Hostname: page-component-848d4c4894-m9kch Total loading time: 0 Render date: 2024-05-21T01:02:32.245Z Has data issue: false hasContentIssue false

Organic-walled microfossils from the Ediacaran Sete Lagoas Formation, Bambuí Group, Southeast Brazil: taxonomic and biostratigraphic analyses

Published online by Cambridge University Press:  18 March 2024

Matheus Denezine*
Affiliation:
Institute of Geosciences, University of Brasília, Brasília, Federal District 73105-909, Brazil , Department of Earth Sciences, Virginia Tech, Blacksburg, Virginia 24061, USA
Dermeval Aparecido do Carmo
Affiliation:
Institute of Geosciences, University of Brasília, Brasília, Federal District 73105-909, Brazil ,
Shuhai Xiao
Affiliation:
Department of Earth Sciences, Virginia Tech, Blacksburg, Virginia 24061, USA
Qing Tang
Affiliation:
School of Earth Sciences and Engineering, Nanjing University, Nanjing 210023, China
Vladmir Sergeev
Affiliation:
Geological Institute, Russian Academy of Sciences, Pyzhevskii per. 7, Moscow, 119017 Russia
Alysson Fernandes Mazoni
Affiliation:
Institute of Geosciences, University of Campinas, Campinas, São Paulo 13083-855, Brazil ,
Carolina Zabini
Affiliation:
Institute of Geosciences, University of Campinas, Campinas, São Paulo 13083-855, Brazil ,
*
*Corresponding author.

Abstract

This work presents a detailed taxonomic study on organic-walled microfossils from the Ediacaran Sete Lagoas Formation (Bambuí Group) at the Barreiro section in the Januária area of the São Francisco basin, Brazil. Seven species are described, including Siphonophycus robustum (Schopf, 1968), Ghoshia januarensis new species, Leiosphaeridia crassa (Naumova, 1949), Leiosphaeridia jacutica (Timofeev, 1966), Leiosphaeridia minutissima (Naumova, 1949), Leiosphaeridia tenuissima Eisenack, 1958, and Germinosphaera bispinosa Mikhailova, 1986. These taxa are recovered for the first time in the Sete Lagoas Formation. They occur abundantly in the lower portion of the studied section, but only Ghoshia januarensis is present in the upper part of the studied section, probably due to environmental or taphonomic changes. Leiosphaeridia species, particularly Leiosphaeridia minutissima, dominate the organic-walled microfossil assemblage. Although most taxa described here have long stratigraphic ranges, they are consistent with a terminal Ediacaran age as inferred from detrital zircon data and tubular fossils (e.g., Cloudina and Corumbella) from the Sete Lagoas Formation.

UUID: http://zoobank.org/7f92b900-0176-4da6-93a3-fd51edb22cbf

Type
Articles
Creative Commons
Creative Common License - CCCreative Common License - BY
This is an Open Access article, distributed under the terms of the Creative Commons Attribution licence (http://creativecommons.org/licenses/by/4.0/), which permits unrestricted re-use, distribution and reproduction, provided the original article is properly cited.
Copyright
Copyright © The Author(s), 2024. Published by Cambridge University Press on behalf of Paleontological Society

Non-technical Summary

Seven species occur in shallow-marine limestone of the Sete Lagoas Formation, Bambuí Group, in Januária, Brazil, including Siphonophycus robustum, Leiosphaeridia crassa, Leiosphaeridia jacutica, Leiosphaeridia minutissima, Leiosphaeridia tenuissima, Germinosphaera bispinosa, and a new species named Ghoshia januarensis. In the lower part of the studied section, these occurrences are common, but only Ghoshia januarensis is found in the upper part. This is likely due to changes in the environment or preservation conditions. The Leiosphaeridia species, especially Leiosphaeridia minutissima, dominates the assemblage of organic-walled microfossils. While most described taxa have long stratigraphic ranges, they are consistent with a terminal Ediacaran age, as indicated by detrital zircon data and tubular fossils (e.g., Cloudina and Corumbella) from the Sete Lagoas Formation.

Introduction

The present work provides an updated taxonomic description and biostratigraphic analysis of organic-walled microfossils from the Sete Lagoas Formation, Bambuí Group, from the countryside of Januária Municipality, Minas Gerais State, Brazil. The studied area is an ancient quarry area with an exposure of nearly 70 m of a continuous succession of mixed carbonate and fine-grained siliciclastic rocks. In addition, the Januária area has been the focus of the chronostratigraphic investigation of the Sete Lagoas Formation since Warren et al. (Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014) reported Ediacaran tubular fossils such as Cloudina sp. and Corumbella werneri Hahn et al., Reference Hahn, Hahn, Leonardos, Pflug and Walde1982 from the lower Sete Lagoas Formation in the Januária area.

Several articles on organic-walled microfossils from the Sete Lagoas Formation have been previously published (Sommer, Reference Sommer1971; Simonetti and Fairchild, Reference Simonetti and Fairchild1989, Reference Simonetti and Fairchild2000; Fairchild et al., Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996, Reference Fairchild, Sanchez, Pacheco and Leme2012; Sanchez and Fairchild, Reference Sanchez and Fairchild2018), describing dozens of species from this unit (Table 1), although Sanchez and Fairchild (Reference Sanchez and Fairchild2018) invalidated one species, Bambuites erichsenii Sommer, Reference Sommer1971. It is important to underscore that the investigation conducted by Fairchild et al. (Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996) centered on the analysis of silicified carbonaceous microfossils in petrographic thin sections from the Sete Lagoas Formation in the western portion of the São Francisco craton. Specimens illustrated by Fairchild et al. (Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996) were tentatively compared with seven species, alongside several taxa left under open nomenclature. This study provides a detailed taxonomic description with an analysis of the diversity, abundance, and stratigraphic occurrence of organic-walled microfossils from the Sete Lagoas Formation in the Barreiro section in the Januária area.

Table 1. List of microfossils from the Sete Lagoas Formation, Bambuí Group, published previously and in this study. Articles: 1 = Sommer (Reference Sommer1971); 2 = Marchese (Reference Marchese1974); 3 = Simonetti and Fairchild (Reference Simonetti and Fairchild1989); 4 = Fairchild et al. (Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996); 5 = Simonetti and Fairchild (Reference Simonetti and Fairchild2000); 6 = Fairchild et al. (Reference Fairchild, Sanchez, Pacheco and Leme2012); 7 = Warren et al. (Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014); 8 = Perrella Júnior et al. (Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017); 9 = Sanchez and Fairchild (Reference Sanchez and Fairchild2018); 10 = Denezine et al. (Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022); 11 = this study.

The remarkable diversity of organic-walled microfossils in the Ediacaran Period (Knoll, Reference Knoll1994; Vidal and Moczydłowska-Vidal, Reference Vidal and Moczydłowska-Vidal1997; Huntley et al., Reference Huntley, Xiao and Kowalewski2006) is hypothesized to be associated with the ecological rise of animals (Peterson and Butterfield, Reference Peterson and Butterfield2005), followed by the advent of skeletonized animals as evidenced by fossil cloudinids and corumbellids (Germs, Reference Germs1972; Hahn et al., Reference Hahn, Hahn, Leonardos, Pflug and Walde1982; Hua et al., Reference Hua, Chen, Yuan, Zhang and Xiao2005; Walde et al., Reference Walde, Do Carmo, Guimarães, Vieira, Erdtmann, Sanchez, Adôrno and Tobias2015; Adôrno et al., Reference Adôrno, Do Carmo, Germs, Walde and Denezine2017). The transition from the Ediacaran to the Cambrian Period is marked by a significant turnover of acritarch species (Anderson et al., Reference Anderson, Macdonald, Jones, McMahon and Briggs2017; Grazhdankin et al., Reference Grazhdankin, Nagovitsin, Golubkova, Karlova, Kochnev, Rogov and Marusin2020; Morais et al., Reference Morais, Fairchild, Freitas, Rudnitzki, Silva, Lahr, Moreira, Abrahão Filho, Leme and Trindade2021). The Cambrian is characterized by a diversification of acanthomorphs species compared with the sphaeromorph dominance in the late Ediacaran (Gaucher and Sprechmann, Reference Gaucher and Sprechmann2009). Acanthomorphic acritarchs define four acritarch assemblage zones to recognize the lower Cambrian on the East European Platform (Moczydłowska, Reference Moczydłowska1991): Asteridium tornatumComasphaeridium velvetum Assemblage Zone, Skiagia ornataFimbriaglomerella membranacea Assemblage Zone, Heliosphaeridium dissimilareSkiagia ciliosa Assemblage Zone, and Volkovia dentiferaLiepaina plana Assemblage Zone. These biostratigraphic units have been used to correlate Cambrian successions around the world (Zheng et al., Reference Zheng, Clausen, Feng and Servais2020). Therefore, a systematic study of organic-walled microfossils from the Sete Lagoas Formation not only provides a tool for biostratigraphic correlation but also offers useful data to improve the understanding of Ediacaran evolution.

Geological setting

The Januária Municipality is located in the central part of the São Francisco craton (Fig. 1). The paleogeographic location of the São Francisco craton during the Ediacaran Period has not been precisely constrained, mainly because of poor biostratigraphic and paleomagnetic data. However, Merdith et al. (Reference Merdith, Williams, Collins, Tetley and Mulder2021) placed the São Francisco craton in high latitudes during the Ediacaran Period.

Figure 1. (1) Geological map of the São Francisco basin (red dashed line) in the São Francisco craton, showing its relationship with neighboring Neoproterozoic fold belts. Inset map shows major cratons in the western Gondwana in a Neoproterozoic paleogeographic configuration: A = Amazonian craton; P = Rio de la Plata craton; K = Kalahari craton; WA = West Africa craton; SFC = São Francisco-Congo craton; PC = Paramirim Corridor. Modified from Reis and Alkmim (Reference Reis and Alkmim2015). (2) Stratigraphic position of the studied section in the Bambui Group columnar section. (3) Geological map of the studied area. The purple dot marks the location of the studied section.

The western portion of the São Francisco craton comprises a succession of siliciclastic and carbonate rocks dated between 1.77 Ga and 0.56 Ga (Pimentel et al., Reference Pimentel, Rodrigues, DellaGiustina, Junges, Matteini and Armstrong2011; Alvarenga et al., Reference Alvarenga, Dardenne, Vieira, Martinho, Guimarães, Santos and Santana2012). The Brasília fold belt that bounds the western margin of the São Francisco craton was deformed during the Brasiliano–Pan African orogeny between 790 Ma and 540 Ma (Pimentel and Fuck, Reference Pimentel and Fuck1992). This fold belt borders to the east with the São Francisco craton covered with undeformed Neoproterozoic strata. It consists of a tectonic domain where only the upper 2 km strata are deformed and a domain further west where both the basement and sediment cover are deformed (Alvarenga et al., Reference Alvarenga, Santos, Vieira, Lima and Mancini2014). A thick interval of Mesoproterozoic and Neoproterozoic sedimentary rocks was deposited along the west portion of the São Francisco craton. These strata are separated into three stratigraphic units, including, in ascending stratigraphic order, the Paranoá Group (Barbosa, Reference Barbosa1963), Jequitaí Formation (Oliveira and Leonardos, Reference Oliveira and Leonardos1943), and Bambuí Group (Rimann, Reference Rimann1917).

The Sete Lagoas Formation, which is the main focus of this work, represents the basal unit of the Bambuí Group and consists of a sequence of carbonate-dominated sediments in the São Francisco basin. Those sediments are characterized by a low total organic carbon (TOC) content of less than 2% (Uhlein et al., Reference Uhlein, Uhlein, Pereira, Caxito, Okubo, Warren and Sial2019; Caetano-Filho et al., Reference Caetano-Filho, Sansjofre, Ader, Paula-Santos, Guacaneme, Babinski, Bedoya-Rueda, Kuchenbecker, Reis and Trindade2021) and relatively low thermal maturity (Reis and Suss, Reference Reis and Suss2016). The δ13Ccarb and δ13Corg excursions indicate a disconnection between the São Francisco basin and the global carbon cycles, which would imply marine isolation and paleogeographic shifts driven by the dynamic changes in marginal orogenic systems (Caetano-Filho et al., Reference Caetano-Filho, Paula-Santos, Guacaneme, Babinski and Bedoya-Rueda2019, Reference Caetano-Filho, Sansjofre, Ader, Paula-Santos, Guacaneme, Babinski, Bedoya-Rueda, Kuchenbecker, Reis and Trindade2021; Guacaneme et al., Reference Guacaneme, Babinski, Bedoya–Rueda, Paula–Santos, Caetano–Filho, Kuchenbecker, Reis and Trindade2021).

The depositional age of the Bambuí Group has long been a matter of debate. The Bambuí Group was initially considered to be Cretaceous (Liais, Reference Liais1872 in Couto et al., Reference Couto, Cordani, Kawashita, Iyer and Moraes1981), but recent studies show that it is probably Ediacaran–Cambrian (Pimentel et al., Reference Pimentel, Rodrigues, DellaGiustina, Junges, Matteini and Armstrong2011; Warren et al., Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014; Paula-Santos et al., Reference Paula-Santos, Babinski, Kuchenbecker, Caetano-Filho, Trindade and Pedrosa-Soares2015; Moreira et al., Reference Moreira, Uhlein, Dussin, Uhlein and Misuzaki2020; Sanchez et al., Reference Sanchez, Uhlein and Fairchild2021; DaSilva et al., Reference DaSilva, Pufahl, James, Guimaraes and Reis2022). Geochronological constraints on the Bambuí Group are few and inconclusive. Carbonates of the lower Sete Lagoas Formation yielded Pb–Pb apparent ages of ~740 Ma (Babinski et al., Reference Babinski, Vieira and Trindade2007). However, Caxito et al. (Reference Caxito, Lana, Frei, Uhlein and Sial2021) analyzed samples from crystal-fan-bearing limestone from the base of the Sete Lagoas Formation and obtained U–Pb ages of 615.4 ± 5.9 Ma if both the crystal fans and matrix were considered together, 608.1 ± 5.1 Ma for crystal fans, and 607.2 ± 6.2 Ma for the matrix. The youngest population of detrital zircons from the Sete Lagoas Formation gave U–Pb ages of ~557 Ma (Paula-Santos et al., Reference Paula-Santos, Babinski, Kuchenbecker, Caetano-Filho, Trindade and Pedrosa-Soares2015), and the youngest population of detrital zircons from the Três Marias Formation gave U–Pb ages of ~620 Ma (Rodrigues, Reference Rodrigues2008; Pimentel et al., Reference Pimentel, Rodrigues, DellaGiustina, Junges, Matteini and Armstrong2011), providing maximum age constraints on the host strata. More recently, a zircon U–Pb age of 520.2 ± 5.3 Ma has been reported from a volcanic ash bed in the Serra da Saudade Formation (Moreira et al., Reference Moreira, Uhlein, Dussin, Uhlein and Misuzaki2020), suggesting that the upper Bambuí Group may belong to Stage 2 of the Cambrian System.

The possible occurrence of Cloudina sp. and Corumbella werneri—tubular fossils typically found in terminal Ediacaran rocks—in the lower Sete Lagoas Formation (Warren et al., Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014; Perrella Júnior et al., Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017) and the putative presence of Treptichnus pedum (Seilacher, Reference Seilacher, Schindewolf and Seilacher1955)—a trace fossil whose first appearance is used to define the base of the Cambrian System—in the Três Marias Formation (Sanchez et al., Reference Sanchez, Uhlein and Fairchild2021) also indicate that perhaps the entire Bambuí Group is Ediacaran–Cambrian, although the conflict with the ~740 Ma Pb–Pb age from the Sete Lagoas Formation (Babinski et al., Reference Babinski, Vieira and Trindade2007) remains unresolved.

Materials and methods

The studied Barreiro section is located in the Santa Luzia quarry near the Barreiro Community, western Januária Municipality, Minas Gerais State, Brazil (Fig. 1). The samples were collected from two different mining benches, as well as exposures in the hills where the Santa Luzia quarry is located. The stratigraphic thickness of the Sete Lagoas Formation in the studied area is about 70 m (Fig. 2).

Figure 2. Stratigraphic column and field photographs of the Sete Lagoas Formation at the Barreiro section, Santa Luzia quarry, Januária Municipality, Minas Gerais State, Brazil. (1) Thin-bedded limestone. (2) Intraclastic breccia. Sample horizons are marked with the sample number prefixes MP. Sample numbers in bold mark fossiliferous horizons. The CP- numbers refer to the palynological slides of the illustrated specimens.

The lower 15 m of the studied section consists mainly of dark gray, laminated, microcrystalline lime mudstones with a predominance of parallel bedding with microbial mats. However, there are cross-laminations in layers of fine calcareous grainstones. Microbial mats, silicified ooids, and dolomitic nodules are common at this level (Fig. 2). Intraclastic carbonate breccias, with flat pebbles ranging from <1 to 50 cm and light gray micritic matrix are present at 16 m of the section and above, intercalated with limestones. The top of the section, at around 66 m, is composed of light gray, oolitic, crystalline dolomitic grainstones, sometimes with intraclasts. Such carbonates are cross-stratified. This dolomitic interval presents incipient flat stratification, about 2 cm thick, defined by the changes in the amount of sand-size constituents.

A total of 79 stratigraphic levels were sampled. The curatorship of the rock samples, the remaining organic residues, and the palynological slides followed the protocol presented in Denezine et al. (Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022). Each residual sample was coded with the MP prefix. All specimens recovered from the Sete Lagoas Formation herein illustrated are coded with the CP prefix. Each illustrated specimen is provided with a slide number followed by England Finder coordinates.

Organic-walled microfossils were extracted from thinly laminated lime mudstones and light- to dark-gray fine-grained limestone samples using acid maceration techniques. The samples were dissolved using hydrochloric and hydrofluoric acids. Residues were rinsed repeatedly in distilled water, and after the residues were settled following each rinse, the supernatant was decanted. No centrifugation was used, to minimize mechanical damage to organic-walled microfossils. No oxidative procedure was applied on organic residues. Transmitted-light photomicrographs were acquired using an Axio Imager.A2 microscope equipped with an AxioCam MRc digital camera (both Carl Zeiss). The organic-walled microfossils were also analyzed using epifluorescence microscopy; however, no fluorescence was observed in the specimens recovered.

Size analysis of Leiosphaeridia specimens is based on the measurement of their vesicle diameters. Vesicle diameter, along with vesicle wall thickness, was used to identify the four morphospecies of Leiosphaeridia present in the Sete Lagoas Formation: Leiosphaeridia crassa (Naumova, Reference Naumova1949), Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966), Leiosphaeridia minutissima (Naumova, Reference Naumova1949), and Leiosphaeridia tenuissima Eisenack, 1958.

Abundance data were collected in this study. All palynological slides were examined thoroughly, and complete specimens were counted. Due to their colonial nature or frequent preservation as fragments, the abundance of Siphonophycus robustum (Schopf, Reference Schopf1968) and Ghoshia januarensis n. sp. was not quantified.

Selected organic-walled microfossils from the Sete Lagoas Formation were analyzed using Raman spectroscopy. Specimens were placed on palynological slides and analyzed on a HORIBA JobinYvon LabRAM HR800 Raman microprobe equipped with a high-resolution 600 mm focal length spectrometer and a 514 nm argon laser source in the Department of Geosciences at Virginia Tech. The laser beam was focused to less than 10 μm in diameter with a 40× objective lens. Raman spectra were acquired using the software Labspec 5.0 with an acquisition time of less than one minute for each analysis and an excitation power of 600 mW.

Raman spectroscopy data were processed using Python modules. Baseline correction was applied to the raw data by adjusting a polynomial (third-order) curve using the Raman data from 800 and 2,100 cm–1 that captures the Raman peaks of carbonaceous material. After baseline correction, the four Raman peaks of carbonaceous material (i.e., D1, D2, D3, D4) were decomposed using fitting G of Kouketsu et al. (Reference Kouketsu, Mizukami, Mori, Endo, Aoya, Hara, Nakamura and Wallis2014).

The processed Raman data were subjected to principal component analysis (PCA) in Python. The peak position, peak height, and full width at half maximum of the four Raman peaks of carbonaceous material (D1–D4) were used in PCA. The Python package for PCA is publicly available (Mazoni, Reference Mazoni2021), and PCA in this study used the Python modules Numpy (Harris et al., Reference Harris, Millman, van der Walt, Gommers and Virtanen2020), Scipy (Virtanen et al., Reference Virtanen, Gommers, Oliphant, Haberland and Reddy2020), and Rampy (Le Rosq, Reference Le Rosq2021).

Repository and institutional abbreviation

Types, figured specimens, and other specimens examined in this study are deposited in the Paleontological Collection under the prefix MP in the Museum of Geosciences (MGeo-UnB), University of Brasília, Brasília, Brazil.

Systematic paleontology

The suprageneric taxonomy follows the system of modern cyanobacteria and the informal classification of acritarchs (e.g., Butterfield et al., Reference Butterfield, Knoll and Swett1994; Sergeev and Schopf, Reference Sergeev and Schopf2010). Seven organic-walled microfossil species were recovered: Ghoshia januarensis n. sp., Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986, Leiosphaeridia crassa (Naumova, Reference Naumova1949), Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966), Leiosphaeridia minutissima (Naumova, Reference Naumova1949), Leiosphaeridia tenuissima Eisenack, 1958, and Siphonophycus robustum (Schopf, Reference Schopf1968) (Figs. 3, 4). Two of them, Siphonophycus robustum and Ghoshia januarensis, are considered cyanobacteria. Four of them, Leiosphaeridia crassa, Leiosphaeridia jacutica, Leiosphaeridia minutissima, and Leiosphaeridia tenuissima, are sphaeromorph acritarchs traditionally regarded as protists. The phylogenetic affinity of Germinosphaera bispinosa is uncertain.

Figure 3. Organic-walled microfossils from the Sete Lagoas Formation at the Barreiro section. Slide number and England Finder coordinates (in parentheses) are given for each illustrated specimen. (1–3, 7, 10) Leiosphaeridia minutissima: (1) CP962 (S32); (2) CP962 (F48); (3) CP918 (K22); (7) CP964 (P29); (10) CP963 (F33). (4, 8, 11) Germinosphaera bispinosa, all in slide CP917 (EF coordinates: S26, I43, and O28, respectively). (5) Leiosphaeridia jacutica, CP913 (Y23). (6) Leiosphaeridia crassa, CP964 (H29). (9, 13) Leiosphaeridia tenuissima, all in slide CP914 (EF coordinates: Q30 and R23, respectively). (12) Siphonophycus robustum, CP960 (I50).

