melo, J H G de, Streel, M, Isbell, J L, 2008. Late Devonian and Early Frakes, L A, Crowell, J C, 1970. Late Paleozoic glaciation: Part II, Africa, exclusive of the
Carboniferous glacial records of South America. In: Fielding, C R, Frank, T D, Isbell, Karroo Basin:. GSA Bull. 81, 2261–2286.
CO

J L (Eds.), Resolving the Late Paleozoic Ice Age in Time and Space, 441. Geological Geologic Time Scale, 2008. Commission for the Geological Map of the World.
Society of America Special Paper, Boulder, Colorado, pp. 161–173. Geldsetzer, H H J, Goodfellow, W D, McLaren, D J, Orchard, M J, 1987. Sulfur-isotope
Caputo, M V, Vasconcelos, D N N, 1971. Possibilidades De Hidrocarbonetos No Arco Purus. anomaly associated with the Frasnian-Famennian extinction, Medecine Lake, Alberta,
Petrobras Internal Report, Belém, Brazil. No 644-A, 21 Pp. Canada. Geology 15, 393–396.
Carozzi, A V, 1979. Petroleum exploration of the Paleozoic clastics of the Middle Amazon Góes, A M O, Feijó, F L, 1994. Bacia do Parnaíba. Boletim de Geociências da Petrobras 8
Basin. Brazil. J. Petrol. Geol. 2 (1), 55–74. (1), 57–67.
Carozzi, A V, 1980. Tectonic control and petroleum geology of the Paleozoic clastics of the Gouldey, J C, Saltzman, M R, Young, S A, Kaljo, D, 2010. Strontium and carbon isotope
Maranhão Basin. Brazil. J. Petrol. Geol. 2 (4), 389–410. stratigraphy of Llandovery (Early Silurian): Implications for tectonics and weathering.
Carozzi, A V, Falkenhein, F V H, Carneiro, R G, Esteves, F R, Contreiras, C J A, 1975. Palaeogeogr. Palaeoclimatol. Palaeoecol. 296, 264–275.
UN

Análise ambiental e evolução tectônica sinsedimentar da seção Siluro-Eocarbonífera Grahn, Y, Caputo, M V, 1992. Early Silurian glaciations in Brazil. Palaeogeogr.
da Bacia do Maranhão. Ciência-Técnica-Petróleo, Seção Exploração de Petróleo, Palaeoclimatol. Palaeoecol. 99, 9–15.
Publicação no 7, 89 p. Petrobras. Grahn, Y, Gutíerrez, P R, 2001. Silurian and Mddle Devonian chitinozoa from Zapla and
Carozzi,, A V, Pamplona, H R P, Castro, J C de, Contreiras, C J A, 1973. Ambientes Santa Barbara Ranges, Tarija Basin, Northwestern Argentina. Ameghiniana. Rev. Asoc.
deposicionais e evolução tecto-sedimentar da seção clástica paleozoica da Bacia do Paleontol, Argentina, Buenos Aires 38 (1), 35–50.
Médio Amazonas: An... do XXVII Congresso Brasileiro de Geologia. SBG, Aracaju, Grahn, Y, Loboziak, S, Melo, J H G de, 2001. Integrated miospore-chitinozoan biozonation
Brasil, pp. 279–314. of the Parnaíba Basin and its correlation with Petrobras (Müller, 1962) Silurian-Lower
Carvalho, R R de, Neumann, V H, Fambrini, G L, Assine, M L, Vieira, M M, Rocha, D E G A Carboniferous palynozones. In: Melo, J H G de, Terra, G J S (Eds.), Correlação De
da, Ramos, G M S, 2018. The basal siliciclastic Silurian-Devonian Tacaratu Formation Seqüências Paleozóicas Sul-Americanas. Ciência-Técnica-Petróleo. Seção, Exploração
of the Jatobá Basin: Analysis of facies, provenance and palaeocurrents. J. South Am. De Petróleo, Vol. 20. Petrobras, Rio de Janeiro, pp. 81–89.
Earth Sci. 88, 94–106. Grahn, Y, Melo, J H G de, 2005. Middle and Late Devonian chitinozoa and biostratigraphy
Cecil, C B, 2002. Evidence for Late Devonian and Early Carboniferous global cooling in the of the Parnaíba and Jatobá basins, Northeastern Brazil. Paleotographica Abt. B 272
Appalachian Basin. In: Abstract. GSA Denver annual Meeting. 219 (7). Lfg 1-6,p. 1-50.
Chappell, J, Humphreys, M, 1970. Glacial sedimentation in the Lower Karoo, Mid-Zambezi Grahn, Y, Melo, J H G de, Loboziak, S, 2006. Integrated Middle and Late Devonian
Valley, Rhodesia. In: Haughton, H S (Ed.), Gondwana Palaeontology and Stratigraphy. miospore and chitinozoan of the Parnaíba Basin, Brazil: An update. Revista Brasileira
CSIR, Pretoria, pp. 501–510. de Paleontologia 9 (3), 81–89.

