You are on page 1of 18

Marine and Petroleum Geology 126 (2021) 104917

Contents lists available at ScienceDirect

Marine and Petroleum Geology


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

Changes in depositional paleoenvironment of black shales in the Permian


Irati Formation (Paraná Basin, Brazil): Geochemical evidence and
aromatic biomarkers
Caroline Adolphsson do Nascimento a, Eliane Soares de Souza a, *, Laercio Lopes Martins a,
Hélio Jorge Portugal Severiano Ribeiro a, Victor Hugo Santos a, René Rodrigues b
a
Laboratory of Petroleum Engineering and Exploration (LENEP), North Fluminense State University (UENF). Rodovia Amaral Peixoto, Km 163, Avenida Brennand,
Imboassica. Macaé, Rio de Janeiro, 27925-535, Brazil
b
Department of Stratigraphy and Paleontology, Faculty of Geology, Rio de Janeiro State University (UERJ). Rua São Francisco Xavier 524, Maracanã, Rio de Janeiro,
Rio de Janeiro, 20550-013, Brazil

A R T I C L E I N F O A B S T R A C T

Keywords: Physicochemical changes that occurred during deposition of Irati black shales were investigated via outcrop
Irati Formation samples collected in the northeastern and central-eastern Paraná Basin. Various traditional geochemical in­
Source rock dicators were used, in addition to newly detected biomarkers: C32 lanostane, simonellite, sempervirane, and
Lanostane
C32–C35 benzohopanes. Bulk geochemistry parameters, mineralogy, and conventional saturated biomarkers, such
Multivariate statistical analysis
Simonellite
as tetracyclic polyprenoids (TPPs), reflect restricted and hypersaline marine deposition with freshwater input
Sempervirane and increased terrigenous contribution toward the top of the studied outcrops. This led to stratification of the
Benzohopanes water column, establishing euxinic conditions in the photic zone and reducing conditions at the bottom of the
Aryl isoprenoids water column, which enhanced preservation of organic matter. Principal component and hierarchical cluster
analyses were performed to group the samples according to their similar geochemical characteristics: prevailing
marine environment (basal outcrop samples) and higher freshwater inflow with greater terrigenous contribution
(top outcrop samples). These geochemical observations were corroborated by aromatic biomarkers such as tri­
aromatic steroids, trimethylnaphthalenes, simonellite, and sempervirane. The aryl isoprenoid ratio (AIR) in­
dicates persistence of photic zone euxinia (PZE) during Irati black shale deposition. The presence of C32 lanostane
and C32–C35 benzohopanes is directly related to the hypersaline depositional environment of the analyzed
samples. Finally, the C32-33/C32-35 benzohopane ratio is proposed as a thermal maturity proxy because variations
in the ratio can be linked to the presence of a diabase sill and temperature catalyst minerals near one of the
outcrops.

1. Introduction with type I/II organic matter content reaching over 20 wt% (Mello et al.,
1993; Reis et al., 2018). Although these black shales have been the
The intracratonic Paraná Basin is situated in the central-eastern subject of several studies, there is still disagreement regarding the
portion of South America, and its evolutionary history consists of cy­ physicochemical conditions of the paleoenvironment during their
cles of accelerated subsidence induced by orogenic events (Zalán et al., deposition. While many works have proposed a normal to hypersaline
1990). As a consequence of these cycles, marine incursions into the marine environment based on faciological analyses, fossil content, and
paleocontinent allowed deposition of organic-rich sediments in an organic and inorganic geochemistry parameters (Rodrigues and Quad­
anoxic epicontinental sea, such as the black shales of the Irati Formation ros, 1976; Cerqueira and Santos Neto, 1986; Hachiro, 1997), others have
(Araújo et al., 2001; Milani et al., 2007). suggested deposition of the Irati Formation in fresh or brackish waters
The Irati Formation exemplifies some of the most favorable based on the almost complete absence of marine diagnostic elements in
geochemical conditions for excellent potential source rock in the world, the palynology, geochemical parameters, and stable isotope data

* Corresponding author.
E-mail address: eliane@lenep.uenf.br (E.S. Souza).

https://doi.org/10.1016/j.marpetgeo.2021.104917
Received 22 May 2020; Received in revised form 5 January 2021; Accepted 7 January 2021
Available online 16 January 2021
0264-8172/© 2021 Elsevier Ltd. All rights reserved.
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

(Mendes et al., 1966; Amaral, 1971; Petri and Coimbra, 1982; Correa da matter source input, degree of maturity, and paleoredox conditions.
Silva and Cornford, 1985; Faure and Cole, 1999). Aromatic compounds first detected in the Irati Formation show that
According to Araújo et al. (2001), oscillations of the paleoenvir­ freshwater inflows into the Irati Sea promoted algal blooms, water col­
onmental redox potential, the organic primary productivity rate, and umn stratification, and photic zone euxinia (Martins et al., 2020b), and
relative sea-level led to compositional heterogeneities, which are re­ carried terrigenous organic matter into the Irati Sea (Martins et al.,
flected in different organic facies in this unit. Geochemical indicators 2020c).
based on organic carbon content, stable carbon isotope ratios, Rock-Eval The aim of this research was to interpret the physicochemical con­
pyrolysis, and saturated biomarker results have been used to interpret ditions during deposition of the Irati black shales with outcrop samples
these depositional conditions (e.g., Reis et al., 2018; Nascimento et al., from two different and distant sites at the northeastern border and the
2018; Martins et al., 2020a). More recently, Goldberg and Humayun central east of the Paraná Basin, in addition to samples from the Serra
(2016) used redox-sensitive elements to infer the degree of oxygenation Alta Formation. A variety of geochemical indicators, including some
of the water column during deposition of the Irati black shales. newly detected biomarkers, were used to retrace the depositional con­
In addition to saturated biomarkers used in conventional geochem­ ditions that led to organic facies with some of the highest observed
ical analyses, aromatic biomarkers, although little used in Irati Forma­ organic carbon contents in the world.
tion studies, can play an important role in the interpretation of organic

Fig. 1. Paraná Basin location, with outcrop area and thickness of the supersequence Gondwana I and outcrops of Irati black shales [Amaral Machado (1) and São
Mateus do Sul quarries (2)] used in this research (Modified from Milani, 2004).

2
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

2. Geological settings average thickness is 40 m, reaching 70 m in the basin depocenter


(Mendes et al., 1966; Corrêa and Pereira, 2005). The end of deposition of
The intracratonic Paraná Basin is located in the central-eastern these bituminous shales is marked by non-bituminous gray laminated
portion of the South American continent and covers approximately 1.5 shales of the Serra Alta Formation (Gama Jr., 1979). The base of this
million square kilometers in Brazil, Paraguay, Argentina, and Uruguay formation was deposited in a highstand systems tract marked by a
(Fig. 1). Paleozoic, Mesozoic, and Cenozoic sedimentary rocks in the pronounced decrease in TOC and sulfur (S) values (Araújo et al., 2001).
basin are associated with Cretaceous volcanic rocks (Zalán et al., 1990; The Irati black shales are the source rock of the atypical petroleum
Milani et al., 2007). system Irati-Rio Bonito/Pirambóia(!) (Zalán et al., 1990; Milani and
Milani et al. (2007) recognized six supersequences in the stratig­ Zalán, 1999). These black shales are generally immature because they
raphy of the Paraná Basin. These are defined as distinct rock layers were not sufficiently buried to enter the oil window (Zalán et al., 1990;
delimited by regional unconformities, which represent long periods of Milani and Zalán, 1999; Milani et al., 2007). However, it has been
erosion and non-deposition during the geological history of this basin. suggested that Cretaceous igneous intrusions of the Serra Geral Forma­
The supersequences are characterized by (i) Paleozoic tion provided the necessary temperature for oil generation in some lo­
transgressive-regressive sea-level cycles, represented by the super­ cations of the Paraná Basin (Cerqueira and Santos Neto, 1986; Araújo
sequences Rio Ivaí (Ordovician to Silurian), Paraná (Devonian), and et al., 2000; Corrêa and Pereira, 2005; Thomaz Filho et al., 2008; Santos
Gondwana I (Carboniferous to Lower-Triassic), and (ii) cycles of conti­ et al., 2009).
nental systems deposition associated with strong magmatism, repre­
sented by the supersequences Gondwana II (Middle to Upper-Triassic), 3. Materials and methods
Gondwana III (Upper-Jurassic to Lower-Cretaceous), and Bauru
(Upper-Cretaceous) (Fig. 2). 3.1. Samples
According to Milani et al. (2007), the supersequence Gondwana I, at
approximately 2300 m thick, includes a complete A total of 20 samples from the Irati Formation were collected from
transgressive-regressive cycle that was the result of the invasion and outcrops in two cities in the Paraná Basin (Fig. 1) - seven from a quarry
subsequent withdrawal of the Panthalassa Ocean over the Gondwana in the Amaral Machado Mining Company area near Saltinho City (1,
continent. The cycle started with periglacial deposits from the Itararé Fig. 1; São Paulo State), and the other 13 from two outcrops 700 m from
Group, passed through the coastal and platformal deposits from the each other representing two organic-rich layers, the Lower and the
Guatá Group and the base of the Passa Dois Group, and finally reached Upper Shale (designated SM Lower and SM Upper outcrops), at the
continental conditions with deposition of lacustrine, fluvial, and eolian Petrobras Shale Industrialization Unit (SIX) quarry in the São Mateus do
sediments on the top of the Passa Dois Group. Sul City (2, Fig. 1; Paraná State). Additionally, four samples were
The Permian (Artinskian-Kungurian) Irati Formation (Santos et al., collected in the outcrop of Amaral Machado (AMa) above the horizon of
2006; Rocha-Campos et al., 2011) is subdivided into Taquaral and black shales in order to understand the conditions at the end of Irati
Assistência Members (Barbosa and Gomes, 1958). The Taquaral Member Formation deposition. These samples represent dark gray shales that are
(at the base) is characterized by a normal salinity marine environment interpreted here as belonging to the Serra Alta Formation. Table S1
with a restricted connection to the ocean, whereas the Assistência (Supplementary Material) displays the code for samples and heights in
Member (at the top) has carbonates interbedded with bituminous shales, relation to the base of the outcrops. Although the lower and upper shales
which are related to southwest-inclined ramps, connected to the Pan­ were collected from two outcrops in the São Mateus do Sul (SM) SIX
thalassa only in the southernmost region of the paleocontinent (Santos quarry, these horizons stratigraphically overlap and are separated by a
Neto and Cerqueira, 1993; Araújo et al., 2001; Rodrigues et al., 2010). 5-m carbonate layer (Alferes et al., 2011), while all samples from AMa
Reis et al. (2018) divided these sediments into chemostratigraphic units were interbedded with thin carbonate layers in outcrop.
based on their total organic carbon (TOC) and insoluble residue values
as well as their saturated biomarker contents. According to this study, 3.2. TOC and Rock-Eval pyrolysis
the bituminous shales in the Irati Formation are located in units E and H,
which represent a transgressive systems tract in two fourth-order The TOC and sulfur content were determined using 1 g of pulverized
sequences. rock samples (<80 mesh) in a LECO 628CN instrument, and the results
The Irati Formation outcrop extends approximately 1700 km and its are expressed as a mass percentage relative to the total mass of the

Fig. 2. Simplified stratigraphic chart of the Paraná Basin with emphasis on supersequence Gondwana I, showing the Taquaral and Assistência Members from the Irati
Formation (modified from Milani et al., 2007) and its chemostratigraphic units based on Reis et al. (2018).

