You are on page 1of 270

UNIVERSIDADE FEDERAL DA BAHIA

INSTITUTO DE GEOCIÊNCIAS
PROGRAMA DE PESQUISA E PÓS-GRADUAÇÃO EM GEOLOGIA
ÁREA DE CONCENTRAÇÃO:
GEOLOGIA MARINHA, COSTEIRA E SEDIMENTAR

TESE DE DOUTORADO

SEDIMENTOLOGIA, ESTRATIGRAFIA E GEOQUÍMICA DOS

FOSFORITOS E IRONSTONES FOSFÁTICOS DO DEVONIANO

INFERIOR/MÉDIO DA BACIA DO PARNAÍBA

MAISA BASTOS ABRAM

SALVADOR
2020
SEDIMENTOLOGIA, ESTRATIGRAFIA E GEOQUÍMICA DOS
FOSFORITOS E IRONSTONES FOSFÁTICOS DO DEVONIANO
INFERIOR/MÉDIO DA BACIA DO PARNAÍBA

Maisa Bastos Abram

Orientador: Prof. Michael Holz (UFBA)

Tese de Doutorado apresentada ao Programa de Pós-


Graduação em Geologia do Instituto de Geociências da
Federal da Bahia como requisito à obtenção do Título
de Doutor em Geologia, Área de Concentração
Geologia Marinha, Costeira e Sedimentar.

SALVADOR

2020
Ficha catalográfica elaborada pela Biblioteca Universitária de Ciências e Tecnologias
Prof. Omar Catunda, SIBI - UFBA.

A161 Abram, Maisa Bastos


Sedimentologia, estratigrafia e geoquímica dos fosforitos e
ironstones fosfáticos do Devoniano Inferior/Médio da Bacia do
Parnaíba/ Maisa Bastos Abram. – Salvador, 2020.

276 f.

Orientador: Prof. Dr. Michael Holz

Tese (Doutorado) – Universidade Federal da Bahia. Instituto de


Geociências, 2020.

1. Geoquímica. 2. Sedimentologia. 3. Geologia estratigráfica. I.


Holz, Michael. II. Universidade Federal da Bahia. III. Título.
CDU 550.4
MAISA BASTOS ABRAM

SEDIMENTOLOGIA, ESTRATIGRAFIA E GEOQUÍMICA DOS

FOSFORITOS E IRONSTONES FOSFÁTICOS DO DEVONIANO

INFERIOR/MÉDIO DA BACIA DO PARNAÍBA

Tese apresentada ao Programa de Pós-


Graduação em Geologia da
Universidade Federal da Bahia, como
requisito parcial para a obtenção do
Grau de Doutor em Geologia na área de
concentração em Geologia Marinha,
Costeira e Sedimentar em 27/10/2020.

TESE APROVADA PELA BANCA EXAMINADORA:

Dr. Michael Holz


Orientador – PPPGG/UFBA

Dra. Cleide Regina Moura da Silva


Examinadora Externa – CPRM

Dr. Sérgio Bergamaschi


Examinador Externo – UERJ

Dr. Jose Maria Landim Dominguez


Examinador Interno – UFBA

Dr. Aroldo Misi (UFBA)


Examinador Interno – UFBA

Salvador – BA
2020
Dedico este trabalho a todas as
maravilhosas pessoas que me acompanharam,
apoiaram, incentivaram e inspiraram nesta minha
caminhada evolutiva e que tanto fizeram bem
para a minha mente e para o meu coração, às
quais sou muito grata, em especial a minha
família querida.
“People say walking on water is a miracle,
but to me walking peacefully on Earth is
the real miracle”
Thich Naht Hanh
AGRADECIMENTOS

Agradeço ao Grande Espírito toda a caminhada de minha vida, incluindo todos os aprendizados e
experiências que tive na realização deste trabalho e que serviram para me transformar a cada dia como ser.

Agradeço a toda minha família, especialmente Francisco, Bia, Lucas e Clara, pela compreensão e pelo
apoio amoroso nos momentos mais difíceis que passei. Agradeço muito a meu marido por ter me
acompanhado no período em que morei no Canadá, inclusive abrindo mão de seus compromissos
profissionais durante este período. Agradeço também a minha mãe, meu irmão, minha irmã, sogra,
cunhados, cunhadas, sobrinhos e sobrinhas, pela compreensão quanto aos momentos ausentes devido à
realização deste trabalho. Agradeço imensamente por ter sido através de minha ida ao Canadá que um novo
portal tenha sido aberto para meu filho em direção a uma nova aventura na descoberta dos segredos do
nosso Planeta.

Agradeço ao meu orientador Prof. Michael Holz, sempre presente em todas as etapas do meu trabalho,
atento a todas as minhas dificuldades e por ter propiciado importantes momentos de discussão de grande
valia para o meu crescimento profissional. Agradeço também pela sua amizade e pelas suas palavras de
força nos momentos que precisei.

Agradeço ao meu querido amigo e colaborador Peir K. Pufahl, que despertou em mim o amor pelos fosfatos
sedimentares, pelos importantes momentos compartilhados, pelos grandes ensinamentos, por ter batalhado
para que eu fosse realizar parte do meu trabalho no Canadá, na Acadia University, por ter conseguido para
mim um auxílio financeiro pelo ELAP, que me ajudou a pagar os custos laboratoriais no Canadá e por ter
facilitado com sua influência o meu acesso ao Laboratório de Isótopos da Queens University. Agradeço
também pelo apoio e acolhimento que ele e sua esposa Christa dispensaram a mim e a minha família
durante todo o período que estivemos no Canadá e dos momentos que tivemos ao lado de seus lindos filhos:
o doce Euan e o esportivo Callum.

Agradeço imensamente a CPRM por grande parte da minha vida e pelo maior suporte financeiro deste
trabalho. Um especial agradecimento a Manoel Barreto que enquanto presidente me apoiou nesta
empreitada e ao presidente Esteves Conalgo e diretores de Geologia e Recursos Minerais Roberto Ventura,
José Andriotti e Márcio Remédio, Chefe do Departamento Marcelo Almeida, pelo período que me
liberaram para a realização do doutorado e pelo suporte financeiro autorizado. Também estendo meus
agradecimentos às Superintendências Regionais de Salvador e Recife pelo apoio operacional. Aos queridos
Saulo, Alcemir, Silas e Jefte meu muito obrigado pelo auxílio na manipulação dos testemunhos e na minha
condução para o galpão da ANM Recife. À Cleide pelas valiosas sugestões no meu exame de qualificação.

O presente trabalho foi realizado com apoio da Coordenação de Aperfeiçoamento de Pessoal de Nível
Superior – Brasil (CAPES – Código de Financiamento 001). Agradecemos este apoio, concedido através de
bolsa doutorado sanduíche. Esta bolsa me possibilitou realizar 6 meses de estudo no Canadá. Agradeço
também ao governo do Canadá através do “The Emerging Leaders in the Americas Program (ELAP) ” pelo
apoio financeiro concedido que deu suporte aos estudos laboratoriais no Canadá.
Agradecimento especial ao professor Leonardo Borghi (LAGESED-UFRJ), Marco Klotz (PETROBRAS) e
pessoal da Universidade Federal do Rio de Janeiro (UFRJ) e CENPES/PETROBRAS que me permitiram o
acesso aos furos ARN-1-TO, LT-1-TO e NA-1-TO e também na coleta de amostras dos mesmos.

Agradecimento aos colegas Marcelo Ferreira da Silva e Joseneusa Brilhante Rodrigues por todo o apoio na
realização das análises de microssonda na UFGO.

Agradecimento especial também ao querido amigo e professor Claudio Porto pelos ensinamentos e apoio
emocional na segunda etapa do meu trabalho.

Aos meus amigos queridos da DIPEME, Ioná, Rogério, Nivia, Mada, Tamara, que muito me ajudaram nas
deliberações na CPRM, no encaminhamento de amostras para análise, na elaboração de figuras, auxílios na
geoquímica e por todo o apoio psicológico. Vocês que compartilham comigo momentos importantes em
minha vida. Agradeço também a Danilo, Liz e Eliane pelo auxílio no preenchimento de bases de dados e
elaboração de figuras.

Agradeço aos colegas e amigos Sarah Dunn, Jackson Malone, Pam Frail e Haixin Xu que tanto ajudaram
com o meu estabelecimento no Canadá, no desenvolvimento do meu trabalho, na realização de minhas
lâminas e análises de DRX e MEV. Também agradeço pelos bons momentos que tivemos juntos, nas
grandes confraternizações com outros amigos conquistados neste período como Gabe, Peggy, Mike,
Crystal, Matthew, Caleb, Rick que nos acolheram com muito carinho e seus amigos que se tornaram nossos
amigos.

Agradeço também Flordelinda e às amigas do Japão, especialmente Maika, pelos bons momentos
compartilhados no Canadá.

Agradeço a todos os queridos amigos do GETA/UFBA, Caio, Vinicius, Paulo, Edric, Lorena, Priscilla,
Deco, Ana Clara, Cora, Wellington, Adevilson, Flavio, Matheus, Maria, Maya, Hellen, Carol que tanto me
apoiaram na realização deste trabalho e com quem compartilhei prazerosos momentos.

Minha gratidão a todos os professores e colegas do Programa de Pós-graduação em Geologia que


participaram direta ou indiretamente deste processo, como os professores Simone Cruz, Cícero Paixão, José
Maria Landim, Carlson Leite, Reinaldo Brito, Aroldo Misi e Ruy Kenji.

Agradeço também ao professor Eric Hiatt (University of Winscosin – Oshkosh) pelos ensinamentos sobre
diagênese e pelos generosos esclarecimentos em momentos de dúvidas.

Aos colegas Marcos Cristovão Baptista e Rafael Costa da Silva pelo auxílio na identificação de icnofósseis
e palinomorfos.

Agradeço aos gestores, professores e funcionários de apoio da Acadia University, em especial a Anna
Reden, Suzie Currie, Rob Raeside, Ian Spooner, Sandra Barr, Cliff Stanley e Lynn Graves pelo
acolhimento e por terem facilitado meu acesso a Acadia University.
Aos meus amigos e amigas que compartilham comigo importantes momentos da minha vida e que
dividiram comigo as angústias de realizar um doutorado num momento de início da menopausa, impactada
pelas fortes alterações hormônais.

Finalmente a todas as pessoas que direta ou indiretamente contribuíram para a realização desta tese que se
constitui um novo ciclo na minha formação profissional.
RESUMO
O Devoniano é um período de mudanças climáticas significativas e mares epicontinentais, apresentando
sucessões fosfáticas, mas com pouca importância econômica, como na porção noroeste do Gondwana no
Brasil. O objetivo deste trabalho é aprofundar o conhecimento sobre esta temática, estabelecendo modelos
fosfogenéticos e implicações acerca da evolução do planeta e potencial mineral. Investigamos as
sucessões ricas em fosforito e ironstones fosfáticos entre os intervalos do Praguiano ao Givetiano da Bacia
do Parnaíba, Grupo Canindé, formações Itaim e Pimenteira. Este estudo envolveu análise de estratigrafia
de sequências integrada com a caracterização paleoambiental, estudos diagenéticos (DRX, MEV e
microssonda), litoquímicos (elementos maiores e traços em rocha total, COT / S total, padrões TRY),
mineraloquímicos e isotópicos em sideritas e francolitas (δC13, δO18) das sucessões bioelementais. Quatro
sequências deposicionais e 19 parasequências foram definidas, relacionadas a ambiente marinho raso, tipo
plataforma influenciada e dominado por ondas e também trends progradacionais fluvio-deltaicos.
Fosforitos se relacionam a zonas condensadas associadas às superficies de inundação máxima do Eifeliano
tardio e do Givetiano. Foram formados próximo ao limite de base de ondas por tempestade, em
associações autigênicas francolita-siderita e francolita-glauconita. Ironstones foram observados em
intervalos episódicos do Praguiano-Emsiano ao Givetiano, caracterizados por ferrosilicatos (odinite,
chamosite/clinochlore) e cutículas de francolita formando coated grains, pirita, Mg sideritas, além de
sideritas mais puras, precipitadas numa fase mais tardia. Ironstones associam-se à regressão normal em
tratos de sistemas de nível alto e nível baixo. Em geral, quatro estágios de evolução foram descritos:
detrítico, autigênico+retrabalhado, soterramento raso e intemperismo. Mo, razões V/Cr e V/(V+Ni), além
de variações em enxofre total e COT indicam para os intervalos ricos em ironstones fosfáticos variações
de condições subóxicas a anóxicas, com pouco H2S dissolvido. Destacam-se padrões enriquecidos em TR
medianas característicos de condições eodiagenéticas de subóxicas a anóxicas e também comuns para o
Devoniano. Dados isotópicos para as sideritas (δO18 -12,0 a 2,5 e δC13 -21,5 a -0,8) foram agrupados em
clusters, interpretados como relacionados à zonas de degradação sulfato-redutoras e ferro oxi-redutoras.
Variações em δO18 ressaltaram as possíveis variações de salinidade/temperatura nas fases mais tardias de
precipitação. Para as francolitas, os valores isotópicos estiveram mais agrupados (δO18 -14,1‰ a -4,6‰ e
δC13 -2.9‰ a -8.1‰), refletindo padrões de intemperismo, embora valores de δC13 coincidem com
fosforitos relacionados à zonas de upwelling. Águas óxicas preponderaram e episódicos fenômenos de
upwelling são interpretados como fontes de Fe-Si-P em uma plataforma distal a mediana. Fontes
continentais adicionais de nutrientes e ferro são também consideradas. A estratificação da coluna d’água
não foi uma feição estável, estando periodicamente submetida a variações devido a ventilação por ação de
tempestades e descargas fluviais, refletindo contribuições marinhas e terrígenas, aumento da produtividade
e processos de degradação microbiana. As descargas fluviais com grande contribuição de matéria orgânica
terrígena associada aos processos de upwelling devem ter contribuído para a expansão dos processos de
anoxia que, por sua vez, inibiram a eficiência na precipitação de P, implicando em baixo potencial
econômico dos depósitos devonianos. As TR constituem bem mineral com importância econômica
atrelada a este sistema.

Palavras-chave: Formação Pimenteira – Formação Itaim - Fosforito – Ironstone - Devoniano.


ABSTRACT

The Devonian is a period of climate change and epeiric seas, with phosphate successions, but with little
economic importance, such as the northwestern areas of Gondwana margin in Brazil. The objective of this
work is to deepen the knowledge about the occurrences of Devonian phosphates, in order to approach this
theme, with the establishment of phosphogenetic models and implications on the evolution of the planet
and mineral potential. Pragian to Givetian successions in the Canindé group, Itaim and Pimenteira
formations, Parnaíba Basin (Brazil) were investigated. Phosphorite and phosphatic ironstone rich
successions and their host rocks were investigated in an integrated study based on sedimentology, high
resolution sequence stratigraphy with petrographical, mineralogical and geochemical (whole rock major
and trace element, δ13C and δ18O isotope, total organic carbon/total sulfur and Corg/P ratio) data to
highlight paleoproductivity, redox, recycling, diagenesis and phosphogenetic processes. Lithofacies
stacking patterns indicate that deposition occurred during marine transgressions resulting in four
depositional sequences (B, C, D and E) and 19 parasequences. Phosphorite occurs is in the form of
concretions or pristine layers in offshore environment, near SWB (outer/inner shelf) precipitated under
suboxic to anoxic conditions. It also occurs as glauconite francolite nodules in hetherolitic lithofacies
precipitated in oxic to suboxic zones. Ironstone and phosphatic ironstone were deposited in more proximal
settings, near FWWB. Molibdenium, V/Cr and V/(V+Ni) ratios for phosphatic ironstone and ironstone
intervals indicated variations from dysoxic/suboxic conditions to more anoxic and minimal dissolved H2S
in phosphatic ironstone intervals. Total sulfur versus TOC indicated variations from dysoxic, anoxic and
even euxinic conditions for organic lithofacies. Distinctive MREE enriched patterns are similar to other
Devonian phosphorites patterns, even for samples with authigenic francolite precipitation near surface
water. The variation for the siderites was of δ18O (VPDB) -12.0 to 2.5 and δ13C-21.5 to -0.8. The variation
indicates early diagenesis with pore-water sensitive to organic degradation that evoluted from sulfate
reduction to iron reduction zones and possible variations in salinity due to mixing freshwater. Francolite
presented more clustered data points then siderites (δ18O VPDB from -14.11‰ to -4.6‰) / (δ13C from -
2.9‰ to -8.1‰), although δ18O VPDB data were strongly affected by weathering. We recorded that the
limit inner/outer shelf waters were more enriched in P concentrations then shorewards regions, with the
development of phosphogenesis along the shelf generating an oxygen minimum zone with geochemical
data like modern upwelling sites. River discharges probably affected physical characteristics of the upper
water column and produced large spatial variability in the Parnaiba basin system probably influencing
vertical stratification and salinity or temperature leading to gradients of surface nutrient concentrations.
Not only upwelling ferruginous water but terrestrial contribution of OM contributed to enhance shallower
anoxia resulting in reduced total phosphorus burial efficiencies. The phosphorus cycle was probably
affected by the global low oxygen and anoxic conditions in the Devonian. This may have affected the
burial of P across the entire shelf, implicating in low economic potential for Early and Middle Devonian
phosphorites, although rare earth elements and yttrium concentrations represent a potentially economic
byproduct.

Keywords: Pimenteira Formation – Itaim Formation - Phosphorite – Ironstone - Devonian


SUMÁRIO
CAPÍTULO 1 – INTRODUÇÃO GERAL ………………………………………… 13

CAPÍTULO 2 – ARTIGO 1: EARLY TO MIDDLE DEVONIAN IRONSTONE 30


AND PHOSPHORITE IN THE NORTHWESTERN GONDWANA
PARNAÍBA BASIN, BRAZIL: A RECORD OF AN EPEIRIC MARGIN
PALEOCEANOGRAPHIC CHANGES.

CAPÍTULO 3 – ARTIGO 2: EARLY AND MIDDLE DEVONIAN 59


PHOSPHOGENESIS IN NORTHERN GONDWANA (PARNAÍBA BASIN,
BRAZIL): PETROGRAPHIC, GEOCHEMICAL AND ISOTOPIC
EVIDENCE FOR PHOSPHORUS CYCLING.

CAPÍTULO 4 – CONCLUSÕES ……………………………………………… 150

APÊNDICE A – JUSTIFICATIVA DA PARTICIPAÇÃO DOS CO-AUTORES 156

APÊNDICE B – MAPA COM LOCALIZAÇÃO DOS FUROS E …………… 157


AFLORAMENTOS DESCRITOS

APÊNDICE C – PERFIS DOS FUROS DE SONDAGEM DESCRITOS ……… 160

APÊNDICE D – RESULTADOS DAS ANÁLISES POR DIFRATOMETRIA 183


DE RAIOS X

APÊNDICE E – ESTUDOS POR MICROSSONDA – COMPOSIÇÃO DAS 213


SIDERITAS

APÊNDICE F – TRATAMENTO DOS DADOS GEOQUÍMICOS E OUTROS 217


DIAGRAMAS GERADOS

ANEXO A – REGRAS DE FORMATAÇÃO DA REVISTA SEDIMENTARY 242


GEOLOGY

ANEXO B – REGRAS DE FORMATAÇÃO DA REVISTA JOURNAL OF 256


SOUTH AMERICAN EARTH SCIENCES.

ANEXO C – COMPROVANTE DE ACEITE DO ARTIGO 1 ............................... 269

ANEXO D – COMPROVANTE DE SUBMISSÃO DO ARTIGO 2 .................... 270


13

CAPÍTULO 1
INTRODUÇÃO GERAL

Os fosforitos são sedimentos bioelementais frequentemente associados a rochas ricas em matéria

orgânica. O fosforito é uma rocha sedimentar bioquímica rica em fósforo, que contém um mínimo de 18%

em peso de P2O5, e seu uso está predominantemente relacionado à indústria dos fertilizantes, constituindo

elemento fundamental e sem substituto na agroindústria, portanto essencial para a segurança alimentar

para a humanidade (Glenn et al., 1994; Trappe, 1998; Pufahl, 2010). O fósforo também constitui um

elemento crucial para a vida e para manter a produtividade do ecossistema e consequentemente para

regulagem do clima global. A disponibilidade na biosfera de fósforo resulta da interação entre clima,

soerguimento tectônico, transporte atmosférico e oceânico e ciclos bióticos (Follmi, 1996; Filipelli, 2011).

Periodicamente, através do tempo geológico, fosforitos são formados, em resposta às

condições oceânicas específicas, acumulado em concentrações suficientes para formar espessas camadas

estratigráficas de extensão regional. Este processo quando ocorre não é apenas episódico, mas

geralmente é de amplitude mundial. Nos modelos fosfogenéticos a decomposição de organismos e

a produtividade orgânica são componentes extremamente importantes. Muitos depósitos de fosforito

atuais ocorrem em áreas de ressurgência (upwelling), onde o fósforo, regenerado da reciclagem da matéria

orgânica em oceanos profundos é levado a áreas de costa alimentando organismos e elevando a

produtividade primária (Pufahl, 2010; Papineau, 2010) (Fig.1). Nestes ambientes de upwelling o fósforo

é extraído da superfície dos oceanos pelo fitoplâncton e autigenicamente convertido em carbonato

fluorapatita na lama rica em matéria orgânica depositada juntamente com os sedimentos. O fósforo

também encontra-se adsorvido em óxidos e hidróxidos de ferro em zonas episódicas de ressurgência ou de

não ressurgência. O fósforo da matéria orgânica ou adsorvido em óxidos hidróxidos de ferro são então
14

concentrados por variados mecanismos que incluem atividade microbiana, Fe redox (Heggie et al., 1990) e

processos de concentração interporos (autigênicos-diagenéticos), na superfície ou abaixo da interface

sedimento/água (Nelson et al., 2010; Pufahl, 2010; Papineau, 2010). A matéria orgânica é degradada

segundo uma sequência de reações redox, resultando num consumo crescente de oxigênio. As reações se

processam em ordem segundo uma redução da produção de energia, da seguinte forma: respiração óxica,

denitrificação, redução de óxidos de metais, redução de sulfato e metanogênese (Froelich et al., 1979).

Com isto são distinguidas zonas autigênicas distintas no sedimento que correspondem ao perfil de

concentração de O2, NO3-, Mn2+, Fe2+, e SO42- na água interporo. As zonas sulfato redutoras respondem

por importantes concentrações de fosfato e são em parte mediadas por bactérias que metabolizam enxofre

sob condições anóxicas (Bailey et al., 2013). A precipitação de francolita ocorre logo abaixo do assoalho

oceânico, nos primeiros 5 a 20 cm da interface água sedimento em associação com a redução microbial de

nitrato, óxidos de Mn, óxidos de ferro e sulfato, com a liberação de polifosfato (Janhke et al., 1983). Com

o soterramento a precipitação de francolita fica prejudicada em maiores profundidades devido a ausência

de F que deriva da difusão da água do mar e por causa das altas alcalinidades com a degradação

progressiva da matéria orgânica. O aumento da produtividade primária também controla a razão de

dióxido de carbono que é removido da atmosfera e depositado como matéria orgânica sendo, em larga

escala, um importante mecanismo de feedback regulador do clima (Föllmi, 1996). Sheldon (1980)

demonstrou uma forte correlação entre os maiores episódios fosfogenéticos e os ciclos globais de

primeira ordem quando o nível do mar está muito alto na maior parte das regiões costeiras do mundo

(Highstands). Daí a importante correlação de depósitos de fosfato com as superfícies de inundação

máxima e do entendimento estratigráfico das sequências. Por sua vez, depósitos econômicos são

formados como resultado da concentração hidráulica de fosfato retrabalhado formando estratos granulares

com mínima concentração de ganga (Pufahl e Groat, 2017). Ao longo da superfície máxima de inundação

nas zonas de ressurgência, camadas enriquecidas são amalgamadas em zonas estratiformes de minério

(Pufahl e Groat, 2017).


15

O Devoniano é representado por dois principais supercontinentes Gondwana e Euramerica, que

se distibuiam relativamente próximos em um único hemisfério e separados pelo estreito oceano Rheic,

enquanto no resto do globo existia o vasto oceano Panthalassa (Scotese, 2001; Markello et al., 2008;

Blakey, 2019) (Fig.2). Do começo ao meio do Devoniano não houveram mudanças significativas na

paleogeografia, alterada mais significativamente para o final do Devoniano, quando inicia o fechamento

progressivo do oceano Rheico, causando possivelmente variações na circulação oceânica e flutuaçoes do

nível do mar. O CO2 atmosférico variou de 3300 ppm no Devoniano Inferior (caracterizando condições de

greenhouse), com baixa concentração de oxigênio, para 350 ppm no Devoniano Superior, resultando em

condiçoes de icehouse (Royer, 2006; Algeo e Ingall, 2007) (Fig.3). Essa mudança e o aumento de

temperatura provavelmente influenciou as concentrações marinhas de oxigênio dissolvido e promoveu

mudanças na química do oceano (Algeo e Ingall, 2007; Dahl et al., 2010). Os processos geoquímicos e

climáticos globais foram possivelmente afetados por essas condições de efeito estufa, e também pela

crescente evolução das plantas vasculares terrestres, que juntamente com o efeito estufa provavelmente

intensificaram o intemperismo e a formação de solos (Algeo et al., 1995). Isso contribuiu

progressivamente para o aumento da oxigenação através do sequestro de carbono orgânico em sedimentos

como mecanismo de retroalimentação. A recorrência de episódios prolongados de anóxia oceânica marcou

o registro estratigráfico global, representado por várias excursões positivas a isótopos de carbono e

eventos de sedimentação de folhelhos orgânicos de primeira e terceira ordem (Becker et al., 2012, 2016)

(Fig.3). As condições de greenhouse provavelmente contribuíram para aumentar o fluxo de nutrientes aos

oceanos, onde o desenvolvimento dos solos pode ter desencadeado um aumento na biodisponibilidade do

fósforo (Follmi, 1996; Algeo e Scheckler, 1998; Guidry e Mackenzie, 2000; Percival et al., 2019), já que

os dados isotópicos do estrôncio para o Devoniano permitem inferir que o fluxo de material terrígeno e

possivelmente fósforo reativo no oceano deve ter sido equiparável a outros períodos fosfogenéticos

econômicos (Fig.3) (Veizer et al., 1999).

Portanto, assim como em períodos com importantes depósitos econômicos do Permiano e do

Cretáceo, o Devoniano também é marcado por mares epíricos e condições de greenhouse (Hallam, 1992;
16

Vail et al., 1977; Scotese, 2001; Markello et al., 2008; Gradstein et al., 2012). Apesar de algumas

condições semelhantes e diferenças relacionadas à evolução deste tempo geológico, alguns intervalos

sedimentares devonianos registrados no mundo são marcados por processos de fosfogênese, entretanto,

com reconhecido menor potencial econômico (Dominikovskiy et al., 1964; Cathart e Schmidt, 1977;

Coles e Varga, 1988; Humphreys e Smith, 1989; Correia-Junior et al., 2016; Silveira et al., 2016).

Destacam-se depósitos econômicos de baixa tonelagem do Devoniano Superior (<80 Mt de recursos),

como os depósitos de Chattanooga nos EUA (sul da Euramerica ou Laurentia) e da Formação Geirud no

Irã e na Armênia, no norte de Gondwana (Van Straten, 2002; Boardman, 2009; Morad et al., 2012;

Ghorbani, 2013; Li e Schieber, 2015; Orris et al., 2015; Salama et al., 2018).

Figura 1 – Esboço para a formação de fosforitos e folhelhos pretos em margens continentais atuais através
de processos de ressurgência (upwelling). Destaca-se a explosão de fitoplâncton devido a disponibilidade
de nutrientes, com acumulação de matéria orgânica no assoalho oceânico e degradação microbiana, com a
formação de uma zona mínima em oxigênio. Neste caso os fosforitos associam-se a plataforma externa.
Folhelhos pretos também se associam com ressurgências, mas também formam em ambientes costeiros
17

ricos em nutrientes como em lagunas. Mudanças nas concentrações em elementos traços ao longo da
plataforma também são apresentadas. Depleção em U e Cr elevado próximo ao litoral significam
condições óxicas e subóxicas. Elevado U, V, Cu, Cd, Zn, Mo e Ni reflete deposição em porções anóxicas
profundas. SWB – base de ondas por tempestade; FWB – base de ondas de tempo bom. Fonte: Pufahl e
Hiatt (2012).

No Devoniano, a precipitação mineral de fosfato em parte exibe uma estreita associação com

ironstones (Albuquerque et al., 1972; Oliveira e Barros, 1976; Guerrak, 1988; Andreeva e Chatalov, 2011;

Marshall, 2011; Oliveira e Pereira, 2011; Yilmaz et al., 2015; Denayer, 2016; Dreesen et al., 2016; Abram

e Holz, 2020), que constituem rocha sedimentar bioelementar marinha relacionada ao Fanerozoico que

contém mais que 15% em peso de Fe (Young, 1989; Pufahl, 2010) e que também possui uma importância

econômica em algumas partes do mundo (McGregor et al., 2010).

Como a precipitação de P e Fe está intimamente ligada a componentes orgânicos e seus ciclos

biogeoquímicos são fortemente interconectados (Young, 1989; Pufahl, 2010), esses sedimentos não são

simplesmente registros de processos geológicos, mas também estão intimamente envolvidos na interação

entre processos biológicos, químicos e sedimentológicos e, portanto, podem evidenciar aspectos da

evolução global (Pufahl, 2010; Pufahl e Hiatt, 2012). Tanto os fosforitos quanto os ironstones contém

minerais que precipitam diretamente das águas intersticiais como fases autigênicas guardando

características importantes do ambiente original e do processo de formação, tais como fases minerais e

elementos sensíveis a processos de oxidação-redução e níveis de produtividade orgânica. Aliado a estudos

estratigráficos podem fornecer informações importantes sobre as condições da precipitação ao longo da

plataforma servindo também para a reconstrução das condições paleoambientais de precipitação dos

fosfatos e da história diagenética progressiva. A deposição das fases autigênicas podem nos dizer sobre

possíveis variações na química do oceano, incluindo disponibilidade de fosfato, ferro, processos oceânicos

redox / euxínicos, além de condições biológicas e de pH (Taylor e Macquaker, 2011; Poulton e Canfield,

2011).
18

Figura 2 – Localização da área pesquisada inserida no contexto da margem noroeste do paleocontinente


Gondwana. (A) Reconstrução paleogeográfica para o Emsiano (400 Ma) (Blakey, 2019) e padrão de
circulação oceânica para as águas superficiais (adaptada de Parrish, 1982 e Markello et al. 2008).
Correntes oeste mais frias ems etas azuis e correntes mais aquecidas nas setas vermelhas. (B) Porção norte
da América do Sul mostrando a localização da Bacia do Parnaíba e da Bacia do Amazonas. (C, D)
Localização das áreas estudadas (leste e oeste da bacia), os furos de sondagem descritos em amarelo e da
ANP-PETROBRAS reinterpretados em azul, mapa geológico do Grupo Canindé (east and west side),
Parnaiba basin (modified from Schobbenhaus et al., 2004).

Várias ocorrências de fosfato, em alguns casos com ironstones fosfáticos associados, são

identificadas no Devoniano no Brasil, especialmente nas bacias do Amazonas (Formações

Curiri/Barreirinha) e Parnaíba (Itaim/Pimenteira e Longá), relacionadas à porção noroeste do Gondwana.