Figure 4. Ghoshia januarensis n. sp. from the Sete Lagoas Formation in the Barreiro section. (1) Holotype: CP916 (E46). Note dark spots inside cells indicated by white arrows. Yellow arrows indicate slightly deflated and deformed cells. (2–8) Paratypes: (2–4) CP919 (E18); (3) magnified view of the upper right part of (2), showing slightly deflated and deformed cells; (4) magnified view of the lower left part of (2), showing dark spot in terminal cell (arrow); (5, 6) CP919 (J16); (6) dark-field view of the central part of (5), showing a polyhedral cell (arrow in 5). (7) CP919 (J26); note polyhedral cell at branching point. (8) CP920 (N18/3), showing pointed terminal cell (arrow). (9) Specimen identified in a petrographic thin section of the Sete Lagoas Formation at the Barreiro section in the Januária area. Reproduced from Perrella Júnior et al. (Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017) with permission.

Kingdom Eubacteria Woese and Fox, Reference Woese and Fox1977
Phylum Cyanobacteria Stanier et al., Reference Stanier, Sistrom, Hansen, Whitton and Castenholz1978
Class Hormogoneae Thuret, Reference Thuret1875
Order Oscillatoriales Elenkin, Reference Elenkin1949
Family Oscillatoriaceae Kirchner, Reference Kirchner, Engler and Prantl1900
Genus Siphonophycus Schopf, Reference Schopf1968

Type species

Siphonophycus kestron Schopf, Reference Schopf1968 (holotype: Paleobotany Collection Harvard University no. 58469, stage coordinates 33.6 × 101.4) from the black chert facies in the middle third of the late Precambrian Bitter Springs Formation, exposed on the south slope of the ridge about 1 mile north of Ross River Tourist Camp (Love's Creek Homestead), 40 miles northeast of Alice Springs, Northern Territory, Australia, by original designation.

Other species

Siphonophycus thulenema Butterfield in Butterfield et al., Reference Butterfield, Knoll and Swett1994; Siphonophycus septatum (Schopf, Reference Schopf1968); Siphonophycus robustum (Schopf, Reference Schopf1968); Siphonophycus typicum (Hermann, Reference Hermann and Timfeev1974); Siphonophycus kestron Schopf, Reference Schopf1968; Siphonophycus solidum (Golub, Reference Golub and Sokolov1979); Siphonophycus punctatum Maithy, Reference Maithy1975; and Siphonophycus gigas Tang et al., Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013.

Original diagnosis by Schopf (Reference Schopf1968)

“Thallus broad, tubular, nonseptate, unbranched, commonly quite long, finely rugose in surface texture. Thallus cylindrical, somewhat tapered toward apices, solitary, straight to slightly bent, up to 180 μm long (incomplete specimen), occasionally folded and distorted. Apices apparently capitate, more-or-less constricted adjacent to expanded, broadly conical, bluntly pointed terminus. Thallus quite broad, 8.3–15.00 μm wide, commonly about 12.5 μm wide (based on five specimens), ornamented and ringed by finely punctate surficial ridges regularly spaced out 2/3 μm apart. Reproductive structures unknown.”

Emended diagnosis by Knoll et al. (Reference Knoll, Swett and Mark1991)

“Tubular, filamentous microfossils, nonseptate and unbranched, with little or no tapering toward filament termini; tubes truncated and open at ends or with closed, more or less hemispherical terminations; walls typically preserved as chagrenate to finely reticulate organic matter, but may be preserved as carbonate rinds.”

Remarks

The genus Siphonophycus is characterized by smooth and thin wall filaments without ornamentation. The taxon is traditionally interpreted as representing empty sheaths of filamentous cyanobacteria, but because of simple morphology, it could include a range of bacterial and eukaryotic organisms (Butterfield et al., Reference Butterfield, Knoll and Swett1994). Although it is here placed under cyanobacteria, we recognize that Siphonophycus is a form taxon, and several other genera of filamentous microfossils (e.g., Eomycetopsis, Tenuofilum, and Leiotrichoides) are regarded as synonyms of Siphonophycus (Knoll et al., Reference Knoll, Swett and Mark1991).

Siphonophycus robustum (Schopf, Reference Schopf1968) emend. Knoll et al., Reference Knoll, Swett and Mark1991
Figure 3.12

Reference Schopf1968

Eomycetopsis robusta Schopf, p. 685, pl. 82, figs. 2, 3, pl. 83, figs. 1–4.

Reference Schopf1968

Eomycetopsis filiformis Schopf, p. 685, pl. 82, figs. 1, 4, pl. 83, figs. 5–8.

Reference Knoll and Golubic1979

Eomycetopsis robusta; Knoll and Golubic, p.149, fig. 4a, b.

Reference Mendelson and Schopf1982

Eomycetopsis robusta; Mendelson and Schopf, p. 59, pl. 1, figs. 9, 10.

Reference Sergeev1984

Eomycetopsis robusta; Sergeev, p. 436, fig. 2a–ã.

Reference Hofmann and Jackson1991

Eomycetopsis robusta; Hofmann and Jackson, p. 367, fig. 5.1–5.3, 5.8.

Reference Knoll, Swett and Mark1991

Siphonophycus robustum (Schopf, Reference Schopf1968); Knoll et al., p. 565, fig. 10.3, 10.5.

Reference Zang and Walter1992

Eomycetopsis robusta; Zang and Walter, p. 314, pl. 17, figs. g–I, p. 308, pl. 18, fig. g.

Reference Sergeev1992

Eomycetopsis robusta; Sergeev, p. 93, pl. 7, figs. 9, 10, pl. 16, figs. 3, 6, 7, 10; pl. 19, figs. 1, 5–10, pl. 24, fig. 7.

Reference Golovenok and Belova.1993

Eomycetopsis robusta; Golovenok and Belova, pl. 2, fig. å.

Reference Butterfield, Knoll and Swett1994

Siphonophycus robustum; Butterfield et al., p. 64, fig. 26a, g.

Reference Hofmann and Jackson1994

Siphonophycus robustum; Hofmann and Jackson, p. 10, fig. 11.5.

Reference Sergeev1994

Siphonophycus robustum; Sergeev, p. 250, fig. 8f.

Reference Sergeev, Knoll, Kolosova and Kolosov.1994

Siphonophycus robustum; Sergeev et al., pl. 3, fig. 6.

Reference Kumar and Srivastava1995

Siphonophycus robustum; Kumar and Srivastava, p. 114, fig. 14c–e.

Reference Zang1995

Siphonophycus robustum; Zang, p. 172, figs. 26a, 32l, m.

Reference Sergeev, Knoll and Petrov1997

Siphonophycus robustum; Sergeev et al., p. 230, fig. 14a.

Reference Kumar and Venkatachala1998

Siphonophycus robustum; Kumar and Venkatachala, p. 63, fig. 6c.

Reference Sergeev2001

Siphonophycus robustum; Sergeev, p. 442, fig. 7.8, 7.9.

Reference Sergeev and Lee2001

Siphonophycus robustum; Sergeev and Lee, p. 6, pl. 1, figs. 1, 2, 7, 11, 12.

Reference Samuelsson and Butterfield2001

Siphonophycus robustum; Samuelsson and Butterfield, p. 240, figs. 2b, 9h.

Reference Gaucher, Boggiani, Sprechmann, Sial and Fairchild2003

Siphonophycus robustum; Gaucher et al., fig. 6c, d.

Reference Gaucher and Germs2003

Siphonophycus robustum; Gaucher and Germs, fig. 7.12.

Reference Sharma and Sergeev2004

Siphonophycus robustum; Sharma and Sergeev, figs. 3c, 4a, 6b, e, 7c, f, 9e, 11f.

Reference Sergeev and Lee2004

Siphonophycus robustum; Sergeev and Lee, pl. 2, fig. 4.

Reference Tiwari and Pant2004

Siphonophycus robustum; Tiwari and Pant, fig. 3i, n.

Reference Prasad, Uniyal and Asher2005

Siphonophycus robustum; Prasad et al., pl. 1, fig. 7, pl. 5, fig. 12.

Reference Sergeev2006

Siphonophycus robustum; Sergeev, p. 213, pl. 6, figs. 9, 10, pl. 17, fig. 1, pl. 19, figs. 8, 9, pl. 22, figs. 1, 2, 7, 8, 11, 12, pl. 25, figs. 1, 3, pl. 27, figs. 4, 5, pl. 28, fig. 2, pl. 36, figs. 1, 2, pl. 44, figs. 1–7, 13, pl. 46, figs. 7–10, pl. 48, fig. 4.

Reference Kumar and Pandey2008

Siphonophycus robustum; Kumar and Pandey, fig. 3a, b.

Reference Sergeev, Sharma and Shukla2008

Siphonophycus robustum; Sergeev et al., pl. 6, figs. 1, 5, 6, pl. 9, figs. 1–3, 5–7.

Reference Tiwari and Pant2009

Siphonophycus robustum; Tiwari and Pant, fig. 6a–c.

Reference Dong, Xiao, Shen, Zhou, Li and Yao2009

Siphonophycus robustum; Dong et al., p. 30, fig. 6.12.

Reference Sergeev and Schopf2010

Siphonophycus robustum; Sergeev and Schopf, p. 387, fig. 6.4.

Reference Sergeev, Sharma and Shukla2012

Siphonophycus robustum; Sergeev et al., p. 309, pl. 21, figs. 2, 4, 8–10.

Reference Pandey and Kumar2013

Siphonophycus robustum; Pandey and Kumar, p. 504, fig. 4e.

Reference Knoll, Wörndle and Kah2013

Siphonophycus robustum; Knoll et al., fig. 4c.

Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013

Siphonophycus robustum; Tang et al., fig. 13b, m.

Reference Babu, Singh and Mehrotra2014

Siphonophycus robustum; Babu et al., fig. 3q.

Reference Liu, Xiao, Yin, Chen, Zhou and Li2014

Siphonophycus robustum; Liu et al., fig. 110.1.

Reference Vorob'eva, Sergeev and Petrov2015

Siphonophycus robustum; Vorob'eva et al., fig. 9.14.

Reference Tang, Pang, Yuan, Wan and Xiao2015

Siphonophycus robustum; Tang et al., fig. 18c.

Reference Schopf, Sergeev and Kudryavtsev2015

Siphonophycus robustum; Schopf et al., p. 716, fig. 11.11.

Reference Porter and Riedman2016

Siphonophycus robustum; Porter and Riedman, p. 837, fig. 16.4.

Reference Sergeev, Knoll, Vorob'eva and Sergeeva2016

Siphonophycus robustum; Sergeev et al., fig. 8.4.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Siphonophycus robustum; Baludikay et al., fig. 11n.

Reference Tang, Hughes, McKenzie, Myrow and Xiao2017

Siphonophycus robustum; Tang et al., fig. 8a, c, d.

Reference Shi, Feng, Khan and Zhu2017a

Siphonophycus robustum; Shi et al., fig. 6.3, 6.5.

Reference Shi, Feng, Khan, Awramik and Zhu2017b

Siphonophycus robustum; Shi et al., p. 721, fig. 3e, f.

Reference Javaux and Knoll2017

Siphonophycus robustum; Javaux and Knoll, p. 212, fig. 5.11.

Reference Beghin, Storme, Blanpied, Gueneli, Brocks, Poulton and Javaux2017

Siphonophycus robustum; Beghin et al., pl. 3, fig. i.

Reference Sergeev, Vorob'eva, Petrov and Semikhatov2017b

Siphonophycus robustum; Sergeev et al., p. 290, fig. 5.10, 5.11.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Siphonophycus robustum; Li et al., fig. 15h.

Reference Loron, Rainbird, Turner, Greenman and Javaux2019

Siphonophycus robustum; Loron et al., fig. 3f.

Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019

Siphonophycus robustum; Arrouy et al., fig. 6f.

Reference Knoll, Germs, Tankard and Welsink2020

Siphonophycus robustum; Knoll et al., p. 6, fig. 3n, o.

Reference Arvestål and Willman2020

Siphonophycus robustum; Arvestål and Willman, p. 22, fig. 10f.

Reference Shukla, Sharma and Sergeev2020

Siphonophycus robustum; Shukla et al., p. 496, fig. 5e.

Reference Miao, Moczydłowska and Zhu2021

Siphonophycus robustum; Miao et al., p. 17, fig. 9e.

Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022

Siphonophycus robustum; Denezine et al., fig. 11.6.

For additional synonyms, see Butterfield et al. (Reference Butterfield, Knoll and Swett1994).

Holotype

Paleobotanical collections, Harvard University (thin section Bit. Spr. 10-1, number 58491), from Neoproterozoic Bitter Springs Formation, Amadeus Basin, Australia (Schopf, Reference Schopf1968, pl. 83, fig. 1).

Original diagnosis by Schopf (Reference Schopf1968)

“Filaments commonly solitary, occasionally in groups of a few entangled filaments, rarely showing plectenchymatous organization. Lateral walls approximately1/3–3/4 μ thick, markedly coriaceous, coarsely and irregularly granular in surface texture. Filaments up to 135 μ long (incomplete filament), more-or-less regularly cylindrical with a variance in diameter of less than 0.8 μ from the widest to the most narrow portion of the filament; 2.8–4.2 μ in diameter with an average width (20 filaments measured of 3.5 μ). Septate portions of filament vary in length, commonly less than 25 μ long, with filaments commonly constricted or overlapping at the septa; overlapping portions commonly with rounded ends. Reproductive structures unknown.”

Emended diagnosis by Knoll and Golubic (Reference Knoll and Golubic1979)

“Filaments cylindrical; unbranched; tubular (nonseptate); bent, sinuous and tortuous; partially flattened, circular to elliptical in cross section; intertwined to form more or less dense meshworks; long. Surface coarsely to irregularly granular in texture. Occasional cylindrical and evenly spaced inclusions, homogeneously filled with fine-grained carbonaceous matter and centrally located in the ‘bore’ of the tube. Filaments tubular with average diameters expressed as mean ± standard deviation 2.95 ±μm (range 2.0–4.4, n = 60). Occasional long cylindrical inclusions, 1.09 ± 0.36 μm (n = 8) in diameter, 3–4 μm long located centrally within tubular filaments.”

Emended diagnosis by Knoll et al. (Reference Knoll, Germs, Tankard and Welsink2020)

“A species of Siphonophycus with tubes 2–4 μm in cross-sectional diameter.”

Occurrence in the studied section

MP2985, MP2995, MP3040, MP3708, MP3709, and MP3710.

Illustrated specimen

CP960 (3 μm in diameter).

Remarks

Filamentous microfossils from the Sete Lagoas Formation are scarce and restricted to Siphonophycus robustum (Schopf, Reference Schopf1968).

Order Stigonematales Geitler, Reference Geitler1925
Family Capsosiracea Geitler, Reference Geitler1925
Genus Ghoshia Mandal and Maithy in Mandal et al., Reference Mandal, Maithy, Barman and Verma1984

Type species

Ghoshia bifurcata Mandal and Maithy in Mandal et al., Reference Mandal, Maithy, Barman and Verma1984.

Original diagnosis presented by Mandal and Maithy in Mandal et al. (Reference Mandal, Maithy, Barman and Verma1984)

“Thallus heterotrichous, erect filaments arising from basal horizontally creeping thallus, densely packed, truly laterally branched, with cells in one or two series; sheath absent; reproduction not observed.”

Ghoshia januarensis new species
Figure 4

Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017

Fossil filaments consisting of aligned rounded cells, Perrella Júnior et al., p. 138. fig. 7h.

Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022

Goshia sp.; Denezine et al., fig. 11.5.

Type specimens

Holotype: CP916. Paratypes: CP919 and CP920. Specimens are housed in the Research Collection, Museum of Geosciences, Institute of Geosciences, University of Brasília, Federal District, Brazil.

Type locality

Sete Lagoas Formation, Bambuí Group, Santa Luzia quarry, Municipality of Januária, Minas Gerais State, Brazil.

Type horizon

Intraclastic breccia from the Sete Lagoas Formation, Bambuí Group. Stratigraphic level: between 31.5 and 36.4 m.

Diagnosis

A species of Ghoshia characterized by spherical to doliform cells that are 3–10 μm in diameter. Cells are organized to form uniserial chains that branch irregularly.

Occurrence in the studied section

MP2980, MP3013, MP3015, MP3040, MP3710, MP3714, MP3718, MP3723, and MP3724.

Description

Uniserial cell chains that branch irregularly. Cells are spherical (Fig. 4.1), doliform (Fig. 4.24.6, 4.9), or polyhedral (Fig. 4.7), with smooth cell walls. Side branches arise more or less perpendicularly to the main branches. Cells at the branching points are often polyhedral (Fig. 4.7). Cells 3–10 μm in diameter. Deformation folds, likely resulting from compression, are present in some cells (Fig. 4.2, 4.7, 4.8).

Etymology

In reference to the Municipality of Januária, Minas Gerais State, Brazil.

Illustrated specimens

CP916, CP919, and CP920.

Remarks

The Sete Lagoas specimens are somewhat similar to Arctacellularia German in Timofeev et al., Reference Timofeev, Hermann and Mikhailova1976 in their uniserial filaments consisting of spherical, doliform, and polyhedral cells. However, unlike the Sete Lagoas specimens, Arctacellularia does not branch. The Sete Lagoas specimens are also similar to the Devonian cyanobacteria Langiella Croft and George, Reference Croft and George1959, Kidstoniella Croft and George, Reference Croft and George1959, and Rhyniella Croft and George, Reference Croft and George1959 in having branching filaments. However, these Devonian genera can be distinguished by the presence of morphologically differentiated heterocysts and akinetes or by the presence of a sheath (Croft and George, Reference Croft and George1959). The Sete Lagoas specimens are best placed in the genus Ghoshia, which is characterized by branching uniserial filaments consisting of largely undifferentiated cells. The new species proposed here, Ghoshia januarensis, resembles Ghoshia bifurcata Mandal and Maithy in Mandal et al., Reference Mandal, Maithy, Barman and Verma1984 in cell size but differs in its more variable cell shape; Ghoshia januarensis has spherical, doliform, and polyhedral cells, whereas Ghoshia bifurcata is said to have “drum-shaped to rectangular” cells (Mandal et al., Reference Mandal, Maithy, Barman and Verma1984). In addition, some specimens of Ghoshia bifurcata (including the holotype; Mandal et al., Reference Mandal, Maithy, Barman and Verma1984, pl. 4, fig. 30) seem to have cell aggregates that are not uniserially organized.

A specimen from the Sete Lagoas Formation in the Januária area illustrated as “fossil filaments consisting of aligned rounded cells” (Perrella Júnior et al., Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017, fig. 7H) shares the same characteristics of Ghoshia januarensis, including uniserial and branching filaments consisting of spherical cells. Thus, this specimen is here identified as Ghoshia januarensis. It is important to point out that the specimen illustrated in Perrella Júnior et al. (Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017) was observed in a petrographic thin section, ruling out the possibility of modern contamination.

Raman spectra of the analyzed microfossils (Fig. 5) display well-developed D1 and D2 bands positioned at 1,350 cm–1 and 1,620 cm–1, respectively. These characteristics are typical of organic matter spectra (Kouketsu et al., Reference Kouketsu, Mizukami, Mori, Endo, Aoya, Hara, Nakamura and Wallis2014). The Raman data show that the four analyzed specimens of Ghoshia januarensis, including the holotype, are distinct from other organic-walled microfossils from the Sete Lagoas Formation (Fig. 5). Relative to other organic-walled microfossils from the Sete Lagoas Formation, Ghoshia januarensis specimens exhibit broader peaks (Fig. 5.1). PCA analysis of Raman parameters also shows that Ghoshia januarensis specimens are separate from other organic-walled microfossils from the Sete Lagoas Formation (Fig. 5.2). Therefore, it is possible that specimens of Ghoshia januarensis have different thermal history from other organic-walled microfossils in the Sete Lagoas Formation, indicating that the former could be contaminations. However, the three specimens of Ghoshia januarensis that were analyzed for Raman spectroscopy in Fig. 5.2 overlap with the other organic-walled microfossils largely along the second PCA axis; the difference is mainly along the primary PCA axis. A similar situation is found in the four specimens of Leiosphaeridia minutissima that were analyzed for Raman spectroscopy (labeled as A, B, G, and H in Fig. 5.2): they exhibit a limited range and overlap with other organic-walled microfossils from the Sete Lagoas Formation along the primary PCA axis, but two specimens (labeled A and B in Fig. 5.2) are distinct from all other specimens along the secondary PCA axis. Although subtle differences in carbonaceous material Raman characteristics could be taken as evidence for different degrees of thermal maturation (Kouketsu et al., Reference Kouketsu, Mizukami, Mori, Endo, Aoya, Hara, Nakamura and Wallis2014), recent studies show that such differences can result from differences in organic precursors (Qu et al., Reference Qu, Engdahl, Zhu, Vajda and McLoughlin2015; Pang et al., Reference Pang, Tang, Wu, Li, Chen, Wan, Yuan, Bodnar and Xiao2020). Considering that Ghoshia januarensis specimens showed no fluorescence as would modern organic contaminations, and that Ghoshia januarensis has been found in a petrographic thin section of the Sete Lagoas Formation (Perrella Júnior et al., Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017), we conclude that Ghoshia januarensis is indigenous to the Sete Lagoas Formation.

Figure 5. Raman spectroscopic data of organic-walled microfossils and amorphous organic matter from the Sete Lagoas Formation at the Barreiro section. (1) Baseline-corrected and fitted Raman spectra. Legends are shown in Figure 5.2. Note that Raman spectra of Ghoshia januarensis (J from holotype and I, K from paratypes) have broader peaks of carbonaceous matter around 1,350 cm–1 and 1,600 cm–1 relative to other Sete Lagoas organic-walled microfossils. (2) Principal component analysis of deconvolved Raman data. Samples: A–B and J, CP916; C–D, CP917; E, I, K, CP920; F, H, MP3728; G, MP3723.