3
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

analyzed sample. Prior to these analyses, the samples were treated with constant flow of 1 mL min− 1. The mass spectrometer was operated on
4 M HCl to remove the carbonates. Rock-Eval pyrolysis was performed full scan (50–550 Da) and selected ion monitoring (SIM) modes in
using the Rock-Eval 6 instrument (Vinci Technologies), at programed splitless mode. For the saturated fraction, the oven temperature program
temperatures between 300 and 650 ◦ C, where the S1, S2, S3, and TMAX was initially held at 60 ◦ C for 2 min, then heated at 22 ◦ C min− 1 to
values were determined. It was also possible to calculate the hydrogen 200 ◦ C, held for 3 min, and heated again at 3 ◦ C min− 1 to 300 ◦ C, fol­
index (HI) and the oxygen index (OI) using the TOC, S2, and S3 values. lowed by an isothermal period of 25 min (based on the procedure
described by Martins et al., 2017). β-carotane, regular hopanes and
3.3. Extraction and separation gammacerane, steranes, and tetracyclic polyprenoid were examined
using m/z 125, 191, 217, and 259, respectively, in SIM mode. The initial
Organic extracts were obtained from 90 to 150 g of each sample oven temperature for the aromatic fractions was 80 ◦ C. It was heated at
using a Soxhlet system for 48 h with dichloromethane as the solvent 2 ◦ C min− 1 to 300 ◦ C, held for 20 min, then heated at 10 ◦ C min− 1 to
(300 mL). The organic extract was separated into three fractions by 310 ◦ C, and held for 1 min (based on Nytoft et al., 2016). Aryl iso­
liquid chromatography using silica gel in a glass column 10 mm in prenoids, trimethylnaphthalenes, cadalene, benzohopanes, retene, C18
diameter and 25 cm in length. For this fractionation, 40 mg of each 1,1,7,8-tetramethyl-1,2,3,4- tetrahydrophenanthrene, triaromatic ste­
sample was separated into fractions of saturated hydrocarbons, aromatic roids, and simonellite and sempervirane were examined using m/z 133,
hydrocarbons, and polar (NSO) compounds using the mobile phases 170, 183, 191, 219, 223, 231 and 237, respectively, in SIM mode. The
hexane (30 mL), hexane:dichloromethane 8:2 v/v (30 mL), and injector and transfer line temperatures were 300 ◦ C for both methods.
dichloromethane:methanol 9:1 v/v (30 mL), respectively, as previously The biomarker ratios were calculated based on peak areas from the
described by Nascimento et al. (2018). Serra Alta shales were pretreated selected mass chromatograms.
with colloidal copper to remove crystals of sulfur (Blumer, 1957)
precipitated after extraction with dichloromethane. 3.8. Principal component and hierarchical cluster analyses

3.4. X-ray diffraction Multivariate statistical analysis of the geochemical data was per­
formed using R software (R Core Team, 2013) for data processing. Ten
X-ray diffraction analyses were performed using 5 g of whole rock geochemical parameters were used in the analyses, including TOC, S2,
powder (<270 mesh) through a D2 Phaser diffractometer (Bruker AXS) carbon isotope ratio (δ13C), tetracyclic polyprenoid ratio (TPP/Dia27),
using CuKα radiation (30 kV/10 mA), with patterns of 0.02◦ 2θ and hopane/sterane (Hop/Ste), C29/C27 steranes (C29/C27 Ste), gammacer­
collection time of 2 s per step, from 5 to 70◦ 2θ (Martins et al., 2020c). ane/C30 hopane (Gam/C30H), β-carotane/C30 hopane (βC/C30H), and
After the acquisition of diffractograms, the qualitative analysis was pyrite and sulfur percentages. Principal component analysis (PCA) was
conducted using Diffrac.EVA 4.0 software, which enables identification performed from a 24 × 10 correlation matrix (24 samples x 10 param­
of crystalline substances by comparing the diffractogram with individ­ eters), where Eigenvalues and eigenvectors were extracted from the
ual phase diffractometric patterns provided by the International Center original data to construct the new variables or principal components,
for Diffraction Data (ICDD). DIFFRACplus TOPAS version 4.2 (Bruker which are linear combinations of all the independent variables. For the
AXS), which is based on the Rietveld process for phase refinement, was hierarchical cluster analysis (HCA), it was used the Ward.D grouping
used to obtain the results in quantitative terms. method, and the Canberra metric distance.

3.5. Stable carbon isotope composition 4. Results and discussion

The stable carbon isotope composition of 1 mg of the whole organic 4.1. Conventional geochemical analyses
extract was analyzed using a Thermo Finnigan Mass Spectrometer (Delta
V Advantage), and the δ13C (‰) was reported in relation to the inter­ 4.1.1. Bulk parameters
national PDB (Peedee Formation Belemnitella americana, from Upper The studied Irati sample set consists of bituminous shales with high
Cretaceous, South Carolina, USA) standard. The gases used were helium TOC content (3.2–25.4 wt%), high values of S2 (13.7–135.8 mg HC/g
as the carrier, O2 for combustion, and CO2 as the standard. Isodat 2.5 rock), and HI (366–716 mg HC/g TOC), pointing to excellent potential
software was used for data acquisition and treatment. for oil generation, although TMAX values below 430 ◦ C suggest thermally
immature organic matter (Fig. 3; Tissot and Welte, 1984; Espitalié et al.,
3.6. Gas chromatography with flame ionization detector 1986). These results can be related to the Reis et al. (2018) studies using
well samples, in which the two organic-rich layers from the Irati For­
The whole organic extracts of the Irati black shales and Serra Alta mation were classified as units E, which corresponds to the SM Lower
shales were analyzed by gas chromatography with flame ionization outcrop in the current study, and H, which corresponds to the SM Upper
detector to assess the n-alkanes and the isoprenoids pristane and outcrop and the AMa Irati black shales. On the other hand, Serra Alta
phytane. The analysis was performed using an Agilent 6890N gas shales AMa 25, 26, 28, and 29 at the top of the AMa outcrop contain very
chromatograph, equipped with an HP-5 fused silica capillary column low TOC (<0.5%), low values of S2 (<0.5 mg HC/g rock), and HI (<90
(30 m × 0.32 mm x 0.25 μm). Helium was used as the carrier gas in mg HC/g TOC; Fig. 3), relating to non-bituminous shales with low
splitless mode at a constant flow of 2.2 mL min− 1. The oven temperature oil-generating potential. TOC and Rock-Eval pyrolysis data are available
program was initially held at 40 ◦ C for 1 min, then heated at 6 ◦ C min− 1 in Table 1. According to the modified Van Krevelen diagram (Espitalié
to 310 ◦ C, followed by an isothermal period of 16 min. The injector and et al., 1986, Fig. 4), the Irati black shales contain type I and II organic
detector temperatures were 290 and 320 ◦ C, respectively. matter likely formed from a mixture of autochthonous organic matter
(phytoplankton, zooplankton, and bacteria), which may represent
3.7. Gas chromatography coupled to mass spectrometry lacustrine or marine facies (Tissot and Welte, 1984). In contrast, Serra
Alta shales from the top of the AMa outcrop contain type III organic
The saturated and aromatic fractions from the organic extracts of the matter, probably formed from continental plants and detritus, typical of
Irati black shales and Serra Alta shales were analyzed using an Agilent highstand systems tract (Araújo, 2001).
7890A gas chromatograph equipped with an HP-5 MS column (30 m ×
0.25 mm x 0.25 μm film thickness) coupled to an Agilent 5975C-MSD 4.1.2. Thermal maturity
selective mass detector. Helium was used as the carrier gas at a In order to further evaluate the maturity of the samples on the

4
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 3. TOC and Rock-Eval pyrolysis data for the Irati black shales and Serra Alta shales.

molecular level, the parameters C29 20S/(20S + 20R), C29 ββ/(ββ + αα), correspond to the AMa outcrop base. Therefore, the sum of the igneous
and C31 22S/(22R + 22S) were assessed based on the isomerization of effect of the intrusive rock together with the presence of clay minerals is
the C29 steranes and C31 homohopanes (Fig. 5; calculation procedures in the most plausible explanation for the greater thermal maturity
the Supplementary Material Table S2). Generally, the endpoint values observed in these samples.
for these parameters range from 0.52 to 0.55, 0.67 to 0.71, and 0.57 to
0.62, respectively (Seifert and Moldowan, 1980, 1986). It was observed 4.1.3. Depositional paleoenvironment
that most of the samples were below these values except for samples The depositional paleoenvironment and organic matter input for the
AMa 1 and 5, which were more thermally evolved than the others, Irati black shales and Serra Alta shales were investigated using δ13C, key
having values within or above the endpoint range for the C31 22S/(22R saturated biomarker parameters including TPP/Dia27, Hop/Ste, C29/
+ 22S) ratio. C27 Ste, pristane/nC17 (Pr/nC17), phytane/nC18 (Ph/nC18), Gam/C30H,
Although samples AMa 1 and 5 had the lowest TMAX values (404 and βC/C30H, and by their mineralogical composition (Figs. 6 and 7;
405 ◦ C, respectively), Martins et al. (2020c) noted the presence of more Table 1). All of these parameters show a decreasing trend in marine
clay minerals, mainly montmorillonite (from the smectite group), in the input with a concomitant increase in terrigenous input toward the top of
basal samples of the AMa outcrop (see also the x-ray diffractograms in the outcrops, except for sample SM 3.7.
Fig. 6). Clay-rich rocks can influence pyrolysis experiments (Horsfield Analyses of δ13C values suggest an increase in terrigenous input to­
et al., 1983; Yang and Horsfield, 2016) and according to Dembicki ward the top of the three outcrops since the values decrease from
(1992), montmorillonite can act as a catalyst, decreasing the tempera­ approximately − 22 to − 28‰ at the AMa outcrop and − 22 to − 25‰ at
ture required for kerogen cracking in Rock-Eval pyrolysis, and conse­ the SM outcrops and, according to Meyers (1994), δ13C values decrease
quently the TMAX values. In addition, studies carried out by Seifert from marine algae (− 16 to − 23‰) to freshwater algae (− 26 to − 30‰;
(2013) in well FP-12 (Anhembi, SP) in the same region of the AMa Fig. 7). Araújo (2001) previously suggested that the most positive δ13C
outcrop showed a 2.4 m diabase sill below the samples that would values (− 22 to − 25‰) in the Irati black shales are associated with