25
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

Grahn, Y, Pereira, E, Borgamashi, S, 2000. Silurian and Lower Devonian chitinozoan Lima, E de A M, 1978. Relacionamento geológico dos supostos diamictitos glaciais com
biostratigraphy of the Paraná Basin in Brazil and Paraguay. Palynology 24, 147–176. falhamentos sindeposicionais da Bacia do Meio Norte. An… Congr. Bras. de Geol, 30°.
Grahn, Y, Melo, J H G de, Steemans, P, 2005. Integrated Chitinozoan and miospore Recife, Bol. 1, Resumos das Comunicações. SBG, 153.
zonation of the Serra Grande Group (Silurian-Lower Devonian), Parnaíba Basin, Lisboa, M A R, 1914. The Permian geology of northern Brazil. American Journal of
Northeast Brazil. Revista Española de Micropaleontologia. Instituto Geológico y Sciences 221, 425–442 New Haven, Series 4, v. XXXVII (W. N. CLXXVII), no.
Minero de Espanã, pp. 183–204 37 (2). Liu, J, Algeo, T J, Qie, W, Saltzman, M T, 2018. Intensified oceanic circulation during
Grobbelaar, M, 2015. A comparison between diamictites at the Wittenberg-Dwyka contact Early Carboniferous cooling events: evidence from carbon and nitrogen isotopes.
in Southern South Africa. Thesis (MSc). Stellenbosch University, Faculty of Science. Palaeogeogr. Palaeoclimatol. Palaeoecol.. doi:10.1016/j.palaeo.2018.10.021.

F
Dept of Earth Science, South Africa 141 p. Lobato, G, 2007. Estratigrafia de alta resolução no intervalo do limite formacional Longá/
Hambrey, M J, Kluyver, H M, 1981. Evidence of Devonian or Early Carboniferous Poti (Neodevoniano/Eocarbonífero) em testemunhos de sondagem da Bacia do
glaciation in the Agades region of Niger. In: Hambrey, M J, Harland, W B (Eds.), Parnaíba, Monografia de graduação, 111p. UFRJ, Rio de Janeiro.
Earth’s pre-Pleistocene Glacial Record. Cambridge Univ. Press, pp. 188–190. Lobato, G, 2010. Estratigrafia de alta resolução do intervalo Neodevoniano-Mississipiano

OO
Heckel, P H, Witzke, B J, 1979. Devonian world palaeogeography determined from da Bacia do Parnaíba em testemunhos de sondagem. Dissertação de Mestrado
distribution of carbonates and related lithic palaeoclimatic indicators. In: The (Geologia). Rio de Janeiro. UFRJ, 97 p.
Devonian System: Special Papers in Palaeontology, 23. Univ. Press, Oxford, England, Lobato, G, Borghi, L, 2007. Análise estratigráfica de alta resolução do limite formacional
pp. 99–123. Longá/Poti - em testemunhos de sondagem da Bacia do Parnaíba.. 4° Congresso
Higgs, K, Clayton, G, Keegan, J B, 1988. Stratigraphic and systematic palynology of the Brasileiro de Pesquisa e Desenvolvimento em Petróleo e Gás (4o PDPETRO). ABGP,
Tournaisian rocks of Ireland. Geol. Surv. Irel. (Special paper) 7, 1–93. Campinas São Paulo, pp. 1–10.
Holser, W T, 1977. Catastrophic chemical events in the history of the ocean. Nature 267, Lobato, G, Borghi, L, 2014. Estratigrafia de sequência do contato formacional Longá/Poti
403–408. doi:10.1038/267403a0. (Carbonífero Inferior) em testemunhos de sondagens da Bacia do Parnaíba. Boletim de
House, M R, 1975. In: Facies and Time in Devonian Tropical Areas, 40. Proc. Yorks, Geociências da Petrobras. 22, 213–235.
Geological Society, pp. 233–288. Loboziak, S, Melo, J H G, Quadros, L P, Daemon, R F, Barrilari, I M R, 1994.