Group Acritarcha Evitt, Reference Evitt1963
Subgroup Sphaeromorphitae Downie et al., Reference Downie, Evitt and Sarjeant1963
Genus Leiosphaeridia Eisenack, 1958

Type species

Leiosphaeridia baltica Eisenack, 1958 (in Eisenack, Reference Eisenack1958b).

Other species

Fensome et al. (Reference Fensome, Williams, Bars, Freeman and Hill1990) revised all Leiosphaeridia species and listed 167 valid species.

Original diagnosis presented by Eisenack (Reference Eisenack1958b) in German

“Hohlkugelförmige, dünnwandige und aus einer sehr widerstandsfähigen, hellgelb bis dunkelrotbraun durchscheinenden organischen Substanz bestehende Organismenreste, die oft in scheibenförmig zusammengepreßtem Zustande oder auch unregelmäßig verfaltet überliefert sein können. Wand, auch in erwachsenem Zustande, stets ohne Wandporen (Unterschied zu Tasmanites). Pylome vorhanden.”

Translation of original diagnosis presented by Eisenack (Reference Eisenack1958b)

Hollow spherical, thin-walled organism residues consisting of a very resistant, light yellow to dark red-brown translucent, organic substance, which can often be preserved as a disc-shaped compressed state or an irregularly folded structure. Wall, even when fully grown, always without wall pores (in contrast to Tasmanites). Pylome present.

Emended diagnosis by Downie and Sarjeant (Reference Downie and Sarjeant1963)

“Spherical to ellipsoidal bodies without processes, often collapsed or folded, with or without pylomes. Walls granular, punctate or unornamented, thin. Without divisions into fields and without transverse or longitudinal furrows or girdles.”

Emended diagnosis by Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) in Russian

“Сфероидальные оболочки с гладкой, точеечной или зеррнистой поверхиостью размером от 2–3 до 750 мкм. Толщина стенки от долей микрометра до 3–10 мкм. В ископаемом состоянии сплющены и осложнены складками смятия различной формы и размеров.”

Translation of emended diagnosis by Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989)

Spheroidal vesicle with a smooth, punctate, or granular surface ranging in size from 2–3 to 750 μm. The wall thickness varies from fractions of a micrometer to 3–10 μm. The specimens are flattened and can have compressional folds of various shapes and sizes.

Remarks

A great number of species of the genus Leiosphaeridia have been reported from the Proterozoic, and many of them have very long stratigraphic ranges, e.g., from the Paleoproterozoic to the Mesozoic (Lamb et al., Reference Lamb, Awramik, Chapman and Zhu2009). There are even reports of Leiosphaeridia species from the Miocene (Hannah et al., Reference Hannah, Wilson and Wrenn2000). Because of its simple morphologies, the genus Leiosphaeridia is regarded as a form taxon with diverse phylogenetic affinities, and it is classified in the Acritarcha (Jankauskas et al., Reference Jankauskas, Mikhailova and German1989; Grey, Reference Grey2005; Sergeev and Schopf, Reference Sergeev and Schopf2010), although Sergeev and Schopf (Reference Sergeev and Schopf2010) consider this taxon as belonging to the Kingdom Protista, a proposition followed here. It is important to point out that some authors relate Leiosphaeridia species to chlorophyceaens (Moczydlowska et al., Reference Moczydlowska, Landing, Zang and Palacios2011; Moczydłowska, Reference Moczydłowska2016). Downie and Sarjeant (Reference Downie and Sarjeant1963) emended the diagnosis of the genus Leiosphaeridia to exclude the reference of the vesicle color since it could reflect diagenetic features. Moreover, the maceration protocol could affect the color of organic vesicles due to the use of oxidizing solutions. Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) specified that the diameter of the vesicle of Leisophaeridia species ranges from 2–3 to 750 μm. Furthermore, Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) divided the smooth-walled Leiosphaeridia species into four species according to vesicle diameter and wall thickness, a form-taxonomical scheme followed in the present work. Butterfield et al. (Reference Butterfield, Knoll and Swett1994) suggested that Leiosphaeridia should be restricted to spheroidal fossils with vesicle walls less than 2 μm thick so that it can be differentiated from Chuaria circularis Walcott, Reference Walcott1899, which has thicker vesicle walls (2–3 μm single-wall thickness).

Leiosphaeridia crassa (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989
Figure 3.6

Reference Naumova1949

Leiotriletes crassus Naumova, p. 54, pl. 1, figs. 5, 6, pl. 2, figs. 5, 6.

Reference Pykhova1973

Leiopsophosphaera crassa; Pykhova, p. 99, pl. 2, fig. 3.

Reference Jankauskas, Mikhailova and German1989

Leiosphaeridia crassa (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., p. 75, pl. 9, figs. 5–10.

Reference Zang and Walter1992

Leiosphaeridia crassa; Zang and Walter, p. 289, pl. 9, figs. a–k, pl. 12, fig. k, pl. 14, figs. e, h.

Reference Butterfield, Knoll and Swett1994

Leiosphaeridia crassa; Butterfield et al., p. 40, figs. 16f, 23k.

Reference Hofmann and Jackson1994

Leiosphaeridia crassa; Hofmann and Jackson, p. 22, fig. 1.19–1.29.

Reference Knoll1994

Leiosphaeridia crassa; Knoll, fig. 4b.

Reference Zang1995

Leiosphaeridia crassa; Zang, p. 166, figs. 21d, 28c, d.

Reference Yin and Guan1999

Leiosphaeridia crassa; Yin and Guan, p. 131, figs. 3.8, 4.5, 5.3, 5.5, 5.7, 5.11, 6.2–6.6, 6.9, 6.12.

Reference Javaux, Knoll and Walter2004

Leiosphaeridia crassa; Javaux et al., fig. 4e–i.

Reference Sergeev and Lee2004

Leiosphaeridia crassa; Sergeev and Lee, p. 21, pl. 3, figs. 4, 5.

Reference Tiwari and Pant2004

Leiosphaeridia crassa; Tiwari and Pant, p. 1736, fig. 3v.

Reference Grey2005

Leiosphaeridia crassa; Grey, p. 179, figs. 63a–c, 64a–d.

Reference Marshall, Javaux, Knoll and Walter2005

Leiosphaeridia crassa; Marshall et al., fig. 1e.

Reference Prasad, Uniyal and Asher2005

Leiosphaeridia crassa; Prasad et al., pl. 1, figs. 1, 2, pl. 4, fig. 16, pl. 5, fig. 18, pl. 9, figs. 10, 11.

Reference Javaux and Marshal2006

Leiosphaeridia crassa; Javaux and Marshal, fig. 3.4–3.6.

Reference Sergeev and Seong-Joo2006

Leiosphaeridia crassa; Sergeev and Seong-Joo, p. 15, pl. 2, figs. 2a–c, 5.

Reference Moczydłowska2008a

Leiosphaeridia crassa; Moczydłowska, p. 84, figs. 7a, 8g.

Reference Moczydłowska2008b

Leiosphaeridia crassa; Moczydłowska, fig. 2g.

Reference Sergeev, Sharma and Shukla2008

Leiosphaeridia crassa; Sergeev et al., pl. 7, figs. 5, 6.

Reference Yin, Yang, Peng and Kong2009

Leiosphaeridia crassa; Yin et al., figs. 3a, 3h, 3l, 4d, 4f, 4h, 5a, 5c.

Reference Tiwari and Pant2009

Leiosphaeridia crassa; Tiwari and Pant, figs. 7d, e, 8h, 8o, p.

Reference Stanevich, Maksimova, Kornilova, Gladkochub, Mazukabzov and Donskaya2009

Leiosphaeridia crassa; Stanevich et al., p. 32, pl. 3, figs. 3, 4.

Reference Sergeev and Schopf2010

Leiosphaeridia crassa; Sergeev and Schopf, p. 395, fig. 15.3–15.6.

Reference Strother, Battison, Brasier and Wellman2011

Leiosphaeridia crassa; Strother et al., fig. 1a, e.

Reference Couëffé and Vecolii2011

Leiosphaeridia crassa; Couëffé and Vecolii, figs. 6.2, 7.1, 7.7.

Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013

Leiosphaeridia crassa; Tang et al., fig. 4b.

Reference Lottaroli, Craig and Thusu2014

Leiosphaeridia crassa; Lottaroli et al., fig. 10.2.

Reference Babu, Singh and Mehrotra2014

Leiosphaeridia crassa; Babu et al., fig. 3f.

Reference Tang, Pang, Yuan, Wan and Xiao2015

Leiosphaeridia crassa; Tang et al., fig. 4d.

Reference Nagovitsin and Kochnev2015

Leiosphaeridia crassa; Nagovitsin and Kochnev, fig. 1.55, 1.56.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Leiosphaeridia crassa; Baludikay et al., fig. 8a–c.

Reference Porter and Riedman2016

Leiosphaeridia crassa; Porter and Riedman, p. 833, fig. 13.2, 13.6.

Reference Sergeev, Knoll, Vorob'eva and Sergeeva2016

Leiosphaeridia crassa; Sergeev et al., fig. 4.2.

Reference Javaux and Knoll2017

Leiosphaeridia crassa; Javaux and Knoll, p. 209, fig. 4.6.

Reference Agić, Moczydłowska and Yin2017

Leiosphaeridia crassa; Agic et al., p. 110, fig. 8a–c.

Reference Sergeev, Vorob'eva and Petrov2017a

Leiosphaeridia crassa; Sergeev et al., fig. 3.14.

Reference Sergeev, Vorob'eva, Petrov and Semikhatov2017b

Leiosphaeridia crassa; Sergeev et al., pl. I, fig. 6.

Reference Beghin, Storme, Blanpied, Gueneli, Brocks, Poulton and Javaux2017

Leiosphaeridia crassa; Beghin et al., pl. 2, figs. c, d.

Reference Suslova, Parfenova, Saraev and Nagovitsyn2017

Leiosphaeridia crassa; Suslova et al., fig. 3.1–3.4.

Reference Anderson, McMahon, Macdonald, Jones and Briggs2019

Leiosphaeridia crassa; Anderson et al., p. 510, fig. 8a–e.

Reference Riedman, Porter and Calver2018

Leiosphaeridia crassa; Riedman et al., fig. 5.15.

Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019

Leiosphaeridia crassa; Arrouy et al., fig. 6d, e.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Leiosphaeridia crassa; Li et al., fig. 4f.

Reference Arvestål and Willman2020

Leiosphaeridia crassa; Arvestål and Willman, p. 11, fig. 6j, k, m.

Reference Knoll, Germs, Tankard and Welsink2020

Leiosphaeridia crassa; Knoll et al., p. 6, fig. 3g.

Reference Shukla, Sharma and Sergeev2020

Leiosphaeridia crassa; Shukla et al., p. 502, fig. 6g.

Reference Pang, Tang, Wu, Li, Chen, Wan, Yuan, Bodnar and Xiao2020

Leiosphaeridia crassa; Pang et al., fig. 2m.

For additional synonyms, also see Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) and Zang and Walter (Reference Zang and Walter1992).

Type material

Naumova (Reference Naumova1949) did not designate a holotype for Leiotriletes crassus. Subsequently, Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) designated one specimen of Leiotriletes crassus published by Naumova (Reference Naumova1949) as a “holotype” (Naumova, Reference Naumova1949, pl. 1, fig. 3). In addition, he designated another specimen from a different locality and a different stratigraphic unit as a “lectotype” (Jankauskas et al., Reference Jankauskas, Mikhailova and German1989, LitNIGRI, N 16-800-2942/9, specimen 2, table 9, fig. 5). By so doing, the selection of a holotype by Jankauskas can, according to the International Code of Nomenclature for Algae, Fungi, and Plants, be taken as the designation of a lectotype (Turland et al., Reference Turland, Wiersema, Barrie, Greuter and Hawksworth2018). In addition, the specimen designated by Jankauskas as a “lectotype” should be regarded as a neotype. According to the same code, a lectotype always takes precedence over a neotype. However, the lectotype designated by Jankauskas was a specimen of Leiotriletes simplicissimus (Naumova, Reference Naumova1949), a species he synonymized with a different species of Leiosphaeridia, Leiosphaeridia minutissima. Thus, the lectotype designated by Jankauskas is not valid, and the neotype designated by Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) is here considered the valid type specimen of Leiosphaeridia crassa.

Original diagnosis presented by Naumova (Reference Naumova1949) in Russian

“В очертании спора округлой или округло-овальной формы. Поверхность экзины гладкая, экзина очень толстая и плотная. Форма имеет складки смятия, щель разверзания, простая. Широко распространена в нижнем кембрии Прибалтики.”

Translation of original diagnosis presented by Naumova (Reference Naumova1949)

In outline, the spore is round or round–oval. The surface of the exine is smooth, very thick, and dense. The form has compressional folds, an opening gap, and is simple. Widespread in the lower Cambrian of the Baltic.

Emended diagnosis by Javaux and Knoll (Reference Javaux and Knoll2017) and Knoll et al. (Reference Knoll, Germs, Tankard and Welsink2020)

“A species of Leiosphaeridia with smooth, pliant walls with lanceolate folds and a modal diameter of less than 70 μm.”

Occurrence in the studied section

Fourteen specimens were recovered. They range from ~18 to ~62 μm in diameter: MP3719 and MP3720.

Illustrated specimen

CP964 (diameter ~27 μm).

Remarks

Leiotriletes crassus Naumova, Reference Naumova1949 was originally published with only a description, without a diagnosis. Although the International Code of Nomenclature for Algae, Fungi, and Plants states that either a description or a diagnosis is sufficient for the valid publication of a name (Turland et al., Reference Turland, Wiersema, Barrie, Greuter and Hawksworth2018, Art. 38.1), it is strongly recommended that both the diagnosis and description be presented when describing a new species (Hassemer et al., Reference Hassemer, Prado and Baldini2020). Later, Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) reviewed some species of Leiosphaeridia and transferred Leiotriletes crassus Naumova, Reference Naumova1949 to the genus Leiosphaeridia. When Leiotriletes crassus was transferred to the genus Leiosphaerida, the species epithet was changed to crassa, so the gender of the epithet agrees with the gender of the genus name. Thus, this species became Leiosphaeridia crassa (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989. In addition, Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) did not include in their synonym list the species Leiopsophosphaera crassa Pykhova, Reference Pykhova1973. Finally, Fensome et al. (Reference Fensome, Williams, Bars, Freeman and Hill1990), also in a work of taxonomic revision, transferred Leiopsophosphaera crassa Pykhova, Reference Pykhova1973 to Leiosphaeridia crassa (Pykhova, Reference Pykhova1973). However, Fensome et al. (Reference Fensome, Williams, Bars, Freeman and Hill1990) did not consider the study of Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) and regarded Leiotriletes crassus Naumova, Reference Naumova1949 as taxonomically uncertain (Grey, Reference Grey2005). Thus, Leiosphaeridia crassa (Pykhova, Reference Pykhova1973) is a junior homonym of Leiosphaeridia crassa (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989. Nonetheless, Leiopsophosphaera crassa Pykhova, Reference Pykhova1973 is considered by several authors (Yin and Guan, Reference Yin and Guan1999; Grey, Reference Grey2005) as a synonym of Leiosphaeridia crassa (Naumova, Reference Naumova1949), a synonymy followed in this study. Leiosphaeridia crassa differs from Leiosphaeridia minutissima in its thicker vesicle wall, and it differs from Leiosphaeridia tenuissima and Leiosphaeridia jacutica in vesicle size (Jankauskas et al., Reference Jankauskas, Mikhailova and German1989).

Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966) Mikhailova and Jankauskas in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989
Figure 3.5

Reference Timofeev1966

Kildinella jacutica Timofeev, p. 30, pl. 7, fig. 2, pl.19, fig 9, pl. 61, fig. 5, pl. 67, fig. 8, pl. 72, fig. 1.

Reference Jankauskas, Mikhailova and German1989

Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966) Mikhailova and Jankauskas in Jankauskas et al., p. 77, pl. 12, figs. 3, 7, 9.

Reference Butterfield and Chandler1992

Leiosphaeridia jacutica; Butterfield and Chandler, fig. 5e.

Reference Butterfield, Knoll and Swett1994

Leiosphaeridia jacutica; Butterfield et al., p. 42, fig. 16h.

Reference Hofmann and Jackson1994

Leiosphaeridia jacutica; Hofmann and Jackson, p. 22, fig. 17.1–17.4.

Reference Kumar and Srivastava1995

Leiosphaeridia jacutica; Kumar and Srivastava, p. 106, fig. 11k.

Reference Sergeev2001

Leiosphaeridia jacutica; Sergeev, p. 444, fig. 8.7–8.10.

Reference Javaux, Knoll and Walter2004

Leiosphaeridia jacutica; Javaux et al., fig. 4a–d, 4m.

Reference Grey2005

Leiosphaeridia jacutica; Grey, p. 183, fig. 63g.

Reference Marshall, Javaux, Knoll and Walter2005

Leiosphaeridia jacutica; Marshall et al., fig. 1c.

Reference Prasad, Uniyal and Asher2005

Leiosphaeridia jacutica; Prasad et al., pl. 3, figs. 13, 14, pl. 4, fig. 12, pl. 9, fig. 25, pl. 10, fig. 6.

Reference Sergeev and Seong-Joo2006

Leiosphaeridia jacutica; Sergeev and Seong-Joo, p. 14, pl. 2, fig. 6.

Reference Javaux and Marshal2006

Leiosphaeridia jacutica; Javaux and Marshal, fig. 3.1–3.3.

Reference Stanevich, Maksimova, Kornilova, Gladkochub, Mazukabzov and Donskaya2009

Leiosphaeridia jacutica; Stanevich et al., p. 32, pl. 3, fig. 2.

Reference Vorob'eva, Sergeev and Knoll2009

Leiosphaeridia jacutica; Vorob'eva et al., p. 185, fig. 14.13.

Reference Nemerov, Stanevich, Razvozzhaeva, Budyak and Kornilova2010

Leiosphaeridia jacutica; Nemerov et al., fig. 6.8, 6.9.

Reference Prasad, Asher and Borgohai2010

Leiosphaeridia jacutica; Prasad et al., pl. 1, fig. 3.

Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013

Leiosphaeridia jacutica; Tang et al., fig. 4d.

Reference Babu, Singh and Mehrotra2014

Leiosphaeridia jacutica; Babu et al., fig. 3l.

Reference Chiglino, Gaucher, Sial and Ferreira2015

Leiosphaeridia jacutica; Chiglino et al., p. 643, fig. 5b.

Reference Tang, Pang, Yuan, Wan and Xiao2015

Leiosphaeridia jacutica; Tang et al., figs. 4f, g, 5a.

Reference Nagovitsin and Kochnev2015

Leiosphaeridia jacutica; Nagovitsin and Kochnev, fig. 4.43.

Reference Vorob'eva, Sergeev and Petrov2015

Leiosphaeridia jacutica; Vorob'eva et al., fig. 7.6.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Leiosphaeridia jacutica; Baludikay et al., fig. 8d.

Reference Porter and Riedman2016

Leiosphaeridia jacutica; Porter and Riedman, p. 833, fig. 13.3.

Reference Sergeev, Knoll, Vorob'eva and Sergeeva2016

Leiosphaeridia jacutica; Sergeev et al., fig. 4.1, 4.6, 4.7.

Reference Singh and Sharma2016

Leiosphaeridia jacutica; Singh and Sharma, p. 80, pl. 1, figs. 9, 10.

Reference Javaux and Knoll2017

Leiosphaeridia jacutica; Javaux and Knoll, p. 209, fig. 4.4, 4.5.

Reference Sergeev, Vorob'eva and Petrov2017a

Leiosphaeridia jacutica; Sergeev et al., fig. 3.1, 3.9–3.11.

Reference Sergeev, Vorob'eva, Petrov and Semikhatov2017b

Leiosphaeridia jacutica; Sergeev et al., pl. I, fig. 5.

Reference Beghin, Storme, Blanpied, Gueneli, Brocks, Poulton and Javaux2017

Leiosphaeridia jacutica; Beghin et al., pl. 2, fig. e.

Reference Tang, Hughes, McKenzie, Myrow and Xiao2017

Leiosphaeridia crassa; Tang et al., fig. 3c.

Reference Tang, Hughes, McKenzie, Myrow and Xiao2017

Leiosphaeridia jacutica; Tang et al., fig. 3d.

Reference Anderson, McMahon, Macdonald, Jones and Briggs2019

Leiosphaeridia jacutica; Anderson et al., p. 12, fig. 8f–k.

Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019

Leiosphaeridia jacutica; Arrouy et al., fig. 6b, c.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Leiosphaeridia jacutica; Li et al., fig. 4h.

Reference Arvestål and Willman2020

Leiosphaeridia jacutica; Arvestål and Willman, p. 11, fig. 6i, 6l.

Reference Knoll, Germs, Tankard and Welsink2020

Leiosphaeridia jacutica; Knoll et al., p. 6, fig. 2g.

Reference Shukla, Sharma and Sergeev2020

Leiosphaeridia jacutica; Shukla et al., p. 502, fig. 6l.

Reference Pang, Tang, Wu, Li, Chen, Wan, Yuan, Bodnar and Xiao2020

Leiosphaeridia jacutica; Pang et al., fig. 2f.

Reference Han, Chen, Li, Pang and Wang2021

Leiosphaeridia jacutica; Han et al., fig. 3a–d.

Holotype

IGD Russian Academy of Sciences no. 451/1, from upper Riphean, Lakhanda Group, Neryuen Formation, Siberia (Timofeev, Reference Timofeev1966, pl. 7, fig. 2).

Diagnosis by Javaux and Knoll (Reference Javaux and Knoll2017) and Knoll et al. (Reference Knoll, Germs, Tankard and Welsink2020)

“A species of Leiosphaeridia characterized by smooth, pliant walls with lanceolate folds and a modal diameter greater than 70 μm.”

Occurrence in studied section

Four specimens were recovered. They range from ~74 to ~98 μm in diameter: MP2990, MP3719, and MP3714.

Original description by Timofeev (Reference Timofeev1966) in Russian

“Оболочки диаметром 150–250 мк, сферические, толстые, однослойные, с поверхностью от гладкой до грубошагреневой, с резко очерченными, крупными, серповидными, иногда угловатымц складками. Цвет темно-желтый, желто-коричневый.”