Estas ocorrências foram originalmente reveladas nos Projetos São Miguel do Tapuio (Albuquerque et al.,
19

1972, Oliveira & Barros, 1976), São Nicolau (CPRM, 1978), Sulfetos de Altamira (Macambira et al.,

1977), Geologia dos Rios Tapajós, Juruena e Teles Pires (Reis 2006), Fosfato Brasil Partes I e II (Abram

et al., 2011; Abram et al., 2016), trabalhos de empresas privadas (Du Solo, Vale e Talon Metals) e estudos

realizados pela UERJ, neste caso com foco na caracterização de rocha fonte para a indústria do petróleo

(Pereira et al., 2009, 2010). Em todos os casos citados inexistem trabalhos que aprofundem o

conhecimento no que diz respeito às ocorrências de fosfato que possam caracterizar, ou não, um potencial

efetivo para depósitos econômicos. Dentre as lacunas existentes no conhecimento, destacaria como

principais questões: Quais são as fácies fosfáticas? Como é a distribuição do fosfato nas seções

devonianas e qual a relação com a evolução estratigráfica da bacia? Por que algumas seções das bacias

mostram melhores resultados que outras? Quais foram as condições no Devoniano para fosfogênese, tais

como nível do mar, paleoambiente, paleogeografia, nutrientes disponíveis, paleoceanografia, condições

físico-químicas? As ocorrências de fosfato nessas bacias devonianas são similares às que ocorrem no

mundo, neste período? Qual é o potencial do Devoniano para fosfato econômico? Qual o significado da

associação dos ironstones com os intervalos fosfáticos? O que estas ocorrências podem elucidar sobre os

processos globais que ocorreram no Devoniano? Pela nossa hipótese de trabalho, processos de mudança

do nível do mar, aporte sedimentar, disponibilidade de nutrientes, eutrofismo, anoxia, paleoambitente,

paleogeografia, paleoceanografia, influenciaram a fosfogênse em bacias devonianas, constituindo

parâmetros importantes para a definição de potenciais econômicos para fosfato.


20

Figura 3 – Esboço evolutivo do Devoniano.

Assim, o objetivo deste trabalho foi aprofundar o conhecimento sobre as ocorrências de fosfato

devonianas, de forma a se aproximar desta temática, com o estabelecimento de modelos fosfogenéticos e

implicações acerca da evolução do planeta e potencial mineral. Para tanto, pela maior disponibilidade de

acesso à dados, investigamos as sucessões ricas em fosforito e ironstones fosfáticos entre os intervalos do
21

Praguiano ao Givetiano da Bacia do Parnaíba, envolvendo o Grupo Canindé, formações Itaim e

Pimenteira (Figura 4).

Figura 4 – Carta estratigráfica da Bacia Parnaíba adaptada de Vaz et al. (2007). Setas em vermelho
indicam as unidades estudadas.

A bacia do Parnaíba, aqui pesquisada, e stá representada por sequências epicontinentais

espessas, principalmente de natureza siliclástica, com uma espessura total de 3.500 m no depocentro,

apresentando uma margem assimétrica, mais estruturada a oeste e mais suave a leste (Grahn e Caputo,

1992; Vaz et al., 2007; Daly et al., 2014; Tribaldos e White, 2018). É classificada como uma bacia

intracratônica (Gabaglia e Figueiredo, 1986) que jaz sobre 3 blocos pré-cambrianos (Amazonas/Araguaia,

Parnaíba e o da Borborema; Daly et al., 2014) com uma forma subcircular que cobre cerca de 600 mil

km2. A história tectônica desta bacia iniciou-se numa fase de rifts termicamente induzidos (Castro et al.,

2014; Daly et al., 2014; Tribaldos e White, 2018). Em seguida, instalou-se a fase intracratônica que, por

sua vez, possui uma longa história de subsidência (200Ma) com desenvolvimento de um mar

epícontinental, a uma taxa muito baixa (em geral < 1m por Ma), relacionada à estabilização da plataforma

sul-americana (Araújo, 2015; Menzies et al., 2018; Tribaldos e White, 2018). O registro sedimentológico

compreende seis mega sequências de segunda ordem delimitadas por inconformidades regionais geradas

por flutuações e pelo ciclo epeirogênico laurasiano (Caputo e Lima, 1984; Vaz et al., 2007) (Fig. 4). É
22

digno de nota que alguns autores (por exemplo, Caputo e Santos, 2018; Menzies et al., 2018) desafiam

este modelo estratigráfico, afirmando que as inconformidades em toda a bacia e a influência dos eventos

tectônicos laurasianos são mal interpretados. Discordâncias locais no grupo Canindé são sugeridas por

Caputo e Santos (2018) e Menzies et al. (2018). De acordo com Menzies et al. (2018), as

inconformidades são mais comuns em áreas marginais, onde quedas discretas de nível de base podem

refletir em inconformidades subaéreas apenas na borda da bacia, formando conformidades correlativas em

direção à bacia.

Segundo Vaz et al. (2007), a sucessão de rochas sedimentares e magmáticas da Bacia do

Parnaíba pode ser disposta em cinco supersequências: Siluriana (correspondendo litoestratigraficamente ao

Grupo Serra Grande), Mesodevoniana-Eocarbonífera (Grupo Canindé), Neocarbonífera-Eotriássica (Grupo

Balsas), Jurássica (Fm. Pastos Bons) e Cretácea (formações Codó, Corda, Grajaú e Itapecuru),

delimitadas por discordâncias regionais. Os sedimentos da Formação Itaim possuem idade Praguiana a

Emsiana (Vaz et al., 2007; Grahn et al., 2008; Trindade et al., 2015), compreendendo a unidade basal do

Grupo Canindé. Exibe folhelhos bioturbados em seções basais e uma sucessão progradacional com

granulação fina a média de face de costa e arenitos de amnbientes de maré e delta no topo (Carozzi et al.,

1975; Góes e Feijó, 1994). A Formação Pimenteira tem idade que varia de Eifeliana para Frasniana

(Grahn et al., 2008; Trindade et al., 2015) e é composta por folhelhos de cores cinza-preto a esverdeado,

orgânicos, parcialmente bioturbados, com estratificação cruzada tipo hummocky subordinada (Della

Fávera, 1990; Vaz et al., 2007). A área de pesquisa se situa no lado leste desta bacia, na região centro-sul

do estado do Piauí, proximidades das cidades de Picos, Pimenteiras e São Miguel do Tapuio, e no lado

oeste da bacia, na porção leste do estado do Tocantins, próximo às cidades de Lagoa do Tocantins e

Novo Acordo (Figura 3).

Este estudo teve portanto como foco as unidades Itaim e Pimenteira que foram objeto de uma

análise estratigráfica de sequências de alta resolução envolvendo um estudo integrado das sucessões

bioelementais, incluindo os fosforitos e os ironstones, para caracterização dos intervalos fosfáticos,

definição dos estágios de evolução da bacia no intervalo considerado, reconstrução paleoambiental,


23

reconstrução das variações químicas do oceano, sua influência nos modelos fosfogenéticos e na evolução

do Devoniano. Para tanto foram realizadas etapas de campo (apenas na porção leste da bacia), descrição

de furos de sondagem (22 furos na porção leste da bacia e 3 furos na porção oeste da bacia),

reinterpretação de dados de 4 furos da ANP, estudos petrográficos, difratometria de Raios X, MEV,

microssonda, estudos diagenéticos, estudos litogeoquímicos, mineraloquímicos e isotópicos. Esses estudos

envolveram a determinação de elementos maiores e traços em rocha total, carbono orgânico total, enxofre

total, razão Corg/P, composições elementares de siderita e francolita eodiagenética, investigação de

padrões elementos Terras Raras e Ítrio (ETRY), além de isótopos estáveis (δC13 e δO18) em siderita e

francolita. O objetivo foi destacar processos de paleoprodutividade e paleoredox, de forma a fornecer

informações mais detalhadas sobre ambientes deposicionais e diagênese (Pufahl e Hiatt, 2012; Phan et al.,

2019). Os isótopos estáveis foram utilizados para estimar a composição original da água, degradação

microbiana e processos redox (Mozley, 1989; Mozley e Wersin, 1992; Jarvis et al., 1994; Morad, 1998;

Mozley e Carothers, 1992; Wilkinson et al., 2000: Pufahl e Hiatt, 2012). Também foi combinado com a

paragênese mineral para inferir a variação na química das águas e a implicação das condições

paleoambientais / paleooceanográficas durante a evolução da bacia, que poderiam justificar ou até prever

o potencial econômico dos depósitos de fosfato.

Este trabalhou culminou na elaboração de dois artigos científicos apresentados nos capítulos II e

III desta tese. O primeiro deles tem como título “Early to Middle Devonian ironstone and phosphorite in

the northwestern Gondwana Parnaíba Basin, Brazil: A record of an epeiric margin paleoceanographic

changes” e foi submetido e publicado pela revista Sedimentary Geology, que possui qualis A2 para

Geociências, pelo Sistema Qualis Capes, fator de impacto 2,728. O segundo trabalho que compõem o

capítulo III é entitulado “Early and Middle Devonian Phosphogenesis in Northern Gondwana (Parnaíba

Basin, Brazil): Petrographic, Geochemical and Isotopic Evidences for Phosphorus Output Mechanisms” e

foi submetido a Revista Journal of South American Earth Science, com qualis B1 para Geociências, pelo

Sistema Qualis Capes, fator de impacto 1,704.


24

As normas das respectivas revistas constam nos anexos A e B e as cartas/e-mails de submissão

nas duas últimas revistas nos anexos C e D.

Referências

Abram, M.B., Cunha, I.A., Almeida, R.C., 2016. Projeto Fosfato Brasil: Parte I. Salvador: Cprm, (org.) Il.;
1 Dvd Anexo. Informe de Recursos Minerais. Série Insumos Minerais Para Agricultura 17,
Programa Geologia do Brasil (PGB), 1480p.
Abram, M.B., Cunha, I.A., Porto, C.G., Brito, R.S.C., 2011. Projeto Fosfato Brasil: Parte I. Salvador:
Cprm, (org.) 529 P. Il.; 1 Dvd Anexo. Informe De Recursos Minerais. Série Insumos Minerais
Para Agricultura 13, Programa Geologia do Brasil (PGB).
Abram, M.B., Holz, M., 2020. Early to Middle Devonian ironstone and phosphorite in the northwestern
Gondwana Parnaíba Basin, Brazil: A record of an epeiric margin paleoceanographic changes.
Sedimentary Geology 402, 105646.
Albuquerque, H.J.T.R.; Coelho, J.M.; Farias, C.E.G., 1972. Projeto São Miguel do Tapuio. Recife:
CPRM. 2v. Relatório interno. Convênio CNEN/CPRM.
Algeo, T. J., Berner, R. A., Maynard, B. J., Scheckler, S. E., 1995. Late Devonian oceanic anoxic events
and biotic crisis: “Rooted” in the evolution of vascular land plants. GSA Today 5, 45, 63-66.
Algeo, T.J., Ingall, E., 2007. Sedimentary Corg/P ratios, paleocean ventilation, and Phanerozoic
atmospheric pO2. Palaeogeography, Palaeoclimatology, Palaeoecology 256, 130–155.
Algeo, T.J., Scheckler, S.E., 1998. Terrestrial-marine teleconnections in the Devonian: links between the
evolution of land plants, weathering processes and marine anoxic events. Philosophical
Transactions of the Royal Society of London, series B: Biological Sciences 353, 113–130.
Andrade, C.L.N., 2015. Arcabouço palinoestratigráfico e geoquímica orgânica da Formação Pimenteiras
(Devoniano), na borda oeste da Bacia do Parnaíba, Brasil, Tese de doutorado, Universidade
Federal da Bahia, Salvador.
Andreeva, P., Chatalov, A., 2011. Origin of the Eifelian ironstone from Well R-119 Kardam, Northeastern
Bulgaria. Comptes Rendu de l’ Académie Bulgare des Science 64, 91-101.
Araújo, D.B., 2015. Bacia do Parnaíba. Sumário geológico e setores em oferta. Agência Nacional do
Petróleo, Gás Natural e Biocombustíveis - Décima Terceira Rodada de Licitações.
Superintendência de Definição de Blocos Sdb, 21pp.
Bailey, J.V., Corsetti, F.A., Greene, S.E., Crosby, C.H., Liu, P., Orphan, V.J., 2013. Filamentous sulfur
bacteria preserved in modern and ancient phosphatic sediments: implications for the role of
oxygen and bacteria in phosphogenesis. Geobiology 11, 397-405.
Becker, R.T., Gradstein, F.M., Hammer, O., 2012. The Devonian Period. In: Gradstein, F.M., Ogg, J.G.,
Schmitz, M., Ogg, G. (Eds). The Geologic Time Scale 2012, 2. Elsevier, Amsterdam, 559–601.
Becker, R.T., Kȍnigshof, P., Brett, C.E., 2016. Devonian Climate, Sea Level and Evolutionary Events. In:
Becker, R. T., Kȍnigshof, P., Brett, C.E. (Eds.). Devonian Climate, Sea Level and Evolutionary
Events. Geological Society, London, Special Publications 423, 1–10.
Blakey, R., 2019. Global Paleogeography and Tectonics in Deep Time Series
http://deeptimemaps.com/global-paleogeography-and-tectonics-in-deep-time-series.
25

Boardman, D.R. 2009. Preliminary Analysis of Phosphate Nodules in the Woodford Shale, Late Devonian
– Early Mississippian, Southern Oklahoma. Faculty of The Graduate College of the Oklahoma
State University, Masters in Science, Eua, 77p.
Caputo, M.V., Lima, E.C., 1984. Estratigrafia, idade e correlação do Grupo Serra Grande - Bacia do
Parnaíba. Anais do XXXIII Congresso Brasileiro de Geologia, Rio de Janeiro, 15pp.
Caputo, M.V., Santos, R.O.B., 2019. Stratigraphy and ages of four Early Silurian through Late Devonian,
Early and Middle Mississippian glaciation events in the Parnaíba Basin and adjacent areas, NE
Brazil. Earth-Science Reviews 7, 103002, doi.org/10.1016/j.earscirev.2019.103002.
Carozzi, A.V., Falkenhein, F.U.M., Carneiro, R.G., Esteves, R.P., Contreiras, C.J.A., 1975. Análise
ambiental e evolução tectônica sinsedimentar da seção siluroeocarbonífera da Bacia do Maranhão.
PETROBRAS, Rio de Janeiro, Brasil, 48 pp.

Castro, D.L.; Fuck, R.A.; Phillips, J.D.; Vidotti, R.M.; Bezerra, F.H.R.; Dantas, E.L. 2014. Crustal
structure beneath the Paleozoic Parnaíba Basin revealed by airborne gravity and magnetic data,
Brazil. Tectonophysics 614, p.128–145.

Cathart, J.B., Schmidt, D.L., 1977. Middle Paleozoic sedimentary phosphate in the Pensacola Mountains,
Antarctica. Contributions to the geology of Antarctica. USGS Professional Paper 456-E.
Coles, K.S., Varga, R.J., 1988. Early to middle Paleozoic phosphogenic province in terranes of the
southern Cordillera, western United States. American Journal of Science 288, 891-924.
Companhia de Pesquisa de Recursos Minerais – CPRM, 1978. Projeto São Nicolau: Fosfato.
Correia-Junior, F.C., Abram, M.B., Ferreira, M.V., Cunha, I.A., Neves, M.P., 2016. A pesquisa para
fosfato na Bacia do Amazonas, Estado do Pará. In: Abram, M.B., Bahiense, I.C., Almeida, R.C.
(Eds.), Projeto Fosfato Brasil: parte II, Informe de Recursos Minerais, Série Insumos Minerais
para Agricultura 17, Salvador, CPRM, 891-834.
Dahl, T.W, Hammarlund, E.U., Anbar, A.D., Bond, D.P.G., Gill, B., Gordon, G.W., Knoll, A.H., Nielsen,
A.T., Schovsbo, N.H., Canfield, D.E., 2010. Devonian Rise in atmospheric oxygen correlated to
the radiations of terrestrial plants and large predatory fish. PNAS, Proceedings of the National
Academy of Sciences of the United States, 107, 42, 17911-17915.
Daly, M.C.; Andrade, V.; Barousse, C.A.; Costa, R., McDowell, K.; Piggott, N.; Poole, A.J. 2014.
Brasiliano crustal structure and the tectonic setting of the Parnaíba basin of NE Brazil: Results of
a deep seismic reflection profile. Tectonics, American Geophysical Union 33, 2102–2120.
Della Fávera, J.C., 1990. Tempestitos da bacia do Parnaíba: um ensaio holístico. Tese de Doutorado, IG-
UFRGS, Porto Alegre. 243p.
Denayer, J., 2016. Iron ores of Southern Belgium: much more than hematite. Antrophologica et
Præhistorica 126, 39-49.
Dominikovskiy, V.N., Drogunov, V.I., Librovich, V.L., 1964. Facies of phosphorite deposits in the
Siberian platform. International Geology Review 6, 5, 830-835.
Dreesen, R., Savary, X., Goemaere, E., 2016. Definition, classification and microfacies characteristics of
oolitic ironstones used in manufacturing of red ochres: A comparative petrographical analysis of
Palaeozoic samples from France, Belgium and Germany. Antrophologica et Præhistorica 125,
203-223.
Filippelli, G.M. 2011. Phosphate rock formation and marine phosphorus geochemistry: the deep time
perspective. Chemosphere 84, 6, 759-66.
Follmi, K., 1996. The phosphorus cycle, phosphogenesis and marine phosphate-rich deposits. Earth-
26

Science Reviews 40, 55–124.


Froelich, P.H., Klinkhammer, G.P., Bender, M.L., Luedtke, N.A., Heath, G.R., Cullen, D., Dauphin, P.,
Hammond, D., Hartman, B., Maynard, V. 1979. Early oxidation of organic matter in pelagic
sediments of the eastern equatorial Atlantic: suboxic diagenesis. Geochimica et Cosmochimica
Acta 43, 1075-1090.
Gabaglia, G.P.R., Figueiredo, A.M.F., 1990. Evolução dos conceitos acerca das classificações de bacias
sedimentares. In: Gabaglia, G.P.R. & Milani, E.J. (Eds.), Origem e evolução de bacias
sedimentares. Rio de Janeiro, Petrobras, 31-45.
Ghorbani, M. 2013. The economic geology of Iran--Mineral deposits and natural resources: New York,
Springer, 569 p.
Glenn, C.R., Ollmj, K.F., Gr, K.A., Trappe, J.O., Abed, A.M., Galli, C., Garr, R.E., 1994. Phosphorus and
phosphorites: Sedimentology and environments of formation. Eclogae Geologicae Helvetiae 87,
747-788.
Goemare, E., Katsch, A., Eschghi, I., Dreesen, R. 2016. Geological record and depositional setting of
Palaeozoic Oolitic ironstones in Western Europe. Antrophologica et Præhistorica 125/2014, 23-
43.
Góes, A.M.O., Feijó, F.J., 1994. Bacia do Parnaíba. Rio de Janeiro. Boletim de Geociências da Petrobras
8, 1, 57–67.
Goncuoglu, M.C., Boncheva, I., Yilmaz, I.O., Saydam-Demiray, D., 2008. Devonian Ironstone
Formations within the Paleozoic Carbonate Platform, in NW Anatolia. Palaeozoic Climates.
International Congress.
Gradstein, M., Ogg, J. G., Gabi, M.S. 2012. The Geologic Time Scale 2012 - first Edition.
ISBN: 9780444594488.
Grahn, Y., Caputo, M.V., 1992. Early Silurian glaciations in Brazil. Palaeogeography, Palaeoclimatology,
Palaeoecology 99, 9-15.

Grahn, Y., Young, C., Borghi, L., 2008. Middle Devonian chitinozoan biostratigraphy and sedimentology
in the eastern outcrop belt of the Parnaíba Basin, northeastern Brazil. Revista Brasileira de
Paleontologia 11, 137–146.
Guerrak, S., 1988. Geology of the Early Devonian oolitic iron ore of the Gara Djebilet field, Saharan
Platform, Algeria. Ore Geology Reviews 3, 333-358.
Guidry, M.W., Mackenzie, F.T., 2000. Apatite weathering and the Phanerozoic Phosphorus Cycle.
Geology 28, 631-634.
Hallam, A., 1992. Phanerozoic Sea-Level Changes. Columbia University Press, New York. 266 p.
Hangari, K.M. Ahmad, J.N., Perry, E.C. 1980. Carbon and oxygen isotope ratios in diagenetic siderite and
magnetite from U. Devonian ironstone, Wadi Shatti district, Libya. Economic Geology 75, 538-
545.
Haq, B.U., Schutter, S.R., 2008. A chronology of Paleozoic sea-level changes. Science 322, 64–68.
Heggie, D.T. Skyring, G.W., O’brien, G.W., Reimers, C., Herczeg, A., Moriarty, D.J.W., Burnett, W.C.,
Milnes, A.R., 1990. Organic carbon cycling and modern phosphorite formation on the East
Australian continental margin: an overview, In: Notholt, A.J.G., Jarvis, I. (Eds.). Phosphorite
Research and Development. London: Geological Society, Special Publication 52, 5287-117.
Humphreys, B., Smith, S.A., 1989. The distribution and significance of sedimentary apatite in Lower to
Middle Devonian sediments east of Plymouth Sound. Proceedings of the Ussher Society 7, 118-
124.
27

Johnson, J.G., Klapper, G., Sandberg, C.A., 1985. Devonian eustatic fluctuations in Euramerica:
Geological Society of America Bulletin 96, 567–587.
Jahnke, R.A., Emerson, S.R., Roe, K.K., Burnett, W.C., 1983. The present day formation of apatite in
Mexican continental margin sediments. Geochimica et Cosmochimica Acta 47, 259-266.
Jarvis, I., Burnett, W.C., Nathan, Y., Almbaydin, F.S.M., Attia, A.K.M., Castro, L.N., Flicoteaux, R.,
Hilmy, M.E., Husain, V., Qutawnah, A.A., Serjani, A., Zanin, Y.N., 1994. Phosphorite
geochemistry - state-of-the-art and environmental concerns. Eclogae Geologicae Helveticae 87,
643–700.
Li, Y., Schieber, J., 2015. On the origin of a phosphate enriched interval in the Chattanooga Shale (Upper
Devonian) of Tennessee - A combined sedimentologic, petrographic, and geochemical study.
Sedimentary Geology 329, 40–61.
Macambira, E.M.B., Rezende, N.G.A.M, João, X.S.J, Assis, N.P., Calderaro, R.C.B., 1977. Projeto
Sulfetos de Altamira-Itaituba. Relatório Final. Belém: CPRM, 7v. Convênio DNPM/CPRM.
Markello, J.R., Koepnick, R.B., Waite, L.E., Collins, J.F., 2008. The Carbonate Analogs Through Time
(CATT) hypothesis and the global atlas of carbonate fields - A systematic and predictive look at
Phanerozoic carbonate systems. In: Lukasik, J., Simo, J.A.T. (Eds.), Controls on carbonate
platform and reef development. SEPM Special Publication 89, 15–45.

Marshall, L.A., 2011. Sedimentology and authigenesis of the lower Devonian Torbrook Formation
ironstone, Torbrook, Nova Scotia, Canada. Bachelor of Science Thesis, Acadia Univeristy,
Wolfville, Nova Scotia, Canada.
McGregor, F., Ramanaidou, E., Wells, M., 2010. Phanerozoic ooidal ironstone deposits – generation of
potential exploration targets. Applied Earth Science 119, 60–64.

Menzies, L.A., Carter, A., MacDonald, D.I.M., 2018. Evolution of a cratonic basin: insights from the
stratal architecture and provenance history of the Parnaíba Basin. In: Daly, M.C., Fuck, R.A.,
Julià, J., Macdonald, D.I.M., Watts, A.B. (Eds.), Cratonic Basin Formation: A Case Study of the
Parnaíba Basin of Brazil. Geological Society, London, Special Publications 472, 157–179.

Morad, S. 1998. Carbonate cementation in sandstones: distribution patterns and geochemical evolution.
In: Carbonate Cementation in Sandstones (Ed. S. Morad), Int. Association of Sedimentolologists.
Special Publication 26, 1–26.

Morad, S., Ketzer, J.M., De Ros, L.F., 2012. Linking Diagenesis to Sequence Stratigraphy: An Integrated
Tool for Understanding and Predicting Reservoir Quality Distribution. In: Morad, S., Ketzer,
J.M., De Ros, L.F. (Eds.), Linking Diagenesis to Sequence Stratigraphy. IAS Special Publications
45, 1–36.
Mozley, P.S., 1989. Relation between depositional environment and the elemental composition of early
diagenetic siderite. Geology, 17, 704-706.
Mozley, P.S., Carothers, W.W., 1992. Elemental and isotopic composition of siderite in the Kuparuk
Formation, Alaska: effect of microbial activity and water/sediment interaction on early pore-water
chemistry. Journal of Sedimentary Petrology, 62, 681–692.
Mozley, P.S., Wersin, P., 1992. Isotopie composition of siderite as an indicator of depositional
environment. Geology, 20, 817-820.
28

Nelson, G.J., Pufahl, P.K., Hiatt, E.E., 2010. Paleoceanographic constraints on phosphate-rich deposits.
Earth-Science Reviews 40, 55-124.
Oliveira, J.C., Barros, F.L., 1976. Projeto Fosfato de São Miguel do Tapuio. Recife: CPRM, 1976. 4v.
Convênio DNPM/CPRM. Relatório interno.
Oliveira, L.C., Pereira, E., 2011. Ocorrência de Ironstones no Devoniano da Bacia do Paraná. Revista
Brasileira de Geociências, 4, 447-462.

Orris, G.J., Dunlap, P., Wallis, J.C., 2015, Phosphate occurrence and potential in the region of
Afghanistan, including parts of China, Iran, Pakistan, Tajikistan, Turkmenistan, and Uzbekistan,
with a section on geophysics by Jeff Wynn: U.S. Geological Survey Open-File Report, 1121, 70
p. http://dx.doi.org/10.3133/ofr20151121.
Papineau, D., 2010. Global biogeochimical changes at both ends of the Proterozoic: insights from
phosphorites. Astrobiology 10, 165-181.
Parrish, J.T., 1982. Upwelling and petroleum source beds, with reference to Paleozoic. American
Association of Petroleum Geologists Bulletin, 66, 750-774.
Percival, L.M.E., Selby, D., Bond, D.P.G., Rakociński, M., Racki, G., Marynowski, L., Adatte, T.,
Spangenberg, J.E., Föllmi, K.B., 2019. Pulses of enhanced continental weathering associated with
multiple Late Devonian climate perturbations: Evidence from osmium-isotope compositions.
Palaeogeography, Palaeoclimatology, Palaeoecology, 524, 240–249.
Pereira, E., Rodrigues, R., Souza, M.S.P., Bergamaschi, S., 2009. Chemostratigraphy Applied to Paleozoic
Black-Shale Intervals. AAPG International Conference and Exhibition, Rio de Janeiro, Brazil, 15-
18.
Pereira, E., Rodrigues, R.; Bergamaschi, S.; Souza, M.S.P. 2010. Caracterização Químioestratigráfica do
Devoniano Inferior da Bacia do Paraná. Memorias - Museu e Laboratório Mineralógico e
Geológico, Faculdade de Ciências, Universidade do Porto, 14, 349-353.
Phan, T.T., Hakala, J.A., Lopano, C.L., Sharma, S., 2019. Rare earth elements and radiogenic strontium
isotopes in carbonate minerals reveal diagenetic influence in shales and limestones in the
Appalachian Basin. Chemical Geology 509, 194–212.
Poulton, S.W., Canfield, D.E., 2011. Ferruginous Conditions: A dominant feature of the ocean through
Earth’s history. Elements, 7, 107–112.
Pufahl, P.K., 2010. Bioelemental Sediments. In: James, N.P.; Dalrymple, R.W. (Eds.) Facies Models 4.
Canada, Geological Association of Canada, 477-504.
Pufahl, P.K., Hiatt, E.E., 2012. Oxygenation of the Earth’s atmosphere–ocean system: a review of
physical and chemical sedimentologic responses. Marine and Petroleum Geology, 32, 1–20.
Pufahl, P.K.; Groat, L.A., 2017. Sedimentary and Igneous Phosphate Deposits: Formation and
Exploration: An Invited Paper. Economic Geology 112, 3, 483-516.
Reis N.J., 2006. Projeto Rochas Carbonáticas de Apuí – AM. Informe de Recursos Minerais. Série
Insumos Minerais para a Agricultura. CPRM – Serviço Geológico do Brasil, Manaus, 50 p., il.
Royer, D.L., 2006. CO2-forced climate thresholds during the Phanerozoic. Geochimica et Cosmochimica
Acta, 70, 5665–5675.
Salama, W., Khirekesh Z., Amini, A., Bafti, B.S., 2018. Diagenetic evolution of the upper Devonian
phosphorites, Alborz Mountain Range, Northern Iran. Sedimentary Geology, 376, 90-112.
Sandberg, C.A.; Morrow, J.R.; Ziegler, W. 2002. Late Devonian sea-level changes, catastrophic events,
and mass extinctions. Geological Society of America, Special Paper 356, 473-487.
Scotese, C.R. 2001. Atlas of Earth History. PaleoMap Project. University of Texas at Arlington.
29

Sheldon, R.P. 1980. Episodicity of phosphate deposition and deep ocean circulation - an hypothesis.
Society of Paleontologists and Mineralogists, Special Publication 29, 239-248.
Silveira, D.A., Sachs, L.L.B., Moraes Filho, J.C.R., Bahiense, I.C., Batista, I.H., Silva, R.C., 2016. A
pesquisa para fosfato na Bacia Parnaíba, Área Picos – PI. In: Abram, M.B., Bahiense, I.C.,
Almeida, R.C. (Eds.), Projeto Fosfato Brasil: parte II, Informe de Recursos Minerais, Série
Insumos Minerais para Agricultura 17, Salvador, CPRM, 891-834.
Taylor, K.G., Macquaker, J.H.S., 2011. Iron minerals in marine sediments record chemical environments.
Elements, 7, 113–118.
Trappe, J., 1998. Phanerozoic Phosphorite Depositional Systems: A Dynamic Model for a Sedimentary
Resource System. Lecture Notes Earth Science 74, 316pp.
Tribaldos, V.R., White, N., 2018. Implications of preliminary subsidence analyses for the Parnaíba
cratonic basin. In: Daly, M.C., Fuck, R.A., Julià, J., Macdonald, D.I.M., Watts, A.B. (Eds.),
Cratonic Basin Formation: A Case Study of the Parnaíba Basin of Brazil. Geological Society,
London, Special Publications 472, 147–156.
Trindade, V.S.F., Carvalho, M. de A., Borghi, L., 2015. Palynofacies patterns of the Devonian of the
Parnaíba Basin, Brazil: paleoenvironmental implications. Journal of South American Earth
Sciences, 62, 164–175.
Vail, P. R., Mitchum, R. M Jr., Todd, R. G„ Widmier, J. M„ Thompson, S„ III, Sangree, J. B., Bubb, J. N.,
and Hatlelid, W. G., 1977. Seismic stratigraphy and global changes of sea level: American
Association of Petroleum Geologists Memoir 26, 49-212.
Van Houten, F.V., Karasek, R.M. 1981. Sedimentologic framework of late Devonian oolitic iron
formation, Shatti Valley, West-Central Libya. Journal of Sedimentary Petrology 51, 2, 0415-0427.
Van Straaten, P. 2002. Rocks for Crops: Agrominerals of sub-Saharan Africa. ICRAF, Nairobi, Kenya,
338pp.
Vaz, P.T., Rezende, N.G.A.M, Wanderley Filho, J.R, Travassos, W.A.S., 2007. Bacia do Parnaíba.
Boletim de Geociências da Petrobras 15, 289-297.

Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.K.F., Diener, A., Ebneth, S.,
Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., Strauss, H. 1999. 87Sr/86Sr, δ
13
C and δ 18O evolution of Phanerozoic seawater. Chemical Geology, 161, 59–88.
Wilkinson, M., Haszeldine, R.S., Fallick, A.E., Osborne, M.J., 2000. Siderite zonation with the Brent
Group: microbial influence or aquifer flow? Clay Minerals, 35, 107–117.
Yilmaz, İ.Ö., Göncüoğlu, M.C., Demiray, D.G., Gedik, I., 2015. An approach to paleoclimatic conditions
for Devonian (upper Lochkovian and middle Givetian) ironstone formation, NW Anatolian
carbonate platform. Turkish Journal of Earth Sciences, 24, 21-38.
Young, C.G.K., 2006. Estratigrafia de alta resolução da Formação Pimenteira (Devoniano, Bacia do
Parnaíba). Programa de Pós-graduação em Geologia, Universidade Federal do Rio de Janeiro,
Dissertação de Mestrado, 174 pp.
Young, T.P., 1989. Phanerozoic ironstones: an introduction and review. In: Young, T.P., Taylor, W.E.G.
(Eds.), Phanerozoic Ironstones. Geological Society, London, Special Publications 46, 9-25.
30

CAPÍTULO 2
ARTIGO 1
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59

CAPÍTULO 3
ARTIGO 2

EARLY AND MIDDLE DEVONIAN PHOSPHOGENESIS IN NORTHERN GONDWANA


(PARNAÍBA BASIN, BRAZIL): PETROGRAPHIC, GEOCHEMICAL AND ISOTOPIC
EVIDENCE FOR PHOSPHORUS CYCLING

Maisa B. Abram 1,2, Claudio G. Porto 3 Michael Holz 4

1
CPRM - The Geological Survey of Brazil, Bahia State, Brazil, Av. Ulisses Guimarães, 2862, Sussuarana,

Salvador-Bahia, Brazil, CEP 41213-000. maisa.abram@cprm.gov.br

2
Universidade Federal da Bahia, Pós-Graduação em Geologia

3
Universidade Federal do Rio de Janeiro, Av. Athos da Silveira Ramos, 274 - Cidade Universitária - Ilha do

Fundão, Rio de Janeiro – RJ. porto@geologia.ufrj.br

4
Universidade Federal da Bahia, Pós-Graduação em Geofísica e PROSEDGEO – Consultoria e Treinamento,

Brazil, Campus de Ondina, Salvador, Brazil. michael.holz@ufba.br

* Corresponding author: Av. Ulisses Guimarães, 2862, Sussuarana, Salvador-Bahia, Brazil, CEP 41213-000.

maisa.abram@cprm.gov.br, Phone: +55 (71) 992047284 / +55 (71) 21017359.