PR
House, M R, 1996. An Eocanites fauna from the Early Carboniferous of Chile and its Biocronoestratigrafia dos palinomorfos do Devoniano Médio-Carbonífero Inferior das
palaeogeographic implications. Ann. Soc. Geol. Belg. 117 (1), 95–105. bacias do Solimões e Parnaíba: Estado da arte. Sintex II—2o Seminário de
Iannuzzi, R, 1994. Reavaliação da flora carbonífera da Formação Poti, Bacia do Parnaíba. Interpretação Exploratória, Rio de Janeiro, Atas. Petrobras, pp. 51–56.
Tese (Mestrado). USP, São Paulo, 233 p. Loboziak, S, Caputo, M V, Melo, J H G de, 2000. Middle Devonian–Tournaisian miospore
Iannuzzi, R, Pfefferkorn, H W, 2002. A pre-glacial, warm-temperate floral belt in biostratigraphy in the southwestern outcrop belt of the Parnaíba Basin, north-central
Gondwana (Late Viséan, Early Carboniferous). Palaios 17, 571–599. Brazil. Rev. Micropaleontol. 43 (4), 301–318. doi:10.1016/S0035-1598(00)90154-5.
Lanzoni, E, Magloire, L, 1969. Associations palinologiques e leurs "applications Loboziak, S, Melo, J H G de, Quadros, L P, Daemon, R F, Barrilari, I M R, Streel, M, 1994.
stratigraphiques dans le Devonien Superieur et Carbonifère du Grand Erg Occidental Devonian-Dinantian miospore biostratigraphy of the Solimões and Parnaíba basins
(Sahara Algerien)’’. Revue de L`Institut Français du Pétrole 24 (4), 441–469. (with considerations on the Devonian of the Paraná Basin). Petrobras/Cenpes, 2 Vols.
Isaacson, P E, Hladil, J, Jian-Wei, S, Kalvoda, J, Grader, G, 1999. Late Devonian (Petrobras Internal Report). Rio de Janeiro. Petrobras.

D
(famennian) glaciation in South America and marine offlap in other continents. In: Ludwig, G, 1964. Nova divisão estratigráfico-faciológica do Paleozóico da Bacia
Feist, R, Talent, J A, Daurer, A (Eds.), Abhandlungen Der Geologischen Bundesanstalt, Amazônica. Petrobras/Cenap. Brazil. Monografia I..
54. pp. 239–257. Lüning, S, Craig, J, Loydell, D K, Storch, P, Fitches, B, 2000. Lower Silurian “hot shales”
Isaacson, P E, Hladil, J, Shen, J-W, Kalvoda, J, Díaz-Martínez, E, Grader, G, 1999 b. in North Africa and Arabia: Regional distribution and depositional model. Earth. Rev.
Late Devonian glaciation in Gondwana: Setting the stage for Carboniferous eustasy. 49, 121–200.
TE
Subcomission on Devonian stratigraphy. Newsletter 16, 37–46. Lüning, S, Loydell, D K, Storch, P, Shahin, Y, Craig, J, 2006. Origin, sequence stratigraphy
Junner, N R, 1939. Sekondian System:. Report Gold Coast Geo Survey 13–15 1938-39,. and depositional environment of an Upper Ordovician (Hirnantian) deglacial black
Kaiser, S I, Steuber, T, Becker, R T, Joachimski, M M, 2006. Geochemical evidence for a shale, Jordan - Discussion. Palaeogeogr. Palaeoclimatol. Palaeoecol. 230, 352–355.
major environmental change at the Devonian-Carboniferous boundary in the Carnic Mabesoone, J M, 1975. Desenvolvimento paleoclimático do Nordeste brasileiro. Boletim
Alps and Rhenish Massif. Palaeogeogr. Palaeoclimatol. Palaeoecol. 240, 146–160. Núcleo Nordeste. Soc. Brasil, Geol. 5, 75–93.
Kaiser, S I, Becker, R T, Steuber, T, Aboussalan, S Z, 2011. Climate-controlled mass Mabesoone, J M, 1977. Paleozoic-Mesozoic deposits of the Piauí-Maranhão Syneclise
extinction, facies and sea-level changes around the Devonian-Carboniferous boundary (Brazil): Geological history of a sedimentary basin. Sedim. Geol. 19, 7–38.
EC