Translation of original description by Timofeev (Reference Timofeev1966)

The vesicles are 150–250 microns in diameter, spherical, thick, single-layered, with a smooth to coarse shagreen surface that bears sharply defined, large, crescent-shaped, sometimes angular folds. Color dark yellow, yellow-brown.

Illustrated specimen

CP913 (diameter ~81 μm).

Remarks

Timofeev (Reference Timofeev1966) described the new species Kildinella jacutica Timofeev, Reference Timofeev1966 and designated a holotype with the description of this species, but no diagnosis was provided. Later, Mikhailova and Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) proposed that Kildinella jacutica should be transferred to Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966). They also designated a neotype for Leiosphaeridia jacutica. This neotype is not valid since Timofeev (Reference Timofeev1966) had designated a holotype in its publication, and there is no report of the loss of this holotype. Leiosphaeridia jacutica differs only by the larger size compared with Leiosphaeridia crassa (Jankauskas et al., Reference Jankauskas, Mikhailova and German1989). The specimen illustrated by Tang et al. (Reference Tang, Hughes, McKenzie, Myrow and Xiao2017, fig. 3) may not be Leiosphaeridia crassa but Leiosphaeridia jacutica because its diameter is around 90 μm. Leiosphaeridia jacutica differs from Bambuites erichsenii in its sphaeromorphic vesicle without processes.

Leiosphaeridia minutissima (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989
Figure 3.13.3, 3.7, 3.10

Reference Naumova1949

Leiotriletes minutissimus Naumova, p. 52, pl. 1, figs. 1, 2, pl. 2, figs. 1, 2.

Reference Jankauskas, Mikhailova and German1989

Leiosphaeridia minutissima (Naumova, Reference Naumova1949) Jankauskas in Jankauskas et al., p. 79, pl. 9, figs. 1–4, 11.

Reference Butterfield and Chandler1992

Leiosphaeridia minutissima; Butterfield and Chandler, fig. 3a, i.

Reference Hofmann and Jackson1994

Leiosphaeridia minutissima; Hofmann and Jackson, p. 21, fig. 23.9–23.15.

Reference Gaucher, Boggiani, Sprechmann, Sial and Fairchild2003

Leiosphaeridia minutissima; Gaucher and Germs, fig. 6.10–6.12.

Reference Grey2005

Leiosphaeridia minutissima; Grey, p. 184, fig. 63d.

Reference Blanco and Gaucher2005

Leiosphaeridia minutissima; Blanco and Gaucher, fig. 11b.

Reference Gaucher, Poire, Peral and Chiglino2005b

Leiosphaeridia minutissima; Gaucher et al., fig. 6d.

Reference Prasad, Uniyal and Asher2005

Leiosphaeridia minutissima; Prasad et al., pl. 9, figs. 1, 3.

Reference Gaucher, Chiglino, Blanco, Poiré and Germs2008

Leiosphaeridia minutissima; Gaucher et al., p. 491, fig. 3a.

Reference Moczydłowska2008a

Leiosphaeridia minutissima; Moczydłowska, p. 84, fig. 8h.

Reference Moczydłowska2008b

Leiosphaeridia minutissima; Moczydłowska, figs. 2f, 6d.

Reference Nemerov, Stanevich, Razvozzhaeva, Budyak and Kornilova2010

Leiosphaeridia minutissima; Nemerov et al., fig. 6.7.

Reference Couëffé and Vecolii2011

Leiosphaeridia minutissima; Couëffé and Vecolii, fig. 7.3.

Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013

Leiosphaeridia minutissima; Tang et al., fig. 4a.

Reference Chiglino, Gaucher, Sial and Ferreira2015

Leiosphaeridia minutissima; Chiglino et al., p. 642, fig. 5a.

Reference Tang, Pang, Yuan, Wan and Xiao2015

Leiosphaeridia minutissima; Tang et al., fig. 4c.

Reference Nagovitsin and Kochnev2015

Leiosphaeridia minutissima; Nagovitsin and Kochnev, fig. 4.57, 4.58.

Reference Schopf, Sergeev and Kudryavtsev2015

Leiosphaeridia minutissima; Schopf et al., p. 724, fig. 13.10.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Leiosphaeridia minutissima; Baludikay et al., fig. 8e.

Reference Porter and Riedman2016

Leiosphaeridia minutissima; Porter and Riedman, p. 834, fig. 13.1, 13.5.

Reference Javaux and Knoll2017

Leiosphaeridia minutissima; Javaux and Knoll, p. 210, fig. 4.7, 4.8.

Reference Shi, Feng, Khan and Zhu2017a

Leiosphaeridia minutissima; Shi et al., fig. 11.6, 11.7.

Reference Beghin, Storme, Blanpied, Gueneli, Brocks, Poulton and Javaux2017

Leiosphaeridia minutissima; Beghin et al., pl. 2, figs. g, h.

Reference Tang, Hughes, McKenzie, Myrow and Xiao2017

Leiosphaeridia minutissima; Tang et al., fig. 3a.

Reference Suslova, Parfenova, Saraev and Nagovitsyn2017

Leiosphaeridia minutissima; Suslova et al., fig. 3.6–3.11.

Reference Agić, Moczydłowska and Yin2017

Leiosphaeridia minutissima; Agic et al., p. 110, fig. 8g, h.

Reference Yin, Singh, Bhargava, Zhao, Negi, Meng and Sharma2018

Leiosphaeridia minutissima; Yin et al., fig. 4h, 4j, 4l.

Reference Javaux and Lepot2018

Leiosphaeridia minutissima; Javaux and Lepot, fig. 2e.

Reference Lei, Shen, Algeo, Servais, Feng and Yu2019

Leiosphaeridia minutissima; Lei et al., fig. 3.13, 3.14.

Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019

Leiosphaeridia minutissima; Arrouy et al., fig. 5a–g, 5j.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Leiosphaeridia minutissima; Li et al., fig. 4e.

Reference Shang, Liu and Moczydłowska2019

Leiosphaeridia minutissima; Shang et al., p. 24, fig. 21a.

Reference Arvestål and Willman2020

Leiosphaeridia minutissima; Arvestål and Willman, p. 11, fig. 6c–g.

Reference Knoll, Germs, Tankard and Welsink2020

Leiosphaeridia minutissima; Knoll et al., p. 6, fig. 2a, 2c.

Reference Shukla, Sharma and Sergeev2020

Leiosphaeridia minutissima; Shukla et al., p. 502, fig. 6e, k, m.

Reference Pang, Tang, Wu, Li, Chen, Wan, Yuan, Bodnar and Xiao2020

Leiosphaeridia minutissima; Pang et al., fig. 2n.

Reference Loron, Halverson, Rainbird, Skulski, Turner and Javaux2021

Leiosphaeridia minutissima; Loron et al., fig. 6.2.

Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022

Leiosphaeridia minutissima; Denezine et al., fig. 11.1, 11.2.

Type material

Naumova (Reference Naumova1949) did not designate a holotype for Leiotriletes minutissimus. Afterward, Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) designated one of the specimens published by Naumova (Reference Naumova1949, pl. 1, fig. 1) as the “holotype”. In addition, he designated another specimen from a different locality and a different stratigraphic unit as a “lectotype” (Jankauskas et al., Reference Jankauskas, Mikhailova and German1989, LitNIGRI, N 16-800-2942/9, table 9, fig. 1). According to the International Code of Nomenclature for Algae, Fungi, and Plants (Turland et al., Reference Turland, Wiersema, Barrie, Greuter and Hawksworth2018), the “holotype” selected by Jankauskas should be regarded as a lectotype and the “lectotype” regarded as a neotype. According to the same code, a lectotype takes precedence over a neotype. Thus, the lectotype designated by Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) (Naumova, Reference Naumova1949, pl. 1, fig. 1) is the valid type specimen of Leiosphaeridia minutissima.

Diagnosis presented by Javaux and Knoll (Reference Javaux and Knoll2017)

“A species of Leiosphaeridia characterized by smooth walls with sinuous folds and a modal diameter less than 70 μm.”

Emended diagnosis presented by Knoll et al. (Reference Knoll, Germs, Tankard and Welsink2020)

“A species of Leiosphaeridia characterized by thin, smooth walls with sinuous folds and a modal diameter less than 70 μm.”

Occurrence in studied section

A total of 359 specimens were recovered. They range from ~4.4 to ~68.2 μm in diameter: MP3719, MP2977, MP2979, MP2980, MP2983, MP2985, MP2986, MP2987, MP2988, MP2992, MP2993, MP2994, MP2995, MP2998, MP2999, MP 3002, MP 3004, MP 3005, MP 3006, MP 3007, MP 3011, MP 3012, MP 3013, MP 3015, MP 3016, MP3028, MP 3030, MP3031, MP3033, MP3034, MP3035, MP3036, MP3705, MP3707, MP3708, MP3709, MP3710, MP3712, MP3713, MP3714, MP3715, MP3716, MP3719, and MP3720.

Description presented by Naumova (Reference Naumova1949) in Russian

“Очертание споры округлое. Экзина очень тонкая, прозрачная, наблюдаются многочисленные складки смятия. Поверхность экзины гладкая. Щель разверзания трехлучевая, простая, плохо различимая нз-за складок смятия.”

Translation of description presented by Naumova (Reference Naumova1949)

The outline of the vesicle is round. The exine is very thin, transparent, with numerous compressional folds. The surface of the exine is smooth. The opening slit is three-beam, simple, poorly distinguishable due to the compressional folds.

Illustrated specimens

CP918 (diameter ~32 μm), CP962 (diameters 32~ μm and ~39 μm), CP963 (diameter ~38 μm), and CP964 (diameter ~45 μm).

Remarks

The basionym of Leiosphaeridia minutissima (Naumova, Reference Naumova1949) is Leiotriletes minutissimus Naumova, Reference Naumova1949. As for Leiotriletes crassus, Naumova (Reference Naumova1949) did not present a diagnosis for this species but provided a detailed description. Subsequently, Jankauskas in Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) transferred this species to Leiosphaeridia minutissima (Naumova, Reference Naumova1949) without presenting a diagnosis. When Leiotriletes minutissimus was transferred to the genus Leiosphaerida, the epithet was changed to minutissima, so the gender of the epithet agrees with the gender of the genus name. The first formal diagnosis for Leiosphaeridia minutissima was presented by Javaux and Knoll (Reference Javaux and Knoll2017), emended later by Knoll et al. (Reference Knoll, Germs, Tankard and Welsink2020).

Leiosphaeridia tenuissima Eisenack, 1958
Figure 3.9, 3.13

Reference Eisenack1958a

Leiosphaeridia tenuissima Eisenack, p. 391, pl. 1, figs. 2, 3.

Reference Eisenack1958b

Leiosphaeridia tenuissima; Eisenack, pl. 2, figs. 1, 2.

Reference Jankauskas, Mikhailova and German1989

Leiosphaeridia tenuissima; Jankauskas et al., p. 81, pl. 9, figs. 12, 13.

Reference Butterfield, Knoll and Swett1994

Leiosphaeridia tenuissima; Butterfield et al., p. 42, fig. 16i.

Reference Hofmann and Jackson1994

Leiosphaeridia tenuissima; Hofmann and Jackson, p. 22, fig. 15.16–15.18.

Reference Zhang, Yin, Xiao and Knoll1998

Leiosphaeridia tenuissima; Zhang et al., p. 32, fig. 9.7.

Reference Zhang, Yin, Xiao and Knoll1998

Leiosphaeridia spp. div.; Zhang et al., p. 32, fig. 9.8, 9.9

Reference Turnau and Racki1999

Leiosphaeridia tenuissima; Turnau and Racki, p. 267, pl. 5, fig. 1.

Reference Gaucher2000

Leiosphaeridia tenuissima; Gaucher, p. 68, pl. 11, fig. 5.

Reference Gaucher and Germs2003

Leiosphaeridia tenuissima; Gaucher and Germs, fig. 6.6.

Reference Javaux, Knoll and Walter2004

Leiosphaeridia tenuissima; Javaux et al., fig. 4j–l.

Reference Gaucher, Chiglino and Peçoits2004

Leiosphaeridia tenuissima; Gaucher et al., fig. 4d.

Reference Gaucher, Frimmel and Germs2005a

Leiosphaeridia tenuissima; Gaucher et al., p. 549, fig. 8g–h.

Reference Gaucher, Poire, Peral and Chiglino2005b

Leiosphaeridia tenuissima; Gaucher et al., fig. 6a–b, 6e–h.

Reference Blanco and Gaucher2005

Leiosphaeridia tenuissima; Blanco and Gaucher, fig. 11a.

Reference Grey2005

Leiosphaeridia tenuissima; Grey, p. 184, fig. 63h.

Reference Marshall, Javaux, Knoll and Walter2005

Leiosphaeridia tenuissima; Marshall et al., fig. 1d.

Reference Prasad, Uniyal and Asher2005

Leiosphaeridia tenuissima; Prasad et al., pl. 1, fig. 3, pl. 2, fig. 10, pl. 3, fig. 15, pl. 4, fig. 17, pl. 8, figs. 16, 17.

Reference Gaucher and Germs2006

Leiosphaeridia tenuissima; Gaucher and Germs, p. 207, figs. 7d, f, g, 8b–f.

Reference Javaux2007

Leiosphaeridia tenuissima; Javaux, fig. 1.18, 1.19.

Reference Gaucher, Chiglino, Blanco, Poiré and Germs2008

Leiosphaeridia tenuissima; Gaucher et al., p. 491, fig. 3b–i.

Reference Stanevich, Maksimova, Kornilova, Gladkochub, Mazukabzov and Donskaya2009

Leiosphaeridia tenuissima; Stanevich et al., p. 32, pl. 3, fig. 5.

Reference Prasad, Asher and Borgohai2010

Leiosphaeridia tenuissima; Prasad et al., pl. 1, fig. 1.

Reference Buick2010

Leiosphaeridia tenuissima; Buick, fig. 1e.

Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013

Leiosphaeridia tenuissima; Tang et al., fig. 4c.

Reference Liu, Xiao, Yin, Chen, Zhou and Li2014

Leiosphaeridia tenuissima; Liu et al., fig. 101.

Reference Vorob'eva and Petrov2014

Leiosphaeridia tenuissima; Vorob'eva and Petrov, fig. 6b.

Reference Schopf, Sergeev and Kudryavtsev2015

Leiosphaeridia tenuissima; Schopf et al., p. 724, fig. 13.9.

Reference Nagovitsin and Kochnev2015

Leiosphaeridia tenuissima; Nagovitsin and Kochnev, fig. 4.59.

Reference Chiglino, Gaucher, Sial and Ferreira2015

Leiosphaeridia tenuissima; Chiglino et al., p. 640, fig. 4a–c.

Reference Tang, Pang, Yuan, Wan and Xiao2015

Leiosphaeridia tenuissima; Tang et al., fig. 4e.

Reference Vorob'eva, Sergeev and Petrov2015

Leiosphaeridia tenuissima; Vorob'eva et al., fig. 7.8.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Leiosphaeridia tenuissima; Baludikay et al., fig. 8f.

Reference Porter and Riedman2016

Leiosphaeridia tenuissima; Porter and Riedman, p. 834, fig. 13.4.

Reference Sergeev, Knoll, Vorob'eva and Sergeeva2016

Leiosphaeridia tenuissima; Sergeev et al., fig. 4.2.

Reference Singh and Sharma2016

Leiosphaeridia tenuissima; Singh and Sharma, p. 81, pl. 1, figs. 12, 15.

Reference Beghin, Storme, Blanpied, Gueneli, Brocks, Poulton and Javaux2017

Leiosphaeridia tenuissima; Beghin et al., pl. 2, fig. j.

Reference Tang, Hughes, McKenzie, Myrow and Xiao2017

Leiosphaeridia tenuissima; Tang et al., fig. 3b.

Reference Agić, Moczydłowska and Yin2017

Leiosphaeridia tenuissima; Agic et al., p. 112, fig. 8d, f.

Reference Suslova, Parfenova, Saraev and Nagovitsyn2017

Leiosphaeridia tenuissima; Suslova et al., fig. 3.13, 3.14.

Reference Sergeev, Vorob'eva and Petrov2017a

Leiosphaeridia tenuissima; Sergeev et al., fig. 3.12.

Reference Sergeev, Vorob'eva and Petrov2017a

Leiosphaeridia minutissima; Sergeev et al., fig. 3.13.

Reference Sergeev, Vorob'eva, Petrov and Semikhatov2017b

Leiosphaeridia tenuissima; Sergeev et al., pl. 1, figs. 7, 9.

Reference Anderson, McMahon, Macdonald, Jones and Briggs2019

Leiosphaeridia tenuissima; Anderson et al., p. 512, figs. 8l, m, 15k.

Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019

Leiosphaeridia tenuissima; Arrouy et al., figs. 6a, 7a–d.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Leiosphaeridia tenuissima; Li et al., fig. 4g.

Reference Tang, Hu, Xie, Yuan and Wan2019

Leiosphaeridia tenuissima; Tang et al., fig. 1.2–1.5.

Reference Wan, Tang, Pang, Wang, Bao, Meng, Zhou, Yuan, Hua and Xiao2019

Leiosphaeridia tenuissima; Wan et al., fig. 4f.

Reference Arvestål and Willman2020

Leiosphaeridia tenuissima; Arvestål and Willman, p. 12, fig. 6a, b.

Reference Shukla, Sharma and Sergeev2020

Leiosphaeridia tenuissima; Shukla et al., p. 502, fig. 6a–d, 6f.

Reference Pang, Tang, Wu, Li, Chen, Wan, Yuan, Bodnar and Xiao2020

Leiosphaeridia tenuissima; Pang et al., fig. 2c.

Reference Han, Chen, Li, Pang and Wang2021

Leiosphaeridia tenuissima; Han et al., fig. 3e.

Reference Tang, Pang, Li, Chen, Yuan, Sharma and Xiao2021

Leiosphaeridia tenuissima; Tang et al., fig. 9a.

Reference Loron, Halverson, Rainbird, Skulski, Turner and Javaux2021

Leiosphaeridia tenuissima; Loron et al., fig. 6.1, 6.3.

Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022

Leiosphaeridia tenuissima; Denezine et al., fig. 11.3.

Holotype

Preparation A3, 3 number 4, from the Dictyonema shales of the Ordovician Baltic, Nikolskaya on the Tossna, southeast Leningrad (Eisenack, Reference Eisenack1958a, pl. 1, fig. 2).

Original diagnosis presented by Eisenack (Reference Eisenack1958a) in Germany

“Wand äußerst dünn und zart, glasklar durchscheinend, ohne Wandporen; nur in flachgedrücktem Zustand in Form von fast kreisrunden Scheibchen überliefert. Pylome nicht beobachtet. Ø um rd 100 μm schwankend.”

Translation of original diagnosis presented by Eisenack (Reference Eisenack1958a)

Wall extremely thin and delicate, crystalline translucent, without wall pores; only preserved in the flattened state in the form of almost circular disks. Pyloma not observed. Size around 100 μ.

Emended diagnosis by Javaux and Knoll (Reference Javaux and Knoll2017)

“A species of Leiosphaeridia characterized by smooth walls with sinuous folds and a modal diameter (rather than maximum diameter) greater than 70 μm; the wall color is not a diagnostic criteria.”

Occurrence in the studied section

Fourteen specimens were recovered. They range from ~72 to ~126 μm in diameter: MP3002, MP3007, MP2994, MP3707, MP3709, MP3013, MP3714, MP3719, and MP3720.

Illustrated materials

CP914 (diameters: ~81 μm in Fig. 3.9 and ~86 μm in Fig. 3.13).

Remarks

Both Leiosphaeridia tenuissima Eisenack, 1958 and Leiosphaeridia minutissima (Naumova, Reference Naumova1949,) are simple sphaeromorphs and have a thin and translucent wall less than 0.5 μm thick. However, Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989) differentiated them on the basis of vesicle size, defining specimens smaller than 70 μm in diameter as Leiosphaeridia minutissima and specimens larger than 70 μm as Leiosphaeridia tenuissima. The specimen illustrated by Sergeev et al. (Reference Sergeev, Vorob'eva and Petrov2017a, fig. 3.13) as Leiosphaeridia minutissima is better identified as Leiosphaeridia tenuissima due to its greatest diameter of about 105 μm.

Subgroup Acanthomorphitae Downie et al., Reference Downie, Evitt and Sarjeant1963
Genus Germinosphaera Mikhailova, Reference Mikhailova and Sokolov1986

Type species

Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986.

Other species

Germinosphaera guttaformis Mikhailova in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989; Germinosphaera alveolata Miao et al. (Reference Miao, Moczydłowska, Zhu and Zhu2019).

Original diagnosis presented by Mikhailova (Reference Mikhailova and Sokolov1986) in Russian

“Оболочки округлые, округло-овальные, плотные, толстые, гладкие или шагреневые, проросшие. Отростки, которые могут ветвиться, наблюдаются на одном или двух поолюсах.”

Translation of original diagnosis presented by Mikhailova (Reference Mikhailova and Sokolov1986)

The shells are round or round-oval, dense, thick, smooth or shagreen, sprouted. Processes are observed at one or two poles.

Emended diagnosis by Butterfield et al. (Reference Butterfield, Knoll and Swett1994)

“Spheroidal vesicles with 1–6 open-ended, tubular, and occasionally branched processes that communicate freely with the vesicle. Multiple processes usually restricted to a single ‘equatorial’ plane, but otherwise non-uniformly distributed on the vesicle.”

Emended diagnosis by Miao et al. (Reference Miao, Moczydłowska, Zhu and Zhu2019)

“Vesicle spheroidal, teardrop-shaped to slightly irregular outline, having psilate or low relief sculptured alveolar wall surface and bearing a single to multiple processes. Processes are simple tubular or occasionally branching, and open-ended. Processes are distributed [irregularly] on the vesicle wall, if multiple, and may be predominantly, but not exclusively, distributed in the equatorial plane of the vesicle.”

Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986
Figure 3.4, 3.8, 3.11

Reference Mikhailova and Sokolov1986

Germinosphaera bispinosa Mikhailova, p. 33, fig. 6.

Reference Mikhailova and Sokolov1986

Germinosphaera unispinosa Mikhailova, p. 33, fig. 5.