Abstract

Although the Devonian was a greenhouse period characterized by epeiric seas with elevated

phosphurus availability, it is uncommon for phosphate deposits of this age to be of notable concentration

and/or scale. This contrasts to the economic-scale phosphorites deposited during other intervals of Earth

history which shared the climatic and oceanographic conditions that typified the Devonian. Delineating

the processes responsible for limited phosphogenesis within nutrient-rich Devonian seas requires
60

examination of the petrology of phosphate-bearing successions. Praguian to Givetian stage strata with

minorly enriched phosphate intervals in the Parnaíba Basin of Brazil were investigated to reveal the

mechanisms of phosphorus burial within the ferruginous seawater of the Rheic Ocean. Phosphorite,

phosphatic ironstone, and organic-rich siliciclastic successions were investigated within a stratigraphic

framework via petrographic, mineralogical, and geochemical techniques (whole rock major and trace

elements, rare earth elements concentrations, siderite and francolite δ13C and δ18O, total organic carbon

and sulfur, and Corg/P ratios) to reconstruct the paleoproductivity, redox conditions, OC and P recycling,

and their impact on phosphogenesis. Outer/inner shelf facies in the Parnaíba basin, near storm wave base,

host cm-scale phosphorite and sideritic concretions, pristine phosphorite laminae, and also contain

glauconite/francolite nodules reflecting variations from oxic-suboxic to suboxic-anoxic porewaters. In

addition, proximal facies (near fairweather wave base) of the Parnaíba contain ironstones that are

occasionally phosphatic. Mo (43 to 53), V/Cr (1 to 7.9), V/(V+Ni) (0.63 to 0.98), Corg/P ratios, Cu/Mo

and total sulfur versus total organic carbon plots indicate variable oxic /anoxic environments. Distinctive

MREE enrichment patterns are similar to results from other Devonian phosphorites reflecting extended

anoxia conditions. The isotopic variation of siderite is -12.0 to 2.5 ‰ (average -2.6) for δ18OVPDB and -

21.5 to -0.8 ‰ (average -8.2) for δ13C and indicate pore-water was sensitive to organic degradation.

Sulfate reduction and iron reduction zones and possible variations in salinity due to mixing freshwater are

suggested. Isotopic results for francolite were more tightly clustered than the siderite with δ18OVPDB from -

14.1‰ to -4.6‰ (average -10.1‰), and δ13C from -2.9‰ to -8.1‰ (average -6.3‰). The inner/outer

shelf was more enriched in P than shoreward regions and phosphogenesis developed along the shelf within

an oxygen minimum zone with geochemical data like modern upwelling sites. We interpret that river

discharge probably affected the characteristics of the upper water column and produced spatial variability

in the Parnaíba Basin system probably influencing vertical stratification and salinity leading to gradients

of surface nutrient concentrations. These concentrations contributed to shallower anoxia resulting in

reduced total phosphorus and enhanced organic carbon burial efficiencies. The anoxia enhanced

preservation of organic C and diffusive loss of remineralized organic P and probably affected the
61

phosphorus cycle in the Devonian. This may have influenced the burial of P across the entire shelf,

resulting in low economic potential for Early to Middle Devonian phosphorites, although rare earth

elements and yttrium concentrations represent a potentially economic byproduct.

Keywords: Parnaiba basin – phosphorite – ironstone – geochemistry – oxygen and carbon

isotopes

1. Introduction

Phosphate minerals are the primary phosphorus sink in modern continental margin sediments and

phosphorite formation played a significant role in the phosphorus cycle in ancient oceans, particularly in

epeiric seas (e.g., Ruttenberg, 2003; Hiatt and Budd, 2003; Pufahl, 2010). In the Phanerozoic, phosphatic

deposits are frequently associated with sites of upwelling where nutrient-laden waters stimulate primary

production and the subsequent sedimentation of large amounts of organic matter (Riggs and Sheldon,

1990; Follmi, 1996). Commonly, phosphorus is extracted from nutrient-rich waters by a variety of

phytoplankton whose remains accumulate on the shelf where their degradation liberates phosphate into the

sediment and creates conditions conducive for the formation of phosphate minerals (phosphogenesis;

Follmi, 1996). Authigenic francolite (a highly substituted carbonate-fluorapatite mineral,

Ca5(PO4)2.5(CO3)0.5F; Nathan, 1984) is formed and is believed to be primarily facilitated by microbially

mediated degradation of organic matter, as well as microbial processes that modify bottom and porewater

conditions, including redox gradients, pH and nutrient concentrations (Glenn et al., 1994). Phosphogenic

environments often require rapid, partial to complete degradation of organic matter that releases sufficient

quantities of phosphate for francolite nucleation and mineralization (Trappe, 1998). Inhibiting the release

of phosphate back to overlying waters neccesitates rapid burial, microbial overgrowth, or the formation of

a stable redox-boundary that lies near the sediment-water interface (Trappe, 1998; Pufahl, 2010). The

latter process acts as a primary phosphate supply in environments with low organic matter fluxes because

iron oxyhydroxides can scavenge phosphate in the water column (Heggie et al., 1990). The reductive
62

dissolution of these oxyhydroxides and reprecipitation at the oxic/anoxic boundary can lead to elevated

phosphate concentrations as well as iron accumulation (Heggie et al., 1990). In addition, a granular

deposit can be produced by the hydraulic concentration of francolite peloids derived from pristine

lithofacies (precipitated in situ without reworking or winnowing; Follmi et al., 1991; Pufahl, 2010).

During stratigraphic condensation, cycles of erosion-deposition create thin granular phosphorite horizons

interbedded with pristine phosphorite and this is a significant process in producing economic-scale,

amalgamated granular phosphorite beds (Pufahl et al., 2003; Pufahl and Groat, 2017). Although these

phosphogenetic processes are common in the Phanerozoic, not all periods with an important evolution of

epeiric seas resulted in significant deposits, implying potential differences in this phosphorus cycle.

Phanerozoic phosphorites represent the major economic-scale phosphorus deposits that are mainly

associated with transgressive episodes within Cambrian, Permian, Cretaceous and Eocene epeiric seas

(Cook and McElhinny, 1979; Hiatt and Budd, 2003; Markello et al., 2008; Pufahl, 2010). A majority of

these phosphorites were likely associated with high surface-ocean productivity maintained across

platforms by evaporation-driven lagoonal circulation that supplied phosphorus to near-shore environments

by drawing it away from upwelling fronts and produced enough organic matter to drive phosphogenesis

(Pufahl and Groat, 2017). Devonian Period strata recorded epeiric seas with significant climatic shifts,

prevailing greenhouse conditions, transgressive events, and thick organic-rich successions (Johnson et al.,

1985; Becker et al., 2016), yet these oceanographic conditions conducive for phosphogenesis yielded few

successions with economic significance (Coles and Varga, 1988; Humphreys and Smith, 1989; Correia-

Junior et al., 2016; Silveira et al., 2016). Late Devonian phosphorite intervals, such as the Chattanooga

Shale in the eastern United States (southern Euramerica or Laurentia) and the Geirud Formation in Iran

and Armenia (northern Gondwana), that reach economic concentrations, are often minor in scale (<80 Mt

of resources) (Li and Schieber, 2015; Orris et al., 2015; Salama et al., 2018).

The Devonian evolution is marked by a transition from greenhouse (and probably sluggish

oceanic circulation) in early, middle and part of the late Devonian to icehouse conditions over the final

stage of the late Devonian that is in general alignment with models that approximate atmospheric CO2
63

concentrations that decreased from ca. 3300 ppm in the early Devonian to ca. 350 ppm in the late

Devonian (Royer, 2006; Algeo and Ingall, 2007). These climatic shifts likely influenced marine

concentrations of dissolved oxygen as a consequence of temperatures and air-sea exchange promoting

changes in ocean chemistry (Algeo and Ingall, 2007; Meyer and Kump, 2008). Global continental

weathering rates were potentially greater during greenhouse conditions during the early Devonian due to

higher temperatures, further accelerated by the evolution of land plants, intensifiying physical and

chemical weathering via root growth and organic acids and developing extensive, deeply weathered

mantles (Algeo et al., 1995). Under these conditions the delivery of nutrients (e.g., phosphorus) to the

oceans were enhanced while soil development may have increased the bioavailability of phosphorus

(Follmi, 1996; Algeo et al., 1995; Percival et al., 2019). The increase of nutrients enhanced productivity

and anoxia, with the recurrence of extended episodes of ocean anoxia recorded by black shale

sedimentation events (Becker et al., 2016). Strontium isotopic data values (which is a proxy to the waxing

and waning of Sr input from continental sources flux in contrast with the input from submarine

hydrothermal systems) indicates that the influx of nutrients in the Devonian resembles other periods of

economic phosphorite deposition probably indicating that reactive phosphorus input into the oceans was

similar (Veizer et al., 1999).

Devonian phosphorites occasionally display close associations with ironstones (Guerrak, 1988;

Andreeva and Chatalov, 2011; Oliveira and Pereira, 2011; Yilmaz et al., 2015; Denayer, 2016). As

biogeochemical cycles of P are often interconnected with Fe cycles (Young, 1989; Pufahl, 2010),

phosphatic ironstone petrology can be useful in paleoenvironmental reconstructions as their formation is

often indicative of variations in ocean chemistry such as iron availability, oceanic redox state, and

biological and pH conditions (Taylor and Macquaker, 2011; Poulton and Canfield, 2011). Brazilian

Paleozoic intracratonic basins (northern Gondwana) contain Devonian phosphorites and phosphatic

ironstones mainly within the Amazonas and Parnaíba basins but their economic potencial still needs to be

uncovered (Albuquerque et al., 1972; CPRM, 1978; Oliveira and Barros, 1976; Silveira et al., 2016,

Correia-Junior et al., 2016).


64

The presence of phosphate rich successions in the Devonian of the Parnaíba Basin were first

described by the São Miguel do Tapuio (Albuquerque et al., 1972; Oliveira and Barros, 1976), São

Nicolau (CPRM, 1978) and Phosphate Brazil Project (Silveira et al., 2016) developed in the eastern part of

the basin without a comprehension about the environmental and deposition controls. Ironstones in this

basin were also described in the researchs of Ribeiro and Dardenne (1978), Ribeiro (1984), further

petrographically and chemically characterized by Amaro et al. (2012) without an understanding of

stratigraphic constrains and phosphate relation. Abram and Holz (2020) detailed the environmental

controls of phosphorite and phosphate ironstone deposition in the Parnaíba Basin but still lacking a greater

detail of the chemical characteristics of the phosphogenetic process that can establish modes of iron and

phosphorus burial during the Devonian.

To examine phosphorus and iron burial in the Rheic Ocean during the Devonian, we investigated

Praguian to Givetian stage phosphorite and phosphatic ironstone-bearing successions from the Parnaíba

Basin within a regional stratigraphic framework using petrographic, mineralogical and geochemical

techniques to reconstruct the paleoceanographic conditions during their formation and the influence

diagenesis had on chemical signatures. Authigenic minerals are products of the environmental conditions

they formed in (Pufahl and Hiatt, 2012; Phan et al., 2019) but can be altered by diagenetic processes that

affect their mineralogy, stable isotopic composition, and REY patterns urging cauting when interpreting

primary porewater composition, microbial processes, and redox conditions from these geochemical data

(Mozley, 1989; Mozley and Wersin, 1992; Mozley and Carothers, 1992; Jarvis et al., 1994; Pufahl and

Hiatt, 2012). Construction of a paragenesis using petrographic techniques bolstered paleoenvironmental

interpretations by providing a basis to distinguish primary and secondary mineral phases and geochemical

signatures. The integration of these approaches allows for a holistic depiction of phosphorus burial in the

ferruginous Rheic Ocean during the Early to Middle Devonian.


65

2. Geological setting

The Parnaíba Basin includes the northwestern Gondwana margin, originally connected to the

adjacent Amazonas and Saltpond-Keta basins (Villeneuve, 2005; Menzies et al., 2018). The Parnaíba

basin has a subcircular shape with an areal extent of ca. 600,000 km2 and contains thick epicontinental

sequences dominated by siliciclastics, with a maximum thickness of ca. 3.5 km at the depocenter (Vaz et

al., 2007; Tribaldos and White, 2018). Prolonged (>200 Ma), low rate (<1 m Ma−1) subsidence related to

the stabilization of the South American Platform indicates an epicratonic stage for the Parnaíba Basin and

allowed the development of an epeiric sea that experienced depositional cycles from the late Ordovician to

the late Cretaceous (Araújo, 2015; Tribaldos and White, 2018; Menzies et al., 2018). Six megasequences

comprise the epicratonic evolution which are bounded by regional unconformities that were generated by

eustatic sealevel fluctuations and Laurasian isostatic cycle (Caputo and Santos, 2019; Vaz et al., 2007):

the pre-Silurian, the Silurian, the Devonian/Mississippian, the Permo/Triassic, the Jurassic and the

Cretaceous.

The Devonian to Mississippian megasequence is represented by the Canindé Group, which

hosts the phosphatic units studied in this work and is exposed in the eastern and southwestern regions of

the basin (Fig.1). According to Vaz et al. (2007)’s analysis, this megasequence can be traced throughout

the entirety of the Parnaíba Basin and unconformably overlies the Silurian Serra Grande Group. The

Canindé Group was deposited during a stage of greater subsidence (22m/Ma; Menzies et al. 2016; Castro

et al. 2016) in the Parnaíba Basin and is divided into five formations: Itaim, Pimenteira, Cabeças, Longá

and Poti (Vaz et al., 2007). In general, the Devonian interval of the basin is characterized by fluvial and

proximal to distal shelf facies (Itaim, Pimenteira and Longá formations) and a late Devonian glacial

deposit (Famennian Cabeças Formation; Caputo and Santos, 2019). Early to middle Devonian phosphatic

rich sections described here are located near the eastern and western borders of the Parnaíba and include

the Itaim and Pimenteira Formations which span the Praguian-Emsian to Frasnian stages (Loboziak et al.,
66

2000; Grahn et al., 2008; Breuer and Grahn, 2011; Trindade et al., 2015) that form the base of the Canindé

Group (Vaz et al., 2007) (Fig. 1).

The Itaim Formation was deposited during the Pragian to Emsian stages (Vaz et al., 2007;

Trindade et al., 2015) and have been interpreted as shelf and delta facies with structures indicative of tides

and storms (Carozzi et al.,1975; Góes and Feijó, 1994). Palynological results are suggestive of marine

environments with absolute phytoclast content indicating proximity to a fluvial-deltaic source (Trindade et

al., 2015). Stratigraphically above the Itaim Formation is the Pimenteira Formation which consists mainly

of grayish-black to greenish shales that are organic rich, partially bioturbated and subordinated sandstones,

likely deposited within a storm dominated epeiric sea (Della Fávera, 1990; Vaz et al., 2007). According to

Cunha (1986), the Pimenteira Formation presents the maximum thicknesses of 423 and 526 m along two

depositional axes that indicates a migration of the depocenter from east to northwestern from Eifelian to

Frasnian (Martins, 2019). Palynofacies data indicate that in the Givetian, this Pimenteira Formation was

represented by a mud-dominated oxic distal shelf and, in the Frasnian, a second order maximum flooding

surface was suggested reflecting a global sea level rise event (Rodrigues, 1995; Trindade et al., 2015;

Andrade, 2015). A radioactive interval is registered over 60 m with an average TOC content of 2 - 2.5%

(Rodrigues, 1995), and maximum of 6% (Araújo, 2015). According to Rodrigues (1995) and Melo

(2002), the highest TOCs are located at three intervals above the base of the Devonian: at 13 m during the

Eifelian/Givetian (387 Ma), 17 m during the Givetian (387-382 Ma) and 40 m during the Frasnian (382-

372 Ma). The organic matter (OM) of the Pimenteira Formation is immature and likely related to marine

organisms evidenced by the presence of acritarchs, Prasynophicae and Chitinozoa with variable

contributions of terrestrial organic matter (phytoclasts, pollen, spores, and cuticular tissues) (Amaral et al.,

2013; Martins, 2019).


67

Fig.1. Geological setting of the studied area (black rectangle close to the southpole). (A) Paleogeographic

reconstruction of Emsian (400 Ma) (Blakey, 2019) and hypothetical surface current circulation (after

Parrish, 1982; Markello et al., 2008). Cooler western currents (white arrows) and warmer northeastern

currents (yellow arrows). (B) Location of the Parnaíba Basin. (C, D) Location map of the wells used

during the research plotted on a geological map (modified from Schobbenhaus et al., 2004).

Recent work by Abram and Holz (2020) has described the stratigraphic framework upon which

the present study is based (Fig.2). Five fluvio-deltaic and 10 shallow marine lithofacies were described for
68

the early to middle Devonian studied sections. A simplified description of the depositional and diagenetic

features observed for these 10 shallow marine lithofacies where the phosphorite and phosphatic intervals

are located are presented in Table 1. Based on Abram and Holz (2020) four depositional sequences (B, C,

D and E), form the early to middle Devonian successions of the Canindé Group. The deposition of these

sequences likely reflects mainly shallow marine sediments with transgressive and regressive events

and smaller-scale sea-level fluctuations. The lowermost unit (sequence B), of Praguian to Emsian age

(Trindade et al., 2015), is composed of storm influenced shelf sediments and a prograding coastal (with

tide structures) and fluvio-deltaic succession (with a transgressive and highstand systems tract) and is

correlated with the Itaim Formation. It is unconformably overlain by sequence C with an estimated hiatus

of < 4 Ma (based on data from Grahn et al., 2008; Trindade et al., 2015; Menzies et al., 2018). Sequence C

was deposited during late Eifelian to Givetian stages (cf. Grahn et al., 2008; Breuer and Grahn, 2011)

containing mainly storm dominated shoreface and shelf sediments and is represented by an interpreted

thin lowstand systems tract (LST), in its lowermost section, drowned by the flooding of a transgressive

system tract (TST) marked by a thinning upward succession with minor sea-level fluctuations, which

reflect the alternation of more oxic and anoxic shelf conditions. According to Abram and Hoz (2020), this

TST is followed by a highstand systems tract (HST) characterized by shoreface-offshore transition zone,

lower shoreface and then upper shoreface zone on top. The sequence boundary separating the topmost

prograding succession of sequence C from the basal lower shoreface to the lower shoreface-offshore

deposits of sequence D was interpreted by these authors (op.cit.) as a minor base-level fall. Sequence D, of

early to late Givetian, has a highstand systems tract truncated by a basal surface of forced regression

(BSFR-D), with an upper shoreface/foreshore association and deltaic deposits (distal mouth bar

subsystem), interpreted as a falling stage systems tract (FSST). This drop of relative sea level marks the

onset of Sequence E.

All four sequences record phosphorite, phosphatic siliciclastic rocks, phosphatic ironstone and

organic rich sediments (TOC 1- 3%). These are related to offshore (outer shelf), offshore-lower shoreface

transition zone (inner shelf) to lower shoreface depositional environments that were recognized in the
69

transgressive and normal regressive successions (Figs. 2A, B). Some of the lithofacies originally described

in Abram and Holz (2020) are specially associated to phosphorites and phosphatic ironstones intervals and

related to this shallow marine system. These are summarized in Table 1 as follows: hummocky cross

stratified sandstones (Shcs), Skolithos ichnofacies bioturbated sandstones (Sb-Sk), Cruziana ichnofacies

bioturbated sandstones (Sb-Cz), organic matter rich sandstones (Sborg), ripple laminated sandstones (Sr),

bioturbated wavy heteroliths (Hewb), silty heteroliths (Hesty), bioturbated mudstones (Fb), gray to black

laminated mudstones (Fm), and massive ironstones to phosphatic ironstones (Ip) (Figs. 3, 4 and 5).

Phosphorite rich sections are characterized by bioturbated mudstones and bioturbated heteroliths

and deposited in the shelf in condensed zones near storm wave base (SWB), related to TST and early

HST. Phosphorites are associated with late Eifelian and Givetian maximum flooding surfaces in sequences

C and D (Abram and Holz, 2020). Ironstone and phosphatic ironstone are deposited in more proximal

settings, associated with wavy heteroliths, Cruziana ichnofacies and hummocky cross stratification

sandstones, near the fairweather wave base (FWWB) and are associated with early highstand systems tract

and also late lowstand systems tract, reflecting a slightly higher energy setting. These phosphatic

ironstones were recognized in Praguian, late Eifelian and Givetian intervals (sequences B, C and D).
70

Fig.2. (A) Base level curve; (B) Composite stratigraphic section of the studied interval (Itaim and
Pimenteira formations). Sequence A - Jaicós formation; sequence B - Itaim formation, sequences C to E -
(Pimenteira formation). LST - lowstand systems tract, TST - transgressive systems tract, HST-highstand
systems tract, FSST – falling stage systems tract. C = clay; S = silt; VFS= very fine-grained sand; FS=
fine-grained sand; MS= medium-grained sand; CS= coarse-grained sand. Ps = parasequences. SB –
Sequence Boundary; BSFR - Basal Surface of Forced Regression; CC - Correlative Conformity; MRS -
Maximum Regressive Surface; MFS - Maximum Flooding Surface; TSME - Transgressive Surface of
Marine Erosion. Adapted from Abram and Holz (2020), and lithofacies symbols as in this work.
71

Fig.3. (A) Progradational succession from the inner shelf (normal highstand regression of sequence D of
Abram and Holz, 2020), with phosphorites and high gamma-ray near the maximum flooding surface,
defining a condensed zone (MFS-D, Fig. 2); (B) Interbedded bioturbated heterolithic lithofacies (Hewb)
with hummocky cross stratified sandstone (Shcs) recovered by amalgamated Shcs in the lower
shoreface/offshore transition zone near the fair weather wave base (storm dominated shelf). (C) from left
to right: Sb (Sk) - Skolithos ichnofacies bioturbated sandstone, with Ophiomorpha ichnofossil (Ophi);
Shcs – Very fine-grained hummocky cross stratified sandstones, as proximal tempestites; Ss2 – very fine
grained structureless sandstones; Sb (Cz) - Bioturbated sandstones of the Cruziana ichnofacies (iron rich
phosphatic cement- Ip); (D) from left to right: Shcs – low thickness very fine-grained hummocky cross
stratified sandstones, exhibits Fugichinia ichnofossil, recovered by bioturbated mudstones; bioturbated of
a wavy heterolithic lithofacies (Hewb); Sborg/Ip – Organic rich sandstones and massive phosphatic
ironstone; Hewb - wavy heterolithic and bioturbated lithofacies; Hewb/Ip - wavy heterolithic and poor
bioturbated with thin laminations of phosphatic ironstone; Fm – laminated mudstone; Fb - bioturbated
mudstone with planolites tracefossils; Fb/Ip – Bioturbated (planolites tracefossils) and siderized mudstone
(Py – weathered pyrite spheres). All cores have a 5 cm diameter.
72

Table 1 – The shallow marine facies association related to phosphorites and phosphatic ironstone
successions (lithofacies description and diagenetic features summarized from Abram and Holz, 2020).
LITHOFACIES DESCRIPTION DIAGENETIC FEATURES
Facies Shcs – hummocky Very fine to fine-grained quartzose white sandstones with Sometimes with sideritic cement on top representing
cross stratified sandstones hummocky cross stratification marked by sharp lower contacts. a hardground. Above these hardgrounds, reworked
Variations from large amplitudes in proximal settings and isotropic ironstone grains are common.
small amplitudes in distal sections (Figs.3C, D).
Facies Sb (Sk) – Skolithos Fine-grained to medium-grained quartzose sandstones, moderately A siderite cement is not so common but can be
ichnofacies bioturbated sorted, rare discontinuous clay blades, with Skolithos ichnofacies distinguished mainly into bioturbated ichnotrams and
sandstone (Fig. 3C). also scattered at the top of the sets.

Facies Sb (Cz) – Cruziana Very fine to fine-grained and muddy sandstones, with prevailing These sandstones commonly host an iron and
ichnofacies bioturbated Cruziana ichnofacies, and bioturbation index from 3 to 5 (Droser phosphorus cement with 1% to 9.2% P2O5 (Figs.3 and
sandstones and Bottjer, 1993). 5). Green and black iron silicate coated grains are
present, sometimes stained by a red color associated
with later siderite cementation.
Facies Sborg – Organic Black very fine to fine-grained sandstones. It can be intercalated In some samples siderite is also present. Framboidal
matter rich sandstone with hummocky cross stratification or ripple sandstones. It exhibits pyrite crystals present a diversified size, dominating
a high content of quartz (55%), authigenic smectite/ironsilicate mean diameter from 6 μm–10 μm. Some pyrite
(35%), pyrite (7% - 20%), organic matter (3%) and rare detrital crystals present cubic shape. This lithofacies also
feldspars. Dark grey to black color reflects high OM content and includes the presence of gradual and sharp contact
preservation. Prasynophicae algae (like Tasmanite) is present and intercalated ironstones and phosphatic ironstones on
partially to totally degraded to pyrite, preserving wall cells. top and bottom (0.2 to 3.77% P2O5) (Fig.3D).
Trindade and Carvalho (2018) defined the presence of terrestrial
and marine palynofacies.
Facies Sr - Ripple Very fine-grained white quartzose sandstones, composed of
laminated sandstones abundant quartz with minimal potassium and plagioclase feldspar,
< 20% of interbedded mud, with ripple and parallel laminated
bedding. The bedding is thin, often not more than 3 cm thick.
Facies Hewb – This lithofacies is characterized by smectite rich mudstone and 1 to Phosphorite nodules and concretions are mostly
Bioturbated wavy 3 cm thick fine-grained to very fine-grained quartz sandstones or associated with the interface of these heteroliths with
heteroliths gray siltstones with small and symmetrical oscillation wave-ripple mudstones in an offshore environment (Fig.3D). The
structures (Fig. 3D). In general, it is bioturbated with a great phosphorite nodules (5 cm in Some phosphatic
diversity of trace fossil styles that characterize Cruziana ichnofacies ironstones (1 to 10% P2O5) are also associated with
(Young, 2006; Abram and Holz, 2020). this lithofacies, located at the interface between these
and hummocky and bioturbated sandstones in a
shallower position.
Facies He silty – Silty These are normal graded bedded siltstones interbedded with < 2 cm Reddish, mainly sideritic, iron cement is common
heteroliths thick sandstones (Fig. 3D). It exhibits a low index of bioturbation along the silty laminaes.
(ii2), with scarce Planolites trace fossils. This facies shows thin
bedding with parallel laminae or discrete, very small wave ripples.
Facies Fb - Bioturbated This lithofacies is generally divisible into black mudstones and Phosphorite, phosphatic and sideritic concretions are
mudstone green to gray silty mudstones and clay siltstones with lenticular present. They are mainly discrete or can evidence
very fine-grained sandstone (linsen beddings). The clay is a coalescing features (Fig.4H) and generate a deflection
smectite/nontronite, sometimes associated with illite. The dominant of the mudstones around. Traces of framboidal pyrite
bioturbation pattern is horizontal mainly represented by Planolites replaced by goethite are observed (size of 8 to 30
trace fossil. μm).
Facies Fm - Gray to black This lithofacies is represented by dark-gray-to-reddish laminated Small framboydal (<0.02 mm) or large grains of
laminated mudstone mudstones or muddy siltstones (Fig.3D). It exhibits poor or no radial pyrite overgrowths (8 mm) are common.
bioturbation features. Smectite is the main clay mineral. Silty sized Sometimes they present a reddish color related to
quartz (15%) and detrital muscovite (10%) are also present. It can iron/phosphate cementation. In this case, siderite is
be intercalated with lenticular (linsen structure) very fine-grained enveloping microorganisms such as Acritarchs. The
quartz rich sandstones. reddish color is occasionaly associated with clays.
Facies Ip - Phosphatic/ Very fine to fine-grained massive lithofacies (Fig.5). It can be Fe>18% and also P2O5 between 1-16% when
Non phosphatic ironstone associated to sandstones, organic rich sandstones or heterolithic phosphatic. Odinite, chamosite/clinochlore, francolite
lithofacies. (and possibly greenalite) are present in coated detrital
quartz (12%) or as coated grains (15 - 27%). These
are mainly hybrid and inserted in a matrix with
authigenic chamosite (or clinochlore) (24%), Mg
siderite/siderite (23%), pyrite framboids (2%) and
partially degraded OM (1-2%).
73

3. Methods

The sequence stratigraphic framework used in this study was that presented in Abram and Holz

(2020). It was constructed using 26 outcrop sections and 22 drill cores located at the eastern and western

border of the basin by the Geological Survey of Brazil-CPRM and by the Federal University of Rio de

Janeiro (UFRJ), Conexão Project oil bias (PROFEX/CENPES/PETROBRAS). A total of 202

representative core samples were obtained from one well in the western part of the basin (ARN-1-TO), 9

wells in the eastern part (PM-07-PI, PM-06-PI, PM-08-PI, PM-10-PI, PM-11-PI, PM-21-PI, ST-04-PI,

ST-15-PI, VL-03-PI) and 46 samples from strategically chosen outcrops. Modal compositions were

determined for 60 uncovered thin sections using an abundance classification system of rare (1-5%),

uncommon (6-25%), common (26-50%) and abundant (>50%) for mineral and grain constituents. Thin

sections were analyzed with a Nikon OPTIPHOT-POL petrographic microscope located at Acadia

University in Wolfville, Nova Scotia. Petrographic studies were combined with powder X-ray

diffractometry (XRD; 104 samples) and scanning electron microscopy (SEM; 18 thin sections) to

determine mineralogies and petrographic relationships, respectively. XRD was conducted with an

Empyrean Philips powder diffractometer (PANalyticalBV) at Acadia University and an X´PERT PRO

MPD (PANalytical), PW3050/60 (Theta/Theta) goniometer at CPRM laboratory, Brazil, both using a Cu

target and from 5° to 70° 2-theta. SEM analyses were carried out with a JEOL/EO (v1.1) JSM 5000

located at Acadia University with a back-scattered electron (BSE) detector, working distance of 11-15 mm

and 10 kV of accelerating voltage. Elemental analyses were performed with an EDS (energy dispersive

spectrometry) IMIX X-ray Microanalysis System from Princeton Gamma-Tech, Inc. Whole rock

geochemical analysis was conducted at the certified SGSGeosol laboratory in Brazil. Major elements were

determined by X-ray fluorescence (XRF) and trace elements by inductively-coupled plasma optical

emission spectrometry (ICP-OES) and inductively-coupled plasma mass spectrometry (ICP-MS) after

lithium metaborate fusion. Total organic carbon (TOC) and total sulfur were determined using a LECO

combustion-spectrometric device.
74

Francolite and siderite grains (69 samples) were isolated for mineral geochemistry by first

disaggregating the phosphorite/phosphatic and ironstone rock and crushing the resultant material into

analysing size fractions without purification processes. δ13C and δ18O were determined for francolite

(δ13CCO3 - carbonate site) and siderite at the Queen's Facility for Isotope Research Laboratory in Kingston,

Ontario, with accuracy reported in permil notation (‰) and isotope accuracy for δ13C at 0.1‰ and δ18O at

0.5‰ standard deviation. Replicate analyses indicate a reproducibility of ±0. l‰ for δ13C and δ18O. The

isotope analyses were conducted by reacting approximately 10 mg of powdered material with 100%

anhydrous phosphoric acid at 72°C for 4 hours. The CO2 released was analyzed using a Thermo-Finnigan

Gas Bench coupled to a Thermo-Finnigan DeltaPlus XP Continuous-Flow Isotope-Ratio Mass

Spectrometer. The δ13C and δ18O values are reported using the delta (δ) notation in permil (‰), relative to

Vienna Pee Dee Belemnite (VPDB) standard, with precisions of 0.2‰. These same 69 samples were also

analyzed for REE by lithium metaborate fusion and ICP-MS in the ALS Canada Ltd laboratory in

Vancouver Canada.