in the eastern Anti-Atlas (SE Morocco). Palaeogeogr. Palaeoclimatol. Palaeoecol. 310, Malzahn, E, 1957. Devonisches glazial im staate Piauí (Brasilien), ein neuer Beitrag zur
340–364. Eiszeit des Devons. Beihefte zum Geologischen Jahrbuch 25, 1–30.
Kaiser, S I, Aretz, M, Becker, R T, 2015. The global Hangenberg Crisis (Devonian- Maziani, N, Higgs, K, Streel, M, 1999. Revision of the late Famennian zonation scheme in
Carboniferous transition): Review of a first-order mass extinction. In: Becker, R T, eastern Belgium. Journal of Micropaleontology. 18, 17–25.
Königshof, P, Brett, C E (Eds.), Devonian Climate, Sea Level and Evolucionary Events, Matchen, D L, Kammer, T W, 2006. Incised valley fill interpretation for Mississippian Black
423. Geological Society, London, pp. 387–438 Special Publications. Hand Formation sandstone, Appalachian Basin, USA. Implications for glacial eustasy
Kaljo, D, Martma, T, 2000. Carbon isotopic composition of Llandovery rocks (East Baltic at Kinderhookian-Osagean (Tn2-Tn3) boundary: Sedimentary Geology 191, 89–113.
Silurian) with environmental interpretation. Proceedings of Estonian Academy of doi:10.1016/j.sedgeo.2006.02.002.
RR

Sciences. Geology 49 (4), 267–283. Melo, J H G de, Loboziak, S, 2000. Viséan miospore biostratigraphy and correlation of
Kammer, T M, Matchen, D L, 2008. Evidence for eustasy at the Kinderhookian-Osagean the Poti Formation (Parnaíba Basin, northern Brazil). Rev. Palaeobot. Palynol. 112,
(Mississippian) boundary in the United States: Response to the Late Tournaisian 147–165. doi:10.1016/S0034-6667(00)00043-9.
glaciation? In: Fielding, C R, Frank, T D, Isbell, J L (Eds.), Resolving the Late Paleozoic Melo, J H G de, Loboziak, S, 2001. New miospore zonation of Devonian - Early
Ice Age in Time and Space, 441. Geological Society of America Special Paper, Boulder, Carboniferous strata in the Amazon Basin: A preliminary account. In: MELO, J.H.G.
CO, pp. 261–274. de & TERRA, G.J.S. (Eds.) Ciência Técnica do Petróleo. Seção: Exploração do Petróleo,
Kegel, W, 1953. Contribuição para o estudo do Devoniano da Bacia do Parnaíba. Dept. Correlação de Seqüências Paleozóicas Sul-americanas. Rio de Janeiro, 20:99-107.
Nac. Prod. Min., Div. de Geol. e Min., Rio de Janeiro, Brazil, Boletim 135, 38. Petrobras. Ciência Técnica do Petróleo 1, 99–107.
Kegel, W, 1954. Lamelibrânquios da Formação Poti (Carbonífero Inferior) do Piauí. Dept. Melo, J H G de, Loboziak, S, 2003. Devonian–Early Carboniferous miospore
CO