Reference Jankauskas, Mikhailova and German1989

Germinosphaera bispinosa; Jankauskas et al., p. 142, pl. 47, fig. 2.

Reference Jankauskas, Mikhailova and German1989

Germinosphaera tadasii Weis in Jankauskas et al., p. 143, pl. 47, figs. 3–5.

Reference Jankauskas, Mikhailova and German1989

Germinosphaera unispinosa Jankauskas et al., p. 143, pl. 47, fig. 1.

Reference Knoll, Swett and Mark1991

Germinosphaera sp.; Knoll et al., p. 557, fig. 19.6.

Reference Yan and Liu1993

Gemmispora rudis Yan in Yan and Liu, pl. I, figs. 6, 7.

Reference Butterfield, Knoll and Swett1994

Germinosphaera fibrilla (Ouyang et al., Reference Ouyang, Yin and Zaiping1974); Butterfield et al., p. 38, fig. 17a–h.

Reference Butterfield, Knoll and Swett1994

Germinosphaera bispinosa; Butterfield et al., p. 38, fig. 16d, e.

Reference Butterfield, Knoll and Swett1994

Germinosphaera jankauskasii Butterfield in Butterfield et al., p. 38, fig. 16a–c.

Reference Zang1995

Germinosphaera sp. cf. G. unispinosa; Zang, p. 164, fig. 26k, l.

Reference Yin and Guan1999

Germinosphaera unispinosa; Yin and Guan, p. 128, fig. 5.2, 5.4, 5.6, 5.9.

Reference Prasad, Uniyal and Asher2005

Germinosphaera bispinosa; Prasad et al., p. 44, pl. 11, fig. 3.

Reference Prasad, Uniyal and Asher2005

Germinosphaera unispinosa; Prasad et al., p. 44, pl. 11, figs. 1, 2.

Reference Yin and Yuan2007

Germinosphaera unispinosa; Yin and Yuan, fig. 2.11.

Reference Vorob'eva, Sergeev and Knoll2009

Germinosphaera sp.; Vorob'eva et al., p. 191, fig. 13.13–13.15, 13.17.

Reference Baludikay, Storme, François, Baudet and Javaux2016

Germinosphaera bispinosa; Baludikay et al., fig. 6a–c.

Reference Loron and Moczydłowska2017

Germinosphaera bispinosa; Loron and Moczydłowska, p. 24, pl. 1, fig. 3.

Reference Li, Pang, Chen, Zhou, Han, Yang, Wang, Yang and Yin2019

Germinosphaera bispinosa; Li et al., fig. 10c–g.

Reference Loron, Rainbird, Turner, Greenman and Javaux2019

Germinosphaera bispinosa; Loron et al., fig. 8e–f.

Reference Miao, Moczydłowska, Zhu and Zhu2019

Germinosphaera bispinosa; Miao et al., p. 187, fig. 5d, f.

Reference Miao, Moczydłowska and Zhu2021

Germinosphaera bispinosa; Miao et al., p. 14, fig. 5d, e.

Reference Denezine, Adôrno, Do Carmo, Guimarães, Walde, Alvarenga, Germs, Antonietto, Gianfranco and Nunes Junior2022

Germinosphaera bispinosa; Denezine et al., fig. 11.4.

Holotype

Number 882/2 from the Krasnoyarsk region, River Uderei; Upper Riphean, Dashkin Formation (Mikhailova, Reference Mikhailova and Sokolov1986, fig. 6).

Diagnosis by Butterfield in Butterfield et al. (Reference Butterfield, Knoll and Swett1994)

“A species of Germinosphaera with psilate vesicles 13–35 μm in diameter. Processes 2.5–3.5 μm wide and, when multiple, arranged equatorially on the vesicle.”

Emended diagnosis by Miao et al. (Reference Miao, Moczydłowska, Zhu and Zhu2019)

“Spheroidal to slightly elongate or irregular vesicle with one to multiple tubular processes. Vesicle wall psilate. Processes may [be] arranged irregularly or equatorially on the vesicle wall when multiple.”

Occurrence in the studied section

Twenty-three specimens were recovered: MP3036 and MP3714.

Description

Vesicles are 23.4–34.8 μm in diameter, bearing one to two processes. When two processes are present, they are inserted at two opposing ends of the vesicle (Fig. 3.4). Processes typically taper slightly toward their distal end (Fig. 3.4, lower process, 3.8) or are more or less cylindrical (Fig. 3.11). One of the processes in the specimen illustrated in Fig. 3.4 is apparently constricted at the base. However, it is uncertain whether this constriction is a taphonomic feature related to the twisting of the process. Processes are 1–3 μm in maximum diameter and 22.6–123.0 μm in preserved length.

Illustrated material

CP917.

Remarks

Mikhailova (Reference Mikhailova and Sokolov1986) established two species of Germinosphaera, Germinosphaera unispinosa and Germinosphaera bispinosa. Two additional species were published by Jankauskas et al. (Reference Jankauskas, Mikhailova and German1989), Germinosphaera guttaformis Mikhailova in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989 and Germinosphaera tadasii Weiss in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989. These species were distinguished by the number of processes and the psilate versus shagrinate nature of vesicle walls. However, Butterfield et al. (Reference Butterfield, Knoll and Swett1994) considered the possibility that the processes in Germinosphaera represent growth structures in vegetative stages, analogous to the modern xanthophyte Vaucheria. As such, they emended the diagnosis of Germinosphaera and the diagnosis of G. bispinosa, and they synonymized G. unispinosa with G. bispinosa. Miao et al. (Reference Miao, Moczydłowska, Zhu and Zhu2019) further emended the diagnosis of Germinosphaera and considered shagrinate vesicle walls to represent taphonomic alteration. Furthermore, they noted that the vesicle diameters of different species could overlap each other. Thus, they proposed that G. tadasii and G. jankauskasii, which are characterized by shagrinate vesicle walls, were junior synonyms of G. bispinosa. Following Miao et al. (Reference Miao, Moczydłowska, Zhu and Zhu2019), Germinosphaera currently has three species: Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986, Germinosphaera guttaformis Mikhailova in Jankauskas et al., Reference Jankauskas, Mikhailova and German1989, and Germinosphaera alveolata Miao et al., Reference Miao, Moczydłowska, Zhu and Zhu2019.

Species diversity and abundance

The analysis of 80 samples, 53 of which contained microfossils, yielded a modest diversity of organic-walled microfossils, including seven species of four genera: Siphonophycus robustum (Schopf, Reference Schopf1968), Leiosphaeridia crassa (Naumova, Reference Naumova1949), Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966), Leiosphaeridia minutissima (Naumova, Reference Naumova1949), Leiosphaeridia tenuissima Eisenack, 1958, Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986, and Ghoshia januarensis new species. Following Butterfield et al. (Reference Butterfield, Knoll and Swett1994), the four morphospecies of Leiosphaeridia are differentiated on the basis of their vesicle diameter and wall thickness: Leiosphaeridia minutissima has thin-walled vesicles less than 70 μm in diameter, Leiosphaeridia tenuissima has thin-walled vesicles 70–200 μm in diameter, Leiosphaeridia crassa has thicker-walled vesicles less than 70 μm in diameter, and Leiosphaeridia jacutica has thicker-walled vesicles 70–800 μm in diameter (Fig. 6). Only one species of acanthomorphs is reported, Germinosphaera bispinosa, a smooth vesicle with one or two unbranched processes that are either cylindrical or slightly tapered toward the distal end.

Figure 6. Abundance and size distribution of Leiosphaeridia species from the Sete Lagoas Formation at the Barreiro section.

Filamentous microfossils are common in the Sete Lagoas Formation. Tubular filamentous microfossils recovered in this work are represented by the morphospecies Siphonophycus robustum, which is interpreted as remains of cyanobacterial sheaths. This work follows Knoll et al. (Reference Knoll, Swett and Mark1991), Butterfield et al. (Reference Butterfield, Knoll and Swett1994), and Tang et al. (Reference Tang, Pang, Xiao, Yuan, Ou and Wan2013) in distinguishing Siphonophycus species according to their filament diameter: Siphonophycus thulenema, 0.5 μm in diameter; Siphonophycus septatum, 1–2 μm; Siphonophycus robustum, 2–4 μm; Siphonophycus typicum, 4–8 μm; Siphonophycus kestron, 8–16 μm; Siphonophycus solidum, 16–32 μm; Siphonophycus punctatum, 32–64 μm; and Siphonophycus gigas, 64–128 μm. In addition to tubular filaments, branching filaments of uniserially chained cells from the Sete Lagoas Formation are identified as Ghoshia januarensis new species.

The Sete Lagoas assemblage is numerically dominated by sphaeromorphs. Nearly all fossiliferous samples contain the sphaeromorph genus Leiosphaeridia, and Leioshpaeridia minutissima is the most abundant species (Fig. 6), with 359 specimens (~93% of all Leiosphaeridia specimens) and 1–64 occurrences per horizon in 45 horizons (Fig. 7). By contrast, Leiosphaeridia crassa, Leiosphaeridia jacutica, Leiosphaeridia tenussima, and Germinosphaera bispinosa are rare, represented by 14, 4, 9, and 19 specimens, respectively. About 73% of acritarch specimens recovered from the Sete Lagoas Formation are <40 μm in diameter, highlighting the predominance of small organic-walled microfossils in this unit.

Figure 7. Stratigraphic distribution and relative abundance of organic-walled microfossils from the Sete Lagoas Formation at the Barreiro section.

The organic-walled microfossil assemblage recovered in this study is taxonomically different from those of previous micropaleontological studies of the Sete Lagoas Formation (e.g., Fairchild et al., Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996; Table 1). This difference is likely related to variations in paleoenvironment, paleoecology, taphonomy, and fossil preparation techniques. Previous micropaleontological studies of the Sete Lagoas Formation were focused exclusively on cherts, particularly silicified stromatolites and microbialites. Microfossils recovered in those studies were dominated by benthic microorganisms that constructed or dwelled in microbial mats. For example, Fairchild et al. (Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996) documented abundant filamentous and coccoidal microfossils (e.g., Siphonophycus, Myxococcoides, Gloeodiniopsis) from silicified microbialites of the Sete Lagoas Formation in the State of Goiás (their localities 20–22), more than 350 km to the west of the Barreiro section investigated in this study; only rare acritarchs identified as cf. Leioshpaeridia sp. were reported (see Fairchild et al., Reference Fairchild, Schopf, Shen-Miller, Guimarães, Edwards, Lagstein, Li, Pabst and Melo Filho1996, fig, 5h, table 2). By contrast, our investigation was focused on the lime mudstone and fine-grained limestone facies dominated by planktonic microfossils such as Leioshpaeridia minutissima. It is well known that Precambrian chert and fine-grained siliciclastic facies tend to host taxonomically distinct microfossils, with benthic microbial mat communities dominating the former facies and planktons prevailing in the latter (Butterfield and Chandler, Reference Butterfield and Chandler1992). Thus, the taxonomic difference between this and previous studies of the Sete Lagoas Formation is likely a result of paleoenvironmental and paleoecological differences.

Taphonomic differences may also have played a role in the taxonomic difference between this and previous studies. Silicification of microfossils is fundamentally a three-dimensional cast-and-mold process at the cellular level (Xiao and Tang, Reference Xiao and Tang2021), whereas organic-walled microfossils in fine-grained siliciclastic facies are preserved through two-dimensional compression of recalcitrant organic structures (Butterfield, Reference Butterfield1990), aided by clay mineral coating (e.g., Anderson et al., Reference Anderson, Schiffbauer and Xiao2011). Thus, the taxonomic difference between chert and lime mudstone and fine-grained limestone facies of the Sete Lagoas Formation is related at least partially to taphonomic variations. However, because both paleoecological and taphonomic processes are intertwined with paleoenvironmental conditions, it is impossible to disentangle the paleoecological and taphonomic factors that may have contributed to the observed taxonomic differences.

Finally, methodological differences may also have played a part in the taxonomic difference between this and previous micropaleontological studies of the Sete Lagoas Formation. Organic-walled microfossils preserved in cherts are typically observed in petrographic thin sections, whereas those in fine-grained siliciclastic rocks can be extracted for microscopic analysis. Such methodological difference can lead to biases in microfossil recovery and taxonomic identification, as shown in Ediacaran acritarchs (Xiao et al., Reference Xiao, Jiang, Ye, Ouyang, Banerjee, Singh, Muscente, Zhou and Hughes2023), which have been processed using petrographic thin sections (for those preserved three-dimensionally in cherts), hydrofluoric acid extraction (for those preserved two-dimensionally in shales), and both thin sections and acetic acid extraction (for those preserved three-dimensionally in phosphatic carbonate rocks). Therefore, the taxonomic differences between this and previous studies of the Sete Lagoas Formation likely result from a combination of paleoenvironmental, paleoecological, taphonomic, and methodological factors.

Stratigraphic distribution and biostratigraphy

Organic-walled microfossils in the Sete Lagoas Formation at the Barreiro section range from the base of the measured section to 57.5 m stratigraphic height (Fig. 7). Ghoshia januarensis has the greatest range of all species recovered in this work, occurring at nine stratigraphic levels from 0.4 m to 57.5 m. Leiosphaeridia minutissima is the longest-ranging sphaeromorph, occurring at 45 stratigraphic levels from the base of the section to 36.4 m. Leiosphaeridia tenuissima ranges from 1.3 to 36.4 m and is present in nine horizons, showing almost the same stratigraphic range as Leiosphaeridia minutissima. As a minor component of the assemblage, Leiosphaeridia jacutica was recovered from three levels in the interval of 2.8–31.5 m, Leiosphaeridia crassa from two levels in 31.5–36.4 m, Germinosphaera bispinosa from two horizons in 23.5–26.5 m, and Siphonophycus robustum from four levels in 1.9–28.0 m.

Except for Leiosphaeridia crassa and Germinosphaera bispinosa, all recovered species have their first appearance within 2 m above the base of the studied section, where there is a predominance of lime mudstone. Leiosphaeridia crassa and Germinosphaera bispinosa first emerge in the middle part of the section below the intraclastic breccia beds. The disappearance of organic-walled microfossils in the Sete Lagoas Formation is gradual, although three species (Leiosphaeridia crassa, Leiosphaeridia minutissima, and Leiosphaeridia tenuissima) disappear at approximately 37 m. No organic-walled microfossils other than Ghoshia januarensis were recovered above 37 m, right before a great abundance of intraclastic breccias, which were interpreted as seismic deposits by Okubo et al. (Reference Okubo, Warren, Luvizotto, Varejão, Quaglio, Uhlein and Assine2020).

With the exception of Ghoshia januarensis, organic-walled microfossils from the Sete Lagoas Formation described in this paper have very long stratigraphic ranges when global data are considered. For example, the four Leiosphaeridia species recovered in this work range from the Mesoproterozoic to the Cambrian (Grey, Reference Grey2005). Both Germinosphaera bispinosa and Siphonopycus robustum are known from the late Paleoproterozoic to the Paleozoic (Butterfield et al., Reference Butterfield, Knoll and Swett1994; Sergeev et al., Reference Sergeev, Sharma and Shukla2012; Miao et al., Reference Miao, Moczydłowska, Zhu and Zhu2019).

Because leiospheric sphaeromorphs have rather long stratigraphic ranges globally, they have limited utility for global biostratigraphic correlation, which casts doubt on the biostratigraphic significance of leiosphere-based biozones. Nonetheless, Grey (Reference Grey2005) established the Ediacaran Leiosphere Palynoflora, and Gaucher and Sprechmann (Reference Gaucher and Sprechmann2009) proposed the Early Ediacaran Leiosphere Palynoflora. Both were regarded as early Ediacaran (ca. 635–580 Ma) acritarch biozones. More recent studies from South China, however, recovered abundant and diverse acanthomorphs from early Ediacaran strata (Zhou et al., Reference Zhou, Xie, McFadden, Xiao and Yuan2007; Liu and Moczydłowska, Reference Liu and Moczydłowska2019; Ouyang et al., Reference Ouyang, Zhou, Xiao, Guan, Chen, Yuan and Sun2021), indicating that the leiosphere-based biozones of Grey (Reference Grey2005) and Gaucher and Sprechmann (Reference Gaucher and Sprechmann2009) are controlled by local environmental, regional biogeographical, or taphonomic factors.

It is perceived that the terminal Ediacaran (ca. 550–539 Ma) is characterized by a leiosphere assemblage (Knoll and Walter, Reference Knoll and Walter1992; Gaucher and Sprechmann, Reference Gaucher and Sprechmann2009) (Fig. 17). Gaucher and Sprechmann (Reference Gaucher and Sprechmann2009) presented the Late Ediacaran Leiosphere Palynoflora, which is a low-diversity assemblage characterized by small sphaeromorphs (<150 μm) such as Leiosphaeridia minutissima and Leiosphaeridia tenuissima, among others (Fig. 17). In addition, there are occurrences of Chuaria circularis, as well as Bavlinela faveolata Shepeleva, Reference Shepeleva1962, Soldadophycus bossii Gaucher et al., Reference Gaucher, Sprechmann and Schipilov1996, and small acanthomorphs, such as Asteridium spp. The Late Ediacaran Leiosphere Palynoflora, sensu Gaucher and Sprechmann (Reference Gaucher and Sprechmann2009), has been documented in the Nama Group in Namibia (Germs et al., Reference Germs, Knoll and Vidal1986), the Holgat Formation of the Port Nolloth Group in Namibia (Gaucher et al., Reference Gaucher, Frimmel and Germs2005a), the Mulden Group in Namibia (Gaucher and Germs, Reference Gaucher and Germs2007), the Tent Hill Formation in Australia (Damassa and Knoll, Reference Damassa and Knoll1986), Cijara Formation in Spain (Palacios, Reference Palacios1989), the Cango Caves and Gamtoos groups in South Africa (Gaucher and Germs, Reference Gaucher and Germs2006), the Dengying Formation in South China (Yin and Yuan, Reference Yin and Yuan2007), the Arroyo del Soldado Group in Uruguay (Gaucher, Reference Gaucher2000; Gaucher et al., Reference Gaucher, Boggiani, Sprechmann, Sial and Fairchild2003), the Sierras Bayas Group in Argentina (Cingolani et al., Reference Cingolani, Rauscher and Bonhomme1991; Gaucher et al., Reference Gaucher, Poire, Peral and Chiglino2005b), the La Providencia Group in Argentina (Arrouy et al., Reference Arrouy, Gaucher, Poiré, Xiao, Peral, Warren, Bykova and Quaglio2019), and the Corumbá Group in Brazil (Zaine, Reference Zaine1991; Gaucher et al., Reference Gaucher, Boggiani, Sprechmann, Sial and Fairchild2003; Tobias, Reference Tobias2014). In Namibia, Argentina, Uruguay, and Brazil (Germs et al., Reference Germs, Knoll and Vidal1986; Gaucher et al., Reference Gaucher, Boggiani, Sprechmann, Sial and Fairchild2003, Reference Gaucher, Poire, Peral and Chiglino2005b; Tobias, Reference Tobias2014), the Late Ediacaran Leiosphere Palynoflora occurs in association with biomineralized tubular fossils such as Cloudina lucianoi (Beurlen and Sommer, Reference Beurlen and Sommer1957), Cloudina riemkeae Germs, Reference Germs1972, and Corumbella werneri Hahn et al., Reference Hahn, Hahn, Leonardos, Pflug and Walde1982, which are potential index fossils for the terminal Ediacaran.

Given the predominance of a depauperate leiosphere assemblage in the Sete Lagoas Formation (this study) and the previous report of Cloudina and Corumbella from this unit (Warren et al., Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014), it is tempting to consider that the organic-walled microfossil assemblage reported in this study is correlated with the Late Ediacaran Leiosphere Palynoflora. There are, however, two caveats. First, as discussed by Xiao and Narbonne (Reference Xiao, Narbonne, Gradstein, Ogg, Schmitz and Ogg2020), several recent studies have shown acanthomorphs as a group may extend from the lowermost Ediacaran to the terminal Ediacaran stage in Mongolia (Anderson et al., Reference Anderson, Macdonald, Jones, McMahon and Briggs2017, Reference Anderson, McMahon, Macdonald, Jones and Briggs2019) and in Siberia (Golubkova et al., Reference Golubkova, Zaitseva, Kuznetsov, Dovzhikova and Maslov2015; but see Vorob'eva et al., Reference Vorob'eva, Sergeev and Knoll2009), to rocks considered younger than the Shuram Excursion (Ouyang et al., Reference Ouyang, Guan, Zhou and Xiao2017) or even to the early Cambrian (Grazhdankin et al., Reference Grazhdankin, Nagovitsin, Golubkova, Karlova, Kochnev, Rogov and Marusin2020). Second, a systematic description of the purported Cloudina and Corumbella fossils from the Sete Lagoas Formation (Warren et al., Reference Warren, Quaglio, Riccomini, Simões, Poiré, Strikis, Anelli and Strikis2014; Perrella Júnior et al., Reference Perrella Júnior, Uhlein, Uhlein, Sial, Pedrosa-Soares and Lima2017) is needed to assess their species-level identification and to support their biostratigraphic significance. Nonetheless, considering the maximum age constraint of ~557 Ma provided by detrital zircons from the upper Sete Lagoas Formation (Paula-Santos et al., Reference Paula-Santos, Babinski, Kuchenbecker, Caetano-Filho, Trindade and Pedrosa-Soares2015), the organic-walled microfossil assemblage reported in this paper is consistent with a terminal Ediacaran age interpretation.

Conclusions

A modest diversity of organic-walled microfossils is reported from the Sete Lagoas Formation of Bambuí Group at the Barreiro section in the Januária area of the São Francisco basin, central Brazil. Seven species are described: Siphonophycus robustum (Schopf, Reference Schopf1968), Ghoshia januarensis new species, Leiosphaeridia crassa (Naumova, Reference Naumova1949), Leiosphaeridia jacutica (Timofeev, Reference Timofeev1966), Leiosphaeridia minutissima (Naumova, Reference Naumova1949), Leiosphaeridia tenuissima Eisenack, 1958, and Germinosphaera bispinosa Mikhailova, Reference Mikhailova and Sokolov1986. The first two species are considered cyanobacteria, the four Leiosphaeridia species are considered possible protists, and the phylogenetic affinity of Germinosphaera bispinosa is uncertain. All species occur in the lower part of the studied section, but only Ghoshia januarensis extends to the upper portion of the studied section. The assemblage is numerically dominated by Leiosphaeridia, with Leiosphaeridia minutissima being the most abundant species. The predominance of Leiosphaeridia indicates that the Sete Lagoas organic-walled microfossil assemblage may be correlated with the Late Ediacaran Leiosphere Palynoflora, consistent with a terminal Ediacaran age interpretation inferred from detrital zircon data, Cloudina, and Corumbella from the Sete Lagoas Formation. However, we emphasize that further investigation is needed to test this age interpretation.