5
C.A. Nascimento et al.
Table 1
TOC, sulfur content (S), Rock-Eval, carbon isotope ratio (δ13C), and saturated biomarker data used for the interpretation of the Irati black shales and Serra Alta shales, including the geochemical parameters used in the PCA
and HCA analyses.
Samples TOCa Sa S1 (mg HC/ S2a (mg HC/ S3 (mg HC/ HI (mg HC/ OI (mg CO2/ TMAX PI (wt. δ13Ca Hop/ TPP/ C29/ Gam/ βCIa Pyritea Pr/ Ph/
(%) (%) g rock) g rock) g rock) g TOC) g TOC) (◦ C) ratio) (‰) Stea Dia27a C27 C30Ha (%) nC17 nC18
Stea

AMa 29 0.07 0.40 0.02 0.27 0.20 64 2.86 431 0.07 − 28.30 12.37 1.23 0.62 0.01 0.00 0.33 0.57 0.18
AMa 28 0.05 0.43 0.01 0.17 0.17 53 3.40 432 0.06 − 28.17 10.42 1.06 0.60 0.01 0.00 0.24 0.63 0.19
AMa 26 1.48 1.32 0.02 0.35 0.20 73 0.14 427 0.05 − 28.71 12.21 1.45 0.58 0.01 0.00 0.75 0.35 0.16
AMa 25 0.40 0.96 0.02 0.44 0.21 88 0.53 427 0.04 − 28.16 10.35 2.04 0.60 0.01 0.00 0.28 0.35 0.15
AMa 24 7.18 1.34 1.83 48.31 0.16 716 0.02 439 0.04 − 27.99 5.47 4.02 0.52 0.02 0.01 1.20 3.84 2.27
AMa 23 6.40 5.59 3.66 32.89 0.28 552 0.04 428 0.1 − 26.61 9.32 2.16 0.54 0.03 0.02 5.66 5.08 3.27
AMa 17 5.99 4.57 2.27 29.01 0.19 569 0.03 424 0.07 − 23.06 7.56 0.07 0.49 0.89 0.02 3.33 4.39 5.46
AMa 11 6.38 4.95 3.86 33.73 0.16 620 0.03 421 0.1 − 22.01 3.31 0.07 0.42 0.92 0.03 3.31 3.90 6.19
AMa 7 5.52 4.81 3.94 26.73 0.35 564 0.06 418 0.13 − 22.22 2.53 0.04 0.40 0.96 0.05 2.85 2.89 5.78
AMa 5 4.30 4.50 2.59 13.71 0.20 406 0.05 402 0.16 22.95 2.24 0.03 0.36 0.53 0.05 2.23 2.51 6.77
6


AMa 1 3.83 2.60 4.19 17.37 0.12 534 0.03 397 0.19 − 23.52 2.07 0.02 0.37 0.28 0.05 1.78 2.81 9.48
SM 3.7 23.17 3.43 7.53 102.60 1.08 443 0.05 422 0.07 − 25.43 1.26 0.13 0.42 0.49 0.02 4.57 8.03 13.96
SM 3.6 14.99 4.07 5.61 78.50 0.48 524 0.03 424 0.07 − 22.65 11.08 3.90 0.47 0.50 0.06 5.92 8.34 11.11
SM 3.5 15.37 3.79 6.34 109.49 0.40 712 0.03 423 0.05 − 22.94 9.64 4.64 0.48 0.59 0.08 4.78 8.52 12.00
SM 3.4 12.66 2.72 5.92 88.35 0.32 698 0.03 424 0.06 − 22.36 6.17 1.63 0.49 0.68 0.13 3.36 8.17 11.28
SM 3.3 7.23 3.48 3.45 45.39 0.28 628 0.04 424 0.07 − 23.10 5.04 1.84 0.49 0.88 0.13 4.44 8.16 11.71
SM 3.2 11.89 4.85 5.26 76.16 0.37 641 0.03 425 0.06 − 21.98 4.49 1.80 0.47 1.21 0.14 4.97 5.90 9.67
SM 3.1 9.50 4.91 3.74 59.47 0.37 626 0.04 422 0.06 − 22.10 2.76 1.50 0.46 1.36 0.13 6.33 3.80 7.22
SM 2.6 5.48 5.10 2.48 20.07 0.19 366 0.03 424 0.11 − 25.09 7.20 3.06 0.49 0.04 0.04 4.74 6.87 11.04
SM 2.5 25.38 4.49 9.01 135.80 0.61 535 0.02 428 0.06 − 24.17 4.63 1.08 0.49 0.27 0.01 5.41 3.96 7.14
SM 2.4 22.60 6.87 7.24 114.61 0.49 507 0.02 425 0.06 − 23.57 2.99 1.43 0.42 0.43 0.05 10.22 3.29 6.37
SM 2.3 11.67 8.89 4.67 65.11 0.26 558 0.02 424 0.07 − 22.84 2.21 1.25 0.42 0.80 0.09 6.55 4.29 8.51
SM 2.2 9.10 8.20 4.54 52.77 0.23 580 0.03 421 0.08 − 22.81 1.96 1.86 0.42 0.77 0.13 7.80 4.41 8.69
SM 2.1 6.27 5.33 3.71 33.80 0.21 539 0.03 420 0.1 − 23.39 1.46 1.32 0.41 0.99 0.17 7.08 5.46 9.48

Marine and Petroleum Geology 126 (2021) 104917


a
Geochemical parameters used in the PCA and HCA analyses.
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

like to those of δ13C (Figs. 7 and 9a; top outcrop samples AMa 23 and 24,
SM 2.6, 3.3 to 3.6 with higher TPP/Dia27 and δ13C values), corrobo­
rating freshwater influx into the depositional paleoenvironment, except
for sample SM 3.7, where the low TPP/Dia27 value could characterize a
new episode of a rise in sea level and loss of freshwater input. Following
this reasoning, TPP/Dia27 can be applied to retrace freshwater in­
cursions in a depositional paleoenvironment.
Along with δ13C and TPP/Dia27, the Hop/Ste values increase toward
the top of the outcrops (Fig. 7), corroborating the rising terrigenous
contribution to the organic matter in the Irati black shales, since low
Hop/Ste values are typical of marine organic matter with higher
planktonic contribution and/or benthic algae, while high Hop/Ste ratios
are more common in terrigenous organic matter (Moldowan et al.,
1985). Also, Reis et al. (2018) reported higher values of the Hop/Ste
ratio together with a higher proportion of pollen grains toward the top of
the organic-rich shale layers (units E and H), indicating the influence of
higher plants. Once again, sample SM 3.7 behaved differently from the
others, showing a low value of Hop/Ste, which points to its more marine
character.
The increase in the C29/C27 Ste ratio toward the top of the outcrops
(Fig. 7) also supports freshwater input into the Irati Sea with increasing
higher plant or green algae contribution. The origin of C29 steranes has
Fig. 4. Modified Van Krevelen diagram showing the sample distribution rela­ been attributed to land plants or green algae, the ancestors of land
tive to types I, II, and III organic matter pathways. plants, while plankton-derived organic matter is richer in C27 steranes,
having red algae as its source (Huang and Meinschein, 1979; Hoffmann
autochthonous amorphous organic matter deposited in a transgressive et al., 1984; Volkman et al., 1994; Schwark and Empt, 2006; Kodner
systems tract, while the most negative values (≈− 27‰) are from et al., 2008).
allochthonous terrigenous organic matter deposited in a highstand sys­ Accordingly, more marine contribution is proposed for the basal
tems tract context. outcrop samples, mainly samples AMa 1, 5, 7, 11, and 17 from the AMa
The C30 tetracyclic polyprenoid 18α(H) 21R and 21S isomers (I, TPP; outcrop and sample SM 3.7 from the SM Upper outcrop, with lower
see chemical structure in Appendix A) were detected in all Irati samples values of δ13C, TPP/Dia27, Hop/Ste, and C29/C27 Ste (Fig. 9). However,
(monitoring m/z 259; Fig. 8), and the increase in TPP/Dia27 (Fig. 7; see when these data values for the analyzed samples were higher, greater
calculation procedures in the Supplementary Material Table S2 ac­ freshwater and terrigenous organic matter input into the paleoenvir­
cording to Silva et al., 2011) toward the top of the outcrops may reflect onment was assumed (Fig. 9). The TPP/Dia27 ratio does not separate
an increase of non-marine organisms such as freshwater algae (Holba Serra Alta samples from most of the SM outcrop samples. Several factors
et al., 2000, 2003; Araújo et al., 2018). The values of this ratio behave may influence this response, such as the low organic content of the Serra

Fig. 5. Thermal maturity plots of (a) C29 20S/(20S + 20R) vs. C29 ββ/(ββ + αα) and (b) C29 20S/(20S + 20R) vs. C31 22S/(22R + 22S), and a representative
distribution of biomarkers in the mass chromatograms (c) m/z 191 and (d) m/z 217 of the Irati and Serra Alta shales.

7
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 6. X-ray diffractograms of the Irati black shales, showing the mineralogical differences between Amaral Machado and São Mateus do Sul shales. Sme = Smectite;
Ilt = Illite; Ms = Muscovite; Gp = Gypsite; Qz = Quartz; Py = Pyrite.