Nac. Prod. Min., Div. Geol. Min., Notas Preliminares e Estudos, Rio de Janeiro. 88, 14. biostratigraphy of the Amazon Basin, northern Brazil. Rev. Palaeobot. Palynol. 124,
Klemme, H D, 1980. Petroleum basins – classification and characteristics. J. Petr. Geol. 3 131–202. doi:10.1016/S0034-6667(02)00184-7.
(2), 187–2007. Melo, J H G de, Playford, G, 2012. Miospore palynology and biostratigraphy of
Kiipli, E, Kiipli, T, Kallaste, T, 2004. Bioproductivity rise in the East Baltic sea in the Mississippian strata of the Amazon Basin, northern Brazil. Part Two. AASP
Aeronian (Early Silurian). Palaeogeogr. Palaeoclimatol. Palaeoecol. 205, 255–272. Contribution Series 47, 93–201.
Krebs, W, 1974. Devonian carbonate complexes of Central Europe. In: Laporte, L F (Ed.), Mesner, J C, Wooldridge, L C P, 1964. Maranhão Paleozoic Basin and Cretaceous coastal
Reefs in Time and Space., 18. SEPM Spec. Publ., pp. 155–208. basins, north Brazil. Am. Assoc. Petr. Geol. Bull. 48 (9), 1475–1512.
Lakin, J A, Marshall, J E A, Troth, I, Harding, I C, 2016. Greenhouse to icehouse: A Metelo, C M B S, 1999. Caracterização estratigráfica do Grupo Serra Grande (Siluriano) na
biostratigraphic review of latest Devonian-Mississippian glaciations and their global borda sudeste da Bacia do Parnaíba. Instituto de Geociências M.S thesis. UFRJ, 103 p.
UN

effects. In: Becker, R T, Königshof, P, Brett, C E (Eds.), Devonian Climate, Sea Level Mii, H S, Grossman, E L, Yancey, T E, 1999. Carboniferous isotope stratigraphy of
and Evolutionary Events., 423. Special Publications Geological Society, London, pp. North America: Implications for Carboniferous paleoceanography and Mississippian
439–464. glaciation. GSA Bull. 111, 960–973.
Le Hérissé, A, Melo, J H G, Quadros, L P, Grahn, Y, Steemans, P, 2001. Palynological Moore, B, 1963. Geological reconnaissance of the south-west corner of the Maranhão
characterization and dating of the Tianguá Formation, Serra Grande Group, northern Basin. Petrobras Internal Report, 210. Petrobras, 44 p.
Brazil. In: Melo, J H G, Terra, G J S (Eds.), Correlação de Seqüências Paleozóicas Moura, P de, 1938. Geologia do Baixo Amazonas. Serviço Geológico e Mineralógico do
Sul-americanas., 20. Ciência-Técnica-Petróleo. Seção, Exploração de Petróleo, pp. Brasil, Boletin no 91. Rio de Janeiro, Brasil, 94 p. SGMB.
25–41. Myrrow, P M, Ramezani, J, Hanson, A E, Bowring, S A, Racki, G, Rakocinski, M, 2013.
Le Heron, D P, Meinhold, G, Page, A, Whithan, A, 2013. Did lingering ice sheets moderate High precision U-Pb age duration of the Latest Devonian (Famennian) Hangenberg
anoxia in the Early Paleozoic of Libya. Journal of the Geological Society, London 170, Event and its implications. Terra Nova 0, 1-8.. John Wiley & Sons LTD doi: 10 1111/
327–339. doi:10.1144/jgs2012-108. ter-12090.
Le Heron, D P, Tofaif, S, Melvin, J , 2018. The Early Paleozoic glacial deposits of Nahon, O, Trompette, R, 1982. Origin of Siltstones: Glacial Grinding Versus Weathering:
Gondwana: Overview, chronology, and controversies. Past Glacial environments. Sedimentology, v. 29, no. 1, p. 25-35.
Chapter 3. In: Menzies, J , Meer, J (Eds.). Elsevier Ltd.

26
M.V. Caputo and R.OliverBrasil dos Santos Earth-Science Reviews xxx (xxxx) xxx-xxx

Ojeda, H Y O, Bembom, F da C, 1966. Mapeamento geológico em Semi-detalhe do sudoeste the Wyoming–Lackawanna Valley and its mountain rim, northeastern Pennsylvania. 62nd
de Riachão: Petrobras Internal Report no 260, 74 p. Petrobras. Ann. Conf. PA Geol. PA. Geol. Surv, pp. 34–131.
Ojeda, H Y O, Perillo, I, 1967. Geologia do sudoeste de Carolina. Petrobras Internal Report Small, H L, 1914. In: Geologia e suprimento d’água subterrânea no Piauí e parte do Ceará.
no 269, 86 P. Petrobras. Instituto e Obras Contra Secas, Série I.D. Brasil Publicação no 32, Rio de Janeiro, 146
Oliveira, M A M de, 1961. Reconhecimento geológico no flanco oeste da Bacia do p.
Maranhão. Petrobras Internal Report no 171, 79 p. Petrobras. Smith, A G, Hurley, A M, Briden, J C, 1981. Phanerozoic Paleocontinental World Maps.
Oliveira, M J de, 1997. Caracterização faciológica de sedimentos glaciais da Formação Cambridge Univ. Press 102 p.
Cabeças na borda sudoeste da Bacia do Parnaíba. [M.S. thesis]. Universidade Federal Smith, L B, Read, J F, 2000. Rapid onset of Late Paleozoic glaciation on Gondwana.