Acknowledgments

We thank all institutions that participated in the development of this research: the National Council for Scientific and Technological Development (CNPq), Coordination for the Improvement of Higher Education Personnel (CAPES), Geological Survey of Brazil (CPRM), Brazilian Petroleum Corporation (PETROBRAS), National Agency of Petroleum, Gas, and Biofuels (ANP), and the University of Brasília (UnB). We thank FINATEC for assistance in administrative affairs supporting scientific projects in Brasília. We thank A.H. Giles for grammar review and the undergraduate geology students L.R.O. Gonçalve and A.S. Reis for curating samples at UnB. This study was financed in part by the Coordination for the Improvement of Higher Education Personnel—Brazil (CAPES)—Finance Code 001. S.X. acknowledges the support from the National Science Foundation (EAR-2021207). C. Z. acknowledges the support from São Paulo Research Foundation (2021/12304-0).

Declaration of competing interests

The authors declare none.

Footnotes

Deceased.

References

Adôrno, R.R., Do Carmo, D.A., Germs, G.J.B., Walde, D.H.G., Denezine, M., et al., 2017, Cloudina lucianoi (Beurlen & Sommer, 1957), Tamengo Formation, Ediacaran, Brazil: taxonomy analysis of stratigraphic distribution and biostratigraphy: Precambrian Research, v. 301, p. 1935.CrossRefGoogle Scholar
Agić, H., Moczydłowska, M., and Yin, L., 2017, Diversity of organic-walled microfossils from the early Mesoproterozoic Ruyang Group, North China Craton—a window into the early eukaryote evolution: Precambrian Research, v. 297, p. 101130.CrossRefGoogle Scholar
Alvarenga, C.J.S., Dardenne, M.A., Vieira, L.C., Martinho, C.T., Guimarães, E.M., Santos, R. V., and Santana, R.O., 2012, Estratigrafia da borda ocidental da Bacia do São Francisco: Boletim de Geociências Da PETROBRAS, v. 20, p. 145164.Google Scholar
Alvarenga, C.J.S., Santos, R.V., Vieira, L.C., Lima, B.A.F., and Mancini, L.H., 2014, Meso-Neoproterozoic isotope stratigraphy on carbonates platforms in the Brasília belt of Brazil: Precambrian Research, v. 251, p. 164180.CrossRefGoogle Scholar
Anderson, E., Schiffbauer, J.D., and Xiao, S., 2011, Taphonomic study of Ediacaran organic-walled fossils confirms the importance of clay minerals and pyrite in Burgess Shale-type preservation: Geology, v. 39, p. 643646.CrossRefGoogle Scholar
Anderson, R.P., Macdonald, F.A., Jones, D.S., McMahon, S., and Briggs, D.E.G., 2017, Doushantuo-type microfossils from latest Ediacaran phosphorites of northern Mongolia: Geology, v. 45, p. 10791082.CrossRefGoogle Scholar
Anderson, R.P., McMahon, S., Macdonald, F.A., Jones, D.S., and Briggs, D.E.G., 2019, Palaeobiology of latest Ediacaran phosphorites from the upper Khesen Formation, Khuvsgul Group, northern Mongolia: Journal of Systematic Palaeontology, v. 17, p. 501532.CrossRefGoogle Scholar
Arrouy, M.J., Gaucher, C., Poiré, D.G., Xiao, S., Peral, L.E.G., Warren, L. V., Bykova, N., and Quaglio, F., 2019, A new record of late Ediacaran acritarchs from La Providencia Group (Tandilia System, Argentina) and its biostratigraphical significance: Journal of South American Earth Sciences, v. 93, p. 283293.CrossRefGoogle Scholar
Arvestål, E.H.M., and Willman, S., 2020, Organic-walled microfossils in the Ediacaran of Estonia: biodiversity on the East European Platform: Precambrian Research, v. 341, n. 105626, https://doi.org/10.1016/j.precamres.2020.105626.CrossRefGoogle Scholar
Babinski, M., Vieira, L.C., and Trindade, R.I.F., 2007, Direct dating of the Sete Lagoas cap carbonate (Bambuí Group, Brazil) and implications for the Neoproterozoic glacial events: Terra Nova, v. 19, p. 401406.CrossRefGoogle Scholar
Babu, R., Singh, V.K., and Mehrotra, N.C., 2014, Neoproterozoic age based on microbiotas from the Raipur Group of Baradwar Sub-basin, Chhattisgarh: Journal Geological Society of India, v. 84, p. 442448.CrossRefGoogle Scholar
Baludikay, B.K., Storme, J.Y., François, C., Baudet, D., and Javaux, E.J., 2016, A diverse and exquisitely preserved organic-walled microfossil assemblage from the Meso-Neoproterozoic Mbuji-Mayi Supergroup (Democratic Republic of Congo) and implications for Proterozoic biostratigraphy: Precambrian Research, v. 281, p. 166184.CrossRefGoogle Scholar
Barbosa, O., 1963, Geologia Econômica e Aplicada a Uma Parte Do Planalto Central: Goiânia, DNPM/PROSPEC, 70 p.Google Scholar
Beghin, J., Storme, J.Y., Blanpied, C., Gueneli, N., Brocks, J.J., Poulton, S.W., and Javaux, E.J., 2017, Microfossils from the late Mesoproterozoic–early Neoproterozoic Atar/El Mreïti Group, Taoudeni Basin, Mauritania, northwestern Africa: Precambrian Research, v. 291, p. 6382.CrossRefGoogle Scholar
Beurlen, K., and Sommer, F.W., 1957, Observações estratigráficas e paleontológicas sobre o calcário Corumbá: Boletim Da Divisão de Geologia e Mineralogia, v. 168, p. 135.Google Scholar
Blanco, G., and Gaucher, C., 2005, Estratigrafia, paleontologia y edad de la Formacion Las Ventanas (Neoproterozoico, Uruguay): Latin American Journal of Sedimentology and Basin Analysis, v. 12, p. 109124.Google Scholar
Buick, R., 2010, Ancient acritarchs: Nature, v. 463, p. 885886.CrossRefGoogle ScholarPubMed
Butterfield, N.J., 1990, Organic preservation of non-mineralizing organisms and the taphonomy of the Burgess Shale: Paleobiology, v. 16, p. 272286.CrossRefGoogle Scholar
Butterfield, N.J., and Chandler, F.W., 1992, Palaeoenvironmental distribution of Proterozoic microfossils, with an example from the Agu Bay Formation, Baffin Island: Palaeontology, v. 35, p. 943957.Google Scholar
Butterfield, N.J., Knoll, A.H., and Swett, K., 1994, Paleobiology of the Neoproterozoic Svanbergfjellet Formation, Spitsbergen: Fossils and Strata, v. 34, 84 p.CrossRefGoogle Scholar
Caetano-Filho, S., Paula-Santos, G.M., Guacaneme, C., Babinski, M., Bedoya-Rueda, C., et al., 2019, Sequence stratigraphy and chemostratigraphy of an Ediacaran–Cambrian foreland-related carbonate ramp (Bambuí Group, Brazil): Precambrian Research, v. 331, n. 105365.CrossRefGoogle Scholar
Caetano-Filho, S., Sansjofre, P., Ader, M., Paula-Santos, G.M., Guacaneme, C., Babinski, M., Bedoya-Rueda, C., Kuchenbecker, M., Reis, H.L.S., and Trindade, R.I.F., 2021, A large epeiric methanogenic Bambuí sea in the core of Gondwana supercontinent? Geoscience Frontiers, v. 12, p. 203218.CrossRefGoogle Scholar
Caxito, F.A., Lana, C., Frei, R., Uhlein, G.J., Sial, A.N., et al., 2021, Goldilocks at the dawn of complex life: mountains might have damaged Ediacaran–Cambrian ecosystems and prompted an early Cambrian greenhouse world: Scientific Reports, v. 11, n. 20010.CrossRefGoogle ScholarPubMed
Chiglino, L., Gaucher, C., Sial, A.N., and Ferreira, V.P., 2015, Acritarchs of the Ediacaran Frecheirinha Formation, Ubajara Group, Northeastern Brazil: Anais Da Academia Brasileira de Ciências, v. 87, p. 635649.CrossRefGoogle ScholarPubMed
Cingolani, C.A., Rauscher, R., and Bonhomme, M., 1991, Grupo La Tinta (Precaámbrico y Paleozoico inferior) provinciade Buenos Aires, República Argentina. Nuevos datos geocronoloógicos y micropaleontológicos en las sedimentitas de Villa Cacique, partido de Juarez: Revista Técnica de YPFB, v. 12, p. 177191.Google Scholar
Couëffé, R., and Vecolii, M., 2011, New sedimentological and biostratigraphic data in the Kwahu Group (Meso- to Neo-Proterozoic), southern margin of the Volta Basin, Ghana: stratigraphic constraints and implications on regional lithostratigraphic correlations: Precambrian Research, v. 189, p. 155175.CrossRefGoogle Scholar
Couto, J.G.P., Cordani, U.G., Kawashita, K., Iyer, S.S., and Moraes, N.M.P., 1981, Considerações sobre a idade do Grupo Bambuí com base em análises isotópicas de Sr e Pb: Revista Brasileira de Geociências, v. 11, p. 516.Google Scholar
Croft, W.N., and George, E.A., 1959, Blue-green algae from the Middle Devonian of Rhynie, Aberdeenshire: Bulletin of the British Museum, v. 3, p. 341353.Google Scholar
Damassa, S.P., and Knoll, A.H., 1986, Micropalaeontology of the late Proterozoic Arcoona Quartzite Member of the Tent Hill Formation, Stuart Shelf, South Australia: Alcheringa, v. 10, p. 417430.CrossRefGoogle Scholar
DaSilva, L.G., Pufahl, P.K., James, N.P., Guimaraes, E.M., and Reis, C., 2022, Sequence stratigraphy and paleoenvironmental significance of the Neoproterozoic Bambui Group, Central Brazil: Precambrian Research, v. 379, n. 106710.CrossRefGoogle Scholar
Denezine, M., Adôrno, R.R., Do Carmo, D.A., Guimarães, E.M., Walde, D.H.G., Alvarenga, C.J.S., Germs, G., Antonietto, L.S., Gianfranco, C.G.V.R., and Nunes Junior, O.O., 2022, Methodological development of a combined preparation for micropaleontological and sedimentological studies of samples from the Proterozoic record: Frontiers in Earth Science, v. 10, n. 810406.CrossRefGoogle Scholar
Dong, L., Xiao, S., Shen, B., Zhou, C., Li, G., and Yao, J., 2009, Basal Cambrian microfossils from the Yangtze Gorges area (South China) and the Aksu area (Tarim block, northwestern China): Journal of Paleontology, v. 83, p. 3044.CrossRefGoogle Scholar
Downie, C., and Sarjeant, W.A.S., 1963, On the interpretation and status of some hystrichosphere genera: Palaeontology, v. 6, p. 8396.Google Scholar
Downie, C., Evitt, W.R., and Sarjeant, W.A.S., 1963, Dinoflagellates, hystrichospheres and the classification of the acritarchs: Geological Sciences, v. 7, 16 p.Google Scholar
Eisenack, A., 1958a, Microfossilien aus dem Ordovizium des Baltikums, 1, Markasitschicht, Dictyonema-Scheifer, Glaukonitsand, Glaukonitkalk: Senckenbergian Lethaea, v. 39, p. 389404.Google Scholar
Eisenack, A., 1958b, Tasmanites Newton 1875 und Leiosphaeridia n. g. als Gattungen der Hystrichosphaeridea: Palaeontographica, v. 110, p. 119.Google Scholar
Elenkin, A.A., 1949, (Monographie Algarum Cyanophycearum Aquidulcium et Terrestrium Infinibus URSS Inventarum: Special Part 2): Moscos, Izd. Akademia Nauk SSSR, p. 9851908. [in Russian]Google Scholar
Evitt, W.R., 1963, A discussion and proposals concerning fossil dinoflagellates, hystrichospheres, and acritarchs: Geology, v. 49, p. 158164.Google ScholarPubMed
Fairchild, T.R., Schopf, J.W., Shen-Miller, J., Guimarães, E.M., Edwards, M.D., Lagstein, A., Li, X., Pabst, M., and Melo Filho, L.S., 1996, Recent discoveries of Proterozoic microfossils in south-central Brazil: Precambrian Research, v. 80, p. 125152.CrossRefGoogle Scholar
Fairchild, T.R., Sanchez, E.A.M., Pacheco, M.L.A.F., and Leme, J.M., 2012, Evolution of Precambrian life in the Brazilian geological record: International Journal of Astrobiology, v. 11, p. 309323.CrossRefGoogle Scholar
Fensome, R.A., Williams, G.L., Bars, M.S., Freeman, J.M., and Hill, J.M., 1990, Acritarchs and Fossil Prasinophytes: An Index to Genera, Species and Intraspecific Taxa: American Association of Stratigraphic Palynologist Foundation Contributions, 771 p.Google Scholar
Gaucher, C., 2000, Sedimentology, Palaeontology and Stratigraphy of the Arroyo Del Soldado Group (Vendian to Cambrian, Uruguay): Würzburg, Beringeria, 120 p.Google Scholar
Gaucher, C., and Germs, G.J.B., 2003, Preliminary biostratigraphic correlation of the Arroyo Del Soldado Group (Vendian to Cambrian, Uruguay) with the Cango Caves and Nama Groups (South Africa and Namibia): Revista de La Sociedad Uruguaya de Geología, v. 1, p. 141160.Google Scholar
Gaucher, C., and Germs, G.J.B., 2006, Recent advances in South African Neoproterozoic–early Palaeozoic biostratigraphy: correlation of the Cango Caves and Gamtoos Groups and acritarchs of the Sardinia Bay Formation, Saldania Belt: South African Journal of Geology, v. 109, p. 193214.CrossRefGoogle Scholar
Gaucher, C., and Germs, G.J.B., 2007, First report of organic-walled microfossils from the Otavi and Mulden groups (Neoproterozoic, Namibia): III Symposium on Neoproterozoic–Early Palaeozoic Events in Southwestern Gondwana, p. 1317.Google Scholar
Gaucher, C., and Sprechmann, P., 2009, Neoproterozoic acritarch evolution: Developments in Precambrian Geology, v. 16, p. 319326.CrossRefGoogle Scholar
Gaucher, C., Sprechmann, P., and Schipilov, A., 1996, Upper and middle Proterozoic fossiliferous sedimentary sequences of the Nico Pérez Terrane of Uruguay: lithostratigraphic units, paleontology, depositional environments, and correlations: Neues Jahrbuch für Geologie und Paläontologie, v. 199, p. 339367.CrossRefGoogle Scholar
Gaucher, C., Boggiani, P.C., Sprechmann, P., Sial, A.N., and Fairchild, T.R., 2003, Integrated correlation of the Vendian to Cambrian Arroyo del Soldado and Corumbá Groups (Uruguay and Brazil): palaeogeographic, palaeoclimatic and palaeobiologic implications: Precambrian Research, v. 120, p. 241278.CrossRefGoogle Scholar
Gaucher, C., Chiglino, L., and Peçoits, E., 2004, Southernmost exposures of the Arroyo del Soldado Group (Vendian to Cambrian, Uruguay): palaeogeographic implications for the amalgamation of W-Gondwana: Gondwana Research, v. 7, p. 701714.CrossRefGoogle Scholar
Gaucher, C., Frimmel, H.E., and Germs, G.J.B., 2005a, Organic-walled microfossils and biostratigraphy of the upper Port Nolloth Group (Namibia): implications for latest Neoproterozoic glaciations: Geological Magazine, v. 142, p. 539559.CrossRefGoogle Scholar
Gaucher, C., Poire, D.G., Peral, L.G., and Chiglino, L., 2005b, Litoestratigrafia, Bioestratigrafia y Correlaciones De Las Sucesiones Sedimentarias Del Neoproterozoico-Cambrico Del Craton Del Rio De La Plata (Uruguay y Argentina): Latin American Journal of Sedimentology and Basin Analysis, v. 12, p. 145160.Google Scholar
Gaucher, C., Chiglino, L., Blanco, G., Poiré, D., and Germs, G.J.B., 2008, Acritarchs of Las Ventanas Formation (Ediacaran, Uruguay): implications for the timing of coeval rifting and glacial events in western Gondwana: Gondwana Research, v. 13, p. 488501.CrossRefGoogle Scholar
Geitler, L., 1925, Synoptische Darstellung der Cyanophyceen in morphologischer und systematishcer Hinsicht: Beihefte Zum Botanischen Centralblatt, v. 2, p. 163324.Google Scholar
Germs, G.J.B., 1972, New shelly fossils from Nama Group, South West Africa: American Journal of Science, v. 272, p. 752761.CrossRefGoogle Scholar
Germs, G.J.B., Knoll, A.H., and Vidal, G., 1986, Latest Proterozoic microfossils from the Nama Group, Namibia (South West Africa): Precambrian Research, v. 32, p. 4562.CrossRefGoogle Scholar
Golovenok, V.K., and Belova., M.Y., 1993, The microfossils in the cherts from the Riphean deposits of the Turukhansk Uplift: Stratigraphy and Geological Correlation, v. 1, p. 5161.Google Scholar
Golub, I.N., 1979, A new group of problematic microfossils from Vendian deposits of the Orshan depression (Russian Platform), in Sokolov, B.S., ed., Paleontology of Precambrian and Early Cambrian: Leningrad, Nauka, p. 147155.Google Scholar
Golubkova, E.Y., Zaitseva, T.S., Kuznetsov, A.B., Dovzhikova, E.G., and Maslov, A. V., 2015, Microfossils and Rb–Sr age of glauconite in the key section of the upper Proterozoic of the northeastern part of the Russian plate (Keltmen-1 borehole): Doklady Earth Sciences, v. 462, p. 547551.CrossRefGoogle Scholar
Grazhdankin, D., Nagovitsin, K., Golubkova, E., Karlova, G., Kochnev, B., Rogov, V., and Marusin, V., 2020, Doushantuo–Pertatataka-type acanthomorphs and Ediacaran ecosystem stability: Geology, v. 48, p. 708712.CrossRefGoogle Scholar
Grey, K., 2005, Ediacaran Palynology of Australia: Memoir 31 of the Association of Australasian Palaeontologists, 439 p.Google Scholar
Guacaneme, C., Babinski, M., Bedoya–Rueda, C., Paula–Santos, G. M., Caetano–Filho, S., Kuchenbecker, M., Reis, H.L.S., and Trindade, R.I.F., 2021. Tectonically–induced strontium isotope changes in ancient restricted seas: the case of the Ediacaran–Cambrian Bambuí foreland basin system, east Brazil: Gondwana Research, v. 93, p. 275290, https://doi.org/10.1016/j.gr.2021.02.007.CrossRefGoogle Scholar
Hahn, G., Hahn, R., Leonardos, O.H., Pflug, H.D., and Walde, D.H.G., 1982, Körperlich erhaltene Scyphozoen-Reste aus dem Jungpräkambrium Brasiliens: Geologica et Palaeontologica, v. 16, p. 118.Google Scholar
Han, C.-M., Chen, L., Li, G.J., Pang, K., Wang, W., et al., 2021, First record of organic-walled microfossils from the Tonian Shiwangzhuang Formation of the Tumen Group in western Shandong, North China: Palaeoworld, v. 30, p. 208219.CrossRefGoogle Scholar
Hannah, M.J., Wilson, G.S., and Wrenn, J., 2000, Oligocene and Miocene marine palynomorphs from CRP-2/2A, Victoria Land Basin, Antarctica: Terra Antarctica, v. 7, p. 503511.Google Scholar
Harris, C.R., Millman, K.J., van der Walt, S.J., Gommers, R., Virtanen, P., et al., 2020, Array programming with NumPy: Nature, v. 585, p. 357362.CrossRefGoogle ScholarPubMed
Hassemer, G., Prado, J., and Baldini, R.M., 2020, Diagnoses and descriptions in plant taxonomy: are we making proper use of them?: Taxon, v. 69, p. 14.CrossRefGoogle Scholar
Hermann, T.N., 1974, (Nakhodki massovykh skopleniy trikhomov v rifee), in Timfeev, B.V., ed., Mikrofitofossilii proterozoya i rannego paleozoya SSSR: Leningrad, Nauka, p. 610. [in Russian]Google Scholar
Hofmann, H.J., 1976, Precambrian microflora, Belcher Islands, Canada: significance and systematics: Journal of Paleontology, v. 50, p. 10401073.Google Scholar
Hofmann, H.J., and Jackson, G.D., 1991, Shelf-facies microfossils from the Uluksan Group (Proterozoic Bylot Supergroup), Baffin Island, Canada: Journal of Paleontology, v. 65, p. 361382.CrossRefGoogle Scholar
Hofmann, H.J., and Jackson, G.D., 1994, Shale-Facies Microfossils from the Proterozoic Bylot Supergroup, Baffin Island, Canada: The Paleontological Society Memoir 37, 39 p.CrossRefGoogle Scholar
Horodyski, R.J., 1980, Middle Proterozoic shale-facies microbiota from the lower Belt Supergroup, Little Belt Mountains, Montana: Journal of Paleontology, v. 54, p. 649663.Google Scholar
Hua, H., Chen, Z., Yuan, X., Zhang, L., and Xiao, S., 2005, Skeletogenesis and asexual reproduction in the earliest biomineralizing animal Cloudina: Geology, v. 33, p. 277280.CrossRefGoogle Scholar
Huntley, J.W., Xiao, S., and Kowalewski, M., 2006, 1.3 billion years of acritarch history: an empirical morphospace approach: Precambrian Research, v. 144, p. 5268.CrossRefGoogle Scholar
Jankauskas, T. V., Mikhailova, N.S., and German, T.N., 1989, Mikrofossilii Dokembriya SSSR [Precambrian Microfossils of the USSR]: Leningrad, Nauka, 191 p. [in Russian]Google Scholar
Javaux, E.J., 2007, The early eukaryotic fossil record: Advances in Experimental Medicine and Biology, v. 607, p. 119.CrossRefGoogle ScholarPubMed
Javaux, E.J., and Knoll, A.H., 2017, Micropaleontology of the lower Mesoproterozoic Roper Group, Australia, and implications for early eukaryotic evolution: Journal of Paleontology, v. 91, p. 199229.CrossRefGoogle Scholar
Javaux, E.J., and Lepot, K., 2018, The Paleoproterozoic fossil record: implications for the evolution of the biosphere during Earth's middle-age: Earth-Science Reviews, v. 176, p. 6886.CrossRefGoogle Scholar
Javaux, E.J., and Marshal, C.P., 2006, A new approach in deciphering early protist paleobiology and evolution: combined microscopy and microchemistry of single Proterozoic acritarchs: Review of Palaeobotany and Palynology, v. 139, p. 115.CrossRefGoogle Scholar
Javaux, E.J., Knoll, A.H., and Walter, M.R., 2004, TEM evidence for eukaryotic diversity in mid-Proterozoic oceans: Geobiology, v. 2, p. 121132.CrossRefGoogle Scholar
Kirchner, O., 1900, Shizophyceae, in Engler, A., and Prantl, K., eds., Die Natürlichen Pflanzenfamilien: Leipzig, I Teil, Abteilung Ia, p. 115121.Google Scholar
Knoll, A.H., 1982, Microfossils from the late Precambrian Draken Conglomerate, Ny Friesland, Svalbard: Journal of Paleontology, v. 56, p. 131162.Google Scholar
Knoll, A.H., 1994, Proterozoic and early Cambrian protists: evidence for accelerating evolutionary tempo: Proceedings of the National Academy of Sciences of the United States of America, v. 91, p. 67436750.CrossRefGoogle ScholarPubMed
Knoll, A.H., and Golubic, S., 1979, Anatomy and taphonomy of a Precambrian algal stromatolite: Precambrian Research, v. 10, p. 115151.CrossRefGoogle Scholar
Knoll, A.H., and Walter, M.R., 1992, Latest Proterozoic stratigraphy and Earth history: Nature, v. 356, p. 673678.CrossRefGoogle ScholarPubMed
Knoll, A.H., Swett, K., and Mark, J., 1991, Paleobiology of a Neoproterozoic tidal flat/lagoonal complex: the Draken Conglomerate Formation, Spitsbergen: Journal of Paleontology, v. 65, p. 531570.CrossRefGoogle ScholarPubMed
Knoll, A.H., Wörndle, S., and Kah, L.C., 2013, Covariance of microfossil assemblages and microbialite textures across an upper Mesoproterozoic carbonate platform: Palaios, v. 28, p. 453470.CrossRefGoogle Scholar
Knoll, A.H., Germs, G.J.B., Tankard, A., and Welsink, H., 2020, Tonian microfossils from subsurface shales in Botswana: Precambrian Research, v. 345, n. 105779.CrossRefGoogle Scholar
Kouketsu, Y., Mizukami, T., Mori, H., Endo, S., Aoya, M., Hara, H., Nakamura, D., and Wallis, S., 2014, A new approach to develop the Raman carbonaceous material geothermometer for low-grade metamorphism using peak width: Island Arc, v. 23, p. 3350.CrossRefGoogle Scholar
Kumar, A., and Venkatachala, B.S., 1998, Proterozoic chert microbiota from the Riasi Inlier of the Vaishnodevi Limestone in the Himalayan Foot-hills, Jammu, India: Indian Journal of Petroleum Geology, v. 7, p. 5170.Google Scholar
Kumar, S., and Pandey, S.K., 2008, Discovery of organic-walled microbiota from the black-bedded chert, Balwan Limestone, the Bhander Group, Lakheri area, Rajasthan: Current Science, v. 94, p. 797800.Google Scholar
Kumar, S., and Srivastava, P., 1995, Microfossils from the Kheinjua Formation, Mesoproterozoic Semri Group, Newari area, central India: Precambrian Research, v. 74, p. 91117.CrossRefGoogle Scholar
Lamb, D.M., Awramik, S.M., Chapman, D.J., and Zhu, S., 2009, Evidence for eukaryotic diversification in the ~1800 million-year-old Changzhougou Formation, North China: Precambrian Research, v. 173, p. 93104.CrossRefGoogle Scholar
Lei, Y., Shen, J., Algeo, T.J., Servais, T., Feng, Q., and Yu, J., 2019, Phytoplankton (acritarch) community changes during the Permian–Triassic transition in South China: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 519, p. 8494.CrossRefGoogle Scholar
Le Rosq, C.R., 2021, rampy. https://github.com/charlesll/rampy (accessed Jul 2021).Google Scholar
Liais, E., 1872. Climat, géologie, faune et géographie botanique du Brésil, VIII. Paris, Georges Chamerot.CrossRefGoogle Scholar
Li, G., Pang, K., Chen, L., Zhou, G., Han, C., Yang, L., Wang, W., Yang, F., and Yin, L., 2019, Organic-walled microfossils from the Tonian Tongjiazhuang Formation of the Tumen Group in western Shandong, North China Craton and their biostratigraphic significance: Gondwana Research, v. 76, p. 260289.CrossRefGoogle Scholar
Liu, P., and Moczydłowska, M., 2019, Ediacaran microfossils from the Doushantuo Formation chert nodules in the Yangtze Gorges area, South China, and new biozones: Fossils and Strata, v. 65, 172 p.CrossRefGoogle Scholar
Liu, P., Xiao, S., Yin, C., Chen, S., Zhou, C., and Li, M., 2014, Ediacaran acanthomorphic acritarchs and other microfossils from chert nodules of the Upper Doushantuo Formation in the Yangtze Gorges area, South China: Paleontological Society Memoir 72, 139 p.CrossRefGoogle Scholar
Loron, C., and Moczydłowska, M., 2017, Tonian (Neoproterozoic) eukaryotic and prokaryotic organic-walled microfossils from the upper Visingsö Group, Sweden: Palynology, v. 42, p. 220254.CrossRefGoogle Scholar
Loron, C.C., Rainbird, R.H., Turner, E.C., Greenman, J.W., and Javaux, E.J., 2019, Organic-walled microfossils from the late Mesoproterozoic to early Neoproterozoic lower Shaler Supergroup (Arctic Canada): diversity and biostratigraphic significance: Precambrian Research, v. 321, p. 349374.CrossRefGoogle Scholar
Loron, C.C., Halverson, G.P., Rainbird, R.H., Skulski, T., Turner, E.C., and Javaux, E.J., 2021, Shale-hosted biota from the Dismal Lakes Group in Arctic Canada supports an early Mesoproterozoic diversification of eukaryotes: Journal of Paleontology, v. 95, p. 11131137.CrossRefGoogle Scholar
Lottaroli, F., Craig, J., and Thusu, B., 2014, Neoproterozoic–early Cambrian (Infracambrian) hydrocarbon prospectivity of North Africa: a synthesis: Geological Society, London, Special Publications, v. 326, p. 137156.CrossRefGoogle Scholar
Maithy, P.K., 1975, Micro-organisms from the Bushimay System (late Precambrian) of Kanshi, Zaire: Palaeoobotanist, v. 22, p. 33149.Google Scholar
Maithy, P.K., and Shukla, M., 1977, Microbiota from the Suket Shales, Vindhyan (late Precambrian), Madhya Pradesh: Palaeobotanist, v. 23, p. 176188.Google Scholar
Mandal, J., Maithy, P.K., Barman, G., and Verma, K.K., 1984, Microbiota from the Kushalgarh Formation, Delhi, India: Palaeobotanist, v. 32, p. 119.Google Scholar
Marchese, H.G., 1974, Estromatolitos “Gymnosolenidos” en el lado oriental de Minas Gerais, Brasil: Revista Brasileira de Geociências, v. 4, p. 172190.CrossRefGoogle Scholar
Marshall, C.P., Javaux, E.J., Knoll, A.H., and Walter, M.R., 2005, Combined micro-Fourier transform infrared (FTIR) spectroscopy and micro-Raman spectroscopy of Proterozoic acritarchs: a new approach to Palaeobiology: Precambrian Research, v. 138, p. 208224.CrossRefGoogle Scholar
Mazoni, A.F., 2021, Raman Online. https://github.com/alyssonmazoni/raman_online (accessed Jul 2021).Google Scholar
Mendelson, C.V., and Schopf, J.W., 1982, Proterozoic microfossils from the Sukhaya Tunguska, Shorikha, and Yudoma formations of the Siberian platform, USSR: Journal of the Palaeontological Society of India, v. 56, p. 4283.Google Scholar
Merdith, A.S., Williams, S.E., Collins, A.S., Tetley, M.G., Mulder, J.A., et al., 2021, Extending full-plate tectonic models into deep time: linking the Neoproterozoic and the Phanerozoic: Earth-Science Reviews, v. 214, p. 244.CrossRefGoogle Scholar
Miao, L., Moczydłowska, M., Zhu, S., and Zhu, M., 2019, New record of organic-walled, morphologically distinct microfossils from the late Paleoproterozoic Changcheng Group in the Yanshan Range, North China: Precambrian Research, v. 321, p. 172198.CrossRefGoogle Scholar