Fig. 7. δ13C, TPP/Dia27, Hop/Ste, C29/C27 Ste (as indicative of organic matter input), Gam/C30H, βC/C30H, and percentages of S and pyrite (as indicative of
physicochemical conditions) showing the depositional paleoenvironment behavior of the Irati and Serra Alta shales.

Alta shales and low concentrations of biomarkers, or the possibility of depositional paleoenvironment, which can be illustrated by the Pr/nC17
different types of organic matter, such as greater input of freshwater vs. Ph/nC18 plot (Fig. 10; Peters et al., 1999; see representative gas
algae and bacteria (Correa da Silva and Cornford, 1985) in SM shales chromatograms from GC-FID in the Supplementary Material Fig. S1),
than Serra Alta shales, which contain abundant phytoclasts (Martins where samples from the top of the outcrops, especially AMa 23 and AMa
et al., 2020a). 24, exhibit more oxidizing characteristics and mixed organic matter
Organic matter input is linked to physicochemical conditions of the input (types II/III), while the samples from the base of the outcrops are

8
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 8. Distribution of C27 diasteranes, TPP, and C32 lanostane in mass chromatograms m/z 259 for samples from Amaral Machado and São Mateus do Sul outcrops.

Fig. 9. Distributions of the samples with higher marine or terrigenous contributions based on the plots of (a) TPP/Dia27 vs. Hop/Ste, (b) TPP/Dia27 vs. δ13C, and (c)
TPP/Dia27 vs. C29/C27 Ste.

more reducing and have an algal marine input (type II). In this case, the freshwater. It is interesting to note that Serra Alta samples have Pr/nC17
increase in oxidation and terrigenous input toward the top of the and Ph/nC18 values < 1 (Table 1; Fig. S1 for representative sample AMa
outcrop samples is also presumably linked to the greater influx of 26), resulting in negative logarithmic values. This is due to the greater

9
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

relation to the AMa outcrop. This behavior occurs because paleogeo­


graphically, SM was located in deeper portions of the Paraná Basin with
greater siliciclastic input while AMa was located in an arid and restricted
environment at the edge of the basin, forming part of an internal car­
bonate ramp (Faure and Cole, 1999; Araújo et al., 2001), which is why
the black shales in this outcrop are interbedded with carbonate layers.
Finally, C32 lanostane (II, Lan32; see chemical structure in Appendix
A and the mass spectrum in Appendix B) was detected for the first time
in the Irati black shales (monitoring m/z 259; Fig. 8). The relative
abundance of this compound does not vary as a function of TPP/Dia27,
with high values corresponding to non-marine organisms, although its
presence can be attributed to hypersaline conditions of the depositional
environment since its relative abundance decreases in samples with
higher freshwater input and decreased salinity (AMa 23 and SM 2.6;
Figs. 7 and 8). Chen et al. (1989) first identified the C30–C32 lanostanes
in Tertiary saline sediments in China. These compounds were also re­
Fig. 10. Pr/nC17 vs. Ph/nC18 plot showing the classification of the Irati black ported in low-maturity crude oils (Peng et al., 1998), in fossilized
shales according to their oxicity and organic matter type.
immature organic matter derived from hypersaline lakes (Parfenova,
2011), and in sulfur-rich crude oils (Lu et al., 2011). Although lanosterol
n-alkanes abundance in these samples, compatible with the greater is the most likely precursor of lanostane (Birgel and Peckmann, 2008), a
contribution of terrigenous organic matter (Mello et al., 1988). In direct correlation cannot be made between C32 lanostane and its bio­
addition, Martins et al. (2020b) showed that Pr/Ph ratios < 1 in the logical precursor since lanosterol can be biosynthesized by plants, ani­
basal samples of the AMa outcrop point to a reducing and hypersaline mals, fungi, dinoflagellates, sponges, and methanotrophic bacteria
environment. (Peckmann et al., 2004; Lamb et al., 2007). Like in previous works,
Gam/C30H and βC/C30H responses (Jiang and Fowler, 1986; Mello lanostane in Irati black shales characterizes: (1) immature organic ex­
et al., 1988; ten Haven et al., 1988; Sinninghe et al., 1995), show the tracts since it was not detected in the more mature samples AMa 1 and 5;
same trends of variation of the previous parameters, generally (2) deposition under high salinity conditions since it has lower relative
decreasing toward the top of the outcrops (Fig. 7). This suggests a abundances in the samples with higher freshwater input and was not
decrease in salinity and stratification of the water column as a result of detected in Serra Alta shales; and (3) sulfur-rich environments.
freshwater entering the black shale depositional environment. This
behavior is consistent with increased terrigenous input that probably 4.1.4. Multivariate statistical analysis
resulted from the increased inflow of freshwater toward the top of these Taking into account the conventional geochemical parameters used
outcrops. to analyze the organic input, salinity, and physicochemical conditions at
Stratification of the water column and high salinity at the bottom of the time of deposition of the Irati and Serra Alta shales, PCA was per­
the Irati Sea is the probable cause of the reducing conditions verified in formed to describe patterns in the multivariate data (Syms, 2008) by
the samples from the base of the outcrops through redox-sensitive pa­ graphically displaying groupings of samples with similar geochemical
rameters such as sulfur and pyrite percentages (Fig. 7; Lyons and Sev­ characteristics. PCA was performed using a 24 × 10 correlation matrix
ermann, 2006; Goldberg and Humayun, 2016). The marked presence of (used parameters marked in Table 1) in which principal components 1
pyrite in the Irati black shales from the SM outcrops, particularly for the and 2 (PC1 and PC2) explained 72.7% of the total variation of the data
SM Lower outcrop (4.7–10.2%), was previously demonstrated by Correa and the components were split into four groups (Fig. 11), which repre­
da Silva and Cornford (1985), who related this to extreme reducing sent: (1) Serra Alta shales (AMa 25, 26, 28, and 29) with lower TOC and
conditions in an oxygen-deficient water column. At the top of the SM higher freshwater and terrigenous input; (2) Irati Formation base sam­
Upper outcrop (SM 3.5, 3.6, and 3.7 samples; Fig. 7), however, although ples from the AMa outcrop (AMa 1, 5, 7, 11, and 17) with higher salinity
salinity declined (see Gam/C30H and βC/C30H ratios), reducing condi­
tions remained strong, as shown by the high values of S (3.4–4%) and
pyrite (4.6–5.9%). This is consistent with the reducing conditions in the
Irati Sea due to water column stratification during deposition of the high
TOC Irati black shales (Martins et al., 2020b). On the other hand, Serra
Alta shales show very low percentages of S and pyrite and very low
Gam/C30H values (Fig. 7), suggesting the absence of salinity-related
density stratification during deposition (Tulipani et al., 2015). This
interpretation is consistent with higher abundance of C30 tetracyclic
polyprenoids (higher TPP/Dia27 ratio) and lower values of δ13C, sug­
gesting extensive freshwater incursion during deposition of these sam­
ples. Depositional dysoxic/oxic conditions during deposition of these
shale rocks also are consistent with very low TOC content (<0.5 wt%)
and low oil generation potential.
In addition to pyrite, other minerals can reflect specific character­
istics of the paleoenvironment during deposition of the Irati black shales
(Holanda et al., 2019). Chemical precipitation of the gypsite (Gp) occurs
in saline environments with little water circulation (Droste, 1961;
Warren, 1999), and its occurrence in the basal Irati black shales from the
AMa outcrop (Supplementary Material Table S3) and absence in the top
outcrop samples (AMa 23 and 24; Fig. 6) is consistent with the hyper­
saline environment. Quartz, on the other hand, shows higher percent­
Fig. 11. Principal component analysis chart with principal components 1 and 2
ages in the SM outcrops, reflecting their more siliciclastic character in grouping the samples into four groups based on geochemical characteristics.