F
do Pará, Centro de Geociências, Belém 142 p. Evidence from Upper Mississippian strata of Midcontinent, United States. Geology 28,
Paiva, G de, Miranda, J, 1937. Carvão mineral do Piauí: Serviço Geológico e Mineralógico 279–282.
do Brasil. Fomento Produção Mineral, Bol. no 20, Rio de Janeiro. Brasil. Souza-Lima, W, Farias, R M, 2006. Litoestratigrafia e evolução tectono-sedimentar da
Perez-Loinaze, V S, Limarino, C O, Césari, S N, 2010. Glacial events in Carboniferous Bacia de Sergipe-Alagoas – O Paleozoico: A provável sequência siluro-devoniana.

OO
sequences from Paganzo and Río Blanco basins (northwest Argentina): Palynology and Phoenix 94 (8), 1–4.
depositional setting. Geologica Acta 8 (4), 399–418. doi:10.1344/105.000001579. Souza-Lima, W, 2006. Litoestratigrafia e evolução tectono-sedimentar da Bacia de
Perillo, I, Nahass, S, 1968. Semi-detalhe do sudeste de Pedro Afonso. Petrobras Internal Sergipe-Alagoas. O Paleozoico: A provável seqüência siluro-devoniana. O Paleozoico:
Report no 320, 40 p. Petrobras. A seqüência Carbonífera (II). Phoenix 94 (8), 1–6.
Playford, G, Borghi, L, Lobato, G, Melo, J H G de, 2012. Palynological dating and Souza-Lima, W, Borba, C, Rancan, C C, Cangussu, L P, Costa, M N C, Santos, M R
correlation of Early Mississippian diamictite sections, Parnaíba Basin, northeastern F M, Ribas, N, Pierini, C, Bezerra, C P V, 2014. Formação Karapotó, uma nova
Brazil. Revista Española de Micropaleontologia, Instituto Geológico y Minero de unidade estratigráfica paleozoica na Bacia Sergipe-Alagoas. Boletim de Geociências da
España 44 (1–3), 1–22. Petrobras. 22 (1), 83–112.
Plummer, F D, 1948. Report on Maranhão-Piauí (Geosyncline) in relatório de 1946 do Stapleton, R P, 1977. Carboniferous unconformity in the southern África. Pollen et Spores
Conselho Nacional do Petróleo. Brasil, pp. 87–134. 19 (3), 427–440.

PR
Plumstead, E P, 1964. Gondwana Floras, Geochronology and Glaciation in South Africa: Storch, P, Frýda, J, 2012. Late Aeronian graptolite Sedgwickii Event, associated carbon
22nd International Congress Geology, India, Pt. 9, Proc. Sec. 9, P. 303-319. isotope excursions and facies changes in the Prague Synform (Barrandian, Bohemia).
Ponciano, L C M O, Della Fávera, J C, 2009. Flood-dominated fluvio-deltaic system: A new Geol. Mag. 149, 1089–1106.
depositional model for the Devonian Cabeças Formation, Parnaíba Basin, Piauí. Brazil. Stover, L E, 1967. Palynological dating of the Carrizal Formation of eastern Venezuela.
Anais Academia Brasileira de Ciências 81 (4), 769–780. Associación Venezolana de Geólogos Mineros y Petroleros Boletin; Instituto de
Reyes, F C, 1972. On the Carboniferous and Permian of Bolivia and north-western Estudios de Poblacion y Desarrollo [Dominican Republic] 10, 288–301.
Argentina. Anais da Academia Brasileira de Ciências. in Simpósio Internacional sobre Streel, M, Loboziak, S, 1996. Middle and Upper Devonian miospores. Chapter 18 B. In:
os Sistemas Carbonífero e Permiano na América do Sul 44 (v. 44 Supl.,), 261–277. Jansonius, J, McGregor, D (Eds.), Palynology: Principles and applications, 2. AASP.
Rocha-Campos, A C, 1981. Late Paleozoic tillites of the Sergipe-Alagoas Basin, Rondônia Found.,, pp. 575–587.
and Mato Grosso, Brazil. In: Hambrey, M J, Harland, W B (Eds.), Earth´s Streel, M, Theron, J N, 1999. The Devonian-Carboniferous boundary in South Africa and