All these data were integrated with known sedimentological and sequence stratigraphic

framework to develop a depositional and oceanographic model for phosphogenesis in the Parnaíba Basin

during the Devonian.

4. Phosphorites and phosphatic ironstones

4.1. Mode of occurence

Phosphorites were mainly identified in outcrops in the sequences C, D and E. They are associated

with maximum flooding surfaces and condensed zones of late Eifelian (MFS-C) and Givetian (MFS-D)

successions or in flooding events interpreted by Abram and Holz (2020) (Figs.3, 5). The phosphorites

occur as decimetrics to centimetrics concretions, discrete phosphatic layers, phosphatic nodules,

hardgrounds, authigenic and reworked peloids, near storm wave base into the shelf (4 to 28% P2O5)
75

(Figs.6, 7, 8). They are mainly related to bioturbated mudstones (in intervals close to 1.5 m), containing

phosphorite concretions, or as thin layers or nodules (5 to 10 cm width) associated with bioturbated

heteroliths (Fig.4). The containing phosphorite/phosphatic layers can be metric-scale continuous or with

some lateral discontinuity along outcrops. The thin layers and nodules are also associated with flooding

surfaces in the sequences C, D and E.

Phosphatic ironstones were identified in: (1) sequence B, mainly in the early highstand system

tracts (condensed zones), although also present in flooding surfaces at the upper limit of parasequences on

late highstand system tracts; (2) sequence C, on drowned lowstand system tracts and mainly at the early

highstand system tracts; (3) sequence D, on transgressive system tracts and early highstand system tracts

and; (4) on transgressive surfaces (early TST) at the basal part of sequence E. In most cases, phosphatic

ironstones are related to the lower shoreface-offshore transition zone, near fairweather wave base. These

intermittent layers of phosphatic ironstones sum up to a thickness of more than 6 m.


76

Fig.4. (A) Weathered mudstones (minor heteroliths) and pristine phosphorites (on the lower part),
interpreted as a condensed section, recovered by reworked phosphorites and very fine-grained hummocky
cross stratified sandstones; (B) a detail of the section on A, with granular phosphorite up and pristine
phosphorite mudstone down; (C) a condensed zone with high radiometric (1282 cps) phosphatic mudstone
77

layers; (D) a detail of the phosphatic mudstone layers, iron rich blades interspersed; (E) a condensed zone
with bioturbated and laminated mudstone with siderite and phosphorite disk shaped concretions associated
to the MFS-D; (F) a detail of the phosphorite concretion; (G) nodular phosphorite level associated to
bioturbated heteroliths, with bioturbated thin sand wavy ripples interbedded with mudstones; (H)
Phosphate and siderite rich layer, as coalescing concretions in the condensed zone associated to mudstone
lithofacies Glauconite and francolite mineral associations reflect suboxic conditions. In concretionary
phosphorites, francolite and siderite are the main associations reflecting anoxic conditions. Lithofacies
symbols as described in figure 4. Some photos were obtained from Abram and Holz (2020).

Fig.5. (A) An interval of the sequence B comprising ironstones and phosphatic ironstones associated to a
prograding section immediately above the MFS-B (ps3b on figure 2) interpreted as a lower shoreface-
offshore transition zone environment. It is composed by minor fluctuations of thin mudstones beds (Fb),
78

bioturbated wavy heteroliths with iron silicate and siderite cement, to Cruziana ichnofacies well
bioturbated sandstones (Sb muddy), with organic rich intervals and phosphatic ironstone (Ip) associated.
(B) In detail, one drill core showing the bioturbated ironstone. (C) Ironstone sample with burrows filled by
siderite (sd) on left and ironstone sample with siderite cement being burrowed on the right side. (D) Near
the FWWB, very-fine grained hummocky sandstones (Shcs) recovered by bioturbated sandstones and
phosphatic ironstone reflecting lowering sedimentation rate. P2O5 percentage in white color. Modified
from Abram and Holz (2020).

4. 2. Petrographical aspects and diagenetic constituents

Phosphorite and phosphatic concretions

The dominant phosphorite occurrence (at about 70%) is in the form of concretions making layers

intercalated with siderite concretions enriched layers in offshore (outer to inner shelf) environment

(Abram and Holz, 2020). It is associated with the interface between bioturbated mudstones (Fb) and

heteroliths (Hewb). The phosphorite (>18% P2O5) or phosphate rich concretions (1 to 16% P2O5) are

muddy, discoidal to oblate, 20 to 30 cm in length, 5 to 10 cm in width, and sometimes are encapsulated by

thin siderite band. The concretions are hosted within phosphorus-rich mudstones or siltstones. Most of the

concretions are discrete or show evidence of coalescing features. The phosphorite concretions are

composed of pristine francolite (carbonate-fluorapatite) as the dominant phase followed by siderite

spherulites, less than 5% of quartz (silty sized) and muscovite as detrital components, weathered

framboydal pyrite and relicts of organic matter (Figs. 6A, B, C and D). Framboidal pyrite replaced infills

of Tasmanites (Fig.6H). It can present outer marcasite rims on framboidal pyrites infilling partially

degraded Tasmanites. In hosting bioturbated mudstones pyrite can also occur as macrocrystalline radiating

spheres. Sepiolite (one kind of authigenic smectite) was also detected by XRD in some phosphorites.

Organisms’ cells can be also infilled by siderite. Kaolinite, goethite, hematite are common products of

weathering.
79

The siderite that occurs associated to phosphorite or siderite concretions, or within bioturbated

mudstones (Fb) is here considered the type 1 siderite. In phoshorite concretions it occurs as a poor

crystalline phase formed after francolite precipitation (Fig. 6C, D). In phosphorites, it constitutes a

microcrystalline, siderite filling francolite or nucleus of degraded organic matter.

Phosphorite and Phosphatic nodules

The phosphorite and phosphatic nodules form 5 cm thick strata in the heteroliths (Hewb),

associated with an inner shelf environment (Abram and Holz, 2020). They are pebble-sized (5 cm in

length) exhibiting irregular banding of pristine authigenic francolite developed in different stages with

outer rims containing partially degraded OM (possibly indicating microbial layering) and authigenic

glauconite (Fig. 7B). In some cases, the nodules are bioturbated or even reworked. They may also contain

pyrite (not common), and an iron rich porous cement (due to weathering) with detrital quartz as matrix

(Fig.7C). Francolite nodules include marine palynomorphs, like Prasynophicae algae (some recognized as

Tasmanites, Fig.7D), Bryozoan and mollusk fragments and other organisms such as radiolarians (with no

delicate features, possibly due to increased water circulation pre-phosphorite precipitation) and a diversity

of Acritarchs not individually distinguished (Fig.7C). Tasmanites exhibit preserved wall cells cysts.

Kaolinite, wavellite, and goethite are products of partial weathering. Some nodules show late depositional

burrows or reworking features.

Phosphorite laminae

Phosphorite or phosphate rich layers (1.5-10 cm thick) occur associated with offshore mudstones

and bioturbated heteroliths (Fig. 6). The laminae are composed of francolite, siderite, silt-sized detrital

quartz, feldspar, muscovite. Kaolinite, goethite and aluminum-phosphate are products of weathering.

These layers were defined as stratigraphically related to condensed zones or flooding surfaces in Abram

and Holz (2020).

Phosphorite hardgrounds

These are granular phosphorites (phosphoclast grainstone) represented by detrital and bioclasts

cemented by different francolite generation as a cryptocrystalline mass (Figs. 4, 6). It contains detrital
80

grains of quartz, francolite-coated grains, bioclasts (e.g., bryozoa), micas and lithic fragments of

phosphorite and iron cemented sandstones. Goethite is present as a phase related to weathering. Partially

degraded Tasmanites and possibly Acritarchs are also present. It is related to condensed zones (Abram and

Holz, 2020) and defined as hardgrounds.

Phosphatic ironstones and ironstones

These phosphate ironstones are massive rocks or can occur associated with bioturbated sandstones

(Sb), heteroliths (Hewb) with Cruziana ichnofacies or even to organic rich sandtsones (Sborg) (Table 1). It

is represented by mainly authigenic siderite-rich cement and green and black ironsilicate/francolite coated

grains (1.1 to 9.5 % P2O5). Ironsilicate coated grains with or without detrital quartz-centered grains present

elongate or irregular shapes and hybrid types of coating grains, showing regular coatings or unconformity

bounded (discontinuous coatings) grains, distinguishing respectively minerals formed in situ from those

that evoluted by storm reworking exhumation and abrasion (Pufahl and Grimm, 2003; Dunn, 2020).

Coated grains in ironstones and phosphatic ironstones (0.1-0.5 mm length) exhibit odinite (Fe3+

and Mg richer phase) as an earlier phase in inner parts or as nuclei, but overall, it has an uncommon

abundance (Fig.8). Greenalite was determined by X-ray powder diffraction in one sample but not yet

confirmed by SEM or petrography. Odinite and greenalite are minerals from the serpentine group and

verdine facies. The coated grains are 0.1 to 0.5 mm length. Odinite is followed by authigenic rims or

cortical layers of magnesium and Fe+2 rich chlorite as clinochlore (Mg5Al (AlSi3O10)(OH)8) or chamosite

((Fe2+, Mg, Al, Fe3+)6 (Si, Al)4 (O)10(OH, O)8), as solid solution end members, and francolite mainly in

inner parts (Fig.8). Clinochlore and chamosite are more frequent early authigenic minerals forming the

coated grains and can also present francolite cortical layers. Clinochlore/chamosite (uncommon-to-

common phases) are also present in the groundmass after pyrite and Mg siderite precipitation.

Here siderites were recognized with two different associations and crystallinities. A type 2 siderite

association that are small crystalline spherulites (organogenic), mainly with a Mg richer siderite

composition, associated with ironstone and phosphatic ironstone in the lower shoreface-offshore transition

zone (Figs.8E, G, H). In this case, it shows evidence of precipitation around degraded organic matter
81

partially replaced by framboydal pyrites. These Mg siderites are uncommon-to-common early authigenic

cements precipitated in a step after the degradation of organic matter and pyrite (Mazzullo, 2000). In its

turn, a third type of siderite (type 3) is defined as an authigenic siderite with almost pure FeCO3 that occur

as microspherulites engulfing and replacing other Mg siderites and coated grains. Sometimes this type 3

occurs filling bioturbated trace fossils or is being overprinted by later non-cemented trace fossils (Figs.8C,

D).

The Sborg also associated to this lithofacies show an authigenic-to-shallow burial phase,

represented by a nontronite smectite phase. This smectite occurs as clumpy or dense masses filling

intergranular spaces in Sborg sandstones (Fig. 8F). This is recognized as a marine smectite with higher Fe-

Mg contents and intermediate compositions between the saponite and nontronite fields (Iacoviello et al.,

2012).

Pyrite (rare to uncommon) occurs as very fine-grained framboids replacing organic matter. It can

also occur in associated Hewb (less common) and is more common in organic rich sandstones (Sborg).

Framboidal pyrite replaced infills of Tasmanites or other not recognized palynomorphs, preserving the

organic walls (Figs. 8F, H). Outer marcasite rims on framboidal pyrites infilling partially degraded

Tasmanites are also present but rare.


82
83

Fig.6. (A) Francolite, brown siderite and silty sized quartz in pristine phosphorite mudstone concretion

(PPL) (from Abram and Holz, 2020); (B) quartz, detrital mica (muscovite), some replaced by

siderite+goethite bounded by francolite cement; (C) SEM image of phosphorite concretion with a

pervasive francolite cement and spherulitic siderite replacing very fine-grained organic palynofacies; (D) a

detail of the prior slide with siderite, in some cases surrounding inner partially degraded organic matter;

(E) Granular phosphorite, francolite coated and quartz grains cemented by francolite (PPL) (from Abram

and Holz, 2020); (F) Pellets of authigenic francolite in a granular phosphorite with a quartz and francolite

rich cement (PPL); (G) Backscattered SEM image showing Bryozoa and quartz clasts cemented by

different francolite (Fr2 and Fr3) generation; (H) EDS elemental map (P, Ca and C SEM image, higher

values indicated by bright yellow collor) and backscattered showing partially degraded Prasynophicae

algae filled by pyrite and partially replaced by marcassite in phosphorite mudstone with authigenic

francolite. Symbols: Glau – glauconite, Fr-francolite, Kao-kaolinite, Qz-quartz, Pr algae-Prasinophicae

algae, Sd-siderite, OM – Organic matter. (BEC - Backscattered; BES - Secondary electrons).

4.3. Diagenetic Interpretation

Francolite in phosphorites and phosphatic ironstones precipitated at a very early stage of

diagenesis during periods of non-sedimentation.

For the phosphorite concretions, a slow rate of sedimentation is suggested by coalesced

concretions that are indicative of a sulfate-reducing bacteria zone (Morad et al., 2012). On the other hand,

discrete (mainly sideritic) concretions possibly are related to a slightly more rapid depletion and

exhaustion of oxidants, reflecting possible microenvironments of the interface sulfate-reducing bacteria

zone and methanogenic production (Bojanowski and Clarkson, 2012; Morad et al., 2012). In phosphorites,

partially degraded Tasmanites often occur with destroyed walls and nuclei replaced by siderite. Under this

condition this is also interpreted as a result of degradation under evoluting anoxic zones. Francolite

possibly precipitated in higher productivity areas that created an oxygen minimum zone generated by the

microbial degradation. This can reflect an upwelling environment affected by fluctuations or ventilation
84

processes that could favour more anoxic or oxic waters, as new colonization of organisms and

bioturbation are commonly observed, leading to stabilizing redox conditions. Francolite and glauconite in

phosphorite nodules precipitated within suboxic sediments, sometimes related to the partial microbial

reduction of Fe-(oxyhydr) oxides at the sediment-water interface on the shelf (Glenn and Arthur, 1988).

The preserved palynomorphs inside the nodules resulted from francolite growing cement that engulfed the

organisms and probably avoided the biological destruction. As it requires low sedimentation rates, Abram

and Holz (2020) interpreted glauconite-bearing phosphorites as having formed close to an upwelling front

where microbial respiration of sedimentary organic matter fueled the production of suboxic pore waters

(Froelich et al., 1979; Taylor and Curtis, 1995), favoring the precipitation of francolite and glauconite. So,

bathymetric variations with different oxygenation conditions were suggested for the different associations

observed: (1) phosphorite nodules (associated mainly with bioturbated heterolithic lithofacies) with

francolite-glauconite associated with suboxic conditions near SWB and (2) the concretions (associated

with bioturbated mudstones) with francolite –siderite reflecting evoluting more anoxic conditions in

deeper sections.

For the phosphatic ironstones, the presence of redox-agradded and unconformity bounded coated

grains of ironsilicate/francolite suggest long residence times and episodes of precipitation and reworking

similar to francolite redox aggraded mechanism, confirming an eodiagenetic stage for these grains but a

more instable condition (Pufahl and Grimm, 2003). The cortical layers suggest that these coated grains

record the accumulation of a condensed bed (Pufahl and Grimm, 2003; Squires, 2019). The sediment-

water interface was first oxygenated (considering the primary bioturbation) and affected by fluctuations in

the rate of organic carbon accumulation. Bioturbation may have altered the flux of elements and sea water

surface interaction, increasing chemical reactivity. In sites with slow sedimentation rates the burrows

acted as nuclei for early diagenesis mineralization. The degradation of OM in organic-rich bioturbated

lithofacies favoured the creation of a reduced porewater leading to iron silicate and carbonate precipitation

(Dunn, 2020) and increasing alkalinity. Authigenic ironsilicate formed as a result of saturation of

porewater with both ferrous and ferric iron available (Taylor and Macquaker, 2011). With the continued
85

degradation of OM (mainly marine but also some terrigenous contribution) and stabilization of the

degradation zone, seawater anoxia increasingly favoured redox-aggraded grain coatings near the

sediment-water interface, modifying the redox interface. The comsumption of oxidants was progressively

increased generating more anoxic conditions probably under sulfate-reduction zones.

Odinite precipitated as an early diagenetic phase in these phosphatic ironstones. We consider that

this phase represents a metastable mineral that probably formed very quickly under initial conditions and

was the precursor to authigenic and grain-coating chlorites (Odin and Sen Gupta, 1988; Kronen and

Glenn, 2000; Harding et al., 2014). This precipitation probably occurred in suboxic pore waters because it

is a ferric iron rich phase (Odin and Sen Gupta, 1988). In other basins, odinite was found between depths

of 15 and 200 m associated with moderately condensed intervals (Odin, 1988; Kronen and Glenn, 2000),

as observed here. Greenalite, like odinite, has been interpreted as authigenic or early diagenetic in other

successions in the world (Pecoits et al., 2009; Tosca et al., 2015). This phase is known primarily from

Precambrian rocks (Fischer and Knoll, 2009), although recent sediments in New Caledonia lagoon

(Merrot et al., 2019) also contain greenalite. This suggests that this phase, as odinite, can represent a

metastable phase precursor to authigenic chlorites. In Precambrian rocks, Tosca et al. (2015) and Johnson

et al. (2018) recognized that it forms from reactions between dissolved Fe2+ or Fe (OH)3 and dissolved or

amorphous silica. It normally nucleates from waters with a pH as high as 7.7– 8.3, implicating also

alkalinity (Tosca et al., 2015). Higher rates of organic matter production at the inner shelf could also have

led to a rise in oxygen demand, resulting in the upwards migration of the redox boundary, favouring

shallower authigenic precipitation of reduced phases.

Other ferruginous authigenic minerals in phosphatic ironstones, like chamosite, may also have

precipitated under oxygen deficient conditions. This suggests that Fe2+ could have been supplied to the

inner shelf through upwelling Fe-rich waters. In its turn, when the concentration of iron was not high and

Mg concentrations and alkalinity increased, clinochlore precipitation was probably favoured (Squires,

2019).
86

Siderite precipitates in continental or marine waters by abiotic or biotic processes, although under

very limited conditions: Low Eh, low ∑S=, high Pco (high dissolved inorganic carbon concentration), high

pH, and high Fe2+ concentration (Zymela, 1996). Microbial methanogenic production or Fe-reducing

bacteria are the common biotic processes that form siderite (Mozley and Carothers, 1992; Zymela, 1996).

However, siderite can also be formed directly under sulfate-reducing bacteria via an enzymatic

mechanism as observed in the experiments of Coleman et al. (1993) or limited to zones where the

diffusion of sulfur is not so efficient (Zymela, 1996). The massive siderite concretions denote prevailing

ferruginous conditions, a decrease in sedimentation rate, and also the redox boundary between

oxic/bioturbated and anoxic/non-bioturbated zones (Majewski, 2000; Witkowska, 2012). The common

occurrence of siderite precipitated around degraded organic matter (type 2 siderite) suggests that a

mediated bacteria origin was an important factor in this siderite precipitation. We interpreted that under

sulfate-reduction zones and poor H2S diffusion or near the interface of sulfate-reduction to methanogenic

production, increasing OM degradation and the consumption of H2S with the formation of pyrite probably

conducted to more anoxic conditions and the precipitation of siderite.

On the other hand, in these phosphatic ironstones, siderite as a later common phase (type 3)

following early diagenetic odinite, clinochlore/chamosite, and pyrite can be probably considered to be

associated with an interface of layered microbial communities under iron reduction zones, in a later

resumption to more oxic porewater conditions.

In normal (non-euxinic) terrigenous and anoxic marine sediments, organic matter is the major

control for the formation of pyrite (Berner, 1984; Zymela, 1996). In anoxic, non-euxinic environments,

iron can migrate through the sediment, permitting localized pyrite precipitation and normally produces

pyrite aggregates. Based on classification of pyrite framboids size (Bond et al., 2004; Bond and Wignall,

2010; Liao, 2010), we suggest environments of weakly oxygenated water condition, near the interface

dysoxic/anoxic where the variation of the redox interface was located just below the sediment–water

interface. These aggregates were interpreted as being the result of microbial reduction by sulfate-reducing

bacteria with a source of iron and dissolved sulfate in marine porewaters (Coleman et al., 1993; Schieber
87

and Baird, 2001). So, partially degraded organic matter infilled by pyrite framboids indicate an early

diagenetic phase under an anoxic sulfate reduction zone. However, intermittent reoxidation of earlier

formed iron sulfides with marcasite formation and later bioturbation on ironstone indicates that

reoxygenation occurred probably related to dynamic ocean circulation (cf. Schieber, 2011). A longer

growing time allowed the formation of also larger pyrite framboids. Lower abundance of cubic pyrite

crystals coexisting with framboidal grains just observed in the Sborg lithofacies indicates evolving

shallow burial diagenesis.

4.4. Weathering conditions

Weathering is a common to uncommon characteristic for all lithofacies and was observed in

outcrop and in drill core within the first 10 meters below the surface, although rare to uncommon

weathering features were also recognized at a depth of 30-70 meters. Alteration due to weathering makes

the interpretation of geochemical data more difficult. Kaolinite and goethite are the main weathering

products recorded in most lithofacies. Hydroxylapatite, wavellite (Al3(PO4)2(OH, F)3·5H2O), florencite

(CeAl3(PO4)2(OH)6) and kaolinite (in intense leaching zones) are common weathering products of

francolite. Fluorite and analcime constitute rare minerals in phosphorus and iron rich rocks. Analcime

entirely replaces coated grains. Vivianite (Fe2+3(PO4)2·8H2O) was documented in phosphatic ironstones, as

well as rare calcite and chlorites, that replace coated grains. Iron bearing primary minerals (pyrite, iron

rich silicates, and carbonates dissolution and/or oxidation) may alter to goethite (dominant) and hematite.
88
89

Fig.7. (A) Different generation of authigenic francolite nodules in a matrix of quartz, siderite and goethite

(as weathering product), PPL; (B) Authigenic francolite nodule rimmed by authigenic glauconite (green

color), and a burrow with another phase of francolite infill (PPL); (C) Acritarchs included in authigenic

francolite nodules (PPL); (D) Prasynophicae algae included in francolite nodule; (E) Francolite nodule

showing an outer rim of francolite plus glauconite and detrital quartz in the matrix; (F) X-Ray

diffractogram of a nodular phosphorite associated to heterolithe showing the presence of francolite and

glauconite, suggesting suboxic conditions; (G) SEM images (EDS) showing the elemental map for P, Ca,

Al, Si, Fe and K (K-electron) of the marked view on E (francolite nodule rimmed by glauconite and

francolite rich layer). In concretionary phosphorites, francolite and siderite are the main associations

reflecting anoxic conditions. Symbols: Glau – glauconite, Fr-francolite, Kao-kaolinite, Qz-quartz, Pr

algae-Prasinophicae algae, Sd-siderite, OM – Organic matter.

5. Geochemistry

This chapter links the petrographic characteristics of phosphorites, phosphatic rocks, ironstones

and associated host rocks with whole rock, mineral and isotope geochemistry to understand the processes

that governed their formation. The results permit the interpretation of pore water chemistry at the time of

deposition and possibly bottom water conditions for the studied sections.
90
91

Fig.8. (A) Authigenic odinite coated grain interlayered with francolite and chamosite, centered by detrital
quartz, chamositic and minor Mg siderite cement (PPL); (B) Partially degraded organic matter (sometimes
replaced by pyrite) bordered by post organogenic Mg siderite immersed in a clinochlore/chamosite cement
(PPL); (C) authigenic coated grains (some centered by quartz) of interlayered odinite/francolite and
chamosite in a chamositic and Mg siderite (left side) and an authigenic posterior Fe sideritic cement (right
side) (PPL); (D) SEM image of iron silicate coated grains (right side) and a siderite rich matrix (left side);
(E) SEM image of a coated grain highlighting the coatings of odinite, francolite, clinochlore and organic
matter; (F) SEM image of a Prasynophicae algae (Tasmanites) filled by pyrite with preserved walls in
organic rich sandstone (Sborg) with authigenic smectite; (G) Unconformity bounded chamosite/francolite
coatings nucleated around quartz grain, with organic matter infilling coating walls and Mg siderite crystals
forming rhombohedral outer rims in organic matter centers; (H) SEM image (BEC – Backscattered) of Mg
rich siderite surrounding organic palynomorphs, framboydal pyrites and a Prasynophicae algae with
preserved walls and filled by pyrite; (I) Ironstones or iron-rich-cement sandstones X-Ray diffractograms
showing the main identified assemblages. Odinite recognized by a sharp and asymmetric peak near 7 Å
and a wide near 15 Å (Velde, 2005). Symbols: Mu-muscovite, Sm-smectite, Py-Pyrite, Odi-Odinite, Gre-
Greenalilte, Sd-siderite, Mg-Sd – Mg siderite, Qz-quartz, Ch-chamosite, Cl-Clinochlore, OM-Organic
matter, Pr algae–Prasynophicae algae, c.g.-coated grain Fr-francolite, Mg Sd–magnesium siderite, Goe–
goethite, Py–pyrite, ill- detrital illite, Al PO4–aluminophosphate (BEC - Backscattered; SEI - Secondary
electrons).

5.1 Geochemical associations

Whole rock geochemical analysis for major and trace elements including rare earth elements, total

organic carbon (TOC) and total sulphur (TS) were conducted on 80 selected most fresh samples

representing (i) phosphorite; (ii) phosphatic mudstones/heterolithe; (iii) siderite/phosphorite concretions;

(iv) ironstone; (v) phosphatic ironstone; (vi) siderite concretions and also associated host rocks like (vii)

mudstones and (viii) heterolithes. Results are available on Tables 2, 3 and 4 for major, trace and rare earth

elements respectively.

The geochemical results for 80 samples were treated using principal component analysis (PCA), a

multivariate statistical technique designed to extract a limited number of linear functions from the original

variables that can capture the essential variability of the data set (Jackson, 1991). The selected
92

geochemical variables include major oxides plus LOI and 13 trace elements plus REE, TOC and TS. The

PCA results are shown in “biplots” that display analysis of geochemical variables (R-mode) coupled with

analysis of sample types (Q-mode) (Fig. 9). Geochemical variables are plotted as vectors and sample type

variables as points. The length of the vectors is proportional to the variability of that element along the

principal component axis and the proximity between the vectors is related to the correlation between the

elements. Finally, the direction of the vectors to the sample points indicate how heavily a given sample

type is influenced by that element. The “biplot” of Figures 10 A and B shows PCA of the major oxides

and trace elements (plus Fe2O3, Al2O3 and P2O5 for comparison) respectivelly. The PCA of the major

oxides shows that the first two components account for nearly 40 % of the data variability whereas the

PCA for trace elements the first two components account for nearly 70 % of the data variability.

There are 3 main groups of major oxide variables detected (Fig. 9A). One includes Fe2O3, MgO,

MnO and LOI with negative loadings on PC1 and is more closely associated with ironstone and

phosphatic ironstone samples. The second group shows a distinct close association between P2O5 and CaO

and is characterized by strongly negative loadings on PC2 that clearly indicates phosphorite samples. The

third group is defined by Al2O3, TiO2, SiO2 and K2O with positive loadings on PC1. There is a mixture of

lithologies associated to this group includings ironstones, siderite concretions, phosphatic mudstones and

siderite/phosphatic concretions but the latter two types are also influenced by the the P2O5-CaO groups

reflecting their phosphate component. As for the trace elements it is observed the Fe group is followed by

V, Cr, TOC and secondarily by Mo and TS and Co, Ni. The Al group is related to Zr and the phosphate

group is associated with Y, U, Sr, REE (Fig.9B).


93

Fig.9. 2D multivariate method used for major (A) and chosen trace elements (B) in principal components
analysis (PCA) to display variable analysis (R-mode) and sample analysis (Q-mode) in a single plot.
Samples plot as points and variables as vectors.

5.2. Trace element geochemistry

In general, trace-element enrichments/depletions can reflect paleoredox conditions (U, V, Cd, Cr

and Mo, V/Cr, V/V+Ni, Ni/Co ratios; Lewan and Maynard, 1992; Hatch and Leventhal, 1992; Jones and

Manning, 1994; Tribovilliard et al., 2006; Algeo and Lyons, 2006). Most of these metals are associated to

the Fe group identified above (Fig. 9B). The Al group (Zr, Ti and K) can be used to estimate terrigenous

contributions (Pearce and Jarvis, 1992). The phosphorite group, besides P and Ca, are also associated to
94

REE, Y, U and Sr that are usually related to phosphate minerals. Paleoproductivity is reflected by Ni, Cu

and Ba, P (Tribovilliard et al., 2006).

As weathering can affect trace-element chemistry, the mobility of some elements (in an oxidizing

and more acidic environment, as must be expected for the studied area) was also considered here as a

limiting factor in the use of these proxies (Smith, 2007), as some samples, specially from phosphorites,

were collected just from outcrop. So, more mobile elements like Mo, Co or Cd in largely weathered

phosphorites were avoided in this study. Some drill cores were also affected by sulfide oxidation and

possibly some leaching of OM and in both cases, it may have altered their Ni content (not as mobile as Cd

and Mo), as an example. So, the use of proxies based on trace elements was limited to some drill holes and

selected trace elements.

The use of trace metals as a proxy for redox conditions is based on the premises that under

oxidizing conditions redox-sensitive trace metals are more soluble than under reducing conditions,

resulting in authigenic enrichments in oxygen-depleted sedimentary facies (Tribovillard et al., 2006). The

mechanisms of these trace-elements enrichment/depletion have been widely applied to mudstones and also

ironstones (Algeo and Lyons, 2006; Algeo and Rowe, 2012; Rimstidt et al. 2017; Garnit and Bouhlel,

2017; Rudmin et al., 2019), with associated strengths and limitations, since the understanding of the

primary phases of residence of each element, are crucial for paleoenvironmental inferences (Tribovilliard

et al., 2006).
+2
In francolite, trace elements generally replace Ca site, but the trace elements can also be

transferred to the sediment by absorption onto crystal surfaces, scavenging by organic matter or by

substitution in sulfides (Jarvis et al., 1994; Hiatt and Budd, 2003; Tribovilliard et al., 2006; Pufahl and

Hiatt, 2012). The trace elements associated to the P2O5-CaO group (Y, U, Sr, Ba, REE) are probably

related with francolite precipitation. These trace elements are suitable to fit into the carbonate-rich

fluorapatite/francolite ([Ca5(PO4, CO3)3(F, OH)]) structure. It is worth noting that the correlation of P2O5

with MREE is stronger than for the LREE or HREE.


95

To the Al2O3 group (SiO2, TiO2, K2O, Zr) it may also include other metals that show high

correlation with Al2O3 such as Rb (0.844), Ga (0.873), Cs (0.691), Nb (0.616) and Sn (0.456) indicating a

terrigenous influence. Most of the trace elements associated with phosphorites and phosphatic rocks show

average concentrations similar to other phosphorite in the World (Bech et al., 2010; Zaravansdi et al.,

2019) except for Th (that is slightly higher), As (higher) and Cd (significantly lower). Cd has strong to

moderate mobility in acidic and oxidizing environment (Smith, 2007), this depletion can represent a

secondary process due to weathering.

Considering the limitations due to weathering, the phosphatic ironstone intervals in a more

preserved well representing the HST of sequence B (Praguian to Emsian) and LST of sequence C

(Givetian), were evaluated (Fig.10). The investigated proxies include Mo and the ratios V/V+Ni and V/Cr

for redox potential and Ni and TOC for productivity (Jones and Manning, 1994; Tribovillard et al., 2006).

Molibdenum has a long residence time in seawater due to its conservative behavior under oxic

conditions (under these conditions it occur as Mo+6 and the stable form of Mo is recognized as molybdate

oxyanion, MoO4-2) (Algeo and Lyons, 2006). Under anoxic conditions, Mo is released from organic

matter through decay by sulfate-reducing bacteria, reduced to Mo+5 or Mo+4 and is accumulated into the

sediment as humic substances or from adsorbed Mn-Fe-oxyhydroxides or by uptake in solid solution with

authigenic Fe-sulfides (Adelson et al., 2001; Algeo and Lyons, 2006). The burial flux of Mo is higher

under reducing anoxic conditions under higher H2S (Tribovillard et al., 2006; Anbar et al., 2007). In the

plot of Figure 10 two main peaks of Mo are distinguished (43 to 53) coinciding with phosphatic ironstone

deposition.