10
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

and stronger marine characteristics; (3) Irati Formation top samples the Irati black shales (chemical structure in Appendix A). These bio­
(AMa 23 and 24) plus sample SM 2.6, with lower salinity and higher markers include triaromatic steroids (III, TAS), trimethylnaphthalenes
freshwater and terrigenous input; and (4) the other samples from SM (IV, TMN), simonellite (V, Sim) and sempervirane (VI, Semp). More­
outcrops (SM 2.1 to 2.5 and 3.1 to 3.7), with the highest TOC values and over, detection of a series of benzohopanes (VII, C32 to C35) and aryl
great freshwater input. The ellipses were traced using R software with a isoprenoids (VIII, AI) provides further evidence of the restricted and
95% confidence level. PC1 explained 53.6% of the data variation, being hypersaline depositional paleoenvironment during the deposition of the
influenced more heavily by the parameters δ13C, C29/C27 Ste, Pyrite and Irati black shales, in agreement with their higher amounts of sulfur,
S percentages, Hop/Ste, and Gam/C30H while PC2 explained 19.1% of pyrite, gammacerane, and the presence of the C32 lanostane. These ar­
the data variation, being represented primarily by the TPP/Dia27, S2, omatic biomarkers were not detected (or detected in very low abun­
and TOC, which mainly separated the Irati Formation base samples dance) in the low TOC Serra Alta shales, in accordance with their
(AMa outcrop) from the Irati Formation top samples (AMa outcrop) plus deposition under oxic paleoenvironment conditions.
sample SM 2.6 (Supplementary Material Fig. S2). The triaromatic steroid (TAS) distribution (26S and R, 27S and R,
In order to confirm the characterization of the samples in terms of 28S and R TAS; monitoring m/z 231, Fig. 13a) typically resembles the
organic and freshwater input and better understand the changes in the regular sterane distribution associated with the organic matter input,
depositional paleoenvironment in each outcrop, an HCA was performed meaning that C26 TAS (26S and R TAS) can be related to marine envi­
using the Ward.D method and the Canberra metric distance (Fig. 11). ronments, while C28 TAS is associated with greater terrigenous contri­
HCA provides a tree dendogram that classifies objects based on their bution (Mackenzie et al., 1982). This association can also be made by
similarities (Johnson, 1967). As can also be observed in the PCA compounds of the trimethylnaphthalene family (monitoring m/z 170,
two-dimensional plane chart (Figs. 11), Fig. 12 shows that the strong Fig. 13b). Strachan et al. (1986) proposed the TDE-1 ratio (1,2,
marine characteristics observed in the Irati samples from the base of the 5-TMN/1,2,4-TMN) for differentiating coal samples from marine,
AMa outcrop (group G3, samples AMa 1, 5, 7, 11, and 17) change to lacustrine, and deltaic environments. According to their study, the
more terrigenous toward the top (group G2, samples AMa 23 and 24), relative concentrations of 1,2,5-TMN were higher in oils derived from
reaching dysoxic/oxic conditions and the low TOC Serra Alta shales Cretaceous or younger rocks dominated by large input from higher
(group G1, samples AMa 25, 26, 28, and 29) due to episodes of fresh­ plants. Grice et al. (2001), however, proposed cyanobacterial hopanoids
water inflow into the Irati Sea (Goldberg and Humayun, 2016; Martins as a precursor for 1,2,5-TMN in Permian and Carboniferous torbanites
et al., 2020b). Irati black shales from the SM outcrops also have more (boghead coal), which predate the evolution of angiosperms. The plot of
marine characteristics in the base of the outcrops (group G4; samples SM TDE-1 vs. C28/C26 TAS (Fig. 13c) shows higher values of C28/C26 TAS for
2.1, 2.2, 2.3, 3.1, 3.2, 3.3, and 3.4), changing to a depositional setting the samples with greater terrigenous contribution based on conventional
under more freshwater influence with greater terrigenous input in the geochemical analyses (G2 and G5 samples, plus samples SM 3.3 and
top of the outcrops (group G5; samples SM 2.4, 2.5, 2.6, 3.5, 3.6, and 3.4). However, samples from the AMa outcrop (G2 and G3) show higher
3.7). TDE-1 values than samples from SM outcrops (G4 and G5), which could
Inside the group G5, samples SM 2.4, 2.5, and 3.7 are in a subgroup, suggest that higher plants are not the only precursor of 1,2,5-TMN in
which reflects their transitional characteristics. Accordingly, although these samples, strengthening the cyanobacterial origin suggested by
samples SM 3.3 and 3.4 are in the group G4 due to their likely affinity Grice et al. (2001). The fact that black shales from the AMa outcrop are
with the other samples conforming to some variables such as δ13C and interbedded with carbonates whose microbial mats formed by cyano­
βC/C30H, they show a transitional character, as they also have bacteria (Warren et al., 2017) reinforces that suggestion.
geochemical affinities with samples from the G5 group, as can be seen by In addition to trimethylnaphthalenes, simonellite and sempervirane
the parameters TPP/Dia27 and Gam/C30H (Fig. 7). (V and VI, see the chemical structure in Appendix A and the mass
spectrum in Appendix B) were also detected in all Irati Formation
samples (monitoring m/z 237; Fig. 14a), as well as cadalenes, whose
4.2. Geochemical analyses for aromatic biomarkers
increase toward the top of the outcrops points to greater terrigenous
contribution as already been suggested by Martins et al. (2020c).
Although often used to assess thermal maturity of petroleum source
Simonellite and sempervirane have higher relative abundances in sam­
rocks, aromatic biomarkers have been little used to study the organic
ples with a greater terrigenous contribution (G2 and G5, plus samples
matter source and physicochemical characteristics of the Irati Formation
SM 3.3 and 3.4), as observed in the plot of the proposed ratio (Sim +
(Afonso et al., 1994; Osorio et al., 2017; Martins et al., 2020a, 2020b,
Semp)/C26 TAS vs. C28/C26 TAS (Fig. 14b), while the groups with more
2020c). In the present study, we detected key aromatic biomarkers and
marine characteristics (G3 and G4) have lower values for these ratios.
used some diagnostic ratios to characterize the organic matter source of

Fig. 12. Hierarchical cluster analysis grouping samples in G1, G2, G3, G4, and G5, where groups G1, G2, and G5 present higher terrigenous contribution while
groups G3 and G4 have characteristics of higher marine contribution.

11
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 13. Distribution of the (a) triaromatic steroids in the mass chromatogram m/z 231 and the (b) trimethylnaphthalenes in m/z 170. (c) The plot of TDE-1 vs. C28/
C26 TAS shows higher values in each outcrop for the samples with greater terrigenous contribution interpreted based on conventional geochemical analyses (G2 and
G5 samples).

and the mass spectrum in Appendix B). These compounds are particu­
larly abundant in carbonate and evaporitic environments and it was
suggested that they form in the subsurface from bacteriohopanepolyol
precursors (Hussler et al., 1984; He and Lu, 1990; Oba et al., 2009). In
the present work, the presence of the benzohopane homologues is
consistent with the inferred hypersaline conditions in the stratified
waters of the depositional environment for the Irati black shales.
The distribution of benzohopanes varies among the samples, and this
variation may be related to different degrees of maturity (Nytoft et al.,
2016). In samples AMa 1 and 5, for example, the relative abundances of
C32 and C33 benzohopanes are notably lower than in other samples
(Fig. 15). Therefore, the C32-33/C32-35 benzohopane ratio is proposed
here as a maturation indicator since there is correlation between this
ratio and the maturity parameters C29 ββ/(ββ + αα), C29 20S/(20S +
20R), and C31 22S/(22S + 22R), where samples AMa 1, 5, and 7 are
more mature and have the lowest values for C32-33/C32-35 benzohopanes
Fig. 14. (a) Mass chromatogram m/z 237 showing the relative abundance of
(Fig. 16).
simonellite and sempervirane in samples from Amaral Machado e São Mateus
do Sul and (b) plot of (Sim + Semp)/C26 TAS vs. C28/C26 TAS showing an in­ Finally, a series of aryl isoprenoids ranging from C14 to C29 was also
crease of terrigenous contribution toward G2 and G5 samples. detected in all Irati black shales (monitoring m/z 133, Fig. 17a), with
different relative abundances among the samples. These compounds
appear to have the same origin as the green sulfur bacteria based on
their structure and isotopic composition (Summons and Powell, 1986;
These biomarkers may also suggest higher-plant input since they have Grice et al., 2005; Tulipani et al., 2015). Therefore, the identification of
conifer resins as precursors (Simoneit and Didyk, 1986; Alexander et al., aryl isoprenoids suggests that euxinia reached the photic zone during
1988; Otto and Wilde, 2001; Cox et al., 2007). It was also observed that deposition of the Irati black shales since these compounds are diagenetic
the SM Upper samples (G5) have higher values for the (Sim + Semp)/C26 or catagenetic products of isorenieratene, a photosynthetic pigment in
TAS ratio compared to the AMa top samples (G2). This behavior can be green sulfur bacteria (Schaeffer et al., 1997; Clifford et al., 1998; Sousa
associated with the paleogeography of the outcrops, since the region of et al., 2013; Martins et al., 2020b). This interpretation is corroborated by
the SM outcrops had greater sea depths (restricted deep sea; Araújo C18 1,1,7,8-tetramethyl-1,2,3,4-tetrahydrophenanthrene (THP), which
et al., 2001; Martins et al., 2020b) than the region of the AMa outcrop, was noted for the first time in the Irati black shales by Martins et al.
which was more restricted and arid (restricted shallow sea; Faure and (2020b) and was detected by monitoring m/z 223 for all analyzed Irati
Cole, 1999; Araújo et al., 2001). For this reason, although the AMa re­ black shales herein with great relative abundance (Supplementary Ma­
gion was shallower and proximal and had a contribution of plant debris terial Fig. S3). Sinninghe Damsté et al. (1999) suggested that this com­
(phytoclasts), in SM there was greater humidity and better conditions pound originates from tetrahymanol. The principal source of this lipid
for the proliferation of higher plants, characterized primarily by co­ appears to be bacterivorous ciliates, which occur at the interface be­
nifers, which were dominant in the Paraná Basin during the early tween oxic and anoxic zones in stratified water columns (Sinninghe
Permian (Ricardi-Branco et al., 2013). et al., 1995). This is consistent with salinity-related density stratification
A family of hopanoid-related monoaromatic biomarkers was also in the depositional setting of the Irati black shales.
detected (monitoring m/z 191; Fig. 15) for the first time, as far as we To our knowledge, the aryl isoprenoids ratio (AIR), given by the ratio
know, in the organic extracts of the Irati Formation: the cyclized at C20 of C13-17 to C18-22 AI, was applied to the Irati Formation samples for the
benzohopanes (VIII, C32–C35, see the chemical structure in Appendix A first time in the present work. AIR has values < 1 in all analyzed samples

12
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 15. Mass chromatogram m/z 191 showing the distribution of the C32–C35 benzohopanes in samples from the Amaral Machado e São Mateus do Sul outcrops.

(Fig. 17b), suggesting persistence of the anoxia/euxinia photic zone of deposited organic matter. In addition, freshwater input brought
(PZE; Schwark and Frimmel, 2004; Martinez et al., 2018), and is also greater terrigenous organic matter (G2 and G5 samples) to this deposi­
inversely proportional to the C28/C26 TAS ratio. It is interesting to note tional paleoenvironment, culminating in the low TOC shales and more
that for samples that have a greater terrigenous contribution (G2 and G5 oxidizing conditions interpreted for the Serra Alta Formation (G1
samples, plus samples SM 3.3 and 3.4) AIR decreases, meaning that PZE samples).
persistence is increased, which is consistent with the freshwater inflow The assessed parameters are based on bulk geochemical data,
and stratification of the water column. mineralogical composition, and mostly on saturated and aromatic bio­
markers. The TPP/Dia27 ratio was indicative of non-marine organic
5. Conclusions input and freshwater incursion, in addition to the TDE-1 and C28/C26
TAS ratios. The detection of cadalene corroborates the input of higher
Changes in paleoenvironmental depositional conditions for the Irati plants into the Irati Sea, in addition to simonellite and sempervirane,
black shales were interpreted based on a series of geochemical param­ which suggest greater contribution of conifer resins in the SM Upper
eters, and the investigated samples were separated into groups by outcrop based on the proposed (Sim + Semp)/C26 TAS ratio, consistent
principal component and hierarchical cluster analyses based on their with its paleogeographic position, greater water depth, and humidity
source-related geochemical characteristics. These interpretations sup­ compared to the AMa outcrop samples.
port deposition in a restricted and hypersaline marine paleoenvironment C32 lanostane and the C32-35 benzohopanes were detected for the first
(G3 and G4 samples) with episodes of freshwater input. Freshwater time in organic extracts from the Irati Formation, and their presence
inflow led to stratification of the water column, generating a euxinic supports a hypersaline and sulfur-rich paleoenvironment. The presence
photic zone and reducing bottom conditions that enhanced preservation of tetrahydrophenanthrene (THP) and aryl isoprenoids ratio (AIR) also

13
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Fig. 17. (a) Representative mass chromatograms m/z 133 showing the aryl
isoprenoid distribution in samples with marine (AMa 1) and terrigenous (SM
3.6) contributions and (b) plot of C28/C26 TAS vs. AIR showing G2 + G5
samples with greater terrigenous contribution, deposited under increased PZE
persistence.

supports persistence of photic zone euxinia (PZE) during deposition of


the Irati black shales. Additionally, the proposed C32-33/C32-35 benzo­
hopane ratio seems to be related to different thermal maturities
observed in the Amaral Machado outcrop, with higher values for sam­
ples from its base, which are more thermally evolved than the others due
to igneous intrusive rock and the presence of catalyst clay minerals.