D
pre-Pleistocene Glacial Record.. Cambridge University Press, pp. 838–841. age of the earliest episode of the Dwyka glaciation. Episodes 22 (1), 41–44.
Rocha-Campos, A C, Santos, P R dos, Salvetti, R A P, 2003a. Iceberg scours and associated Streel, M, Caputo, M V, Loboziak, S, Melo, J H G, 2000a. Late Frasnian-Famennian
subglacial furrows and striae on sandstones of the Late Paleozoic Curituba Formation, climates based on palynomorph analyses and the question of Late Devonian
NE Brazil. Latinamerican Congress of Sedimentology, 3, Belém, Pará. Museu Paraense glaciations. Earth. Rev. 52, 121–173.
Emilio Goeldi. Universidade Federal do Pará. Abstract Book, pp. 142–144.
TE
Streel, M, Caputo, M V, Loboziak, S, Melo, J H G, Thorez, J, 2000b. Palynology and
Rocha-Campos, A C, dos Santos, P R, Canuto, J R, 2008. Late Paleozoic glacial deposits sedimentology of laminites from latest Famennian of the Parnaíba Basin. Brazil.
of Brazil: Paraná Basin. In: Fielding, C R, Frank, T D, Isbell, J L (Eds.), Resolving the Geologica Belgica 3 (1–2), 87–96.
Late Paleozoic Ice Age in Time and Space, 441. Geological Society of America Special Streel, M, Caputo, M V, Melo, J H G de, Perez-Leyton, M, 2012. What do latest Famennian
Paper, Boulder, Colorado, pp. 97–114. and Mississippian miospores from South American diamictites tell us? Palaeobio
Rocha-Campos, A C, Santos, P R dos, Tomio, A, Salvetti, R A P, 2003b. Alternative Palaeoenv. 93, 299–316.
paleogeography of Late Paleozoic glacial rocks of NW State of Minas Gerais. Swart, R, 1982. The Stratigraphy and sedimentology of the Kommadagga Subgroup and
Latinamerican Congress of Sedimentology, 3, Belém, Pará. Museu Paraense Emilio contiguous rocks.. Rhodes Univ, Grahamstown, South Africa 120 pp. M.Sc thesis
EC

Goeldi. Universidade Federal do Pará, Abstract Book , pp. 144–146. (unpubl).


Rodrigues, R, 1967. Estudo sedimentológico e estratigráfico dos depósitos Silurianos e Swart, R, Hiller, N, 1982. Soft sediment deformation of sandstone related to the Dwyka
Devonianos da Bacia do Parnaíba: Petrobras Internal Report no 273, 63 p. Petrobras, glaciation in South Africa. Sedimentology 29, 749–751.
Brasil. The ocks of the Sekondi Series of the Gold Coast, 1952. Gold Coast geology Survey Bull.
Rodrigues, R, Vasconcelos, D N N, Caputo, M V, 1971. Sedimentologia das formações 18, 65 p.
pré-pensilvanianas da Bacia do Amazonas. Petrobras internal report 643A. Sistema de Theron, J N, 1994. The Devonian-Carboniferous boundary in South Africa. Annales de la
Informação em Exploração – SIEX- 130-4050). Petrobras 91 p. Societé Géologique de Belgique. T. 116 (fascicule 2), 291–300 1993.
Rosa, E L M, Vesely, F F, Isbell, J L, Kipper, F, Fedorchuk, N D, Souza, P A, 2019. Thompson, J B, Newton, C R, 1989. Late Devonian mass extinction: Episodic climatic
Constraining the timing, kinematics and cyclicity of Mississippian-Early
RR

cooling or warming? In: McMillan, N J, Embry, A F, Glass, D J (Eds.), Devonian of the