Chromium is commonly related to the detrital clastic fraction of the sediments as it can be

incorporated within clays, within ferromagnesian minerals in which Cr replaces Mg or be adsorbed or

occur as chromite (Jones and Manning, 1994). Vanadium in contrast may be incorporated in organic
4+
matter as V and concentrated in sediments deposited under reducing conditions (Jones and Manning,

1994). Vanadium lacks any correlation with Ti or Al implying that this element was not derived from

detrital sources and that it was probably enriched in sediments by reduction from the water column. So,
96

the V/Cr ratio can be a proxy that records redox conditions (Oxic <2; 2< Dysoxic/Suboxic < 4.25; Anoxic

> 4.25; Jones and Manning, 1994). The obtained ratios for this well varied from 1 to 7.9 (Fig.10) but in

the phosphatic ironstone intervals this ratio tends to be higher.

The redox potential can also be evaluated by the V/V+Ni ratio (Lewan and Maynard, 1992; Hatch

and Leventhal, 1992). Schovsbo (2001) considered that sulphur <2.5 wt.% and V/(V+Ni)>0.84 represents

pore water anoxia closer to the sediment/water interface. In general, we observed that V/(V+Ni) ratios

were between 0.63 to 0.98 with low sulfur and intervals with a ratio >0.84 is related to phosphatic

ironstones, suggesting dominating anoxia closer to the sediment/water interface.

The presence of phosphatic ironstone were marked by a tendency of covariation and increased

values for V/V+Ni, V/Cr and Mo indicating more anoxic conditions and higher Ni (that are delivered to

the sediment mainly in association with organic matter and preserved in anaerobic conditions; Rimmer et

al., 2004; Tribovillard et al., 2006) and TOC suggesting the relation with higher productivity zones.

Copper is scavenged by complexation with OM, so with organic matter decay copper maybe

released to pore waters and enriched in the sediments due to increased organic flux being a

paleoproductivity proxy (Tribovillard et al., 2006). Enrichments in Mo suggest anoxic bottom waters. If

enrichments in Mo are also coupled with enrichments in Cu, anoxia is accompanied by a high organic

flux. So, Cu concentration and Cu/Mo ratios are proxies that were used to elucidate if the reducing

conditions were related to increased organic flux due to increased productivity (Goldberg and Humayun,

2016) (Fig.11). In the presence of a high organic flux, Ni and Cu are typically enriched in the sediments,

but not Mo (Tribovillard et al. 2006). Under strongly anoxic and sulfidic conditions, Cu plus Mo are

enriched in the sediments (Tribovillard et al. 2006). Cu concentration and Cu/Mo ratio plot indicate that

Mo concentration were higher for ironstones and phosphatic ironstones accounting to a more planar

distribution semi-parallel to the abscissa axis. This indicates that anoxia influenced these results.

However, for phosphorites and most of the phosphatic rocks the plot displays a more positive co-variation

suggesting that the enrichments in these redox-sensitive elements were possibly due to increased organic

flux, related to increased productivity and episodic fluxes of reactive organic matter to the seafloor
97

(Trabucho-Alexandre et al., 2012). Different rates of productivity, specially for phosphatic mudstones and

concretions, and also anoxia conditions, for ironstones, may have accounted to a more widespread pattern.

Fig.10. Log of the ARN-1-TO. Anoxia conditions are marked by higher values of V/(V+Ni), Mo and V/Cr
ratio. Corg and Ni are plotted showing correlation of productivity with anoxia. Stc – Trough cross bedded
98

sandstones, Sgb – graded bedded sandstones of fluvial environment Jaicós formation. The other lithofacies
symbols are like in figure 4 and Abram and Holz (2020). Gray dashed lines - anoxia limits.

Fig.11. Plot of Cu concentration (in ppm) and Cu/Mo ratio for samples from the B, C and D sequences.

5.3. Total Organic Carbon, Total Sulfur and C/P ratios

It has been observed that higher recorded TOC values (>1.5%, with the highest measured value of

2.8 weigh %) are related to ironstone and phosphatic ironstones. Heteroliths and mudstone and most

samples have TOC less than 1%. In phosphorites we can consider that early diagenesis can probably have

lowered the original total organic carbon values by consuming OM. Total organic carbon and total sulfur

values were used as in Hiatt and Budd (2003) to define dysoxic, anoxic or euxinic conditions (Fig.12).

Dysoxic facies were defined as stratigraphic units with less than 1.5 wt% total organic carbon, anoxic for

facies with greater than 1.5 wt% total organic carbon and TOC/TS ratios greater than 2.0 and euxinic by

total organic carbon values greater than 1.5 wt% and TOC/TS ratios less than 2.0 (Berner and Raiswell,

1983; Hiatt and Budd, 2003). The studied samples have not undergone significant organic burial

diagenesis that, in some cases, can modify the relationship between carbon and sulphur (Jones and

Manning, 1994). However, in general, we observed that groups of lithofacies were not well discriminated.

Most of the lithofacies were associated with a dysoxic condition, with a proximity variation to euxinic

(like the Sorg lithofacies, organic mudstones and heteroliths, in a decreasing order in tendency line from
99

more euxinic to less euxinic) or to more anoxic conditions (like ironstone and phosphatic ironstone, with

similar tendency lines and a positive intercept on the sulphur axis). Phosphorites and phosphatic rocks

presented low sulfur and low TOC probably due to weathering conditions as these samples were collected

from outcrops or eodiagenetic consumed OM. Weathering may result in 60 to nearly 100% loss of the

total organic matter (Petsch et al., 2000), so results for deeply weathered phosphorite samples are not

reliable. Normally larger TOC loss corresponds to limited physical erosion and greater contact time

between OM and oxidizing surface waters (Petsch et al, 2000) and pyrite loss precedes or coincides with

TOC loss indicating that pyrite oxidation is faster than OM weathering (Petsch et al., 2000; Wildman et

al., 2004). Therefore, total sulfur can be fasterly modified than total organic carbon and this can

significantly modify the TOC/TS ratios in outcrop samples.

Corg/PTOT (mol) ratios calculated for host bioturbated mudstone and heteroliths without authigenic

P varied from 1 to 157 (average 61) mostly associated with higher or less bioturbated lithofacies samples

(samples with more than 0.5% TOC) (Table 5). Corg/P ratios for modern normal marine phytoplankton can

vary from ∼50:1 to ∼150:1 (Algeo and Ingall, 2007) or the average value may be close to 117:1

(Anderson and Sarmiento,1994) or near 106:1 as the Redfield ratio (Redfield et al., 1963). Corg/P ratio for

terrestrial OM ranges from 300-1300 because terrestrial plants contain less P (Ruttenberg and Goni,

1997). Laminated shales deposited in oxygen-derived environments typically have Corg/P ratios an order of

magnitude higher than bioturbated shales deposited in oxygenated waters (Sageman et al., 2003; Rimmer

et al., 2004; Algeo and Ingall, 2007). In modern anoxic marine environments, the Corg/P ratios ranges from

∼110-200 (Redfield ratio) and is much smaller than observed for most Devonian shales deposited under

anoxic conditions (>1000) (Ruttenberg and Goni, 1997). Very high average Corg/P ratio can be explained

by the regeneration of P under anoxic conditions, where excess P in the water column is returned to the

photic zone through water column mixing, consequently promoting high primary productivity. These very

high values were not observed in the analyzed samples, what implicates an interface of more oxygenated

waters in the P precipitation zone.


100

Fig.12. Total organic carbon (TOC) and total sulfur data from all lithofacies. Line 1 is trend defined by
modern “normal” marine shales and line 2 is trend for marine shales believed to have been deposited in
euxinic environments (based on Berner and Raiswell, 1983). TL- tendency line. Anoxic and Dysoxic
fields are based in Hiatt and Budd (2003).

5.4. Geochemistry of REE

5.4.1. Whole rock

The same 80 samples were analyzed for REE and the relative abundances of REE were

normalized to NASC (North American Shale Composite) (Gromet et al., 1984). We calculated ΣREE, Ce

and Eu anomalies, and Y/Ho ratios and the relationships between these proxies and trace and major

elements. The results show that the average shale-normalized REE patterns are rather consistent

throughout the dataset (Fig 13). The total REE content of ironstone varies from 59 to 707 ppm (av. 305

ppm), phosphatic ironstone from 229 to 2160 (av. 660 ppm) and phosphorites from 667 to 2992 (av. 1771

ppm). Therefore, phosphorites contain the highest average REE concentrations and it represents samples

from late Eifelian and Givetian ages. Host sandstone, heteroliths and mudstones exhibit a relatively low

average REE concentration (from 106 to 358, av. 222).


101

In general, the patterns for phosphatic ironstone and phosphorite/phosphatic rocks are enriched in

MREE in relation to HREE and LREE, with the phosphorite with more enriched and distinctive patterns

(Fig.13B). LREE vs MREE and MREE vs HREE enrichments were calculated as PrSN/TbSN (0.25-0.55)

and TbSN/YbSN (1.4-2.5), respectively. There is not a notable difference between phosphatic ironstone

from phosphorite concretions, phosphatic braquipod shells or nodule REY patterns, and no Eu anomalies

are observed. Ironstone shows a flatter pattern with some exceptions with enriched MREE. Mudstone and

heteroliths shows a flat pattern, with discrete differences in ∑REE, but with similar REY shape.

The YSN/HoSN ratio reveals decoupling of the geochemical twin’s Y and Ho, which produces YSN

anomalies in REYSN patterns (Bau, 1996). Phosphorite and phosphatic rocks show Y/Ho variations from

0.64 to 1.19 (av. 0.87), varying from discrete negative to discrete positive anomalies, even for the same

sampled interval.

Redox related decoupling of Ce from the other REY where investigated based on the CeSN/CeSN*

ratio where CeSN*= 0.5LaSN + 0.5PrSN (Bau et al., 2014), which produces Ce anomalies in REY SN patterns.

The calculation of (Pr/Pr *) SN ratios, [(Pr/Pr*) SN = Pr SN /(0.5Ce SN + 0.5Nd SN)] were used to help in

defining true Ce anomalies as in Bau and Dulski (1996) and Bau et al. (2014). Some phosphorite,

ironstone and phosphatic ironstone show discrete positive Ce anomalies and for most of the samples the

calculated values ranged between 0.90 and 1.10 and were considered devoid of anomalies (phosphorite -

1.06; phosphatic ironstone -1.02; and ironstone - 1.05).


102

Fig.13. NASC-normalized REE-Y patterns of phosphorites, phosphatic rocks, ironstones, phosphatic


ironstones and host rocks (Table 4).
103

There is not a difference between REY patterns of ironstone collected from drills of the western

side of the basin in relation from those collected in the eastern part and there is not a control related to

more enriched or flat patterns with variable age (Fig.13). The REY pattern from weakly weathered

ironstone samples from the wells when compared with deeply weathered ironstone from outcrops present

a quite similar pattern, instead of LREE that tend to a depletion or show a general impoverishment for

some samples.

5.4.2 REE from separated minerals (siderite/francolite)

REE from francolite, siderite and Mg siderite are shown on Table 6. The total REE content of

francolite varies from 635 to 4092 ppm (av. 1359 ppm). It is characterized by “hat” shaped REY patterns,

with notable enhancement of MREE and prevailing weak Y positive anomaly (0.86 to 1.31 – av. 1.12), a

lack of Ce (CeSN/CeSN* between 0.86 a 1.06, av. 0.97) or Eu (av. 1,19) anomalies, calculated as Eu/Eu* =

[Eu/(Sm2 * Tb)1/3]SN from Lawrence et al. (2006).

Two different patterns were obtained for siderites (Fig.14). One with a conspicuous enrichment in

MREE, with some differences in ∑REE (but with similar shape) and lack of Ce (av. CeSN/CeSN* - 1,07)

and Eu anomalies (EuSN/Eu*SN - 1,03), mainly representing type 3 siderites. Another pattern is represented

by a flat shape with some regular enrichment, exhibiting not a conspicuous difference from Mg siderite

(associated to ironsilicates) and siderite on concretions related to mudstones.


104

Fig.14. NASC-normalized REE-Y patterns of francolites and siderites (Table 6).

5.4.3. REE Interpretation

REE can provide meaningful insights in the marine environment as they are influenced by

seawater, riverine, dust and hydrothermal input, and participate in scavenging, complexation and redox

reactions that regulate trace element abundances (Haley et al., 2004; Lawrence et al., 2006; Bau et al.,

2014). Fe-Mn crusts and phosphorites obtain REY from seawater by different mechanisms. In the Fe-Mn
105

crusts it is by sorption of REY and oxidation of the Ce. In the phosphorites it is by substitution for Ca

during precipitation and to lesser extents by inheritance and sorption (Hein et al., 2016). Preferential

removal of Ce and LREE by oxyhydroxide particles is responsible for the distinctive REE patterns

characteristic of modern seawater (Sholkovitz et al., 1994). In general, ironstone, phosphorite, chert,

carbonates and francolite in oxygenated marine environments show negative Ce anomalies and an

enrichment in HREE that are incorporated during earliest diagenesis, recording the REE composition of

seawater at the time of deposition (Jarvis et al. 1994; Lécuyer et al., 2004).

Under oxidizing conditions, the decoupling of Ce in seawater results from Ce3+ oxidation to Ce4+,

which is less mobile and results in Ce depletion of seawater (Sholkovitz et al., 1994; Bau and Dulski,

1996). In addition, the abundance of REE is generally regulated by scavenging processes (Byrne and Kim,

1990) that involves organic matter, iron coatings and phosphate in early diagenetic processes resulting in

different patterns under different redox conditions (e.g., Haley et al., 2004; Caetano-Filho et al., 2018).

Reducing conditions results in Ce positive anomalies. Seawater pattern can also vary from local input of

terrigenous sources, especially in the proximity of coastal regions (Zang et al., 2019), where siliciclastic

material can influence REE patterns with no Ce anomaly or heavy REE enrichment (Watkins et al., 1995).

The REE patterns can also be influenced by extensive re-crystallization through late diagenetic processes

or late weathering effects (Phan et al., 2019). Reducing conditions also results in Eu positive anomalies as

Eu3+ is reduced to Eu 2+ (Danielson et al., 1992; MacRae et al., 1992; Bau and Dulski, 1996).

The observed conformity and regularity of patterns under varying weathering intensities (from 0

to 25%) suggests that weathering had a minimal effect on the REY patterns. It results only in some

impoverishement, especially in LREE, as observed in weathered ironstone collected in outcrop (Fig. 14

G). This regularity is even observed for the Y/Ho ratios, especially considering outcrop data, that is

normally more affected, as Y loss outweighs its geochemical twin Ho during chemical weathering

(Babechuk et al., 2012).

The typical HREE enrichment and Ce negative anomaly of seawater was not observed in the REY

patterns describe here and, although reported by some authors (Phan et al., 2019), this is not a common
106

pattern observed in Devonian phosphatic sediments (Jarvis et al., 1994; Emsbo et al. 2014). The Eu like

the other MREE exhibits enriched patterns possibly obscuring the discrimination of Eu positive

anomalies.

Two main patterns were observed considering all lithofacies analysed, one with a main MREE

enrichment and another with a flat shape (Fig.13). The close association between P2O5 and ΣREE

indicates that REE were directly influenced by the abundance of scanveged francolite, probably by

complexation of CO3-2 in this phase. The poor correlation of ΣREE with Al and Rb, indicates that it was

not influenced by terrigenous contamination.

Phosphorites and francolite, as well as phosphatic ironstone and type 3 siderite, present MREE

enrichment and “hat shaped” that may reflect porewater composition. Probably these phases where

precipitated under similar porewater conditions. The studied phosphate concretions show similar bulk

REE concentration and shale-normalized distribution patterns as two braquiopod phosphatic shells,

phosphorite nodules and one hardground sample that normally precipitates direct from seawater, reflecting

enriched MREE seawater patterns or diagenetic phosphatic replacement. This result is also similar for

various types of Devonian nodules (McArthur and Walsh, 1984) and phosphate rich skeletal conodonts

and braquiopod shells (Grandjean-Lécuyer, 1993, Jarvis et al. 1994; Holser, 1997). As observed in other

works, Devonian phosphorites present a MREE enrichment with higher ΣREE (Emsbo et al., 2014).

MREE enrichments has been attributed by some authors to different processes, that includes

extensive or late diagenetic processes, inheritance from organic matter decay, early diagenetic processes in

the presence of dissolved iron or a reflection of secular seawater changes (Felitsyn and Morad, 2002;

Haley et al., 2004; Phan et al., 2019). Most authors avoid the idea of enrichment of REE by late diagenesis

and consider that early diagenetic REE scavenging is more likely to alter REE distributions (Banner et al.,

1988; Jarvis et al., 1994; Shields and Stille, 1998). Phosphorite and ironstone with enriched MREE are

early diagenetic (authigenic) phases and the observed prevailing paragenesis does not point to an

extensive recristalization process.


107

Haley et al. (2004) described “hat shaped patterns” in early diagenetic processes in the presence of

dissolved iron. MREE patterns with a bulge shape for francolite was also described for diagenetic

processes under dysoxic/suboxic and anoxic conditions where pore waters lose or weaken the negative Ce

anomaly of overlying oxygenated bottom waters (Shields and Stille, 2001). Therefore, the “hat shaped

patterns” obtained are better explained as developed under reducing conditions in pore waters of organic-

rich muds in the presence of iron dissolved porewater. Organic matter patterns evaluated by Felitsyn and

Morad (2002) show that, similar to francolite (in some phosphorite concretions), the REE in humic acids,

kerogen and carbonaceous fossils results in a pronounced MREE enrichment with “hat shaped REE

patterns”. Mainly mobile compounds (e.g., humic acids) are enriched in MREE. So, a genetic link can also

be suggested between the organic matter and ironstone/phosphate, where the diagenetic decomposition of

organic matter and formation of mobile components (e.g., MREE-rich humic acid) may not only provide

phosphorus and suitable early suboxic/anoxic conditions for francolite precipitation, but was a likely

source of REE (Felitsyn and Morad, 2002). Therefore, the organic matter decay in early diagenesis with

the presence of dissolved iron probably propitiates here the redox interface conditions for MREE

enrichment. On the other hand, hardground samples also suggest certain original MREE bottom seawater

enrichment.

Some authors also relate that differences in REE abundance in phosphate phases constitute a

direct result of secular variations in ocean chemistry (Lécuyer et al., 2004; Emsbo et al., 2014). Lecuyer et

al. (2004) attributed MREE enriched patterns of Early Cretaceous as related to a period of greenhouse

conditions and stratified and poorly oxygenated waters. The Devonian is one of these periods of REE

enrichment associated to phosphorite deposition and MREE enrichment patterns. This has led Emsbo et al.

(2014) to suggest the economic potential for REE associated with phosphorites. The grade of REY oxides

in phosphorites and phosphatic rocks fluctuate between 0.02 to 0.48 wt% RE2O3, (av. 0.15 wt% RE2O3)

confirming this economic potential and suggests more anoxia condition of the oceans during this period.

The relationship between Ce/Ce* and Nd and between Ce/Ce* and YSN/HoSN, and Y negative

anomalies in ironstone and phosphatic ironstone indicates hydrogenetic/diagenetic source of REE (Fig.
108

15) (Bau et al., 2014; Rudmin et al., 2019). The flat pattern of ironstone is similar to a typical “shale”

REE pattern although without a Ce negative anomaly, but with a weak positive Ce anomaly. This typical

“shale” REE patterns lacking negative Ce anomalies can represent also diagenetic types that inherited

their REE in anoxic or dysoxic/suboxic environments associated with bottom and interstitial waters

(Watkins et al., 1995). This also seems to be the case of siderite concretions associated to mudstone

lithofacies (Fig.15).

Bioturbated mudstone, organic mudstone and heteroliths show a weak negative Ce or no

anomalies what suggests exposure to oxygenated waters. Flat REE patterns in mudstones and heteroliths

can probably reflect detrital influence (Kidder et al. 2003). However, the flat REE can be also sourced by

particulate organic matter. According to Haley et al. (2004), as it is remineralized, LREEs are released,

becoming relatively equivalent to the HREEs and making the patterns “flat” (Fig.15), creating a “linear”

pattern in suboxic, anoxic and at depth in oxic sediments.

Y is geochemically similar to the heavy REE Ho but in marine systems evidences different

complexion properties, as Y is less effectively scavenged from seawater then Ho. So, Y/Ho ratios in

marine environments are higher than in shale composites (Bau et al., 1997; Nozaki et al., 1997). Y/Ho

ratios from seawater is generally high (44 to 74), being less high for ocean margin seawater then to open

seawater (Nozaki et al., 1997). Terrigenous materials have lower Y/Ho ratios (28) (Taylor and McLennan,

1985). YSN/HoSN has a good correlation with CaO (0,67), Sr (0,63) and P2O5 (0,68) meaning a scavenging

preference of Y than Ho by francolite. The Y/Ho ratio for phosphorite range from 24 to 40, phosphatic

ironstone from 21 to 40 and 28 to 44 for francolite which suggests a terrigenous influence or possibly

mobility of Y in relation to Ho due to weathering. However, there is a decrease in dissolved seawater

Y/Ho ratios in dysoxic/suboxic and anoxic waters driven by an increase in Ho relative to Y as Fe-rich

particles dissolve (Bau et al., 1997). Considering the regularity of REYpatterns that evidences minor

weathering influence, we can conclude that the observed large variations of Y/Ho suggests suboxic and

anoxic waters, although we can consider also some terrestrial input contribution in marine water.
109

Fig.15. (A) CeSN/CeSN* ratio vs YSN/HoSN ratio discrimination diagram. (B) CeSN/CeSN* ratio vs Nd
concentration discrimination diagram. The samples plot into hydrogenetic and diagenetic genetic types
(fields from Bau et al., 2014; Rudman et al., 2019).

5.5. Isotope pattern

Primary, postdepositional diagenetic alteration and weathering processes where researched by

francolite, siderite and Mg-siderite δ13C and δ18O isotopes (Tables 7 and 8). The 18O isotope composition

in francolite was from the carbonate site since it is more sensitive to diagenetic change, although alteration

by meteoric water can cause isotopic signatures to shift along a mixing line towards groundwater

signatures (McArthur et al., 1986; Jarvis et al., 1994).

5.5.1. Siderite isotope pattern

The results obtained are probably general indicators of the conditions of siderite precipitation,

since individual grains were hardly separated due to the small grain size. There was not enough material to

investigate siderite-phase 1. Mg Siderite phase 2 and Fe siderite-phase 3 were investigated as these were

the most widespread types. Siderite phase 4 was only observed in outcrop associated to concretions, so the

isotope signature of this phase can have been deeply altered chemically due to the weathering influence.

In a general way a wide range of results were obtained for the siderites (Fig.16 and Table 7) and

this is related to the dominant petrographic association of the cement and is in agreement with the
110

different textural patterns and phases of generation. This widespread distribution can be clustered into four

main zones (A, B, C and D). All the values of δ13C were negative. The variation for the siderites is

δ18OVPDB from -12.0 to 2.5, with a mean of -2.6 and δ13C that ranges from -21.5 to -0.8, with a mean of -

8.2. The greatest TOC concentrations in the bulk rock corresponds to lower values of δ13C in siderites,
12
which suggests an influence of OM degradation on these results, as preferentially OM incorporate C

(Rodrigues et al., 2019) and when degraded probably liberate it to porewater decreasing 13C/12C ratios.

Isotope values for Mg siderite from the type 2 presented δ18OVPDB variation from -8.3 to -0.4 ‰

(average -3.3 ‰) and δ13C -15.3 ‰ to -4.9 ‰ (average -11.2). Most of the Mg siderite was associated

with clustered zone A.

Samples of the surficial deeply weathered siderite concretions of type 4 have the lowest δ 18
O

values (δ18OVPDB -12.0 to -10.8) and intermediate δ 13C (δ 13C -8.5 to -5.5).

In the same sample we can observe a shift on isotopic signature from Mg siderite (type 2) to late

iron siderite (type 3). Most of the shift is related to a more negative signature for δ13C and/or a less

negative δ18O.

Siderite istope pattern interpretation

Compiled δ13C and δ18O data from Devonian seawater obtained from low magnesium calcite

brachiopods, belemnites, oysters, foraminifera (Gao, 1993; Veizer et al., 1999; Selleck and Koff, 2008)

indicates that Devonian seawater had a positive δ13C isotope composition (between 0‰ and <5‰) and a

widespread variable δ18O composition (between -2‰ and -10‰), probably associated to secular variations

and increase in temperature, that normally results in decrease of δ18O isotope water signatures (Weissert et

al., 2008; Veizer and Prokoph, 2015). Therefore, as most of the δ18O were not lower than -6 ‰ and exhibit

a variable negative δ13C isotope composition, it suggests that the siderites evolved mainly from marine

waters (Mozley and Wersin, 1992; Zymela, 1996) and associated porewater microbial processes (Chow et

al., 1996).
111

River discharges, high precipitation and ice melting can change salinity conditions, generating

regions with more depleted values of δ18O (Redfield and Friedman, 1965). δ18O and salinity are strongly

correlated (Belem et al., 2019). Here, a variation in salinity can be suggested by the horizontal widespread

δ18O differences of clustered zones B and C. In this case, we can associate lower δ18O sample values with

a decrease in salinity probably associated with river discharges influence and more positive δ18O values as

related to higher salinity conditions.

One sample presents lower δ18O cement (-8.3 ‰). Although this lithofacies show some evidences

of shallow burial diagenesis that could account to a shift of the primary values (Potter et al., 2005), this

lower value can be also considered as related to a meteoric diagenetic influence.

δ13C values in marine carbonates precipitated direct from seawater are a function of the 13C/12C of

dissolved inorganic carbon (DIC) in surface water (Rodrigues et al., 2019). Organisms preferentially use

the lighter isotopic species because of the lower energy "costs", resulting in significant fractionations

between the substrate (heavier) and the biologically mediated product (lighter) (Kendall and Cadwell,

1998). By this way, organic matter is enriched in the light isotope 12C and its removal by burial results in

ocean water richer in 13C. As a result, δ13C from carbonates are normally increased (Scholle and Arthur,

1980). However, early diagenetic processes in porewater, like the oxidation of organic matter that results

in addition of 12C-enriched cement, can influence measured signatures. The fact is that siderite isotopes in

diagenetic fractionation environments with associated organic matter degradation will strongly depend on

the degree of decomposition and the type of the bacterial communities (Goevert and Conrad, 2008). In this
13
case, carbon isotopes record microbial processes and C depends on the type of dominant bacteria. In

biotic processes siderites are usually formed by microbial methanogenic production or by Fe-reducing

bacteria (Mozley and Carothers, 1992; Zymela, 1996). It can be also reduced directly by sulfate reduction

bacteria by an enzymatic mechanism (Coleman et al., 1993) or it may form in the zone of sulfate reduction

when the rate of Fe3+ reduction exceeds the rate of sulfate reduction (Pye et al., 1990).

The common occurrence of siderite associated with degraded organic matter suggests that

mediated bacteria origin have been an important control in the siderites precipitation. As none of the δ13C
112

results were positive, this suggests that precipitation didn’t occur from pore water under fermentation of

organic lithofacies in a broad methanogenic production zone.

Three main zones of 13C (A, B and C) were distinguished from lighter values to near zero values,

what probably indicates variations on prevailing microbial communities related to the degradation of

organic matter. Some siderites have δ13C values that varies from 0 to -6‰ (zone C siderites), which

suggests an origin in the Fe-reduction, suboxic zone (Chow et al., 1996). The reduction of organic matter
+3
and Fe by bacterial processes increased the dissolved inorganic carbon concentration with high Fe+2

concentration of the porewaters so that siderite precipitated. This suggests large amounts of reactive iron

and low sedimentation rates, which would allow a wider zone of iron reduction. However, others have

lower δ13C values (e.g., -10‰ to -15‰ o and also rare less then -15‰), indicating essential derivation of

dissolved carbon from the oxidation of marine organic matter in the bacterial sulfate reduction zone, as

described for the cluster A (Irwin et al., 1977; Abdul-Hadi and Astin, 1994). This finding is supported by

the close association between siderite and authigenic framboidal pyrite in these samples in which pyrite

predates the siderite. Framboidal pyrite precipitates in organic rich sediment from anoxic pore water

where bacterial Fe and sulfate reduction generate Fe2+ and sulfide (Pufahl and Groat, 2017). Therefore,

this suggests that dissolved carbon was derived from the oxidation of marine organic matter in the

bacterial sulfate-reduction zone and also probably near the sulfate-reduction/methanogenesis transition

zone. Most of the siderite and Mg siderite results indicate an evolving porewater from suboxic to more

anoxic conditions, and then the resumption to more suboxic conditions under iron reduction zones (Zhang

et al. 2001). A slightly increase in the sedimentation rate associated to prograding sections can have

produced rapid burial conditions which was able to promote iron reduction processes over sulphate

reduction, possibly also allowing SRB to switch to iron reduction (Pye, 1981; Coleman et al., 1993). The

presence of siderite-cemented rip-up clasts in upper units demonstrates that some of the siderites formed at

very shallow depths (Masterson and Paris, 1987). Lower values of δ13C are also interpreted by some

authors to the oxidation of sulfide (Mortimer et al., 1997). Although others authors interpret low values of

δ13C to enrichment in δ 12C related to early stages of methanogenesis, subsequent to depletion of pore
113

waters in dissolved sulfate (cf. Chow et al. 1996). Very early siderites formed in the transiton methanic

zone/sulfate zone can also occur while pore waters still have very low δ13C values from preceding sulfate

reduction (Curtis and Coleman, 1986). Although, the resumption to bioturbated features favours hinder

iron reduction condition. Cluster B is in an intermediate position, probably with mixing communities.

Based on the previous discussed arguments, we conclude that both the oxygen and carbon isotopic

were sensitive to organic degradation process and possible variations in salinity/temperature. The variation

from zones A to B and then to C, can be likely explained by lighter δ13C values for pore-water evolving to

higher values, probably due to a shift from sulfate reduction to iron reduction microbial communities.

Fig.16. Cross plot of δ18O and δ13C for siderite. Compiled δ13C and δ18O data from Devonian seawater
were obtained based on low magnesium calcite brachiopods, belemnites, oysters, foraminifera (Veizer et
al., 1999), brachiopod shells (Selleck and Koff, 2008; Gao, 1993). δ18O data from low magnesium calcite
were corrected reducing the amount on calcite data by -1.44 (Fractionation factor
for siderite of 1.01025, c.f. Maynard, 1982). Cluster A – gather siderites formed in the sulfate reduction
zone after sulfur consumed by pyrite formation. Cluster C – gather most of the siderite type 3 formation in
an iron reduction zone. Cluster B – gather probably mixing comunnities from A and C. Horizontal
114

variation in A and C are interpreted as variation in salinity. Cluster D – gather samples with weathering
signatures.

5.5.2. Francolite isotope pattern/interpretation

Francolite shows more clustered data points than siderites ranging from δ18O VPDB from -

14.11‰ to -4.6‰, with a mean of -10.08‰ and δ13C ranging from from -2.9‰ to -8.1‰, with a mean of -

6.3 (Fig.17 and Table 8). Most of the francolite results were probably modified by extensive weathering,

characterized by generally lower δ18O values although most of the δ13C are in the range of shelf associated

phosphorites, like Namibia concretions, Peru Sechura and Morrocan (McArthur et al, 1986).

Fig.17. Cross plot of δ18O and δ13C for francolite. Fields like Pufahl and Groat (2017), after Jarvis et al.
(1994) with data from McArthur et al. (1986) and Girard et al. (1993). δ13C signatures for francolites
resembles other samples in upwelling systems. δ18O variations indicates the weathering effect on samples.

6. Discussion

6.1. Paleoenvironmental reconstruction and phosphorus output


115

The petrographic and geochemical data presented here support the model that phosphorites and

ironstones/phosphatic ironstones were not deposited at the entire same conditions. Figure 18 depicts a

general schematic model for the different conditions in the genesis of these rocks. Phosphorites were

probably formed under more oxic bottom waters condition that allowed francolite precipitation into the

sediment related to upwelling nutrient sources near SWB. On the other hand, phosphatic ironstones were

formed under a redox evolution that permitted the preservation of more anoxic conditions and probably

more phosphorus diffusion (Fig.18) resulting in less enriched phosphatic rocks near FWWB.