CRediT author statement

Caroline Adolphsson do Nascimento: Formal analysis, Investiga­


tion, Writing - Original Draft, conducted the experimental work, which
was developed during her PhD; Eliane Soares de Souza: Conceptuali­
zation, Resources, Writing - Review & Editing, Supervision, Project
administration; Laercio Lopes Martins: Writing - Review & Editing,
Supervision; Hélio Jorge Portugal Severiano Ribeiro: Resources, Su­
pervision, Funding acquisition; Victor Hugo Santos: Investigation;
René Rodrigues: Investigation, Writing - Review & Editing.

Declaration of competing interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influence
the work reported in this paper.

Acknowledgments

The authors are grateful for support from the North Fluminense State
University – UENF (Macaé, Brazil), the Laboratory of Petroleum Studies
(LEPETRO) at the Bahia Federal University, and the Laboratory of
Chemostratigraphy and Organic Geochemistry at the Rio de Janeiro
State University (UERJ). LEPETRO provided TOC and Rock-Eval pyrol­
ysis analyses, and the Laboratory of Chemostratigraphy and Organic
Geochemistry at UERJ provided δ13C analysis. We also thank CAPES for
the doctoral scholarship of Caroline Adolphsson do Nascimento. Finally,
the authors thank BG E&P Brasil Ltd. for sponsorship that allowed the
field trip for sample collection through the “Commitment to Research
Fig. 16. Plot of (a) C32-33/C32-35 benzohopanes vs. C29 ββ/(ββ + αα), (b) C29 and Development Investments” with the ANP (National Agency of Pe­
20S/(20S + 20R), and (c) C31 22S/(22S + 22R) showing samples AMa 1, 5, and troleum, Natural Gas and Biofuels). We also thank the Section Editor Hui
7 more thermally evolved than the other samples. Tian, the reviewer Kenneth Peters, and an anonymous reviewer for the
relevant comments that improved the quality of this manuscript.

14
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Appendix A

15
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Appendix B

16
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

Appendix C. Supplementary data

Supplementary data to this article can be found online at https://doi.org/10.1016/j.marpetgeo.2021.104917.

References Hachiro, J., 1997. O Subgrupo Irati (Neopermiano) da Bacia do Paraná. Instituto de
Geociências. Universidade de São Paulo 196. https://doi.org/10.11606/T.44.1997.
tde-11032013-164814.
Afonso, J.C., Schmal, M., Cardoso, J.N., 1994. Hydrocarbon distribution in the Irati shale
He, W., Lu, S., 1990. A new maturity parameter based on monoaromatic hopanoids. Org.
oil. Fuel 73, 363–366. https://doi.org/10.1016/0016-2361(94)90088-4.
Geochem. (Tokyo. 1967) 16, 1007–1013. https://doi.org/10.1016/0146-6380(90)
Alexander, R., Kagi, R., Toh, E., Van Bronswijk, W., 1988. The use of aromatic
90137-O.
hydrocarbons for assessment of thermal histories of sediments. In: Purcell, P.G.,
Hoffmann, C.F., Mackenzie, A.S., Lewis, C.A., Maxwell, J.R., Oudin, J.L., Durand, B.,
Purcell, R.R. (Eds.), The North West Shelf, Australia: Proceedings of the Petroleum
Vandenbroucke, M., 1984. A biological marker study of coals, shales and oils from
Exploration Society of Australia. Petroleum Exploration Society of Australia, Perth,
the Mahakam Delta, Kalimantan,. Indonesia. Chem. Geol. 42, 1–23. https://doi.org/
pp. 559–562.
10.1016/0009-2541(84)90002-0.
Alferes, C., Rodrigues, R., Pereira, E., 2011. Geoquímica orgânica aplicada à Formação
Holanda, W., dos Santos, A.C., Bertolino, L.C., Bergamaschi, S., Rodrigues, R., da
Irati, na área de São Mateus do Sul (PR). Brasil. Geochim. Bras 25, 47–54. https://
Costa, D.F., Jones, C.M., 2019. Paleoenvironmental, paleoclimatic and stratigraphic
doi.org/10.21715/gb.v25i1.331.
implications of the mineralogical content of the Irati Formation, Paraná Basin,
Amaral, S.E., 1971. Geologia e petrologia da Formação Irati (Permiano) no Estado de São
Brazil. J. S. Am. Earth Sci. 94, 102243 https://doi.org/10.1016/j.
Paulo. Boletim do Instituto de Geociências e Astronomia da Universidade de São
jsames.2019.102243.
Paulo 2, 3–81.
Holba, A.G., Tegelaar, E., Ellis, L., Singletary, M.S., Albrecht, P., 2000. Tetracyclic
Araújo, B.Q., Neto, F.R.A., Azevedo, D.A., 2018. Occurrence of extended tetracyclic
polyprenoids: indicators of freshwater (lacustrine) algal input. Geology 28, 251–254
polyprenoid series in crude oils. Org. Geochem. 118, 27–35. https://doi.org/
https://doi.org/10.1130/0091-7613(2000)28<251:TPIOFL>2.0.CO;2.
10.1016/j.orggeochem.2018.01.006.
Holba, A.G., Dzou, L.I., Wood, G.D., Ellis, L., Adam, P., Schaeffer, P., Albrecht, P.,
Araújo, L.M., 2001. Análise da expressão estratigráfica dos parâmetros de geoquímica
Greene, T., Hughes, W.B., 2003. Application of tetracyclic polyprenoids as indicators
orgânica e inorgânica nas sequências deposicionais Irati. PhD thesis. In: Instituto de
of input from fresh-brackish water environments. Org. Geochem. 34, 441–469.
Geociências, vol. 2. Universidade Federal do Rio Grande do Sul.
https://doi.org/10.1016/S0146-6380(02)00193-6.
Araújo, L.M., Rodrigues, R., Scherer, C.M.S., 2001. Sequências deposicionais Irati:
Horsfield, B., Dembicki, H., Ho, T.T.Y., 1983. Some potential applications of pyrolysis to
arcabouço químio-estratigráfico e inferências paleoambientais. Ciência-Técnica-
basin studies. J. Geol. Soc. London 140, 431–443. https://doi.org/10.1144/
Petróleo 20, 193–202.
gsjgs.140.3.0431.
Araújo, L.M., Triguis, J.A., Cerqueira, J.R., Freitas, L.D.S., 2000. The atypical Permian
Huang, W., Meinschein, W.G., 1979. Sterols as ecological indicators. Geochem.
petroleum system of the Paraná Basin, Brazil. In: Mello, M.R., Katz, B.J. (Eds.),
Cosmochim. Acta 43, 739–745. https://doi.org/10.1016/0016-7037(79)90257-6.
Petroleum Systems of South Atlantic Margins, vol. 73. AAPG Memoir, pp. 377–402.
Hussler, G., Connan, J., Albrecht, P., 1984. Novel families of tetra- and hexacyclic
Barbosa, O., Gomes, F.A., 1958. Pesquisa de petróleo na bacia do Rio Corumbataí.
aromatic hopanoids predominant in carbonate rocks and crude oils. Org. Geochem.
DNPM/DGM, Rio de Janeiro. Boletim, 171.
6, 39–49. https://doi.org/10.1016/0146-6380(84)90025-1.
Birgel, D., Peckmann, J., 2008. Aerobic methanotrophy at ancient marine methane seeps:
Jiang, Z.S., Fowler, M.G., 1986. Carotenoid-derived alkanes in oils from northwestern
a synthesis. Org. Geochem. (Tokyo. 1967) 39, 1659–1667. https://doi.org/10.1016/
China. Org. Geochem. (Tokyo. 1967) 10, 831–839. https://doi.org/10.1016/S0146-
j.orggeochem.2008.01.023.
6380(86)80020-1.