Miao, L., Moczydłowska, M., and Zhu, M., 2021, A diverse organic-walled microfossil assemblage from the Mesoproterozoic Xiamaling Formation, North China: Precambrian Research, v. 360, n. 106235.CrossRefGoogle Scholar
Mikhailova, N.S., 1986, Novye nakhodki mikrofitofossilij iz otlozhenij verkhnego rifeya Krasnoyarskogo kraya in Sokolov, B.S., ed., Aktual ’nye Voprosy Sovremennoj: Kyiv, Naukova Dumka, p. 3137.Google Scholar
Moczydłowska, M., 1991, Acritarch Biostratigraphy of the Lower Cambrian and the Precambrian–Cambrian Boundary in Southeastern Poland: Oslo, Universitetsforlaget, 127 p.CrossRefGoogle Scholar
Moczydłowska, M., 2008a, New records of late Ediacaran microbiota from Poland: Precambrian Research, v. 167, p. 7192.CrossRefGoogle Scholar
Moczydłowska, M., 2008b, The Ediacaran microbiota and the survival of Snowball Earth conditions: Precambrian Research, v. 167, p. 7192.CrossRefGoogle Scholar
Moczydłowska, M., 2016, Algal affinities of Ediacaran and Cambrian organic-walled microfossils with internal reproductive bodies: Tanarium and other morphotypes: Palynology, v. 40, p. 83121.CrossRefGoogle Scholar
Moczydlowska, M., Landing, E., Zang, W., and Palacios, T., 2011, Proterozoic phytoplankton and timing of chlorophyte algae origins: Palaeontology, v. 54, p. 721733.CrossRefGoogle Scholar
Morais, L., Fairchild, T.R., Freitas, B.T., Rudnitzki, I.D., Silva, E.P., Lahr, D., Moreira, A.C., Abrahão Filho, E.A., Leme, J.M., and Trindade, R.I.F., 2021, Doushantuo-Pertatataka-like acritarchs from the late Ediacaran Bocaina Formation (Corumbá Group, Brazil): Frontiers in Earth Science, v. 9, n. 787011, https://doi.org/10.3389/feart.2021.787011.CrossRefGoogle Scholar
Moreira, D.S., Uhlein, A., Dussin, I.A., Uhlein, G.J., and Misuzaki, A.M.P., 2020, A Cambrian age for the upper Bambuí Group, Brazil, supported by the first U–Pb dating of volcaniclastic bed: Journal of South American Earth Sciences, v. 99, n. 102503.CrossRefGoogle Scholar
Nagovitsin, K.E., and Kochnev, B.B., 2015, Microfossils and biofacies of the Vendian fossil biota in the southern Siberian Platform: Russian Geology and Geophysics, v. 56, p. 584593.CrossRefGoogle Scholar
Naumova, S.N., 1949, Spores of the lower Cambrian: Izvestiya Akademii Nauk SSSR, v. 4, p. 4956.Google Scholar
Nemerov, V.K., Stanevich, A.M., Razvozzhaeva, E.A., Budyak, A.E., and Kornilova, T.A., 2010, Biogenic sedimentation factors of mineralization in the Neoproterozoic strata of the Baikal–Patom region: Russian Geology and Geophysics, v. 51, p. 572586.CrossRefGoogle Scholar
Nyberg, A.V., and Schopf, J.W, 1984, Microfossils in stromatolitic cherts from the upper Proterozoic Min'yar Formation, southern Ural Mountains: Journal of Paleontology, v. 58, p. 738772.Google ScholarPubMed
Oehler, J.H., 1977, Microflora of the H.Y.C. Pyritic Shale Member of the Barney Creek Formation (McArthur Group), middle Proterozoic of Northern Australia: Alcheringa, v. 1, p. 315349.CrossRefGoogle Scholar
Okubo, J., Warren, L. V., Luvizotto, G.L., Varejão, F.G., Quaglio, F., Uhlein, G.J., and Assine, M.L., 2020, Evidences of seismic events during the sedimentation of Sete Lagoas Formation (Bambuí Group—Ediacaran, Brazil): Journal of South American Earth Sciences, v. 98, n. 102461.CrossRefGoogle Scholar
Oliveira, A.I., and Leonardos, O.H., 1943, Geologia Do Brasil: Rio de Janeiro, Serviço de Informação Agrícola, 813 p.Google Scholar
Ouyang, Q., Guan, C., Zhou, C., and Xiao, S., 2017, Acanthomorphic acritarchs of the Doushantuo Formation from an upper slope section in northwestern Hunan Province, South China, with implications for early–middle Ediacaran biostratigraphy: Precambrian Research, v. 298, p. 512529.CrossRefGoogle Scholar
Ouyang, Q., Zhou, C., Xiao, S., Guan, C., Chen, Z., Yuan, X., and Sun, Y., 2021, Distribution of Ediacaran acanthomorphic acritarchs in the lower Doushantuo Formation of the Yangtze Gorges area, South China: evolutionary and stratigraphic implications: Precambrian Research, v. 353, n. 106005.CrossRefGoogle Scholar
Ouyang, S., Yin, L., and Zaiping, L., 1974, Sinian and Cambrian spores and acritarchs, in Press, S., ed., Handbook of the Stratigraphy and Paleontology of South Western China: Beijing, Science Publishing House, p. 7280, 114–123.Google Scholar
Palacios, T., 1989, Microfósiles de Pared Orgánica Del Proterozoico Superior (Región Central de La Península Ibérica): Memorias del Museo Paleontológico de la Universidad de Zaragoza, 91 p.Google Scholar
Pandey, S.K., and Kumar, S., 2013, Organic walled microbiota from the silicified algal clasts, Bhander limestone, Satna area, Madhya Pradesh: Journal of the Geological Society of India, v. 82, p. 499508.CrossRefGoogle Scholar
Pang, K., Tang, Q., Wu, C., Li, G., Chen, L., Wan, B., Yuan, X., Bodnar, R.J., and Xiao, S., 2020, Raman spectroscopy and structural heterogeneity of carbonaceous material in Proterozoic organic-walled microfossils in the North China Craton: Precambrian Research, v. 346, n. 105818.CrossRefGoogle Scholar
Paula-Santos, G.M., Babinski, M., Kuchenbecker, M., Caetano-Filho, S., Trindade, R.I.F., and Pedrosa-Soares, A.C., 2015, New evidence of an Ediacaran age for the Bambuí Group in southern São Francisco craton (eastern Brazil) from zircon U–Pb data and isotope chemostratigraphy: Gondwana Research, v. 28, p. 702720.CrossRefGoogle Scholar
Perrella Júnior, P., Uhlein, A., Uhlein, G.J., Sial, A.N., Pedrosa-Soares, A.C., and Lima, O.N.B., 2017, Facies analysis, sequence stratigraphy and chemostratigraphy of the Sete Lagoas Formation (Bambuí Group), northern Minas Gerais State, Brazil: evidence of a cap carbonate deposited on the Januária basement high: Brazilian Journal of Geology, v. 47, p. 5977.CrossRefGoogle Scholar
Peterson, K.J., and Butterfield, N.J., 2005, Origin of the Eumetazoa: testing ecological predictions of molecular clocks against the Proterozoic fossil record: Proceedings of the National Academy of Sciences of the United States of America, v. 102, p. 95479552.CrossRefGoogle ScholarPubMed
Pimentel, M.M., and Fuck, R.A., 1992, Neoproterozoic crustal accretion in central: Geology, v. 20, p. 375379.2.3.CO;2>CrossRefGoogle Scholar
Pimentel, M.M., Rodrigues, J.B., DellaGiustina, M.E.S., Junges, S., Matteini, M., and Armstrong, R., 2011, The tectonic evolution of the Neoproterozoic Brasília Belt, central Brazil, based on SHRIMP and LA-ICPMS U-Pb sedimentary provenance data: a review: Journal of South American Earth Sciences, v. 31, p. 345357.CrossRefGoogle Scholar
Porter, S.M., and Riedman, L.A., 2016, Systematics of organic-walled microfossils from the ca. 780–740 Ma Chuar Group, Grand Canyon, Arizona: Journal of Paleontology, v. 90, p. 815853.CrossRefGoogle Scholar
Prasad, B., Uniyal, S.N., and Asher, R., 2005, Organic-walled microfossils from the Proterozoic Vindhyan Supergroup of Son Valley, Madhya Pradesh, India: Palaeobotanist, v. 54, p. 1360.Google Scholar
Prasad, B., Asher, R., and Borgohai, B., 2010, Late Neoproterozoic (Ediacaran)–early Paleozoic (Cambrian) Acritarchs from the Marwar Supergroup, Bikaner-Nagaur Basin, Rajasthan: Journal of the Geological Society of India, v. 75, p. 415431.CrossRefGoogle Scholar
Pykhova, N.G., 1973, Dokembriskie akritarhi Moskovskogo graben I Yuzhnogo: Obshchestva Ispitateli Prirody Otdel Geologicheskii Novaya, v. 48, p. 91107.Google Scholar
Qu, Y., Engdahl, A., Zhu, S., Vajda, V., and McLoughlin, N., 2015, Ultrastructural heterogeneity of carbonaceous material in ancient cherts: investigating biosignature origin and preservation: Astrobiology, v. 15, p. 825842.CrossRefGoogle ScholarPubMed
Reis, H.L.S., and Alkmim, F.F., 2015, Anatomy of a basin-controlled foreland fold-thrust belt curve: the Três Marias salient, São Francisco basin, Brazil: Marine and Petroleum Geology, v. 66, p. 711731.CrossRefGoogle Scholar
Reis, H.L.S., and Suss, J.F., 2016, Mixed carbonate–siliciclastic sedimentation in forebulge grabens: an example from the Ediacaran Bambuí Group, São Francisco Basin, Brazil: Sedimentary Geology, v. 339, p. 83103.CrossRefGoogle Scholar
Riedman, L.A., Porter, S.M., and Calver, C.R., 2018, Vase-shaped microfossil biostratigraphy with new data from Tasmania, Svalbard, Greenland, Sweden and the Yukon: Precambrian Research, v. 319, p. 1936.CrossRefGoogle Scholar
Rimann, E., 1917, A Kimberlita no Brazil: Anais Da Escola de Minas Gerais, v. 15, p. 2732.Google Scholar
Rodrigues, J.B., 2008, Proveniência de sedimentos dos grupos Canastra, Ibiá, Vazante e Bambuí - Um estudo de zircões detríticos e idades modelos Sm-Nd [Ph.D. thesis]: Brasília, Brazil, Universidade de Brasília, 128 p.Google Scholar
Samuelsson, J., and Butterfield, N.J., 2001, Neoproterozoic fossils from the Franklin Mountains, northwestern Canada: stratigraphic and palaeobiological implications: Precambrian Research, v. 107, p. 235251.CrossRefGoogle Scholar
Sanchez, E.A.M., and Fairchild, T.R., 2018, Reavaliação De Fósseis Do Grupo Bambuí: Implicações Paleobiológicas Para O Neoproterozoico Tardio Do Brasil: Geonomos, v. 25, https://doi.org/10.18285/geonomos.v25i2.1076.Google Scholar
Sanchez, E.A.M., Uhlein, A., and Fairchild, T.R., 2021, Treptichnus pedum in the Três Marias Formation, south-central Brazil, and its implications for the Ediacaran–Cambrian transition in South America: Journal of South American Earth Sciences, v. 105, n. 102983.CrossRefGoogle Scholar
Schopf, J.W., 1968, Microflora of the Bitter Springs Formation, late Precambrian, central Australia: Journal of Paleontology, v. 42, p. 651688.Google Scholar
Schopf, J.W., Sergeev, V.N., and Kudryavtsev, A.B., 2015, A new approach to ancient microorganisms: taxonomy, paleoecology, and biostratigraphy of the lower Cambrian Berkuta and Chulaktau microbiotas of South Kazakhstan: Journal of Paleontology, v. 89, p. 695729.CrossRefGoogle Scholar
Seilacher, A., 1955, Spuren und fazies im Unterkambrium, in Schindewolf, O.H., and Seilacher, A., eds., Beitrage zur Kenntnis des Kambriums in der Salt Range (Pakistan), Abhandlungen 10, Mainz: Akademie der Wissenschaften und der Literatur zu Mainz, Mathematisch Naturwissenschaftliche Klasse, p. 373399.Google Scholar
Sergeev, V.N., 1984, Microfossils in the silicified columnar stromatolites from the upper Riphean deposits of the Turukhansk Uplift: Doklady AN SSSR, v. 278, p. 436440.Google Scholar
Sergeev, V.N., 1992, Silicified Microfossils from the Precambrian and Cambrian Deposits of the Southern Ural Mountains and Middle Asia: Moscow, Nauka, 134 p.Google Scholar
Sergeev, V.N., 1994, Microfossils in cherts from the middle Riphean (Mesoproterozoic) Avzyan Formation, southern Ural Mountains, Russian Federation: Precambrian Research, v. 65, p. 231254.CrossRefGoogle ScholarPubMed
Sergeev, V.N., 2001, Paleobiology of the Neoproterozoic (upper Riphean) Shorikha and Burovaya Silicified Microbiotas, Turukhansk Uplift, Siberia: Journal of Paleontology, v. 75, p. 427448.2.0.CO;2>CrossRefGoogle Scholar
Sergeev, V.N., 2006, Precambrian Microfossils in Cherts: Their Paleobiology, Classification, and Biostratigraphic Usefulness: Moscow, Geos, 280 p.Google Scholar
Sergeev, V.N., and Lee, S.-J., 2001, Microfossils from cherts of the middle Riphean Svetlyi Formation, the Uchur–Maya Region of Siberia and their stratigraphic significance: Stratigraphy and Geological Correlation, v. 9, p. 110.Google Scholar
Sergeev, V.N., and Lee, S.-J., 2004, New data on silicified microfossils from the Satka Formation of the lower Riphean Stratotype, the Urals: Stratigraphy and Geological Correlation, v. 12, p. 121.Google Scholar
Sergeev, V.N., and Schopf, J.W., 2010, Taxonomy, paleoecology and biostratigraphy of the late Neoproterozoic Chichkan Microbiota of South Kazakhstan: the marine biosphere on the eve of metazoan radiation: Journal of Paleontology, v. 84, p. 363401.CrossRefGoogle Scholar
Sergeev, V.N., and Seong-Joo, L., 2006, Real eukaryotes and precipitates first found in the middle Riphean stratotype, Southern Urals: Stratigraphy and Geological Correlation, v. 14, p. 118.CrossRefGoogle Scholar
Sergeev, V.N., Knoll, A.H., Kolosova, S.P., and Kolosov., P.N., 1994, Microfossils in cherts from the Mesoproterozoic Debengda Formation, Olenek Uplift, Northeastern Siberia: Stratigraphy and Geological Correlation, v. 2, p. 2338.Google ScholarPubMed
Sergeev, V.N., Knoll, A.H., and Petrov, P.Y., 1997, Paleobiology of the Mesoproterozoic–Neoproterozoic transition: the Sukhaya Tunguska Formation, Turukhansk Uplift, Siberia: Precambrian Research, v. 85, p. 201239.CrossRefGoogle ScholarPubMed
Sergeev, V.N., Sharma, M., and Shukla, Y., 2008, Mesoproterozoic silicified microbiotas of Russia and India—characteristics and contrasts: Palaeobotanist, v. 57, p. 323358.Google Scholar
Sergeev, V.N., Sharma, M., and Shukla, Y., 2012, Proterozoic fossil cyanobacteria: Paleobotanist, v. 61, p. 189358.Google Scholar
Sergeev, V.N., Knoll, A.H., Vorob'eva, N.G., and Sergeeva, N.D., 2016, Microfossils from the lower Mesoproterozoic Kaltasy Formation, East European Platform: Precambrian Research, v. 278, p. 87107.CrossRefGoogle Scholar
Sergeev, V.N., Vorob'eva, N.G., and Petrov, P.Y., 2017a, The biostratigraphic conundrum of Siberia: do true Tonian–Cryogenian microfossils occur in Mesoproterozoic rocks?: Precambrian Research, v. 299, p. 282302.CrossRefGoogle Scholar
Sergeev, V.N., Vorob'eva, N.G., Petrov, P.Y., and Semikhatov, M.A., 2017b, Taxonomic composition and biostratigraphic value of the early Riphean organic-walled microfossil association from the Ust’-Il'ya Formation of the Anabar Uplift, Northern Siberia: Stratigraphy and Geological Correlation, v. 25, p. 241255.CrossRefGoogle Scholar
Shang, X., Liu, P., and Moczydłowska, M., 2019, Acritarchs from the Doushantuo Formation at Liujing section in Songlin area of Guizhou Province, South China: implications for early–middle Ediacaran biostratigraphy: Precambrian Research, v. 334, n. 105453.CrossRefGoogle Scholar
Sharma, M., and Sergeev, V.N., 2004, Genesis of carbonate precipitate patterns and associated microfossils in Mesoproterozoic formations of India and Russia—a comparative study: Precambrian Research, v. 134, p. 317347.CrossRefGoogle Scholar
Shepeleva, E.D., 1962, Rastitel'nye? ostatki neizvestnoj sistematicheskoj prinadlezhnosti iz otlozhenij bavlinskoj serii Volgo-Ural'skoj provintsii: Doklady Akademii Nauk SSSR, v. 142, p. 456457.Google Scholar
Shi, M., Feng, Q., Khan, M.Z., and Zhu, S., 2017a, An eukaryote-bearing microbiota from the early Mesoproterozoic Gaoyuzhuang Formation, Tianjin, China and its significance: Precambrian Research, v. 303, p. 709726.CrossRefGoogle Scholar
Shi, M., Feng, Q., Khan, M.Z., Awramik, S.M., and Zhu, S., 2017b, Silicified microbiota from the Paleoproterozoic Dahongyu Formation, Tianjin, China: Journal of Paleontology, v. 91, p. 369392.CrossRefGoogle Scholar
Shukla, Y., Sharma, M., and Sergeev, V.N., 2020, Organic walled microfossils from the Neoproterozoic Owk Shale, Kurnool Group, South India: Palaeoworld, v. 29, p. 490511.CrossRefGoogle Scholar
Simonetti, C., and Fairchild, T.R., 1989, Paleobiologia de uma nova microflórula silicificada do Grupo Bambuí (Proterozoico Superior), da região de Unaí, MG: Boletim IG, v. 7, https://doi.org/10.11606/issn.2317-8078.v0i7p01-25.Google Scholar
Simonetti, C., and Fairchild, T.R., 2000, Proterozoic microfossils from subsurface siliciclastic rocks of the São Francisco Craton, south-central Brazil: Precambrian Research, v. 103, p. 129.CrossRefGoogle Scholar
Singh, V.K., and Sharma, M., 2016, Mesoproterozoic organic-walled microfossils from the Chaporadih Formation, Chandarpur Group, Chhattisgarh Supergroup, Odisha, India: Journal of the Palaeontological Society of India, v. 61, p. 7584.Google Scholar
Sommer, F.W., 1971, Microfósseis do Calcário Bambuí, de Pedro Leopoldo, Estado de Minas Gerais: Anais Da Academia Brasileira de Ciências, v. 61, p. 7584.Google Scholar
Stanevich, A.M., Maksimova, E.N., Kornilova, T.A., Gladkochub, D.P., Mazukabzov, A.M., and Donskaya, T. V., 2009, Microfossils from the Arymas and Debengda formations, the Riphean of the Olenek Uplift: age and presumable nature: Stratigraphy and Geological Correlation, v. 17, p. 2035.CrossRefGoogle Scholar
Stanier, R.Y., Sistrom, W.R., Hansen, T.A., Whitton, B.A., Castenholz, R.W., et al., 1978, Proposal to place nomenclature of the Cyanobacteria (blue-green algae) under the rules of the International Code of Nomenclature of bacteria: International Journal of Systematic Bacteriology, v. 28, p. 335336.Google Scholar
Strother, P.K., Battison, L., Brasier, M.D., and Wellman, C.H., 2011, Earth's earliest non-marine eukaryotes: Nature, v. 473, p. 505509.CrossRefGoogle ScholarPubMed
Suslova, E.A., Parfenova, T.M., Saraev, S. V., and Nagovitsyn, K.E., 2017, Organic geochemistry of rocks of the Mesoproterozoic Malgin Formation and their depositional environments (southeastern Siberian Platform): Russian Geology and Geophysics, v. 58, p. 516528.CrossRefGoogle Scholar
Tang, Q., Pang, K., Xiao, S., Yuan, X., Ou, Z., and Wan, B., 2013, Organic-walled microfossils from the early Neoproterozoic Liulaobei Formation in the Huainan region of North China and their biostratigraphic significance: Precambrian Research, v. 236, p. 157181.CrossRefGoogle Scholar
Tang, Q., Pang, K., Yuan, X., Wan, B., and Xiao, S., 2015, Organic-walled microfossils from the Tonian Gouhou Formation, Huaibei region, North China Craton, and their biostratigraphic implications: Precambrian Research, v. 266, p. 296318.CrossRefGoogle Scholar
Tang, Q., Hughes, N.C., McKenzie, N.R., Myrow, P.M., and Xiao, S., 2017, Late Mesoproterozoic–early Neoproterozoic organic-walled microfossils from the Madhubani Group of the Ganga Valley, northern India: Palaeontology, v. 60, p. 869891.CrossRefGoogle Scholar
Tang, Q., Hu, J., Xie, G., Yuan, X., Wan, B., et al., 2019, A problematic animal fossil from the early Cambrian Hetang Formation, South China—a reply: Journal of Paleontology, v. 93, p. 12791282.CrossRefGoogle Scholar
Tang, Q., Pang, K., Li, G., Chen, L., Yuan, X., Sharma, M., and Xiao, S., 2021, The Proterozoic macrofossil Tawuia as a coenocytic eukaryote and a possible macroalga: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 576, n. 110485.CrossRefGoogle Scholar
Thuret, G., 1875, Essai de classification des nostocines: Annales Des Sciences Naturelles; Botanique, v. 6, p. 372382.Google Scholar
Timofeev, B. V., 1966, Micropaleophytological Investigation into Ancient Formations: Moscow and Leningrad, Nauka, 238 p.Google Scholar
Timofeev, B.V., Hermann, T.N., and Mikhailova, N.S., 1976, Microphytofossils of the Precambrian, Cambrian and Ordovician: Leningrad, Nauka, 106 p.Google Scholar
Tiwari, M., and Pant, C.C., 2004, Organic-walled microfossils from the Neoproterozoic black phosphatic stringers in the Gangolihat Dolomite, Lesser Himalaya, India: Current Science, v. 87, p. 17331738.Google Scholar
Tiwari, M., and Pant, I., 2009, Microfossils from the Neoproterozoic Gangolihat Formation, Kumaun Lesser Himalaya: their stratigraphic and evolutionary significance: Journal of Asian Earth Sciences, v. 35, p. 137149.CrossRefGoogle Scholar
Tobias, T.C., 2014, Micropaleontologia da Formação Tamengo, Eco Parque Cacimba da Saúde, Ediacarano, Grupo Corumbá, Estado de Mato Grosso do Sul, Brasil [M.Sc. thesis]: Brasília, Brazil, University of Brasília, 78 p.Google Scholar
Turland, N.J., Wiersema, J.H., Barrie, F.R., Greuter, W., Hawksworth, D.L., et al., 2018, International Code of Nomenclature for Algae, Fungi, and Plants (Shenzhen Code) Adopted by the Nineteenth International Botanical Congress Shenzhen, China, July 2017. Regnum Vegetabile 159: Glashütten, Koeltz Botanical Books.Google Scholar
Turnau, E., and Racki, G., 1999, Givetian palynostratigraphy and palynofacies: new data from the Bodzentyn Syncline (Holy Cross Mountains, central Poland): Review of Paleobotany and Palynology, v. 106, p. 237271.CrossRefGoogle Scholar
Uhlein, G.J., Uhlein, A., Pereira, E., Caxito, F.A., Okubo, J., Warren, L.V., and Sial, A.N., 2019, Ediacaran paleoenvironmental changes recorded in the mixed carbonate–siliciclastic Bambuí Basin, Brazil: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 517, p. 3951.CrossRefGoogle Scholar
Vidal, G., and Moczydłowska-Vidal, M., 1997, Biodiversity, speciation, and extinction trends of Proterozoic and Cambrian phytoplankton: Paleobiology, v. 23, p. 230246.CrossRefGoogle Scholar
Virtanen, P., Gommers, R., Oliphant, T.E., Haberland, M., Reddy, T., et al., 2020, SciPy 1.0: fundamental algorithms for scientific computing in Python: Nature Methods, v. 17, p. 261272.CrossRefGoogle ScholarPubMed
Vorob'eva, N.G., and Petrov, P.Y., 2014, The genus Vendomyces Burzin and facies—ecological specificity of the Staraya Rechka Microbiota of the late Vendian of the Anabar Uplift of Siberia and its stratigraphic analogues: Paleontological Journal, v. 48, p. 655666.CrossRefGoogle Scholar
Vorob'eva, N.G., Sergeev, V.N., and Knoll, A.H., 2009, Neoproterozoic microfossils from the northeastern margin of the East European Platform: Journal of Paleontology, v. 83, p. 161196.CrossRefGoogle Scholar
Vorob'eva, N.G., Sergeev, V.N., and Petrov, P.Y., 2015, Kotuikan Formation assemblage: a diverse organic-walled microbiota in the Mesoproterozoic Anabar succession, northern Siberia: Precambrian Research, v. 256, p. 201222.CrossRefGoogle Scholar
Walcott, C.D., 1899, Pre-Cambrian fossiliferous formations: Geological Society of America, v. 10, p. 199244.CrossRefGoogle Scholar
Walde, D.H.G., Do Carmo, D.A., Guimarães, E.M., Vieira, L.C., Erdtmann, B., Sanchez, E.A.M., Adôrno, R.R., and Tobias, T.C., 2015, New aspects of Neoproterozoic–Cambrian transition in the Corumbá region (state of Mato Grosso do Sul, Brazil): Annales de Paléontologie, v. 101, p. 213224.CrossRefGoogle Scholar
Wan, B., Tang, Q., Pang, K., Wang, X., Bao, Z., Meng, F., Zhou, C., Yuan, X., Hua, H., and Xiao, S., 2019, Repositioning the Great Unconformity at the southeastern margin of the North China Craton: Precambrian Research, v. 324, p. 117.CrossRefGoogle Scholar
Warren, L. V., Quaglio, F., Riccomini, C., Simões, M.G., Poiré, D.G., Strikis, N.M., Anelli, L.E., and Strikis, P.C., 2014, The puzzle assembled: Ediacaran guide fossil Cloudina reveals an old proto-Gondwana seaway: Geology, v. 42, p. 391394.CrossRefGoogle Scholar
Woese, C., and Fox, G., 1977, No phylogenetic structure of the prokaryotic domain: Proceedings, National Academy of Sciences of the United States of America, v. 74, p. 50885090.CrossRefGoogle ScholarPubMed
Xiao, S., and Narbonne, G.M., 2020, The Ediacaran Period, in Gradstein, F.M., Ogg, J.G., Schmitz, M.D., and Ogg, G.M., eds., Geologic Time Scale 2020 (first edition): Oxford, Elsevier, p. 521561.CrossRefGoogle Scholar
Xiao, S., and Tang, Q., 2021, Part B, Volume 2, Chapter 7: Microfossils of prokaryotes (Bacteria and Archaea): research history, taphonomy, and paleobiology: Treatise Online, v. 160, p. 137.Google Scholar
Xiao, S., Jiang, G., Ye, Q., Ouyang, Q., Banerjee, D. M., Singh, B. P., Muscente, A. D., Zhou, C., and Hughes, N. C., 2023, Systematic paleontology, acritarch biostratigraphy, and δ13C chemostratigraphy of the early Ediacaran Krol A Formation, Lesser Himalaya, northern India: Journal of Paleontology, v. 97, https://doi.org/10.1017/jpa.2022.7.Google Scholar
Xing, Y.S., and Liu, K.Z., 1973, On Sinian micro-flora in Yenliao Region of China and its geological significance: Acta Geologica Sinica, v. 1, p. 164.Google Scholar
Yan, Y., and Liu, Z., 1993, Significance of eucaryotic organisms in the microfossil flora of Changcheng System: Acta Micropalaeontologica Sinica, v. 10, p. 167180.Google Scholar
Yin, L., and Guan, B., 1999, Organic-walled microfossils of Neoproterozoic Dongjia Formation, Lushan County, Henan Province, North China: Precambrian Research, v. 94, p. 121137.CrossRefGoogle Scholar
Yin, L., and Yuan, X., 2007, Radiation of Meso-Neoproterozoic and early Cambrian protists inferred from the microfossil record of China: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 254, p. 350361.Google Scholar
Yin, L., Yang, R., Peng, J., and Kong, F., 2009, New data regarding acritarch biostratigraphy from the early–middle Cambrian Kaili Formation in Chuandong, Guizhou Province, China: Progress in Natural Science, v. 19, p. 107114.CrossRefGoogle Scholar
Yin, L.M., Singh, B.P., Bhargava, O.N., Zhao, Y.L., Negi, R.S., Meng, F.W., and Sharma, C.A., 2018, Palynomorphs from the Cambrian Series 3, Parahio valley (Spiti), Northwest Himalaya: Palaeoworld, v. 27, p. 3041.CrossRefGoogle Scholar
Zaine, M.F., 1991, Análise dos fósseis de parte da Faixa Paraguai (MS, MT) e seu contexto temporal e paleoambiental [Ph.D. thesis]: São Paulo, Brazil, Universidade de São Paulo, 215 p.Google Scholar
Zang, W., 1995, Early Neoproterozoic sequence stratigraphy and acritarch biostratigraphy, eastern Officer Basin, South Australia: Precambrian Research, v. 74, p. 119175.CrossRefGoogle Scholar
Zang, W., and Walter, M.R., 1992, Late Proterozoic and early Cambrian microfossils and biostratigraphy, northern Anhui and Jiangsu, central-eastern China: Precambrian Research, v. 57, p. 243323.Google Scholar
Zhang, Y., Yin, L., Xiao, S., and Knoll, A.H., 1998, Permineralized fossils from the terminal Proterozoic Doushantuo Formation, South China: Paleontological Society Memoir 72, 52 p.CrossRefGoogle Scholar
Zheng, S., Clausen, S., Feng, Q., and Servais, T., 2020, Review of organic-walled microfossils research from the Cambrian of China: implications for global phytoplankton diversity: Review of Palaeobotany and Palynology, v. 276, n. 104191.CrossRefGoogle Scholar
Zhou, C., Xie, G., McFadden, K., Xiao, S., and Yuan, X., 2007, The diversification and extinction of Doushantuo–Pertatataka acritarchs in South China: causes and biostratigraphic significance: Geological Journal, v. 42, p. 229262.Google Scholar
Figure 0