The main output process of phosphorus with phosphorites and phosphatic rocks precipitation was

triggered by increased organic flux to the sea floor as a consequence of high primary productivity

probably due to migration of nutrient rich upwelled waters that resulted in francolite precipitation into the

sediment near the limit inner/outer shelf (e.g. Follmi, 1996; Pufahl, 2010) with the predominance of

marine phytoplankton (Trindade et al., 2015). This interpretation can be supported by lithofacies

association and the positive correlation between Cu and Cu/Mo ratio, with redox boundary mainly varying

from the sediment-water interface and inside sediment facilitated by burrowing. An increase in Acritarchs

and Prasinophycean algae typical of open marine conditions (Trindade et al., 2015; Trindade and

Carvalho, 2018) was detected in well ARN-1-TO corresponding to the level of maximum flooding surface

(Eifelian/Early Givetian) of the sequence C. In this maximum flooding surface are also concentrated the

phosphorite rich intervals observed in outcrops. This condition is similar to most upwelling processes

associated to transgressions and maximum flooding surfaces (Glenn and Garrison, 2003; Morad et al.,

2012). Increasing anoxia was probably a consequence of the increased organic flux. Ocean stratification

probably separated the anoxic laminated mudstones from oxic bioturbated mudstones representing a

possible chemocline on the shelf (<50 - 100 m), as in modern coastal systems (Trucco-Pignata et al.,

2019) but phosphorites precipitation occurred mainly in the interface of bioturbated mudstones and

bioturbated heteroliths what implicates expansion of oxygen minimum zones due to higher productivity.

As these rocks have low OM content, we presume that most of the organic matter may have been oxidized
116

and eliminated from the sediment before burial. This condition was achieved with marine transgression

and a decrease in sedimentation rate with stratigraphic condensation. Sedimentation rate generally affect

the residence time of reactive organic matter within any particular diagenetic zone within the sediment. In

general, if the rate is too slow, most of the organic matter is oxidized and eliminated from the sediment

before burial (Morad et al., 2012). The obtained C/P ratios are close to the Redfield ratios for three

samples. This suggests that for these samples, the organic-rich mud has preserved its original molar ratio

(Follmi et al., 2005). However, some results were lower what implies general enrichment of the sediment

in phosphorus relative to organic carbon. This implies that phosphogenesis would have been driven by the

degradation of almost the entirely marine organic matter and transfer of organic phosphorus into

phosphates. For some rocks the relation is two low what can imply an additional source of phosphorus,

what suggests in some cases that iron redox pumping (enrichment of phosphate in pore waters by the

release of phosphate sorbed onto Fe oxyhydroxides; Heggie et al., 1990; Pufahl and Grimm, 2003; Pufahl

et al., 2010) could have been another viable mechanism for phosphorus output (Follmi et al., 2005). The

observed C/P ratios from phosphorite host shales and heteroliths are similar from other modern coastal

upwelling sites like the Oman margin (C/P ratio 22), Peru margin (C/P ratio 97), Gulf of California (C/P

ratio 108), Namibian shelf (C/P ratio 52) (Brian, 2009). So, increases in the dissolved phosphate

concentration of pore water by the release of polyphosphate resulted in francolite precipitation inside

sediment prior to compaction of the sediments (by the presence of disruption in the laminar bedding).

Probably this occurred just beneath seafloor, with the formation of phosphorite concretions in more

suboxic to anoxic conditions, as evidenced by siderite precipitation after francolite, the close association

with siderite concretions on outcrops and by significant MREE enriched patterns. The hosting bioturbated

lithofacies, suggests that the activities of burrowing organisms contributed to keep ventilation conditions

inside sediment, allowing degradation and also the release and concentration of phosphorus. This

condition propitiated the main P2O5 concentrations in the Parnaíba Basin, in the MFS of sequences C and

D. Minor variations in the Eh and pH conditions, and concentration of phosphate or excess iron

concentration, favoured the deposition of francolite or siderite. The siderite concretions were formed in a
117

2+
higher availability of reduced iron (Fe ) and a higher concentration of bicarbonates (HCO3–) causing

supersaturation in respect to siderite (Coleman et al., 1983). The variation from discrete concretions to

coalesced concretions indicates that it mainly occurred in sulphate reduction zone and also in the limit

from sulphate reduction to methanogenic zones, possibly in local sites (Bojanowski and Calrkson, 2012).

Probably, these siderite concretions were generated in methanogenic microenvironments developed

entirely within the sulfate reduction zone, although this condition could not be proper evaluated due to the

intense weathering of these rocks (lowering the isotopic C13 and O18 data). The anoxic or dysoxic/suboxic

condition for the siderite concretion is also a feature marked by typical “shale” REE patterns lacking

negative Ce anomaly (Watkins et al., 1995) what implies a different porewater condition from phosphorite

concretion, as different patterns were obtained. The δ13C content for francolite in phosphorites and

phosphatic rocks (-2.9‰ to -8.1‰), is similar from δ13C from other important deposits like the Namibia

shelf, Peru, Senegal and Moroccan deposits (MacArthur et al. 1986; Compton and Bergh, 2016). Highest

U values are also closely associated with these oxygen-restricted zones, in middle to more persistently

anoxic intervals. This corroborate that these francolites formed near the suboxic to anoxic boundary where

intense microbial processes rapidly removed oxygen from the sediment and oxidized hydrogen sulphide

diffusing from below (Compton and Bergh, 2016). This condition suggests that sulfur bacteria can have

also played a significant role in phosphogenesis as in other phosphorite deposits (Bailey et al., 2013).

Phosphorite nodules were formed associated with glauconite coatings and bioturbated heterolithic

lithofacies. Glauconite precipitation occurs in similar low-sedimentation-rate open-marine environment

with limited terrigenous input characteristic of phosphorite deposition (Velde, 2005). These were formed

under suboxic conditions, denoting phosphorite precipitation near surface water and in more oxygenated

seawaters (oxic to suboxic). Internal and external microbial layering is suggested by weathered opaque

minerals and OM linear layers around glauconite and phosphorite nodules what suggests that possible iron

reduction and also microbial conditions were responsible for the reactions for phosphorus release and

francolite precipation. The Fe-redox boundary must have been at the seafloor and moved up and down
118

through the sediment in response to changing biological oxygen demand (Pufahl and Grimm, 2003).

Inside the nodules the preservation of cells of Acritarchs, Prasynophicean and possibly Radiolarians also

record that these were formed under high productivity zones. These phosphorite nodules (and also

phosphate rich braquiopods) present MREE enrichments, what suggests that seawater composition during

this period produced different patterns from modern depleted LREE patterns, expected for this

phosphorite precipitation near seawater. This corroborates a possible variation in global ocean chemistry

in this period as observed for other Devonian patterns around the world (Jarvis et al., 1994; Emsbo et al.,

2014). The coincidence of MREE in samples from the Devonian (Jarvis et al., 1994; Salama et al., 2018),

in general, suggests that the MREE-enriched pattern can have been associated to a more global signature

derived from a major influence of increases in pronounced MREE and Ce transferred to the water column

due to periods of common anoxic conditions as observed by other authors (Jenkyns, 2010).

For these phosphorites, reworking and the development of granular facies indicates that storm

induced processes were able in producing more enriched intervals. Granular beds are cemented by

multigerational pristine phosphate at the sea floor (hardground) that showed that sediment starvation

caused prolonged stagnation of the zone of phosphogenesis (Pufahl and Groat, 2017). However, large

pristine phosphorite production (due to the persistence and intensity of upwelling) and the generation of

thick amalgamated granular beds (produced by increasing storm frequency and intensity with time), was

probably not an efficient mechanism, as for most of the main economic deposits (Föllmi et al.,1991;

Pufahl et al., 2003; Pufahl and Groat, 2017). We observed that these hardground phosphorites also exhibit

MREE enrichments also suggesting a distinct seawater signature.

In its turn, phosphatic ironstones and ironstones contain increasing marine phytoplankton with the

occurrence of synsedimentary Fe and Mg silicates, francolite and framboidal pyrite in the transition zone

offshore – lower shoreface lithofacies. They present REE data that characterizes a limit between

hydrogenetic/diagenetic sources (Fig.15) (Bau et al., 2014; Rudmin et al., 2019). For the precipitation of

these rocks, the evolution history has shown a sequence of steps as indicated by mineral precipitation. In

phosphatic ironstones, odinite and francolite coated grains were earlier precipitated associated with
119

condensed zones (or not) and suboxic environments. Sometimes exposed by reworking and submitted to

new coatings. This must have occurred on flooding events and near surface water. Then these were

succeeded by chamosite or clinochlore (when iron was too low to form chamosite) new coatings,

reflecting available Fe or not. Hybrid coated grains suggests that repeated episodes of precipitation,

exhumation, erosion and reburial produced cortical layers and mixing of oxygen occurred, maintaining

suboxic to anoxic porewaters, although the absence of hematite cortical layers indicate that oxygen mixing

was not so efficient and constantly mixed into sediment as indicated by the prevailing anoxia in the Cu

versus Cu/Mo ratio diagram. Then, framboidal pyrite precipitated in organic rich sediment from anoxic

pore water where bacterial Fe and sulfate reduction generated Fe2+ and sulfide. Progressive OM

degradation led to more anoxic conditions well represented by clustered lower δ13C values - A (e.g., -10‰

to -15‰) indicating essential derivation of dissolved carbon from the oxidation of marine organic matter

in the bacterial sulfate reduction zone (Irwin et al., 1977), by MREE enrichments and by Mo and V/Cr,

V/V+Ni ratios. The evolution from the total consumption of H2S by pyrite formation and excess iron with

Mg siderite precipitation probably occurred in the sulfate reduction zone (Coleman et al., 1993). Higher

alkalinity or Mg concentration may have inhibited continuing phosphate precipitation. The presence of

marine and also some terrestrial organic matter (Souza, 2007; Trindade et al. 2015) also can have

contributed to less P liberation and accumulation. CO2 produced by the decay of organic matter during

diagenesis caused the widespread crystallisation of later siderites possibly with a terrestrial source of iron.

The increased expansion of anoxia probably reduced the amount of P buried and absorbed into sediments

and enhanced the pool of easily mobilized P in the water column and favoured recycling into the water

column with biocoupling (Carmichael et al., 2014). Iron siderites were formed under suboxic to anoxic

conditions in the iron reducing zones in a later phase, exhibiting bored possible hardgrounds due to

stratigraphic condensation, and denoting the resumption to more oxic conditions. A major influence of

river run-off (also supported by the palynofacies studies of Souza, 2007; Trindade et al. 2015; Martins,

2019) probably accounted for variation in salinities or temperature, as suggested by δO18 variations and

also probably contributed to this iron source and to a terrestrial nutrient contribution reaching deeper
120

sections (transition zone) (Young, 1989; Yilmaz et al., 2015; Pufahl et al., 2020). Storm-induced dynamic

processes resulted also in thin accumulations of reworked ironstone and phosphatic ironstone.

Fig.18. Depositional model for Early and Middle Devonian phosphorites and phosphatic ironstones.
Phosphorite zones were formed near SWB with more stable redox-boundaries that lies near the sediment-
water interface, were francolite precipitated in more suboxic then anoxic environments, with prevailing
oxic bottom waters, an efficient process of degradation with less preserved organic matter (Algeo and
Ingall, 2007), resulted in more available phosphorus for phosphorite accumulation into the sediment. The
phosphatic ironstones precipitated near FWWB, with less stable redox-boundaries, in shallower more
anoxic and iron rich environments (ironsilicates and siderites precipitation,with marine siderites and also
later siderites with meteoric signatures reflecting river discharges influence) that allowed organic matter
preservation (a mixing zone of marine and terrestrial organic matter; Souza, 2007, Trindade et al., 2015
and Martins, 2019) and more P diffusion to water column (Algeo and Ingall, 2007).

6.2. Potential economic implication

Devonian paleogeographic reconstructions (Blakey, 2019) coupled with ocean-atmospheric

modelling (Parrish, 1982; Markello et al., 2008) suggests that the Parnaíba Basin margin was favorable
121

positioned for upwelling (Parrish, 1982; Markello et al., 2008) and similar condition must have happened

for the connected Amazonas Basin, although the mechanism for the extension of this productivity in the

epicontinental basin is not well understood. Probably this was not an intense upwelling source,

considering the maximum TOC observed for this studied period in the basin (4%) (Andrade, 2015;

Martins, 2019), although this can have been intensified during the Frasnian/Famenian transgressive

periods in the Parnaíba and Amazonas basins considering the higher TOC’s with algal origin and marine

environment (1 – 8 %) (Pereira et al., 2009; Andrade, 2015). In general, a temperate climate should have

encouraged a more oxygenated water column however with anoxia phases due to events of higher

productivity. We recorded that the limit inner/outer shelf was more enriched in P concentrations then

shorewards regions. This condition reflects phosphogenesis along the shelf developed within an oxygen

minimum zone like modern upwelling sites. In general, maximum flooding surface keeps the potential for

very low rates of sedimentation that commonly produces high-grade, amalgamated phosphorite beds

(Pufahl and Groat, 2017), producing economic potential for phosphorite deposits. However, at the

Parnaíba system, marine primary productivities were not sustained across the entire platform, as observed

in most of the economic phosphorites.

The presence of intermittent increased terrigenous input (Trindade et al., 2015) suggests that there

was also a terrestrial nutrient source to the marine areas, that increased iron and nutrient deliveary,

providing additional shallower anoxia conditions. Inflow and variable runoff river discharges probably

made phosphogenetic process maintenance difficult, considering an additional complexity in stabilizing

the redox boundary with possible variations in alkalinity/anoxia conditions as suggested by the later

siderites precipitation.

Sinking flux was probably variable as suggested by organic productivity rates and may have been

lower in some regions than in others. This productivity although being upwelling driven may also have

been indirectly linked to river inflow producing large-scale estuarine circulation (Kusyk et al., 2010). So,

the Parnaíba Basin system probably was influenced by seawater inflowing and freshwater discharge from

rivers. All the different water and nutrient contributions can have been also mixed and redistributed via
122

probably cyclonic circulation within the basin, in a process like at the modern epeiric Hudson sea, where

primary productivity owes to both riverine nutrients and vertical mixing with nutrient-rich bottom-water

(Algeo et al., 2008).

The output of phosphorus was probably similar to other Phanerozoic mechanisms of

phosphogenesis, considering that organic matter degradation and sulfate reduction and iron redox

processes were possible mechanisms responsible for phosphorus precipitation. However, probably

adsorbed P into iron oxides/hidroxides was not a so effective process of phosphorus transport considering

a more widespread global anoxia condition.

Based on data set it seems clear that the development of marine anoxia was stimulated also by

terrestrial runoff. Even when the direct input of terrigenous material was limited, the influx of continental

runoff to the global ocean elsewhere likely increased the oceanic nutrient inventory. Upwelling of deep

waters to marine shelves can have recycled the terrestrial nutrients to areas that did not experience major

inputs of terrestrial runoff (like observed in the west coast of Peru today; Hutchins et al., 2002), extending

the area of oxygen-depleted water bodies. Consequently, recycling of sedimentary phosphorus in an

oxygen-depleted water column to sustain the levels of primary productivity and eutrophication may have

been more important for the prolongation of anoxic conditions during the Devonian. As a consequence,

shallower anoxia contributed in diffusing P to the water column decreasing the potential for P export to

the sediment.
123

Table 2 – Major oxides, TOC and total S (in weight %) of studied samples from outcrops and wells.
Associated host rock in parentheses. [org] – organic rich.
Fe2O3 TOTAL
Rock type Sample Al2O3 CaO K2 O MgO MnO Na2O P2O5 SiO2 TiO2 LOI TOC
(tot) S
Ironstone HHL-308 13.6 0.35 1.83 1.06 0.05 0.99 0.513 50.8 0.86 21.7 7.82 0.07 0.13
Ironstone HHL-312 4.79 0.14 1.95 0.42 0.14 0.6 0.185 53 0.51 33.2 4.56 0.1 0.04
Ironstone HHL-337 18.6 0.1 1.89 1.04 0.04 5.07 0.184 42.7 0.8 20.3 9.62 0.17 0.005
Ironstone HHL-661 7.6 1.21 1.33 3.66 0.16 1.51 0.37 37.5 0.42 28.3 16.9 0.24 0.32
Ironstone HHL-668 6.77 0.81 1.59 2.68 0.15 1.73 0.18 36.1 0.45 34.2 15.65 0.09 0.39
phosphatic ironstone HHL-593 6.54 1.71 1.09 3.75 0.41 0.18 1.1 29.6 0.36 38.9 16.03 0.26 0.31
phosphatic ironstone HHL-669 6.88 2.54 1.06 2.61 0.14 1.6 1.29 24.5 0.44 38.1 19.85 0.76 0.45
phosphatic ironstone HHL-678 8.5 2.28 1.56 4.07 0.32 0.24 1.19 24.9 0.57 33.8 22.01 0.7 0.33
phosphatic ironstone [org] HHL-686 4.77 4.42 0.89 2.52 0.49 0.55 2.95 12.4 0.32 47.1 22.31 1.08 0.28
Ironstone HHL-148 16.8 0.19 1.57 2.27 0.14 0.05 0.16 37.5 0.8 28.9 10.39 0.56 2.45
phosphatic ironstone HHL-149 14.4 1.25 1.7 2.03 0.16 0.05 1.02 46.1 0.81 23.4 8.71 0.53 0.5
Ironstone HHL-152 17.1 0.13 2.24 1.91 0.02 0.05 0.11 46.5 0.98 19 10.32 0.5 1.51
ironstone [org] HHL-153 12.9 1.03 1.19 2.89 0.38 0.05 0.78 29.4 0.67 37.1 13.94 1.57 0.35
ironstone [org] HHL-158 9.89 0.57 1.33 2.65 0.45 0.05 0.51 35.9 0.61 34.9 13.77 1.77 0.19
ironstone [org] HHL-163 15.4 0.99 1.49 2.51 0.23 0.11 0.8 37.7 0.79 29.3 11.03 1.08 0.49
ironstone [org] HHL-164 9.43 0.44 1.27 1.87 0.22 0.05 0.44 56.4 0.61 21.1 9.12 1.12 0.39
phosphatic ironstone [org] HHL-166 9.62 1.33 0.7 2.81 0.37 0.1 1.04 38 0.39 34.4 12.84 1.71 0.13
ironstone [org] HHL-174 5.06 1.37 0.37 2.33 0.54 0.05 0.97 37.3 0.27 36.6 13.98 2.24 0.12
phosphatic ironstone [org] HHL-175 7.66 1.75 0.36 3.1 0.46 0.05 1.28 25.4 0.42 43.9 15.8 1.7 0.88
phosphatic ironstone [org] HHL-180 10.1 2.06 1.08 2.75 0.34 0.05 1.6 33.2 0.5 34.9 12.5 1.85 0.16
ironstone [org] HHL-181 11.1 0.98 1.11 2.67 0.3 0.18 0.78 35.4 0.53 33.7 12.46 1.69 0.49
Ironstone HHL-609 8.42 1.07 1.35 3.02 0.2 0.82 0.49 49.2 0.53 25.9 9.46 0.28 0.14
Ironstone HHL-617 8.65 1.88 1.19 3.07 0.2 2.93 0.9 31.2 0.44 30.1 18.99 0.34 0.7
Ironstone HHL-646 10.9 1.66 1.67 3.99 0.33 0.25 0.95 25.1 0.57 39.1 15.33 0.16 0.22
Ironstone HHL-665 8.24 0.96 1.58 2.59 0.11 2.04 0.5 42.5 0.81 27.1 12.87 0.57 0.66
Ironstone HHL-671 5.62 1.84 1.16 1.47 0.27 0.92 0.27 31.5 0.73 50.7 4.11 0.24 0.03
Ironstone HHL-688 7.52 0.77 1.29 3.39 0.42 0.34 0.48 29.4 0.37 38.8 15.54 0.17 0.21
Ironstone HHL-416 4.27 1.24 1.27 0.43 0.14 0.05 0.959 40.4 0.35 49.1 2.23 0.025 0.01
ironstone [org] HHL-143 3.52 0.65 0.83 1.87 0.39 0.05 0.45 54.8 0.88 24.9 11 1.25 0.11
Ironstone HHL-182 16.9 0.24 1.58 2.87 0.16 0.05 0.23 37.5 0.72 30.7 10.29 0.95 0.61
Ironstone HHL-613 10 0.96 1.45 5.82 0.29 0.05 0.13 18 0.43 37.5 25.94 0.21 0.86

mudstone (Fb) HHL-624 19.5 0.18 3.14 0.97 0.03 6.48 0.17 57.4 1.2 4.25 7.82 0.59 0.4

Ironstone HHL-645 6.79 0.41 0.84 3.32 0.23 0.15 0.15 40.6 0.37 34.6 13.21 0.14 0.28

mudstone (Fb/Ip) HHL-658 21.8 0.33 2.97 1.68 0.05 0.5 0.16 53.5 1 7.92 10.04 0.47 0.58

mudstone (Fb/Ip) HHL-664 14.4 0.32 2.86 1.94 0.04 2.61 0.09 54.7 0.97 11.1 10.51 0.77 0.76

phosphatic ironstone HHL-650 10.3 5.14 1.67 3.67 0.27 0.22 3.3 23.6 0.5 31.3 20.66 0.13 0.51

phosphatic ironstone [org] HHL-682 10.8 7.41 1.8 3.89 0.18 0.49 5.05 24 0.52 25.7 18.77 1.01 1

phosphatic ironstone HHL-642 3.22 5.82 0.86 0.67 0.17 0.36 3.9 13.1 0.71 67.1 3.92 0.025 0.02
Ironstone HHL-643 6.97 0.74 1.27 1.74 0.1 0.35 0.33 32.4 0.61 51.4 4.74 0.025 0.01
Ironstone HHL-670 10.4 0.15 2.09 0.56 0.05 3.13 0.14 51.1 0.74 29.3 3.25 0.12 0.03
Ironstone HHL-672 6.11 0.69 1.04 3.87 0.34 1.08 0.09 36.3 0.3 32.3 17.7 0.15 0.25

phosphatic ironstone (Sb) HHL-654 5.81 5.49 0.87 3.62 0.3 0.54 3.77 20.7 0.28 40.5 17.66 0.24 0.62
sideritic/phosphatic concretion
(mudstone) HHL-367 14.1 0.52 1.55 0.55 0.2 0.05 2.004 26 0.55 47.6 7.09 0.07 0.02
sideritic/phosphatic concretion
(mudstone) HHL-368 14.9 4.63 1.7 0.81 0.12 0.05 4.6 30.3 0.63 32.2 9.4 0.09 0.03

phosphorite (phosph.grainstone) HHL-387 4.26 17.7 1.15 0.25 0.18 0.43 13.15 46.6 0.57 12.6 2.59 0.025 0.1
124

Table 2 (continuation) – Major oxides, TOC and total S (in weight %) of studied samples from outcrops
and wells. Associated host rock in parentheses. [org] – organic rich.
Fe2O3 TOTAL
Rock type Sample Al2O3 CaO K2 O MgO MnO Na2O P2O5 SiO2 TiO2 LOI TOC
(tot) S

phosphorite (Hewb) HHL-362 10.1 19.3 1.58 0.39 0.05 0.2 16.258 23.5 0.48 17.5 8.43 0.11 0.53

phosphatic mudstone/heterolithe HHL-405 9.81 7.16 2.28 0.97 0.08 3.27 4.58 58.9 0.57 11.1 1.33 0.025 0.07

phosphatic mudstone/heterolithe HHL-194 16.5 1.53 2.82 2.57 0.31 0.13 0.99 48.9 0.93 14.1 10.21 1.02 0.06

phosphatic ironstone HHL-324 12.3 2.19 2.34 1.02 0.18 0.18 2.427 38.9 0.7 32.7 7.76 0.1 0.04

phosphatic ironstone HHL-335 5.65 11.1 0.91 1.51 0.3 0.24 9.157 13.4 0.34 46.5 10.48 0.12 0.02

phosphatic ironstone HHL-333 6.67 2.54 1.4 0.84 0.16 0.52 2.47 52.2 0.6 27.4 5.86 0.025 0.005

phosphatic ironstone [org] HHL-155 8.73 2.52 1.01 2.95 0.48 0.05 1.81 22.4 0.53 41.6 18 2.96 0.44

phosphatic ironstone [org] HHL-159 13.8 2.96 1.17 2.82 0.32 0.05 2.48 23.7 0.63 37 14.98 2.03 0.29

phosphatic ironstone [org] HHL-161 9.63 3.9 0.83 3.28 0.47 0.05 2.89 22.5 0.45 40.5 16.8 2.48 0.13

phosphatic ironstone [org] HHL-170 5.37 6.96 0.85 2.39 0.43 0.13 4.93 25.3 0.32 38 16.07 2.02 2.09

phosphatic ironstone [org] HHL-173 5.01 7.73 0.57 2.04 0.52 0.1 5.61 27.7 0.33 34.2 15.39 2.49 0.64

phosphatic ironstone HHL-177 7.77 5.39 0.66 3.08 0.46 0.17 3.87 18.4 0.43 41.3 16.45 2.3 0.41

phosphatic ironstone HHL-612 14.9 3.71 2.11 4.57 0.22 0.05 2.03 23.3 0.56 28.4 21.81 0.27 0.45

phosphatic ironstone HHL-356 4.84 1.49 0.98 0.32 0.03 0.05 1.534 45.5 0.31 42.3 2.66 0.08 0.02

phosphatic ironstone [org] HHL-142 3.45 3.21 0.7 2.66 0.59 0.05 2.22 33.1 0.44 36.2 17.23 2.86 0.2

phosphatic ironstone [org] HHL-145 7.75 2.19 0.65 3.88 0.67 0.05 1.63 18.2 0.52 45.7 20.31 2.57 0.3

phosphatic mudstone/heterolithe HHL-315 10.8 5.51 3.01 0.46 0.01 1.54 4.664 59.4 0.52 9.9 4.43 0.07 0.32

phosphatic mudstone/heterolithe HHL-407 10.8 12.5 2.45 1.52 0.04 2.33 9.001 50.5 0.66 6.31 3.64 0.025 0.03

phosphatic mudstone/heterolithe HHL-412 7.99 11.2 2.67 1.45 0.05 1.18 8.14 58.1 0.51 5.19 3.1 0.025 0.04
phosphatic mudstone/heterolithe
[nodules] HHL-383 10.7 12.9 1.68 0.56 0.05 0.26 10.28 33.2 0.67 21.8 7.23 0.08 0.15
sideritic/phosphatic concretion
(mudstone) HHL-378 15.1 5.24 1.95 0.68 0.19 0.15 4.242 32.8 0.64 31.5 6.97 0.06 0.06

phosphorite [concretion] HHL-384 7.9 37 1.13 0.43 0.03 0.47 28.015 16 0.46 2.49 4.51 0.13 0.14

phosphorite [nodules] HHL-385 10.9 19.7 1.69 0.65 0.05 0.3 15.345 28.3 0.6 14.3 6.16 0.12 0.09

phosphorite [concretion] HHL-389 12.9 25.9 1.62 0.62 0.02 0.34 20.96 25.2 0.68 3.59 6.4 0.025 0.08

phosphatic ironstone HHL-392 8.72 9.65 1.55 0.39 0.09 0.25 7.928 30.8 0.59 34.1 5.93 0.09 0.05

phosphatic ironstone (Sb) HHL-605 8.61 2 1.51 2.93 0.17 0.98 1.22 46.9 0.68 24.4 11.52 0.7 0.38

phosphatic ironstone (Sb) HHL-662 11.7 1.95 1.81 2.8 0.28 0.45 1.07 34.4 0.66 27.7 18.14 0.28 0.06

phosphatic ironstone (Sb) HHL-689 4.18 6 0.18 3.75 1.59 0.12 4.37 23 0.25 42.7 12.3 0.3 0.95
siderite concretion HHL-316 16 0.21 2.2 0.95 0.23 2.99 0.86 40.3 0.9 24.9 8.53 0.06 0.01
siderite concretion HHL-322 10.4 0.21 1.74 1.14 0.17 0.15 0.78 28.8 0.62 46 9.86 0.025 0.005
siderite concretion HHL-396 16.2 0.52 2.85 0.97 0.06 0.29 0.37 49.6 0.87 21.1 7.52 0.025 0.02
siderite concretion HHL-318 16.9 0.18 2.34 1.35 0.14 0.72 0.246 47.9 0.98 19.3 8.28 0.08 0.03
sideritic/phosphatic concretion
(mudstone) HHL-305 16.9 1.61 2.33 1.32 0.27 0.12 1.339 39.2 0.79 25.5 9.35 0.07 0.005
sideritic/phosphatic concretion
(mudstone) HHL-366 13.7 9.61 1.65 0.55 0.12 0.16 7.78 28.6 0.59 28.9 6.85 0.12 0.06
sideritic/phosphatic concretion
(mudstone) HHL-369 11.8 1.73 2.72 0.37 0.09 0.12 1.547 62.6 0.88 14.2 4.14 0.025 0.02
mudstone (Fb) MA-T-360
-- -- -- -- -- -- -- -- -- -- -- 0,25 0,12
mudstone (Fb) MA-T-324
-- -- -- -- -- -- -- -- -- -- -- 0,3 0,74
MA-T-269
mudstone (Fb) -- -- -- -- -- -- -- -- -- -- -- 1,92 2,46
MA-T-105
mudstone (Fb) -- -- -- -- -- -- -- -- -- -- -- 1,29 1,23
MA-T-107
mudstone (Fb) -- -- -- -- -- -- -- -- -- -- -- 1,51 0,025
125

Table 2 (continuation) – Major oxides, TOC and total S (in weight %) of studied samples from outcrops
and wells. Associated host rock in parentheses. [org] – organic rich.
Fe2O3 TOTAL
Rock type Sample Al2O3 CaO K2 O MgO MnO Na2O P2O5 SiO2 TiO2 LOI TOC
(tot) S
MA-T-111
bioturbated heterolithe -- -- -- -- -- -- -- -- -- -- -- 1,3 0,16

bioturbated heterolithe MA-T-137 -- -- -- -- -- -- -- -- -- -- 1,05 0,66

bioturbated heterolithe MA-T-139 -- -- -- -- -- -- -- -- -- -- -- 1,06 0,1


MA-T-141
bioturbated heterolithe -- -- -- -- -- -- -- -- -- -- -- 2,85 1,27
MA-T-141B
bioturbated heterolithe -- -- -- -- -- -- -- -- -- -- -- 2,88 1,26
ironstone (sandstone) MA-T-326
-- -- -- -- -- -- -- -- -- -- -- 0,15 2,69
organic sandstone/ironstone MA-T-362
-- -- -- -- -- -- -- -- -- -- -- 0,3 0,95

Greenhouse, increased weathering and terrestrial runoff conditions in the global Devonian

probably lead the ocean towards a state whereby oxygen depletion could easily occur across much of the

marine shelf area and/or global ocean (Carmichael et al., 2014) or that anoxic conditions may have arisen

via migration of expanded oxygen-minimum zones during marine transgressions (Bond et al., 2004). The

Corg/P ratios of marine sediments are strongly influenced by benthic redox conditions. Algeo and Ingall

(2007) suggested that a secular variation regulated most of the predominant values of C/P ratios, where

epochs characterized by high Corg/P ratios had poorly ventilated seas, and epochs characterized by low

Corg/P ratios had well-ventilated seas (e.g., Algeo and Ingall, 2007). That was the case for the Devonian

greenhouse conditions where most of the studied samples in different parts of the world revealed high C/P

ratios, even higher than for Cretaceous greenhouse conditions, where huge phosphorite deposits were

described (Cook and McElhinny, 1979; Pufahl et al., 2003). Higher CO2 conditions for the Devonian in

relation to the Cretaceous period is also evidenced by CO2 proxies (Royer, 2006) where oxygen depletion

simultaneously enhances preservation of organic C and diffusive loss of remineralized organic P (Benitez-

Nelson, 2000). Large-scale P cycling has been also invoked to explain unusual periods of intense

productivity (e.g., Föllmi, 1996; Murphy et al., 2000;). Like observed here, global REE data also suggests

that anoxic bottom water conditions prevailed for the Devonian. Because of this, an additional economic

potential for REE is recognized for Devonian phosphorites.