Blumer, M., 1957. Removal of elemental sulfur from hydrocarbon fractions. Anal. Chem.
Johnson, S.C., 1967. Hierarchical clustering schemes. Psychometrika 32, 241–254
29, 1039–1041.
https://doi.org/10.1007/BF02289588.
Cerqueira, J.R., Santos Neto, E.V., 1986. Papel das intrusões de diabásio no processo de
Kodner, R.B., Pearson, A., Summons, R.E., Knoll, A.H., 2008. Sterols in red and green
geração de hidrocarbonetos na Bacia do Paraná. In: 3◦ Congresso Brasileiro de
algae: quantification, phylogeny, and relevance for the interpretation of geologic
Petróleo, Rio de Janeiro, vol. 73, pp. 1–15.
steranes. Geobiology 6, 411–420. https://doi.org/10.1111/j.1472-
Chen, J.H., Philp, R.P., Fu, J.M., Sheng, G.Y., 1989. The occurrence and identification of
4669.2008.00167.x.
C30-C32 lanostanes: a novel series of tetracyclic triterpenoid hydrocarbons. Geochem.
Lamb, D.C., Jackson, C.J., Warrilow, A.G., Manning, N.J., Kelly, D.E., Kelly, S.L., 2007.
Cosmochim. Acta 53, 2775–2779. https://doi.org/10.1016/0016-7037(89)90148-8.
Lanosterol biosynthesis in the prokaryote Methylococcus capsulatus: insight into the
Clifford, D.J., Clayton, J.L., Sinninghe Damsté, J.S., 1998. 2, 3, 6-/3, 4, 5-Trimethyl
evolution of sterol biosynthesis. Mol. Biol. Evol. 24, 1714–1721. https://doi.org/
substituted diaryl carotenoid derivatives (Chlorobiaceae) in petroleums of the
10.1093/molbev/msm090.
Belarussian Pripyat River Basin. Org. Geochem. 29, 1253–1267. https://doi.org/
Lu, H., Sheng, G., Ma, Q., Lu, Z., 2011. Identification of C24 and C25 lanostanes in tertiary
10.1016/S0146-6380(98)00086-2.
sulfur rich crude oils from the jinxian sag, bohai bay basin, northern China. Org.
Correa da Silva, Z.C., Cornford, C., 1985. The kerogen type, depositional environment
Geochem. (Tokyo. 1967) 42, 146–155. https://doi.org/10.1016/j.
and maturity, of the Irati shale, upper permian of Paraná Basin, southern Brazil. Org.
orggeochem.2010.11.009.
Geochem. 8, 399–411. https://doi.org/10.1016/0146-6380(85)90018-X.
Lyons, T.W., Severmann, S., 2006. A critical look at iron paleoredox proxies: new insights
Corrêa, L.M.S.A., Pereira, E., 2005. Estudo da distribuição das intrusões mesozóicas e sua
from modern euxinic marine basins. Geochim. Cosmochim. Acta 70, 5698–5722.
relação com os sistemas petrolíferos da Bacia do Paraná. In: Simpósio de Vulcanismo
https://doi.org/10.1016/j.gca.2006.08.021.
e Ambientes Associados, vol. 3, pp. 21–26.
Mackenzie, A.S., Brassell, S.C., Eglinton, G., Maxwell, J.R., 1982. Chemical fossils: the
Cox, R.E., Yamamoto, S., Otto, A., Simoneit, B.R., 2007. Oxygenated di-and tricyclic
geological fate of steroids. Science 217, 491–504. https://doi.org/10.1126/
diterpenoids of southern hemisphere conifers. Biochem. Systemat. Ecol. 35,
science.217.4559.491.
342–362. https://doi.org/10.1016/j.bse.2006.09.013.
Martinez, A.M., Boyer, D.L., Droser, M.L., Barrie, C., Love, G.D., 2018. A stable and
Dembicki Jr., H., 1992. The effects of the mineral matrix on the determination of kinetic
productive marine microbial community was sustained through the end-Devonian
parameters using modified Rock-Eval pyrolysis. Org. Geochem. 18, 531–539.
Hangenberg Crisis within the Cleveland Shale of the Appalachian Basin, United
https://doi.org/10.1016/0146-6380(92)90116-F.
States. Geobiology 17, 1–16. https://doi.org/10.1111/gbi.12314.
Droste, J.B., 1961. Clay minerals in sediments of owens, China, searles. Panamint,
Martins, C.M., Cerqueira, J.R., Ribeiro, H.J.P., Garcia, K.S., da Silva, N.N., Queiroz, A.F.
bristol, cadiz, and danby lake basins, California. Geol. Soc. Am. Bull. 72, 1713–1721
S., 2020a. Evaluation of thermal effects of intrusive rocks on the kerogen present in
https://doi.org/10.1130/0016-7606(1961)72[1713:CMISOO]2.0.CO;2.
the black shales of Irati Formation (Permian), Paraná Basin, Brazil. J. South Am.
Espitalié, J., Deroo, G., Marquis, F., 1986. La pyrolyse Rock-Eval et ses applications.
Earth Sci 100, 102559. https://doi.org/10.1016/j.jsames.2020.102559.
Troisième partie. Rev. Inst. Fr. Petrol 41, 73–89. https://doi.org/10.2516/ogst:
Martins, L.L., Pudenzi, M.A., Da Cruz, G.F., Nascimento, H.D.L., Eberlin, M.N., 2017.
1986003.
Assessing biodegradation of Brazilian crude oils via characteristic profiles of O1 and
Faure, K., Cole, D., 1999. Geochemical evidence for lacustrine microbial blooms in the
O2 compound classes: petroleomics by negative-ion mode electrospray ionization
vast Permian Main Karoo, Paraná, Falkland Islands and Huab basins of southwestern
fourier transform ion cyclotron resonance mass spectrometry. Energy Fuel. 31,
Gondwana. Palaeogeogr. Palaeoclimatol. Palaeoecol. 152, 189–213. https://doi.org/
6649–6657. https://doi.org/10.1021/acs.energyfuels.7b00109.
10.1016/S0031-0182(99)00062-0.
Martins, L.L., Schulz, H.M., Ribeiro, H.J.P.S., Nascimento, C.A., de Souza, E.S., da
Gama Jr., E., 1979. A sedimentação do Grupo Passa Dois (exclusive Formação Irati): um
Cruz, G.F., 2020b. Organic geochemical signals of freshwater dynamics controlling
modelo geomórfico. Rev. Bras. Geociencias 9, 1–16.
salinity stratification in organic-rich shales in the Lower Permian Irati Formation
Goldberg, K., Humayun, M., 2016. Geochemical paleoredox indicators in organic-rich
(Paraná Basin, Brazil). Org. Geochem. 140, 103958 https://doi.org/10.1016/j.
shales of the Irati Formation, permian of the Paraná Basin, southern Brazil. Brazilian
orggeochem.2019.103958.
J. Geol 46, 377–393. https://doi.org/10.1590/2317-4889201620160001.
Martins, L.L., Schulz, H.M., Ribeiro, H.J.P.S., Nascimento, C.A., de Souza, E.S., da
Grice, K., Audino, M., Boreham, C.J., Alexander, R., Kagi, R.I., 2001. Distributions and
Cruz, G.F., 2020c. Cadalenes and norcadalenes in organic-rich shales of the Permian
stable carbon isotopic compositions of biomarkers in torbanites from different
Irati Formation (Paraná Basin, Brazil): tracers for terrestrial input or also indicators
palaeogeographical locations. Org. Geochem. 32, 1195–1210. https://doi.org/
of temperature-controlled organic-inorganic interactions? Org. Geochem. 140,
10.1016/S0146-6380(01)00087-0.
103962 https://doi.org/10.1016/j.orggeochem.2019.103962.
Grice, K., Cao, C., Love, G.D., Böttcher, M.E., Twitchett, R.J., Grosjean, E., Summons, R.
Mello, M.R., Gaglianone, P.C., Brassell, S.C., Maxwell, J.R., 1988. Geochemical and
E., Turgeon, S.C., Dunning, W., Jin, Y., 2005. Photic zone euxinia during the
biological marker assessment of depositional environments using Brazilian offshore
Permian-Triassic superanoxic event. Science 307, 706–709. https://doi.org/
10.1126/science.1104323.