Pennsylvanian glaciations in the Paraná Basin, Brazil. In: NP Griffis, R Mundil (Eds.). World, 14. Canadian Society of Petroleum Geologists, Memoir, pp. 29–34.
A new time scale constraining the timing of the Late Paleozoic ice age. Sedimentary Torsvik, T H, Cocks, L R M, 2013. Gondwana from top to base in space and time.
Geology, pp. 29–49. Gondwana Res. 24 999-103.
Saltzman, B, 1985. Carbon dioxide and the δ18C record of Late Quaternary climate change. Twenhofel, W H, 1932. Treatise on Sedimentation. Williams & Wilkins Co, 661 p. Illust.
Clim. Dynam. 1, 77–86. Plates, Diagrs.
Saltzman, M R, 2002. Carbon and oxygen stratigraphy of the Lower Mississippian Uesugui, N, Santos, A, 1966. Distribuição dos esporomorfos do Permo-Carbonífero da
(Kinderhookian-Osagean), western United States: Implications for sea water chemistry Bacia de Alagoas-Sergipe. Maceió, Relatório Interno da Petrobras 27. Petrobras.
and glaciation. Geol. Soc. Am. Bull. 114 (1), 96–108. Valsardieu, C, Dars, R, 1971. Presence de Moraines sur la Bordure occidentale de L’aïr a
CO

Santos, M E C M, Carvalho, M S S, 2004. Paleontologia das bacias do Parnaíba, Grajaú e la limite entre le Devonien et le Carbonifère. Region d’Agadès (Republique do Niger):
São Luís – Reconstituições Paleobiológicas. Rio de Janeiro, CPRM 212 p. Bull. Serv. Geol. Als. Lorr. 24 (4), 269–276 Strasbourg.
Santos, M E C M, Carvalho, M S S, 2009. Paleontologia das bacias do Parnaíba, Grajaú e Vaz, P T, Rezende, N de A G das M, Wanderley Filho, J R, Travassos, W A S, 2007. Bacia
São Luís: Reconstituições Paleobiológicas. Serviço Geológico do Brasil CPRM, Rio de do Parnaíba. Boletim de Geociências da Petrobras 15 (2), 253–264.
Janeiro 215 p. Viviani, J B, Rocha-Campos, A C, Ernesto, M, 2000. Late Paleozoic glacial sedimentation
Schermerhorn, L L G, 1971. Upper Ordovician glaciation in northwest África? Discussion: in northeastern Brazil: New results. An. da Acad. Bras. Ciências 72 (4), 598.
GSA Bull. 82, 265–268. Wicander, R, Clayton, G, Marshal, J E A, Troth, I, Racey, A, 2011. Was the latest Devonian
Semtner, A K, Klitzsch, E, 1994. Early Paleozoic paleogeography of the northern glaciation a multiple event? New palynological evidence from Bolivia. Palaeogeogr.
Gondwana margin: New evidence for Ordovician-Silurian glaciation. Geol. Rundsch. Palaeoclimatol. Palaeoecol. 305, 75–83.
UN

83, 743–751. Wilde, P, Berry, W B N, 1984. Desestabilization of oceanic density structure and its
Sevon, W D, 1969. The Pocono Formation in northeastern Pennsylvania. In: 34th Ann. significance to marine extinction events. Palaeogeogr. Palaeoclimatol. Palaeoecol. 48,
Conf. Pennsylvania Geol. Pennsylvania Geological Survey, 129 p. 143–162.
Sevon, W D, 1979. Polymitic diamitite in the Spechty Kopf and Rockwell formations. In: Woodworth-Lynas, C M T, Dowdeswell, J A, 1994. In: Deynoux, M, Miller, J M G,
Devonian shales in south-central Pennsylvania and Maryland. 61–66. Dommack, E W, Eyles, N, Fairchild, I J, Young, G M (Eds.), Soft-sediment striated
Sevon, W D, Woodrow, D L, Costolnick, D E, Richardson, J B, Attrep, JrM, 1997. surfaces and massive diamicton facies produced by floating Ice. In: Earth´s Glacial
Convulsive geologic events and the origin of diamictite in the Spechty Kopf Formation Record. Cambridge University Press, pp. 241–258.
in northeastern Pennsylvania. In: Inners, J D (Ed.), Geology of

1
Note: For space reasons the reader should use Google search, to locate towns, coun-
ties, villages or municipalities mentioned in the text.

27

View publication stats

You might also like