126

Table 3 – Selected trace elements (in ppm) and calculated ratios of studied samples from outcrops and
wells. Associated host rock in parentheses. [org] – organic rich.
Rock type Sample Ba Co Cu Mo Ni Sr Th U Zr Rb Cd Cr V V/Cr V/(V+Ni)

ironstone (Hewb) HHL-308 568 21,5 24 14 54 240 39,6 49,87 643,3 72,5 0,03 41 159 3,878 0,797
ironstone (Hewb) HHL-312 416 7,3 12 4 16 60,7 13,9 8,95 548,1 45,1 0,05 25 132 5,280 0,974
ironstone (Hewb) HHL-337 463 19,2 32 5 47 56,9 12,6 5,79 129,4 86 0,01 35 73 2,086 0,753
ironstone (Hewb) HHL-661 196 33 18 11 45 73 9,6 3,58 198 48,8 0,65 20 230 11,500 0,866
ironstone (Hewb) HHL-668 304 11,9 14 1 18 46 5,6 3,8 141 42,8 0,005 13 145 11,154 0,952
phosphatic ironstone
(Hewb) HHL-593 240 14,1 9 18 24 95 20,7 55,99 101 38,7 0,03 59 370 6,271 0,968
phosphatic ironstone
(Hewb) HHL-669 187 9,1 12 34 19 105 11,7 3,94 303 35,7 0,005 23 226 9,826 0,960
phosphatic ironstone
(Hewb) HHL-678 233 9,5 11 5 39 96 15,6 26,9 108 52,8 0,005 18 118 6,556 0,834
phosphatic ironstone
(Hewb) [org] HHL-686 138 5,7 11 1 20 190 4,9 192,55 67 34,1 0,02 10 51 5,100 0,836
ironstone HHL-148 332 39 19 42 48 120 20,2 3,45 176 63,7 0,005 86 378 4,395 0,919
phosphatic ironstone HHL-149 370 30,4 21 4 45 173 22,2 6,06 342 59,8 0,005 86 348 4,047 0,925
ironstone HHL-152 432 27,2 19 12 45 97 14,9 4,97 229 85,1 0,005 52 126 2,423 0,809
ironstone [org] HHL-153 362 36,3 14 6 49 115 25,6 6,09 218 47,9 0,02 105 506 4,819 0,938
ironstone [org] HHL-158 308 16,6 16 2 33 116 33,2 6,24 370 41,9 0,005 152 652 4,289 0,969
ironstone [org] HHL-163 333 23,9 17 6 38 314 27 6,4 242 65,8 0,02 112 506 4,518 0,947
ironstone [org] HHL-164 368 7,9 9 4 18 428 21,5 3,76 360 48 0,36 70 186 2,657 0,941
phosphatic ironstone [org] HHL-166 142 10,6 7 2 19 152 14,4 4,71 140 31 0,05 110 366 3,327 0,969
ironstone [org] HHL-174 120 7 2,5 1 22 94 25,4 7,18 170 15 0,005 159 890 5,597 0,981
phosphatic ironstone [org] HHL-175 56 33,4 6 2 55 108 34,4 6,4 253 18,8 0,005 257 954 3,712 0,952
phosphatic ironstone [org] HHL-180 162 28,2 12 15 42 134 30,9 7,76 194 47,2 0,005 132 784 5,939 0,953
ironstone [org] HHL-181 263 41 14 38 72 134 26,2 5,13 217 47,9 1,22 142 762 5,366 0,924
ironstone HHL-609 197 18,2 21 49 41 76 14 14,81 401 43,2 0,07 39 161 4,128 0,827
ironstone HHL-617 218 22,4 14 33 33 109 18,8 6,82 181 46 0,03 42 252 6,000 0,923
ironstone HHL-646 209 12,6 22 12 38 148 17,4 3,44 163 58,6 0,03 44 304 6,909 0,926
ironstone HHL-665 309 15,2 13 23 22 64 23,1 5,24 961 47,9 0,005 42 207 4,929 0,939
ironstone HHL-671 555 64 19 11 60 270 17 12,44 146 33,8 0,17 35 515 14,714 0,930
ironstone HHL-688 205 9,1 10 3 24 108 23,6 54,88 180 41,6 0,01 50 184 3,680 0,947
ironstone HHL-416 371 37,3 48 38 44 91,5 24,8 17,68 165,1 36,9 0,1 23 840 36,522 0,987
ironstone [org] HHL-143 207 5,2 2,5 5 14 116 27,1 4,9 1545 22,1 0,005 43 186 4,326 0,969
ironstone HHL-182 311 48,8 19 11 52 135 30,2 4,55 167 75,6 0,005 131 609 4,649 0,926
ironstone HHL-613 135 9,1 14 23 44 53 12,8 2,05 77 54,4 0,14 16 97 6,063 0,802
mudstone (Fb) HHL-624 482 14,2 31 27 38 68 15,6 4,53 191 118,5 0,005 16 29 1,813 0,487
ironstone HHL-645 142 11,1 14 36 23 45 16,1 2,37 153 27,9 0,3 48 280 5,833 0,965
mudstone (Fb/Ip) HHL-658 472 17 33 4 48 123 17,2 4,17 184 120,9 0,01 16 34 2,125 0,566
mudstone (Fb/Ip) HHL-664 452 14,6 21 4 38 113 13,9 3,22 305 91,6 0,02 20 54 2,700 0,685
127

Table 3 (continuation) – Selected trace elements (in ppm) and calculated ratios of studied samples from
outcrops and wells. Associated host rock in parentheses. [org] – organic rich.
Rock type Sample Ba Co Cu Mo Ni Sr Th U Zr Rb Cd Cr V V/Cr V/(V+Ni)
phosphatic ironstone HHL-650 238 8,7 25 7 42 109 14,5 27,95 119 56,2 0,01 20 88 4,400 0,773
phosphatic ironstone
[org] HHL-682 263 15,3 19 4 56 227 17,6 158,2 98 62,1 0,005 22 104 4,727 0,737
phosphatic ironstone HHL-642 252 72,8 23 6 80 596 52,7 49,23 115 25 0,42 54 422 7,815 0,884
ironstone HHL-643 464 36,6 19 50 61 217 13,4 4,53 111 39,8 0,03 37 916 24,757 0,962
ironstone HHL-670 577 27,7 8 25 43 108 18,9 4,29 222 69,2 0,03 94 209 2,223 0,897
ironstone HHL-672 204 135,4 8 46 74 61 6,1 1,59 146 32,2 0,01 39 225 5,769 0,957
phosphatic ironstone (Sb) HHL-654 130 19,4 17 5 23 297 40,1 38,27 83 27,7 0,005 35 446 12,743 0,971
sideritic/phosphatic
concretion (mudstone) HHL-367 2850 14,1 22 1 35 175,9 13,1 34,31 155,5 67,3 0,05 7 192 27,429 0,981
sideritic/phosphatic
concretion (mudstone) HHL-368 2642 28,6 50 3 58 213,7 19,1 22,59 261,3 68,9 0,09 17 89 5,235 0,738
phosphorite
(phosph.grainstone) HHL-387 420 34,4 17 28 73 343 16,7 49 693 31,1 0,02 19 49 2,579 0,475
phosphorite (Hewb) HHL-362 2104 27,8 69 9 81 645,2 14,4 23,81 228,5 53,4 0,67 18 51 2,833 0,477
phosphatic
mudstone/heterolithe HHL-405 553 14,3 37 2 39 182 113 8,37 253 54,1 0,06 43 167 3,884 0,847
phosphatic
mudstone/heterolithe HHL-194 438 15 13 1 34 147 17,7 3,19 213 121,1 0,02 57 106 1,860 0,774
phosphatic ironstone HHL-324 343 15,6 23 2 99 87 16,3 28,05 218 81,1 0,26 9 65 7,222 0,459
phosphatic ironstone HHL-335 438 63,7 20 14 56 396,1 72,5 691,48 140,4 40,7 0,05 37 279 7,541 0,919
phosphatic ironstone HHL-333 441 27,8 8 5 42 160 100,4 27,09 630 38,9 0,03 48 162 3,375 0,824
phosphatic ironstone
[org] HHL-155 226 55,5 10 4 61 99 18,7 6,65 204 36 0,005 92 396 4,304 0,896
phosphatic ironstone
[org] HHL-159 273 21,8 22 5 44 244 34,4 11,91 228 47,9 0,005 219 729 3,329 0,966
phosphatic ironstone
[org] HHL-161 203 21,9 9 15 34 202 27,8 7,56 188 34,9 0,005 124 505 4,073 0,952
phosphatic ironstone
[org] HHL-170 199 23,6 12 6 39 108 7,6 11,49 96 38,9 0,005 31 84 2,710 0,759
phosphatic ironstone
[org] HHL-173 132 31 2,5 14 36 255 8,9 16,46 111 25,3 0,005 46 178 3,870 0,867
phosphatic ironstone HHL-177 155 58,8 11 9 58 244 21 6,78 145 34,1 0,005 102 709 6,951 0,930
phosphatic ironstone HHL-612 183 10,4 24 17 46 96 19,7 12,9 95 74,2 0,02 20 73 3,650 0,743
phosphatic ironstone HHL-356 776 20,8 8 23 43 77,4 10,8 9,92 233,4 32,1 0,06 3 169 56,333 0,890
phosphatic ironstone
[org] HHL-142 175 17,4 2,5 3 26 90 11,2 5,01 586 21 0,02 23 182 7,913 0,921
phosphatic ironstone
[org] HHL-145 164 32,9 13 7 49 127 23,8 7,7 266 26,4 0,005 130 570 4,385 0,947
phosphatic
mudstone/heterolithe HHL-315 984 6,8 19 8 23 270,1 116,4 213,01 243 69,8 0,06 16 47 2,938 0,742
phosphatic
mudstone/heterolithe HHL-407 604 11,2 17 1 20 323,2 319,2 14,18 236,8 55,3 0,08 32 73 2,281 0,840
phosphatic
mudstone/heterolithe HHL-412 698 8,6 18 1 18 294,8 278,9 11,27 200,2 51,7 0,08 22 59 2,682 0,819
phosphatic
mudstone/heterolithe
[nodules] HHL-383 770 25,2 41 7 65 1089,5 19 71,29 381,4 59,8 0,14 16 63 3,938 0,582
sideritic/phosphatic
concretion (mudstone) HHL-378 518 25,1 30 2 112 129,5 13,7 25,63 209,9 76,6 0,11 13 92 7,077 0,561
phosphorite [concretion] HHL-384 866 9,8 17 1 30 596,5 11 156,28 171,6 49,8 0,47 29 55 1,897 0,780
phosphorite [nodules] HHL-385 700 22,6 20 2 96 362,2 11,7 48,98 606,1 60,5 0,43 21 64 3,048 0,635
phosphorite [concretion] HHL-389 3266 9,5 13 2 34 626 12,8 336,77 814 61,1 0,18 32 47 1,469 0,748
phosphatic ironstone HHL-392 337 48,1 66 1 82 794,3 8,7 265,06 209,5 48,5 0,27 6 71 11,833 0,537
phosphatic ironstone (Sb) HHL-605 253 12,2 20 5 30 118 20,6 23,66 398 45,9 0,01 25 144 5,760 0,887
phosphatic ironstone (Sb) HHL-662 302 12 21 14 33 81 12,1 6,04 146 62,7 0,005 21 59 2,810 0,777
phosphatic ironstone (Sb) HHL-689 24 55,1 14 63 56 295 216,1 131,55 141 7,8 0,24 43 478 11,116 0,924
siderite concretion HHL-316 661 22,1 27 7 44 90 11,8 90,11 202 71,3 0,03 21 85 4,048 0,730
siderite concretion HHL-322 386 21 67 4 58 54 8,2 63,79 169 65,5 0,09 24 70 2,917 0,633
siderite concretion HHL-396 544 30,1 70 4 52 197 12,7 14,46 258 85 0,02 42 165 3,929 0,818
siderite concretion HHL-318 544 14,2 26 5 46 82,9 16,3 13,07 196,9 76,2 0,02 23 83 3,609 0,712
sideritic/phosphatic
concretion (mudstone) HHL-305 505 32,4 10 1 56 141,8 17,1 21,62 185,4 90,3 0,05 22 70 3,182 0,652
sideritic/phosphatic
concretion (mudstone) HHL-366 974 22 42 3 45 224 12 22,16 237 61,7 0,09 31 83 2,677 0,772
sideritic/phosphatic
concretion (mudstone) HHL-369 573 17,8 55 1 63 104,4 18,6 8,91 556,7 75,6 0,04 12 42 3,500 0,713
128

Table 4 – Concentrations of rare earth elements (in ppm) of studied samples from outcrops and wells.
Associated host rock in parentheses. [org] – organic rich.
Rock type Sample La Ce Pr Nd Eu Sm Gd Tb Dy Ho Er Tm Yb Lu Y

ironstone (Hewb) HHL-308 114,9 283,1 32,68 127,4 4,89 25,2 21,8 3,22 17,42 3,43 10,14 1,5 8,9 1,2 86,57
ironstone (Hewb) HHL-312 34,5 72,7 9,33 39,1 1,9 9 9,33 1,28 6,97 1,28 3,27 0,47 3 0,45 33,66
ironstone (Hewb) HHL-337 45,4 84,5 9,91 36,3 1,28 6,3 5,49 0,82 4,73 0,97 3,03 0,47 3,1 0,46 25,89
ironstone (Hewb) HHL-661 36,4 96,4 10,63 38,7 1,9 8,2 7,88 1,32 8,77 1,71 1,9 0,81 6 0,87 38,69
ironstone (Hewb) HHL-668 19,3 41 4,94 17,9 0,74 3,9 3,37 0,52 3,08 0,62 0,74 0,28 2 0,33 15,07
phosphatic ironstone
(Hewb) HHL-593 33,5 100,8 10,46 42 2,65 10,8 11,16 2,11 15,28 3,46 2,65 1,82 13 1,95 87,76
phosphatic ironstone
(Hewb) HHL-669 38,6 94,4 11,84 46,9 2,47 10,8 10,55 1,68 10,45 2,04 2,47 0,78 5,5 0,74 48,59
phosphatic ironstone
(Hewb) HHL-678 40,8 91,4 11,49 48 3,49 12,9 17,87 3,15 19,47 3,73 3,49 1,09 5,6 0,79 106
phosphatic ironstone
(Hewb) [org] HHL-686 29,5 57,6 7,25 29,3 1,73 7,2 9,07 1,43 8,58 1,89 1,73 0,68 4 0,58 65,45
ironstone HHL-148 51,5 124,3 12,99 46,8 1,68 8,2 6,44 0,94 5,3 1,07 3,28 0,53 3,8 0,58 26,52
phosphatic ironstone HHL-149 66,7 185,3 20,9 87,1 4,96 21,1 21,7 3,34 18,9 3,46 9,18 1,25 7,9 1,11 81,21
ironstone HHL-152 39,4 83,4 9,77 36,4 1,39 6,5 5,46 0,81 5,19 1,08 3,59 0,59 4,4 0,66 27,94
ironstone [org] HHL-153 72,1 202,5 22,93 94,5 4,99 21,7 21,16 3,16 17,05 2,9 7,7 1,03 6,6 0,92 64,55
ironstone [org] HHL-158 56,3 162,1 18,13 70,5 3,61 16,1 16,01 2,62 16,66 3,24 9,86 1,42 9,6 1,37 76,28
ironstone [org] HHL-163 77,3 210,2 25,57 105,1 4,98 23,2 20,78 3,03 16,44 2,83 7,89 1,04 7,1 0,99 72,06
ironstone [org] HHL-164 54,2 134 15,78 63,8 2,82 13,9 11,67 1,57 7,76 1,31 3,38 0,44 3 0,41 33,4
phosphatic ironstone
[org] HHL-166 53,5 161,9 17,5 74,6 4,33 18,8 18,53 2,82 15,73 2,74 7,01 0,88 5,5 0,76 69,43
ironstone [org] HHL-174 55,4 207 22,74 95,1 5,12 24,2 21,42 3,47 20,1 3,62 10,43 1,46 9,9 1,36 84,88
phosphatic ironstone
[org] HHL-175 90,4 317 34,73 139,1 6,62 30,8 26,77 4,13 24,28 4,55 13,03 1,85 12,6 1,74 109
phosphatic ironstone
[org] HHL-180 67,6 251,4 29,12 131,1 7,9 35,9 33,29 5,3 29,81 5,25 14,16 1,83 11,9 1,57 117
ironstone [org] HHL-181 61,4 196 19,76 77,9 3,82 17,9 17,04 2,86 15,66 3,07 9,03 1,32 9 1,3 70,96
ironstone HHL-609 25,9 64,1 7,71 31,9 1,94 8,4 9,09 1,54 9,91 2 1,94 0,83 5,6 0,77 48,44
ironstone HHL-617 43,9 112,4 13,52 55 3,24 14,2 14,34 2,29 14,01 2,61 3,24 1,03 6,7 0,9 59,25
ironstone HHL-646 55,6 137,3 16,14 64,9 3,7 14,8 15,58 2,43 14,33 2,62 3,7 1,06 6,3 0,82 60,53
ironstone HHL-665 49,1 107,7 13,25 50 2,06 11 10,34 1,61 9,5 1,96 2,06 0,87 6 0,88 46,72
ironstone HHL-671 26,8 71,1 7,73 29,1 1,69 6,6 6,75 1,26 7,94 1,46 1,69 0,51 3,1 0,44 31,75
ironstone HHL-688 22,7 62,9 7,55 31,9 1,87 7,9 8,39 1,39 8,57 1,83 1,87 0,75 5,2 0,73 41,92
ironstone HHL-416 33,4 118 14,91 65,6 4,32 17,4 16,74 2,69 15,05 2,59 6,4 0,8 4,6 0,56 63,66
ironstone [org] HHL-143 59,7 146,7 15,2 57,3 1,81 10,9 9,98 1,5 8,55 1,72 5,13 0,8 5,7 0,86 42,61
ironstone HHL-182 60,2 156,2 14,73 51,1 1,91 9 7 0,99 5,75 1,14 3,93 0,68 5,4 0,79 28,23
ironstone HHL-613 33,1 56 6,22 21,5 0,87 4 3,64 0,52 3,53 0,64 0,87 0,27 2,1 0,29 16,52
mudstone (Fb) HHL-624 54,3 105,6 12,17 45,8 1,75 8,4 6,72 0,95 5,58 1,1 1,75 0,48 3,2 0,47 27,63
ironstone HHL-645 33,5 69,7 7,58 27,3 1,15 5,8 4,74 0,75 4,27 0,77 1,15 0,38 2,5 0,4 17,59
mudstone (Fb/Ip) HHL-658 50,2 97,5 11,48 42,4 1,76 8,6 7,43 1,12 6,62 1,23 1,76 0,55 3,6 0,49 33,24
mudstone (Fb/Ip) HHL-664 45,8 89 10,99 40,5 1,52 8,1 6,95 1,05 5,88 1,11 1,52 0,5 3,5 0,49 29,74
phosphatic ironstone HHL-650 40,1 84,9 10,71 46,7 3,63 13,2 18,97 3,06 19,23 3,92 3,63 1,34 7,6 1 123
phosphatic ironstone
[org] HHL-682 73 187,5 23,45 106,8 8,92 30,8 48,59 8,38 54,08 10,78 8,92 3,23 16,7 2,21 311,9
phosphatic ironstone HHL-642 41,2 160,1 22,19 113,3 10,63 38,6 49,14 8,85 49,8 8,25 10,63 2,3 12 1,64 179,3
ironstone HHL-643 36,4 89,8 9,45 34,3 1,69 7,2 6,73 1,2 7,23 1,43 1,69 0,58 4,1 0,54 31,4
ironstone HHL-670 33,3 75,5 8,79 31,7 1,4 6,2 5,28 0,97 4,98 1,12 1,4 0,57 2,9 0,52 22,78
ironstone HHL-672 31,5 74,1 7,33 23,7 0,82 3,9 2,97 0,52 3,15 0,69 0,82 0,33 2,5 0,37 14,5
phosphatic ironstone
(Sb) HHL-654 39,7 118,2 14,82 64,6 5,84 20,1 28,69 5,91 41,23 8,52 5,84 2,8 15,1 1,9 202,4
sideritic/phosphatic
concretion (mudstone) HHL-367 340,3 833,5 95,7 358,7 14,97 72,4 60,2 8,03 39,96 6,82 16,18 2,01 12,5 1,71 163,6
sideritic/phosphatic
concretion (mudstone) HHL-368 319 915,9 117,42 487,2 23,9 108,5 103,42 14,38 70 12,07 29,67 3,63 21,5 2,79 315,4
phosphorite
(phosph.grainstone) HHL-387 81,7 244 33,71 159,1 10,57 43,4 55,61 7,92 45,21 8,58 22,4 2,59 14,4 2 298,7
phosphorite (Hewb) HHL-362 119,8 394,5 53,31 256,5 17,36 66,2 89,65 12,41 72,1 14,21 35,67 4,25 23 3,12 485,3
129

Table 4 (continuation) – Concentrations of rare earth elements (in ppm) of studied samples from outcrops
and wells. Associated host rock in parentheses. [org] – organic rich.
Rock type Sample La Ce Pr Nd Eu Sm Gd Tb Dy Ho Er Tm Yb Lu Y
phosphatic
mudstone/heterolithe HHL-405 85,5 233,9 33,43 163,1 11,95 47,3 54,6 7,83 40,6 6,75 15,57 1,75 8,9 1,1 199,4
phosphatic
mudstone/heterolithe HHL-194 58,8 139,2 18,05 86,3 7,39 28,4 35,23 4,82 23,81 3,82 8,56 0,88 5,1 0,69 107,5
phosphatic ironstone HHL-324 89,8 193,5 26,31 119,9 8,51 31,8 47,43 7,61 47,2 9,04 22,15 2,54 13,6 1,9 300,6
phosphatic ironstone HHL-335 139,1 456,4 55,01 238,6 14,82 59,4 70,42 12,07 84,19 20,85 82,22 16,15 116,6 17,2 571,1
phosphatic ironstone HHL-333 104 300,6 40,65 176,6 8,95 42,4 42,86 6,48 35,81 6,24 15,26 1,71 8,7 1,09 158,8
phosphatic ironstone
[org] HHL-155 55,9 181,3 24,9 112,2 6,73 29 30,17 4,72 27,83 4,94 12,75 1,52 8,6 1,16 108,5
phosphatic ironstone
[org] HHL-159 108,3 343,2 43,07 192,4 12,33 52,4 53,39 8,22 45,57 7,82 19,67 2,42 14,8 2,01 168
phosphatic ironstone
[org] HHL-161 55 176,7 20,89 98,1 8,3 33,8 37 6,13 35,8 6,58 17,79 2,17 12,9 1,68 172,4
phosphatic ironstone
[org] HHL-170 15 39,6 5,35 30,4 5,93 17,3 34,26 6,5 38,3 7,74 19,34 2,37 12,5 1,56 213,2
phosphatic ironstone
[org] HHL-173 45,9 188,9 27,52 148,2 12,71 49,8 58,7 9,44 54,56 9,34 23,84 2,71 14,9 1,77 252,7
phosphatic ironstone HHL-177 70,2 255 32,52 153,1 10,97 48,6 49,08 7,87 43,83 7,28 18,72 2,3 14 1,8 166,9
phosphatic ironstone HHL-612 53,3 115,7 13,27 55,4 3,86 14,2 20,26 3,38 20,35 3,97 3,86 1,16 6,6 0,86 112,9
phosphatic ironstone HHL-356 90,8 203,3 29,58 120,6 5,23 24,7 22,94 3,18 16,4 2,88 7,28 0,88 5,1 0,65 75,25
phosphatic ironstone
[org] HHL-142 32,8 97,5 11,34 53,9 5,19 18,9 26,94 4,22 25,46 4,48 11,69 1,45 8,3 1,11 112,6
phosphatic ironstone
[org] HHL-145 62,4 206,8 21,32 89,6 5,15 22,2 22,72 3,61 20,53 3,73 10,4 1,43 9,3 1,28 86,24
phosphatic
mudstone/heterolithe HHL-315 35,9 90,7 14,23 75,5 8,94 27,3 48,74 10,96 81,46 16,16 37,02 3,9 19,1 2,39 397,4
phosphatic
mudstone/heterolithe HHL-407 46,1 158,1 17,72 94,2 11,25 37,8 53,47 8,52 47,49 7,97 18,57 2,14 11,1 1,35 229,9
phosphatic
mudstone/heterolithe HHL-412 35,2 113,8 13,45 72,6 9,51 30,4 45,6 7,48 42,25 7,17 16,73 1,93 9,9 1,19 206,3
phosphatic
mudstone/heterolithe
[nodules] HHL-383 78,3 214,7 30,68 156,5 12,46 45,8 62,09 9,24 52,69 9,85 24,96 2,93 16,2 2,27 314,1
sideritic/phosphatic
concretion (mudstone) HHL-378 64,4 166,9 20,95 100,3 7,44 28,1 38,04 5,65 31,63 5,95 14,48 1,71 9,8 1,29 180,2
phosphorite
[concretion] HHL-384 163,6 450,2 63,29 348,1 30,28 106,5 179,6 26,91 154,87 30,47 74,79 8,64 45,7 6,26 964,2
phosphorite [nodules] HHL-385 117,7 304,4 39,22 185 13,42 48,8 73,9 10,49 59,74 12,05 30,02 3,54 19,1 2,6 416,6
phosphorite
[concretion] HHL-389 398,9 1017,9 122,16 564,4 42,2 162,2 237,86 36,6 206,58 36,7 89,88 10,66 59 7,72 1088
phosphatic ironstone HHL-392 46,4 121,6 16,91 79,9 6,51 23,3 36,18 5,54 32,43 6,41 16,24 1,94 10,1 1,44 207
phosphatic ironstone
(Sb) HHL-605 37 82,1 10,39 41,3 2,47 10,4 12,3 2,32 16,15 3,3 2,47 1,13 7,1 1,06 83,1
phosphatic ironstone
(Sb) HHL-662 40,4 85,6 10,8 44,3 2,89 12 13,68 2,04 11,78 2,05 2,89 0,67 3,8 0,56 62,18
phosphatic ironstone
(Sb) HHL-689 63,3 234,3 27,59 127,6 10,97 40,4 56,41 13,82 108,93 23,1 10,97 6,65 33,6 3,92 538
siderite concretion HHL-316 56,9 137,4 17,58 72,9 3,49 15,9 14,59 2,11 11,41 2,06 5,66 0,79 5,5 0,82 50,53
siderite concretion HHL-322 17,5 24,6 2,57 9,2 0,62 2,3 2,76 0,52 3,47 0,65 1,76 0,26 1,8 0,28 13,38
siderite concretion HHL-396 44,3 92,6 11,2 43,7 2,05 9 9,44 1,35 7,67 1,43 4,07 0,53 3,4 0,48 38,39
siderite concretion HHL-318 56,5 84,9 8,93 31,5 1,06 5 4,75 0,75 4,64 0,97 3,19 0,49 3,5 0,56 26,68
sideritic/phosphatic
concretion (mudstone) HHL-305 67,7 150 17,55 73 4,38 18,1 22,72 3,27 17,85 3,21 8,25 1,05 6,1 0,88 91,89
sideritic/phosphatic
concretion (mudstone) HHL-366 133,9 430,6 53,58 259,5 14,99 60,7 76,54 10,31 56,51 10,62 26,45 3,04 16,7 2,36 353,8
sideritic/phosphatic
concretion (mudstone) HHL-369 70,8 152,2 18,08 78,7 3,6 15,7 18,58 2,62 14,66 3,08 8,22 1,14 6,8 0,97 98,89
heterolithe HHL-606 43,7 81,8 9,73 35,3 1,48 7,2 6,33 0,98 6,16 1,23 3,71 0,5 3,3 0,48 31,75
mudstone HHL-590 41,4 78,8 8,96 32,1 1,33 6 5,15 0,82 4,92 0,96 2,52 0,39 2,8 0,37 23,98
mudstone HHL-651 64,4 117,1 13,82 50,5 2 9,2 7,62 1,09 6,7 1,33 4,01 0,56 3,9 0,59 33,77
mudstone HHL-663 31,4 62,5 7,76 28,3 1,15 5,8 5,19 0,85 5,46 1,06 3,39 0,46 3,1 0,44 27,69
mudstone HHL-680 55,1 108,2 13,14 46,8 1,8 9 7,65 1,2 6,92 1,38 4 0,62 3,8 0,54 33,28
mudstone HHL-681 48,4 99,7 12,22 45,1 2,22 9,8 9,12 1,35 7,45 1,42 3,99 0,54 3,6 0,53 34,96
mudstone HHL-687 40,4 84,2 9,69 37,4 1,95 7,9 8,92 1,44 8,52 1,63 4,5 0,56 3,4 0,48 43,95
mudstone HHL-136 53,8 105,6 12,78 49,7 2,25 10,1 9,85 1,44 8,29 1,6 4,44 0,63 4,1 0,6 42,27
mudstone HHL-596 50 101,2 11,93 44,7 1,85 9 7,95 1,16 6,85 1,29 3,89 0,56 3,8 0,54 33,34
130

Table 5 – C/P molar ratios for organic richer host rock sediments.
Lithofacies Sample P (ppm) Corg% C/P ratio (molar)

mudstone (flooding) 4348-MA-T-269 427 1.92 116,0468384


mudstone (flooding) 4348-MA-T-273 223 0.66 76,38340807
bioturbated heterolithe 4348-MA-T-275 244 0.63 66,63627049
bioturbated heterolithe (condensed zone) 4348-MA-T-358 103 0.61 152,8454693
bioturbated heterolithe (condensed zone) 4348-MA-T-359 28400 1.08 0,981443662
heterolithe 4348-MA-T-104 145 1.21 59,45889583
mudstone 4348-MA-T-106 813 1.07 25,23809
heterolithe 4348-MA-T-108 753 0.85 18,56384259
mudstone 4348-MA-T-152 917 0.95 24,3561413
heterolithe 4348-MA-T-164 1759 1.02 6,075439394
bioturbated heterolithe 4348-MA-T-349 471 0.72 39,4522293
bioturbated heterolithe 4348-MA-T-350 100 0.61 157,4308333
bioturbated heterolithe 4348-MA-T-352 361 0.71 50,75877193

7. Conclusions

The Early and Middle Devonian Parnaíba Basin phosphogenic system is represented by episodic

pristine and reworked levels of phosphorite, phosphatic siliciclastic rocks, phosphatic ironstone, ironstone

and associated organic rich sediments (TOC 1- 3%). These are related to offshore (outer shelf), offshore-

lower shoreface transition zone (inner shelf) to lower shoreface depositional environments that were

recognized in the transgressive and normal regressive successions.

Phosphorite occurs as concretions, nodules or layers characterized by francolite-siderite or

francolite-glauconite rich associations related to condensed zones near SWB within late Eifelian and

Givetian intervals. Ironstones /phosphatic ironstones were not deposited under the same conditions. They

are represented by odinite, chamosite/clinochlore, francolite interlayerings as mainly hybrid coated grains,

in a matrix with authigenic chamosite (or clinochlore), Mg siderite, siderite, pyrite framboids and partially

degraded and concentrated near FWWB of Pragian, late Eifelian and Givetian intervals.

REE patterns are marked by MREE enrichments (including phosphorites, phosphatic siliciclastic

rocks, francolite-glauconite nodules, phosphatic ironstones, some ironstones and francolite and early

siderites) that are related to an eodiagenetic stage. Although it can represent suboxic-anoxic conditions, it

is a typical pattern for authigenic phosphorites and brachiopod shells in the Devonian from other locations

globally, implying a possible global seawater signature of more widespread anoxia, as also indicated by

global Corg/P ratios.


131

Table 6 – Concentrations of rare earth elements (in ppm) of separated minerals.