17
C.A. Nascimento et al. Marine and Petroleum Geology 126 (2021) 104917

oils. Mar. Petrol. Geol. 5, 205–223. https://doi.org/10.1016/0264-8172(88)90002- on black shales and limestones of the Permian Irati Formation. J. South Am. Earth
5. Sci. 28, 14–24. https://doi.org/10.1016/j.jsames.2008.12.002.
Mello, M.R., Koutsoukos, E.A., Neto, E.V.S., Telles Jr., A.C.S., 1993. Geochemical and Santos, R.V., Souza, P.A., de Alvarenga, C.J.S., Dantas, E.L., Pimentel, M.M., de
micropaleontological characterization of lacustrine and marine hypersaline Oliveira, C.G., de Araújo, L.M., 2006. Shrimp U–Pb zircon dating and palynology of
environments from Brazilian sedimentary basins. In: Katz, B.J., Pratt, L.M. (Eds.), bentonitic layers from the permian Irati Formation, Paraná Basin. Brazil. Gondwana
Source Rocks in a Sequence Stratigraphic Framework. American Association of Res. 9, 456–463. https://doi.org/10.1016/j.gr.2005.12.001.
Petroleum Geologists, pp. 17–34. Tulsa. Schaeffer, P., Adam, P., Wehrung, P., Albrecht, P., 1997. Novel aromatic carotenoid
Mendes, J.C., Fúlfaro, V.J., Amaral, S.E.D., Landim, P.M.B., 1966. A Formação Irati derivatives from sulfur photosynthetic bacteria in sediments. Tetrahedron Lett. 38,
(Permiano) e fácies associadas. Boletim Soc. Bras. Geol. 15, 23–43. 8413–8416. https://doi.org/10.1016/S0040-4039(97)10235-0.
Meyers, P.A., 1994. Preservation of elemental and isotopic source identification of Schwark, L., Empt, P., 2006. Sterane biomarkers as indicators of palaeozoic algal
sedimentary organic matter. Chem. Geol. 114, 289–302. https://doi.org/10.1016/ evolution and extinction events. Palaeogeogr. Palaeoclimatol. Palaeoecol. 240,
0009-2541(94)90059-0. 225–236. https://doi.org/10.1016/j.palaeo.2006.03.050.
Milani, E.J., 2004. Comentários sobre a origem e evolução tectônica da Bacia do Paraná. Schwark, L., Frimmel, A., 2004. Chemostratigraphy of the Posidonia Black Shale, SW-
In: Mantessoneto, V., Bartorelli, A., Carneiro, C.D.R., Brito-Neves, B.B. (Eds.), Germany: II. Assessment of extent and persistence of photic-zone anoxia using aryl
Geologia do Continente Sul-Americano: Evolução da Obra de Fernando Flávio isoprenoid distributions. Chem. Geol. 206, 231–248. https://doi.org/10.1016/j.
Marques de Almeida. São Paulo, pp. 265–279. chemgeo.2003.12.008.
Milani, E.J., de Melo, J.H.G., de Souza, P.A., Fernandes, L.A., Franca, A.B., 2007. Bacia Seifert, I.A., 2013. Estudo detalhado da Formação Irati na área de Anhembi (Poço FP-12-
do Paraná. Bol. Geociencias Petrobras 15, 265–287. SP) e sua correlação com ocorrências de arenito asfáltico da Fazenda Betumita.
Milani, E.J., Zalan, P.V., 1999. An outline of the geology and petroleum systems of the Dissertação (Mestrado). Universidade do Estado do Rio de Janeiro 107.
Paleozoic interior basins of South America. Episodes-Newsmagazine Int. Union Geol. Seifert, W.K., Moldowan, J.M., 1986. Use of biological markers in petroleum exploration.
Sci. 22, 199–205. Methods Geochem. Geophys. 24, 261–290.
Moldowan, J.M., Seifert, W.K., Gallegos, E.J., 1985. Relationship between petroleum Seifert, W.K., Moldowan, J.M., 1980. The effect of thermal stress on source-rock quality
composition and depositional environment of petroleum source rocks. Am. Assoc. as measured by hopane stereochemistry. Phys. Chem. Earth 12, 229–237 https://doi.
Petrol. Geol. Bull. 69, 1255–1268. https://doi.org/10.1306/AD462BC8-16F7-11D7- org/10.1016/0079-1946(79)90107–1.
8645000102C1865D. Silva, R.S.F., Aguiar, H.G.M., Rangel, M.D., Azevedo, D.A., Aquino Neto, F.R., 2011.
Nascimento, C.A., Souza, E.S., Severiano Ribeiro, H.J.P., Santos, V.H., Martins, L.L., Comprehensive two-dimensional gas chromatography with time of flight mass
2018. Assessing the physicochemical conditions of the depositional spectrometry applied to biomarker analysis of oils from Colombia. Fuel 90,
paleoenvironment of the Irati Formation in two outcrops in São Mateus do Sul City. 2694–2699. https://doi.org/10.1016/j.fuel.2011.04.026.
In: XV Latin American Congress on Organic Geochemistry. Paraná State, Brazil. Simoneit, B.R.T., Didyk, B.M., 1986. Paraffin dirt: bacterial biomass from natural gas, not
Nytoft, H.P., Vuković, N.S., Kildahl-Andersen, G., Rise, F., Životić, D.R., Stojanović, K.A., weathered petroleum. In: workshop on advances in biomarkers and kerogens.
2016. Identification of a novel series of benzohopanes and their geochemical Academia sinica, institute of geochemistry, guiyang. China 131–132.
significance. Energy Fuel. 30, 5563–5575. https://doi.org/10.1021/acs. Sinninghe, Damsté J.S., Kenig, F., Koopmans, M.P., Koster, J., Schouten, S., Hayes, J.M.,
energyfuels.6b00799. de Leeuw, J.W., 1995. Evidence for gammacerane as an indicator of water column
Oba, M., Nakamura, M., Fukuda, Y., Katabuchi, M., Takahashi, S., Haikawa, M., stratification. Geochem. Cosmochim. Acta 59, 1895–1900. https://doi.org/10.1016/
Kaiho, K., 2009. Benzohopanes and diaromatic 8(14)-secohopanoids in some Late 0016-7037(95)00073-9.
Permian carbonates. Geochem. J. 43, 29–35. https://doi.org/10.2343/ Sinninghe, Damsté J.S., Köster, J., Baas, M., Ossebaar, J., Dekker, M., Pool, W.,
geochemj.1.0004. Geenevasen, J.A.J., 1999. A sedimentary tetrahydrophenanthrene derivative of
Osorio, L.L., dos Reis, D.E.S., Rodrigues, R., 2017. Aromatic steroids biomarkers applied tetrahymanol. Tetrahedron Lett. 40, 3949–3952. https://doi.org/10.1016/S0040-
to high resolution stratigraphy: Irati Formation, southern of Paraná Basin. Brazil. J. 4039(99)00618-8.
Sediment. Environ 2, 274–282. https://doi.org/10.12957/jse.2017.32495. Sousa, Júnior G.R., Santos, A.L.S., Lima, S.G., Lopes, J.A.D., Reis, F.A.M., Santos Neto, E.
Otto, A., Wilde, V., 2001. Sesqui-, di-, and triterpenoids as chemosystematic markers in V., Chang, H.K., 2013. Organic geochemistry evidence for euphotic zone anoxia
extant conifers - a Review. Bot. Rev. 67, 141–238. https://doi.org/10.1007/ during the deposition of Aptian source rocks based on aryl isoprenoids in petroleum,
BF02858076. Sergipe-Alagoas Basin, northeastern Brazil. Org. Geochem. (Tokyo. 1967) 63,
Parfenova, T.M., 2011. Lanostanes in cambrian organic matter (southeastern part of the 94–104. https://doi.org/10.1016/j.orggeochem.2013.07.009.
siberian platform). Geochemistry 436, 143–147. https://doi.org/10.1134/ Strachan, M.G., Alexander, R., Kagi, R.I., 1986. Trimethylnaphthalenes as depositional
S1028334X10901180. environment indicators. In: 192nd ACS National Meeting, Div. Of Geochem, p. 49.
Peckmann, J., Thiel, V., Reitner, J., Taviani, M., Aharon, P., Michaelis, W., 2004. Summons, R.E., Powell, T.G., 1986. Chlorobiaceae in Palaeozoic seas revealed by
A microbial mat of a large sulfur bacterium preserved in a Miocene methane-seep biological markers, isotopes and geology. Nature 319, 763–765. https://doi.org/
limestone. Geomicrobiol. J. 21, 247–255. https://doi.org/10.1080/ 10.1038/319763a0.
01490450490438757. Syms, C., 2008. Principal component analysis. Encycl. Ecol 2940–2949.
Peng, P., Morales-Izquierdo, A., Fu, J., Sheng, G., Hogg, A., Strausz, O.P., 1998. ten Haven, H.L., de Leeuw, J.W., Sinninghe Damsté, J., Schenck, P.A., Palmer, S.E.,
Lanostane sulfides in an immature crude oil. Org. Geochem. 28, 125–134. https:// Zumberge, J.E., 1988. Application of biological markers in the recognition of
doi.org/10.1016/S0146-6380(97)00112-5. palaeohypersaline environments. In: Fleet, A.J., Kelts, K., Talbot, M.R. (Eds.),
Peters, K.E., Fraser, T.H., Amris, W., Rustanto, B., Hermanto, E., 1999. Geochemistry of Lacustrine Petroleum Source Rocks, vol. 40. Geological Society Special Publication,
crude oils from. Eastern Indonesia. Am. Assoc. Petroleum Geol. Bull. 83, 1927–1942. pp. 123–130. https://doi.org/10.1144/GSL.SP.1988.040.01.11.
Petri, S., Coimbra, A.M., 1982. Estruturas sedimentares das Formações Irati e estrada Thomaz Filho, A., Mizusaki, A.M.P., Antonioli, L., 2008. Magmatism and petroleum
Nova (Permiano) e sua contribuição para elucidação dos seus paleoambientes exploration in the Brazilian Paleozoic basins. Mar. Petrol. Geol. 25, 143–151.
geradores, Brasil. In: congresso Latinoamericano de Geologia 5. Acta 2, 353–371. https://doi.org/10.1016/j.marpetgeo.2007.07.006.
R Core team, 2013. R: a language and environment for statistical computing. In: Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence. Springer Science &
R Foundation for Statistical Computing, Vienna, Austria. http://R-project.org/. Business Media.
Reis, D.E., Rodrigues, R., Moldowan, J.M., Jones, C.M., Brito, M., da Costa Tulipani, S., Grice, K., Greenwood, P.F., Schwark, L., Böttcher, M.E., Summons, R.E.,
Cavalcante, D., Portela, H.A., 2018. Biomarkers stratigraphy of Irati Formation Foster, C.B., 2015. Molecular proxies as indicators of freshwater incursion-driven
(lower permian) in the southern portion of Paraná Basin (Brazil). Mar. Petrol. Geol. salinity stratification. Chem. Geol. 409, 61–68. https://doi.org/10.1016/j.
95, 110–138. https://doi.org/10.1016/j.marpetgeo.2018.04.007. chemgeo.2015.05.009.
Ricardi-Branco, F., Torres, F., Rösler, O., 2013. Early permian conifers paranocladus and Volkman, J.K., Barrett, S.M., Dunstan, G.A., 1994. C25 and C30 highly branched
buriadia of southern Brazil. Terræ 10, 3–14. isoprenoid alkenes in laboratory cultures of two marine diatoms. Org. Geochem. 21,
Rocha-Campos, A.C., Basei, M.A., Nutman, A.P., Kleiman, L.E., Varela, R., Llambias, E., 407–413. https://doi.org/10.1016/0146-6380(94)90202-X.
Canile, F.M., da Rosa, O., de, C.R., 2011. 30 million years of Permian volcanism Warren, L.V., Quaglio, F., Simões, M.G., Assine, M.L., Alessandretti, L., Luvizotto, G.L.,
recorded in the Choiyoi igneous province (W Argentina) and their source for younger Riccomini, C., Stríkis, N.M., 2017. A Permian methane seep system as a
ash fall deposits in the Paraná Basin: SHRIMP U–Pb zircon geochronology evidence. paleoenvironmental analogue for the pre-metazoan carbonate platforms. Brazilian J.
Gondwana Res. 19, 509–523. https://doi.org/10.1016/j.gr.2010.07.003. Geol. 47, 722–733. https://doi.org/10.1590/2317-4889201720170125.
Rodrigues, R., Pereira, E., Bergamaschi, S., Chaves, H.A.F., Alferes, C.L.F., 2010. Organic Warren, J., 1999. Evaporites. Their Evolution and Economics. Blackwell Science, Oxford,
geochemistry of Irati Formation, lower permian of Paraná Basin. In: Memórias XII p. 438.
Congresso Latinoamericano de Geoquímica Orgânica, pp. 41–43. Yang, S., Holsfield, B., 2016. Some predicted effects of minerals on the generation of
Rodrigues, R., Quadros, L.D., 1976. Mineralogia de argilas e teor de Boro das formações petroleum in nature. Energy Fuel. 30, 6677–6687. https://doi.org/10.1021/acs.
paleozóicas da Bacia do Paraná. In: Anais do Congresso Brasileiro de Geologia, vol. energyfuels.6b00934.
29, pp. 351–379. Zalán, P.V., Wolff, S., Astolfi, M.A.M., Vieira, I.S., Concelcao, J.C.J., Appi, V.T.,
Santos Neto, E.V., Cerqueira, J.R., 1993. Aplicação da geoquímica orgânica na Marques, A., 1990. The parana basin, Brazil. In: Leighton, M.W., Kolata, D.R.,
cronoestratigrafia e paleogeografia da Formação Irati, Bacia do Paraná,. In: Simpósio Oltz, D.F., Eidel, J.J. (Eds.), Interior Cratonic Basins, vol. 51. AAPG Memoir,
sobre Cronoestratigrafia da Bacia do Paraná 1, Rio Claro, Resumes, vol. 71. pp. 681–708.
Santos, R.V., Dantas, E.L., de Oliveira, C.G., de Alvarenga, C.J.S., dos Anjos, C.W.D.,
Guimarães, E.M., Oliveira, F.B., 2009. Geochemical and thermal effects of a basic sill

18

You might also like