Table 1. List of microfossils from the Sete Lagoas Formation, Bambuí Group, published previously and in this study. Articles: 1 = Sommer (1971); 2 = Marchese (1974); 3 = Simonetti and Fairchild (1989); 4 = Fairchild et al. (1996); 5 = Simonetti and Fairchild (2000); 6 = Fairchild et al. (2012); 7 = Warren et al. (2014); 8 = Perrella Júnior et al. (2017); 9 = Sanchez and Fairchild (2018); 10 = Denezine et al. (2022); 11 = this study.

Figure 1

Figure 1. (1) Geological map of the São Francisco basin (red dashed line) in the São Francisco craton, showing its relationship with neighboring Neoproterozoic fold belts. Inset map shows major cratons in the western Gondwana in a Neoproterozoic paleogeographic configuration: A = Amazonian craton; P = Rio de la Plata craton; K = Kalahari craton; WA = West Africa craton; SFC = São Francisco-Congo craton; PC = Paramirim Corridor. Modified from Reis and Alkmim (2015). (2) Stratigraphic position of the studied section in the Bambui Group columnar section. (3) Geological map of the studied area. The purple dot marks the location of the studied section.

Figure 2

Figure 2. Stratigraphic column and field photographs of the Sete Lagoas Formation at the Barreiro section, Santa Luzia quarry, Januária Municipality, Minas Gerais State, Brazil. (1) Thin-bedded limestone. (2) Intraclastic breccia. Sample horizons are marked with the sample number prefixes MP. Sample numbers in bold mark fossiliferous horizons. The CP- numbers refer to the palynological slides of the illustrated specimens.

Figure 3

Figure 3. Organic-walled microfossils from the Sete Lagoas Formation at the Barreiro section. Slide number and England Finder coordinates (in parentheses) are given for each illustrated specimen. (1–3, 7, 10) Leiosphaeridia minutissima: (1) CP962 (S32); (2) CP962 (F48); (3) CP918 (K22); (7) CP964 (P29); (10) CP963 (F33). (4, 8, 11) Germinosphaera bispinosa, all in slide CP917 (EF coordinates: S26, I43, and O28, respectively). (5) Leiosphaeridia jacutica, CP913 (Y23). (6) Leiosphaeridia crassa, CP964 (H29). (9, 13) Leiosphaeridia tenuissima, all in slide CP914 (EF coordinates: Q30 and R23, respectively). (12) Siphonophycus robustum, CP960 (I50).

Figure 4

Figure 4. Ghoshia januarensis n. sp. from the Sete Lagoas Formation in the Barreiro section. (1) Holotype: CP916 (E46). Note dark spots inside cells indicated by white arrows. Yellow arrows indicate slightly deflated and deformed cells. (2–8) Paratypes: (2–4) CP919 (E18); (3) magnified view of the upper right part of (2), showing slightly deflated and deformed cells; (4) magnified view of the lower left part of (2), showing dark spot in terminal cell (arrow); (5, 6) CP919 (J16); (6) dark-field view of the central part of (5), showing a polyhedral cell (arrow in 5). (7) CP919 (J26); note polyhedral cell at branching point. (8) CP920 (N18/3), showing pointed terminal cell (arrow). (9) Specimen identified in a petrographic thin section of the Sete Lagoas Formation at the Barreiro section in the Januária area. Reproduced from Perrella Júnior et al. (2017) with permission.

Figure 5

Figure 5. Raman spectroscopic data of organic-walled microfossils and amorphous organic matter from the Sete Lagoas Formation at the Barreiro section. (1) Baseline-corrected and fitted Raman spectra. Legends are shown in Figure 5.2. Note that Raman spectra of Ghoshia januarensis (J from holotype and I, K from paratypes) have broader peaks of carbonaceous matter around 1,350 cm–1 and 1,600 cm–1 relative to other Sete Lagoas organic-walled microfossils. (2) Principal component analysis of deconvolved Raman data. Samples: A–B and J, CP916; C–D, CP917; E, I, K, CP920; F, H, MP3728; G, MP3723.

Figure 6

Figure 6. Abundance and size distribution of Leiosphaeridia species from the Sete Lagoas Formation at the Barreiro section.

Figure 7

Figure 7. Stratigraphic distribution and relative abundance of organic-walled microfossils from the Sete Lagoas Formation at the Barreiro section.