Mineral SAMPLE La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y

Francolite MA-714B1 88 278 44.5 233 72 23.7 118 16.85 94.7 17.65 43.4 5.5 27 4.01 667
Farncolite MA-714X 625 1445 173.5 714 205 58.7 303 47.6 266 46 109 14.45 74.6 10.5 1370
106.
Farncolite MA-714B2 5 348 51.9 270 83.3 26.8 135.5 19.75 112 20.7 50.5 6.08 29.4 4.19 734
100.
Farncolite DS-R-001 5 316 39.8 184 45.2 13.65 68.6 9.86 57.4 12 32.1 4 20.6 3.13 529
164.
Farncolite DS-R-007C1 5 464 62.7 309 81.8 25.9 150 20.2 115 23.1 58.5 7.22 34.6 5.53 1015
Farncolite MA-R-715E 62.7 186.5 24.6 121.5 38.1 12.45 72.8 11.45 70.2 13.9 35 4.49 23.3 3.38 577
Farncolite MA-R-714D 83 248 36.8 190.5 60.1 19.1 110 16 91.1 17.4 42.7 5.23 26 4.03 682
139.
Farncolite MA-R-714E 5 370 46.8 214 56.2 16.75 91.1 12.8 73.7 14.2 35.1 4.53 22.1 3.35 536
117.
Farncolite MA-T503A(1) 5 414 60.4 304 98.5 29 148.5 24.7 163.5 32.7 88.9 12.6 68.2 10.25 1020
144.
Farncolite MA-R-711G (2) 5 521 72 353 96.9 28.7 140.5 19.45 113 21.4 52.9 6.82 33.2 4.85 787
139.
Farncolite MA- R-715 5 401 56.3 300 112.5 35 199.5 30.1 171.5 30.2 67.5 8.05 37.3 5.26 995
Farncolite MA-715E (2) 98.2 311 47 246 75.6 23.2 121 17.3 100.5 19.05 47.6 5.87 29.2 4.45 736
188.
Farncolite HHE-311 5 541 71.5 346 90.3 26.6 163.5 22.7 133 27.2 67.6 8.59 41.7 6.52 1160
144.
Farncolite HHL-385 5 386 47.8 216 55.5 16.9 94.6 13.05 75.7 14.4 36.1 4.49 22.8 3.42 568
Farncolite MA-R-711G (1) 111 380 51 248 67.4 20.3 106.5 14.75 88.4 18.4 46.1 5.87 29 4.39 752
Farncolite MA-T-460 57.4 184 32.2 199.5 113.5 41.1 267 54.4 345 56.5 117.5 12.7 56.7 7.86 1640
Farncolite MA-R-713E 34.6 59.4 7.26 25.6 5.23 1.3 5.15 0.79 4.65 0.87 2.38 0.37 2.46 0.39 25
Farncolite MA-R-715B 73.5 199.5 28.8 143 44.5 13.35 73.5 11.45 67.9 12.85 31.1 3.97 19.25 2.99 468
Farncolite DS-R-07A 58.9 176 26.8 139 39.7 12.45 65 9.27 53.4 10.2 24.5 3.06 14.5 2.2 405
Mg-Siderite MA-T-149B 45 116.5 10.8 35.8 6.22 1.36 5.11 0.93 6.11 1.33 4.52 0.87 5.89 0.91 32.7
Mg-Siderite MA-T-144B 39.9 163 18.6 76.4 19.55 4.18 19.65 3.07 17.4 3.13 8.05 1.26 7.76 1.15 66.8
Mg-Siderite MA-T-124B 91.5 322 37.5 133.5 24.4 4.97 18.5 3.2 22.6 4.8 15.8 2.75 17.75 2.74 116.5
Mg-Siderite MA-T-146B 59.1 258 31 136.5 38.7 8.5 36.7 5.77 30.1 5.02 12.3 1.72 10.05 1.49 105.5
Mg-Siderite MA-T-129B 142 417 46 168 33.1 6.98 26.2 3.73 19.55 3.57 11.1 1.84 11.95 1.86 84.1
Mg-Siderite MA-T-136B 52.6 149.5 14.5 49.5 8.5 1.77 5.83 0.79 3.6 0.58 1.68 0.3 2.04 0.32 13.6
Mg-Siderite MA-T-130B 116 343 34.9 118.5 22.2 4.75 17.2 3.04 20.5 4.27 13.7 2.47 15.9 2.44 105.5
Mg-Siderite MA-T-123B 68.6 177.5 19.4 69.8 13.25 2.93 10.85 1.56 8.88 1.68 5.1 0.89 5.63 0.85 40.4
124.
Mg-Siderite MA-T-115B 5 403 35.2 123 21.9 4.38 14.85 2.15 12.8 2.78 8.99 1.72 12.2 1.89 65.4
Mg-Siderite MA-T-466B 44.9 104.5 12.95 54.5 15.1 3.95 20.4 3.47 22.3 4.19 10.45 1.36 6.7 0.96 127
Mg-Siderite MA-T-465B 39 83.2 10.5 43.5 12.05 3.22 15.75 2.66 17.2 3.35 8.62 1.07 5.33 0.8 108.5
132

Table 6 (continuation) – Concentrations of rare earth elements (in ppm) of separated minerals.

Mineral SAMPLE La Ce Pr Nd Sm Eu Gd Tb Dy Ho Er Tm Yb Lu Y

Mg-Siderite MA-T-453B 35.6 97.3 11.5 44.2 9.1 2.05 7.3 1.01 6.25 1.17 3.65 0.62 3.74 0.61 26.3
Mg-Siderite MA-T-455B 342 1225 123 447 67.5 12 37.8 5.76 35.7 7.74 24.1 4.32 28.1 4.34 179
Mg-Siderite MA-T-470B 75.5 207 24.5 99.2 24.8 5.41 21.8 3.33 18.95 3.33 8.32 1.15 6.35 0.96 76.9
Siderite MA-T-149A 61 331 52.3 271 89.6 21.4 98.5 15.95 89.3 14.75 34.5 4.5 21.8 2.87 372
Siderite MA-T-147A 104.5 428 55.8 269 89 19.3 90.3 14.6 80 12.85 29.8 3.72 18.95 2.52 291
Siderite MA-T-144A 73.6 255 31.2 144 48 10.4 49.6 8.32 48.1 8.2 21 2.86 15.8 2.31 185
Siderite MA-T-124A 93.5 310 31.5 116.5 23 4.18 18.1 2.72 16.25 3.09 9.79 1.48 10.15 1.34 65.4
Siderite MA-T-146A 38.3 135.5 14.55 67.3 20.4 4.09 21.7 3.42 21.7 3.73 10.55 1.45 9.09 1.22 83
Siderite MA-T-136A 40.2 139 16.25 81.3 27.5 5.69 30.9 4.97 30.3 5.39 14.4 1.75 10.4 1.34 119
Siderite MA-T-131A 59.5 195 21.2 93.1 27.3 6.93 29.9 5.06 27.8 5.27 14.75 1.86 10.95 1.45 132
Siderite MA-T-157A 43.1 94.8 10.9 44.4 9.39 2.31 9.74 1.5 7.7 1.43 4.05 0.56 3.27 0.56 39.9
Siderite MA-T-129A 92.5 311 34.8 153 36.7 9.28 37.3 5.91 29.5 5.42 14.15 1.79 10.55 1.5 118.5
Siderite MA-T-138A 144 388 34.9 144 30.2 6.24 18.85 1.99 7.75 1.19 3.28 0.42 2.77 0.43 28.2
Siderite MA-T-130A 169.5 550 56.6 215 38 7.79 27 3.62 16.15 2.81 7.58 1.18 7.01 1.03 61.7
Siderite MA-T-120A 91.5 406 59.4 374 154.5 44.7 200 34.3 177.5 29.3 65 6.13 26.2 3.06 622
Siderite MA-T-115A 43.8 169 24.7 144.5 50.3 13.5 57.5 9.66 51.7 9.19 23.4 2.86 15.15 2.05 226
Siderite MA-T-451A 15.6 51.3 5.56 25.2 7.16 1.8 7.15 1.27 7.62 1.56 4.44 0.64 3.92 0.5 35.3
Siderite MA-T-119A 61.3 194 24.4 127.5 42.5 11.2 50.2 8.38 45.6 8.52 22.4 2.62 13.15 1.76 216
Siderite MA-T-450A 34.7 125 16.35 83.9 26.1 7.39 32.9 5.8 31.8 5.96 14.75 1.72 8.67 1.17 157.5
Siderite MA-T-452A 36.8 86.2 9.57 38.6 7.53 1.61 6.16 0.98 5.73 1.29 4.24 0.63 4.09 0.56 33.2
Siderite MA-T-453A 43.3 128 13.85 55.8 11.25 2.48 8.38 1.27 6.89 1.37 4.5 0.69 4.84 0.69 32.8
Siderite MA-T-455A 57.9 210 21 86.4 17.1 3.49 11.5 1.96 12.1 2.66 8.99 1.44 9.29 1.46 64.8
Siderite MA-T-454A 44.6 148.5 15.25 63.5 14.25 3.31 11.9 1.74 9.34 1.84 5.35 0.82 5.15 0.81 39.6
Siderite MA-T-468A 23.7 49.4 5.61 22.6 5.25 1.15 5.29 1.01 6.9 1.51 4.75 0.68 3.78 0.52 43.9
Siderite MA-T-473A 21.2 46.8 5.11 20 3.76 0.92 3.38 0.56 3.29 0.63 2.02 0.27 1.89 0.28 16.8
Siderite MA-T-485A 54.5 108 12.25 46.6 8.66 1.78 7.11 1.04 6.28 1.2 3.83 0.53 3.2 0.53 32.3
Siderite MA-T-488A 45.4 97.3 11.35 45.6 9.59 2.21 9.6 1.68 9.34 1.85 5.85 0.86 4.96 0.75 52.1
Siderite MA-T-489A 31 88.8 9.33 35.1 5.72 0.95 3.07 0.5 3.14 0.71 2.42 0.42 3.17 0.48 16.7
Siderite MA-T-491A 55.4 231 39.3 233 93.8 26.7 142.5 26.8 151 28.9 76.6 9.13 44.5 5.71 782
Siderite MA-T-501A 22.7 54.3 6.59 28.5 7.89 1.86 7.53 1.25 6.99 1.27 3.57 0.53 3.05 0.48 34.1
Siderite MA-T-507A 46.3 144.5 18.3 86.1 23.5 6.04 26.5 4.68 26.4 5.15 14.3 1.74 9.96 1.46 139
Siderite MA-T-506A 57.8 132.5 15.55 64.2 13.35 2.5 11.6 1.79 10.05 1.98 6.03 0.81 4.99 0.81 55.7
Siderite MA-T-215A 19.1 32.4 3.94 15.1 3.3 0.8 3.14 0.58 3.65 0.81 2.7 0.47 3.16 0.52 20.8
Siderite MA-T-505A 32.1 71.9 8.38 34.8 7.41 1.29 5.53 0.84 4.46 0.91 2.84 0.42 2.73 0.41 23.7
Siderite MA-T-492A 33.5 75.3 8.89 35.4 7.62 1.38 6.28 1.02 6.13 1.23 4.09 0.6 4.06 0.61 33.4
Siderite MA-T-456A 39.9 94 10.15 39.1 6.83 1.2 3.87 0.56 2.89 0.6 2.09 0.32 2.15 0.35 15
Siderite MA-T-123A 63.3 190.5 19.95 80.6 17.45 3.67 13.4 2.26 11.55 2.21 6.55 0.92 5.42 0.84 50.5
Siderite MA-715V 39.7 81.3 9.26 36.6 6.85 1.38 5.59 0.86 5.02 0.98 2.96 0.45 2.86 0.4 27.9
Siderite MA-714F 36.4 70.1 8.18 32 5.87 1.32 5.19 0.9 5.12 1.05 3.29 0.45 2.93 0.45 29.4
133

Table 7 – Stable carbon and oxygen isotope composition of analyzed siderites.

Sample ID Mineralogy Depth Well δ13C ‰ VPDB δ18O ‰ VPDB

MA-T-485A siderite 40,2 PM-11-PI -0,8 -2,3

MA-T-488A siderite 38,1 PM-11-PI -2,0 -3,0

MA-T-489A siderite 37,3 PM-11-PI -3,5 -1,6

MA-T-491A siderite 27,1 PM-11-PI -5,4 0,1

MA-T-501A siderite 20,3 PM-11-PI -3,5 0,3

MA-T-507A siderite 13,4 PM-11-PI -5,1 1,0

MA-T-506A siderite 15,2 PM-11-PI -3,6 1,6

MA-T-505A siderite 14,9 PM-11-PI -3,9 1,4

MA-T-492A siderite 27,5 PM-11-PI -5,4 0,3

MA-T-149A siderite 99,40 ARN-1-TO -13,7 -3,6

MA-T-147A siderite 101,69 ARN-1-TO -12,3 -4,6

MA-T-144A siderite 103,04 ARN-1-TO -21,5 -2,8

MA-T-124A siderite 145,15 ARN-1-TO -16,6 -2,5

MA-T-146A siderite 102,40 ARN-1-TO -13,2 -4,1

MA-T-136A siderite 132,3 ARN-1-TO -8,8 -2,7

MA-T-131A siderite 140,95 ARN-1-TO -12,1 -3,3

MA-T-157A siderite 92,3 ARN-1-TO -7,2 -5,2

MA-T-129A siderite 142,34 ARN-1-TO -12,9 -2,5

MA-T-138A siderite 122,75 ARN-1-TO -8,2 -4,6

MA-T-130A siderite 142,13 ARN-1-TO -12,2 -2,6

MA-T-120A siderite 147,8 ARN-1-TO -15,2 -2,8

MA-T-115A siderite 152,1 ARN-1-TO -13,1 -2,8

MA-T-119A siderite 148,35 ARN-1-TO -13,5 -3,8

MA-T-123A siderite 145,75 ARN-1-TO -12,5 -3,0

MA-T-149B Mg-rich siderite 99,2 ARN-1-TO -14,0 -8,3

MA-T-144B Mg-rich siderite 103,04 ARN-1-TO -15,3 -3,1

MA-T-124B Mg-rich siderite 145,15 ARN-1-TO -12,8 -3,5

MA-T-146B Mg-rich siderite 102,4 ARN-1-TO -13,3 -3,0

MA-T-129B Mg-rich siderite 142,34 ARN-1-TO -11,4 -4,4

MA-T-136B Mg-rich siderite 132,3 ARN-1-TO -7,8 -3,2

MA-T-130B Mg-rich siderite 142,13 ARN-1-TO -11,5 -3,5

MA-T-123B Mg-rich siderite 145,75 ARN-1-TO -11,6 -4,7

MA-T-115B Mg-rich siderite 152,1 ARN-1-TO -12,4 -4,2

MA-T-451A siderite 68,3 PM-21-PI -4,0 -3,9

MA-T-450A siderite 69,22 PM-21-PI -7,3 -9,8

MA-T-452A siderite 67,3 PM-21-PI -5,4 -1,0


134

Table 7 (continuation) – Stable carbon and oxygen isotope composition of analyzed siderites.

Sample ID Mineralogy Depth Well δ13C ‰ VPDB δ18O ‰ VPDB

MA-T-453A siderite 66,3 PM-21-PI -9,4 1,0

MA-T-455A siderite 65,3 PM-21-PI -8,1 -1,7

MA-T-454A siderite 65,4 PM-21-PI -8,6 0,2

MA-T-473A siderite 23,6 PM-21-PI -1,1 -3,6

MA-T-456A siderite 63,6 PM-21-PI -7,7 2,5

MA-T-466B Mg-rich siderite 50,2 PM-21-PI -4,9 -1,3

MA-T-465B Mg-rich siderite 53,9 PM-21-PI -10,9 -1,2

MA-T-453B Mg-rich siderite 66,3 PM-21-PI -8,5 -0,4

MA-T-455B Mg-rich siderite 65,3 PM-21-PI -10,9 -1,5

Table 8 – Stable carbon and oxygen isotope composition of analyzed francolites.

Sample ID Mineralogy δ13C ‰ VPDB δ18O ‰ VPDB

MA-714B1 francolite -6,8 -8,709106846

MA-714X francolite -7,3 -12,39144031

MA-714B2 francolite -6,9 -8,678842664

DS-R-001 francolite -6,3 -13,91698461

DS-R-007C1 francolite -5,4 -9,239805784

MA-R-715E francolite -7,3 -13,63193551

MA-R-714D francolite -5,2 -8,630936327

MA-R-714E francolite -5,5 -9,156758572

MA-T503A(1) francolite -2,9 -9,425263548

MA-R-711G (2) francolite -7,8 -7,196328009

MA-R-715 francolite -6,3 -13,91698461

MA-715E (2) francolite -6,8 -8,853686449

HHE-311/DS-R-07C francolite -4,7 -10,34050407

HHL-385 francolite -5,5 -9,161061536

MA-R-711G (1) francolite -7,6 -7,11744033

MA-T-460 francolite-braq -5,5 -4,643522687

MA-R-715B francolite -7,3 -11,26004755

DS-R-07A francolite -7,0 -11,26004755

Isotope signatures suggest that siderite from phosphatic ironstones and ironstones are the result of

microbial reduction by sulfate-reducing bacteria with a source of iron and dissolved sulfate in marine

porewaters and also clustered in iron redox zones. Partially degraded organic matter marked by pyrite

framboids confirms anoxic sulfate reduction. Whole rock V/Cr, V/(V+Ni) and Mo values confirm that
135

anoxia/ dysoxia conditions with minimal input of dissolved H2S during formation of phosphatic ironstone.

The variation in δ18O values also indicate possible variations in salinity/temperature due to mixing

seawater and freshwater.

We interpret that upwelling during the transgressions provided the main productivity condition for

generating phosphorites, under low rate of sedimentation. River discharge probably affected physical

characteristics of the upper water column and produced large spatial variability in the Parnaiba Basin

system influencing temperature, vertical stratification and salinity leading to gradients of surface nutrient

concentrations. Terrestrial OM enhanced shallow-water anoxia decreasing bottom and pore-water oxygen

concentrations producing reduced total P burial efficiencies and enhanced organic carbon burial. The

phosphorus cycle was probably affected by the global low oxygen and anoxic conditions in the Devonian.

Although the recycling of nutrients from sediments deposited in oxygen-depleted conditions maintained

enhanced levels of primary productivity, the burial of P was probably affected in coastal regions and this

probably decreased economic potential of the Devonian deposits. On the other hand, strong REE

concentrations represents an aggregate predictive economic potential linked to Devonian phosphorites.

Acknowledgements

The authors acknowledge the Geological Survey of Brazil (CPRM-Phosphate Brazil Project) for

the financial support of this research. Our deepest gratitude to Dr. Peir K. Pufahl for criticisms and

facilitating acces to Acadia University lab. We thank Evelyne Marie Sylvie Leduc for facilitating our

isotope analysis at the Queens University. The editor and anonymous reviewers are thanked for their

suggestions and revisions. We also acknowledge Prof. Leonardo Borghi (LAGESED-UFRJ) and Marco

Klotz (CENPES), for providing access to the Conexão Project drill core samples; Acadia University and

CPRM team for support assistance; Maisa Abram acknowledges CAPES - Brazilian Research Funding

Organization (Coordenação de Aperfeiçoamento de Pessoal de Nível Superior, Brazil) scholarship and the
136

ELAP (The Emerging Leaders in the Americas Program, Canada) funding. Michael Holz acknowledges

CNPq (Brazilian Council for Research) for personal research grant (PQ 304657/15-8).

References

Abdul-Hadi, A.R., Astin, T.R., 1994. Genesis of siderite in the Upper Miocene offshore Sarawak:
constraints on pore fluid chemistry and diagenetic history. Geological. Society Malaysia BuIlletin
37, 595-415.
Abram, M.B., Holz, M., 2020. Early to Middle Devonian ironstone and phosphorite in the northwestern
Gondwana Parnaíba Basin, Brazil: A record of an epeiric margin paleoceanographic changes.
Sedimentary Geology, 402, 105646.
Adelson, J. M, Helz, G. R, Miller, C. V., 2001. Reconstructing the rise of recent coastal anoxia;
molybdenum in Chesapeake Bay sediments. Geochimica et Cosmochimica Acta, 65, 2, 237–252.
Albuquerque, H.J.T.R.; Coelho, J.M.; Farias, C.E.G., 1972. Projeto São Miguel do Tapuio. Recife:
CPRM. 2v. Relatório interno. Convênio CNEN/CPRM.
Algeo, T. J., Berner, R. A., Maynard, B. J., Scheckler, S. E., 1995. Late Devonian oceanic anoxic events
and biotic crisis: “Rooted” in the evolution of vascular land plants. GSA Today 5, 45, 63-66.
Algeo, T.J., Ingall, E., 2007. Sedimentary Corg:P ratios, paleocean ventilation, and Phanerozoic
atmospheric pO2. Palaeogeography, Palaeoclimatology, Palaeoecology, 256, 130–155.
Algeo, T.J., Heckel, P.H., Maynard, J.B., Blakey, R.C., Rowe, H., 2008. Modern and ancient epeiric seas
and the super-estuarine circulation model of marine anoxia. In: Pratt, B.R., Holmden, C., (Eds.).
Dynamics of Epeiric Seas, Special Paper 48: Geological Association of Canada, 7-38.
Algeo, T.J., Lyons, T.W., 2006. Mo–total organic carbon covariation in modern anoxic marine
environments: Implications for analysis of paleoredox and paleohydrographic conditions.
Paleoceanography, 21, PA1016, doi:10.1029/2004PA001112.
Algeo, T.J., Rowe, H., 2012. Paleoceanographic applications of trace-metal concentration data. Chemical
Geology 324-325, 6–18.
Amaral, P.F., Hidalgo, R., Barbosa, R.C.M., Nogueira, A.C.R., 2013. Palinomorfos da Formação
Pimenteiras (Devoniano): contribuição na avaliação do potencial de geração de petróleo da borda
oeste da bacia do Parnaíba, região de Pedro Afonso (TO). Anais do 13º Simpósio de Geologia da
Amazônia, 111-113.
Amaro, G.J.L., Villas, R.N., Kotschoubey, B., 2012. Estudo petrográfico e geoquímico dos
ironstones da base da Formação Pimenteiras, Borda Oeste da Bacia do Parnaíba,
Tocantins, 2012. Revista Brasileira de Geociências, 42, 2, 373-396.
Anbar, A.D., Duan, Y., Lyons, T.W., Arnold, G.L., Kendall, B., Creaser, R.A., Kaufman, A.J., Gordon,
G.W., Scott, C., Garvin, J., Buick, R., 2007. A Whiff of oxygen before the Great Oxidation
Event? Science, 317, 1903 DOI: 10.1126/science.1140325.
Anderson, L.A., Sarmiento, J.L., 1994. Redfield ratios of remineralization determined by nutrient data
analysis. Global Biogeochemical Cycles, 8, 65–80.
Andrade, C.L.N., 2015. Arcabouço palinoestratigráfico e geoquímica orgânica da Formação Pimenteiras
(Devoniano), na borda oeste da Bacia do Parnaíba, Brasil, Phd Thesis, Federal University of
Bahia, Salvador.
Andreeva, P., Chatalov, A., 2011. Origin of the Eifelian ironstone from Well R-119 Kardam, Northeastern
137

Bulgaria. Comptes Rendu de l’ Académie Bulgare des Science, 64, 91-101.


Araújo, D.B., 2015. Bacia do Parnaíba. Sumário geológico e setores em oferta. Agência Nacional do
Petróleo, Gás Natural e Biocombustíveis - Décima Terceira Rodada de Licitações.
Superintendência de Definição de Blocos Sdb, 21pp.
Babechuk, M.G., Kamber, B.S., Widdowson, M., 2012. Yttrium mobility during weathering: implications
for riverine Y/Ho. Mineralogical Magazine, 76, 1443.
Bailey, J.V., Corsetti, F.A., Greene, S.E., Crosby, C.H., Liu, P., Orphan, V.J., 2013. Filamentous sulfur
bacteria preserved in modern and ancient phosphatic sediments: implications for the role of
oxygen and bacteria in phosphogenesis. Geobiology, 11, 397-405.
Banner, J.L., Hanson, G.N., Meyers, W.J., 1988. Rare earth element and Nd isotopic variations in
regionally extensive dolomites from the Burlington-Keokuk Formation (Mississippian):
Implications for REE mobility during carbonate diagenesis. Journal of Sedimentary Petrology, 58,
3, 415-432.
Bau M., Dulski, P., 1996. Distribution of yttrium and rare-earth elements in the Penge and Kuruman iron-
formations, Transvaal Supergroup, South Africa. Precambrian Research, 79, 37–55.
Bau, M., 1996. Controls on the fractionation of isovalent trace elements in magmatic and aqueous
systems: evidence from Y/Ho, Zr/Hf and lanthanide tetrad effect. Contributions to Mineralogy
and Petrology, 123, 323–333.
Bau, M., Mdler P., Dulski P., 1997. Yttrium and lanthanides in eastern Mediterranean seawater and their
fractionation during redox-cycling. Marine Chemistry, 56, 123-131.
Bau, M., Schmidt, K., Koschinsky, A., Hein, J., Kuhn, T., Usui, A., 2014. Discriminating between
different genetic types of marine ferro-manganese crusts and nodules based on rare Earth elements
and yttrium. Chemical Geology, 381,1–9.
Bech, J., Suarez, M., Reverter, F., Tume, P., Sánchez, P., Roca, N., Lansac, A., 2010. Selenium and other
trace element in phosphorites: A comparison between those of the Bayovar-Sechura and other
provenances. Journal of Geochemical Exploration, 107, 146–160.
Becker, R.T., Kȍnigshof, P., Brett, C.E., 2016. Devonian Climate, Sea Level and Evolutionary Events. In:
Becker, R. T., Kȍnigshof, P., Brett, C.E. (Eds.). Devonian Climate, Sea Level and Evolutionary
Events. Geological Society, London, Special Publications 423, 1–10.
Belem A.L., Caricchio C., Albuquerque A.L.S., Venancio I.M., Zucchi M.R., Santos T.H.R., Alvarez
Y.G., 2019. Salinity and stable oxygen isotope relationship in the Southwestern Atlantic:
constraints to paleoclimate reconstructions. Anais Academia Brasileira de Ciências 91,
e20180226. DOI 10.1590/0001-3765201920180226.
Benitez-Nelson, C., 2000. The biogeochemical cycling of phosphorus in marine systems Claudia R.
Earth-Science Reviews, 51, 109–135.
Berner, R.A., 1984. Sedimentary pyrite formation: An update. Geochimica et Cosmochimica Acta, 48,
605–615.
Berner, R.A., Raiswel, R., 1983. Burial of organic carbon and pyrite sulfur in sediments over Phanerozoic
time: a new theory. Geochimica et Cosmochimica Acta 47, 855-862.
Blakey, R., 2019. Global Paleogeography and Tectonics in Deep Time Series
http://deeptimemaps.com/global-paleogeography-and-tectonics-in-deep-time-series.
Bojanowski, M.J., Clarkson, E.N.K., 2012. Origin of siderite concretions in microenvironments of
methanogenesis developed in a sulfate reduction zone: an exception or a rule? Journal of
Sedimentary Research 82, 585–598.
138

Bond, D., Wignall, P.B., 2010. Pyrite framboid study of marine Permian–Triassic boundary sections: a
complex anoxic event and its relationship to contemporaneous mass extinction. Geological
Society of America Bulletin, 122, 1265–1279.
Bond, D.; Wignall, P.B., Racki, G. 2004. Extent and duration of marine anoxia during the Frasnian–
Famennian (Late Devonian) mass extinction in Poland, Germany, Austria and France. Geological
Magazine, 141, 2, 173–19.

Breuer, P., Grahn, Y., 2011. Middle Devonian spore stratigraphy in the eastern outcrop belt of the
Parnaíba Basin, northeastern Brazil. Revista Española de Micropaleontología, 43, 1-2, 19-38.
Brian, T.S., 2009. Variation in C/P ratios in Devonian-Mississippian marine shales: testing the
productivity-anoxia feedback model. University of Kentucky Master's Theses. 609.
https://uknowledge.uky.edu/gradschool_theses/609.
Byrne, R.H., Kim, I-H., 1990. Rare earth element scavenging in seawater. Geochimica B Cosmochemica
Acta, 54, 2645-2656.
Caetano-Filho, S., Paula-Santos, G.M., Dias-Brito, D., 2018. Carbonate REE+Y signatures from the
restricted early marine phase of South Atlantic Ocean (late Aptian – Albian): The influence of
early anoxic diagenesis on shale-normalized REE+Y patterns of ancient carbonate rocks.
Palaeogeography, Palaeoclimatology, Palaeoecology, 500, 69-83.
Caputo, M.V., Santos, R.O.B., 2019. Stratigraphy and ages of four Early Silurian through Late Devonian,
Early and Middle Mississippian glaciation events in the Parnaíba Basin and adjacent areas, NE
Brazil. Earth Science Reviews, 7, 103002, doi.org/10.1016/j.earscirev.2019.103002.
Carmichael, S.K., Waters, J.A. Suttner, T.J., Kido, E., and DeReuil, A.A., 2014. A new model for the
Kellwasser Anoxia Events (Late Devonian): Shallow water anoxia in an open oceanic setting in
the Central Asian Orogenic Belt. Palaeogeography, Palaeoclimatology, Palaeoecology 399, 394-
403.
Carozzi, A.V., Falkenhein, F.U.M., Carneiro, R.G., Esteves, R.P., Contreiras, C.J.A., 1975. Análise
ambiental e evolução tectônica sinsedimentar da seção siluroeocarbonífera da Bacia do Maranhão.
PETROBRAS, Rio de Janeiro, Brasil, 48 pp.

Castro, D.L., Bezerra, F.H.R., Fuck, R.A., Vidotti, R.M., 2016. Geophysical evidence of pre-sag
rifting and post-rifting fault reactivation in the Parnaíba basin, Brazil. Solid Earth Discussions,
doi:10.5194/se-2016-21.

Chow, N., Morad, S., Al-Aasm, I.S. 1996. 24 Origin of authigenic carbonates in Eocene to Quaternary
sediments from the Arctic Ocean and Norwegian-Greenland Sea. In: Thiede, J., Myhre, A.M.,
Firth, J.V., Johnson, G.L., and Ruddiman, W.F. (Eds.). Proceedings Ocean Drilling Program
(ODP), Scientific Results, 151: College Station, TX.

Coleman, M.L., Hedrick, D.B., Lovley, D.R., White, D.C., Pye, K., 1993. Reduction of Fe (III) in
sediments by sulphate-reducing bacteria. Nature, 361, 436-439.

Coles, K.S., Varga, R.J., 1988. Early to middle Paleozoic phosphogenic province in terranes of the
southern Cordillera, western United States. American Journal of Science, 288, 891-924.
Companhia de Pesquisa de Recursos Minerais – CPRM, 1978. Projeto São Nicolau: Fosfato.
139

Compton, J.S.; Bergh, E.W., 2016. Phosphorite deposits of the Namibian shelf. Marine Geology, 380,
290-314.
Cook, P.J., McElhinny, M.W., 1979, A reevaluation of the spatial and temporal distribution of
sedimentary phosphate deposits in the light of plate tectonics. Economic Geology, 74, 315-330.
Correia-Junior, F.C., Abram, M.B., Ferreira, M.V., Cunha, I.A., Neves, M.P., 2016. A pesquisa para
fosfato na Bacia do Amazonas, Estado do Pará. In: Abram, M.B., Bahiense, I.C., Almeida, R.C.
(Eds.), Projeto Fosfato Brasil: parte II, Informe de Recursos Minerais, Série Insumos Minerais
para Agricultura 17, Salvador, CPRM, 891-834.
Cunha, F.M.B., 1986. Evolução paleozóica da bacia do Parnaíba e seu arcabouço tectônico. Dissertação de
Mestrado, IGEO-UFRJ, Rio de Janeiro. 107p.
Curtis, C.D., Coleman, M.L., 1986. Controls on the precipitation of early diagenetic calcite, dolomite, and
siderite concretions in complex depositional sequences. In: Gautier, D.L. (Ed.). Roles of Organic
Matter in Sediment Diagenesis. Society of Economic Paleontologists and Mineralogists Special
Publication 38,23-33.
Danielson, A., Moller, P., Dulski, P., 1992. The europium anomalies in banded iron formations and the
thermal history of the oceanic crust. Chemical Geology, 97, 89-100.
Della Fávera, J.C., 1990. Tempestitos da bacia do Parnaíba: um ensaio holístico. Tese de Doutorado, IG-
UFRGS, Porto Alegre. 243p.
Denayer, J., 2016. Iron ores of Southern Belgium: much more than hematite. Antrophologica et
Præhistorica, 126, 39-49.
Droser, M.L., Bottjer, D.J., 1993. Trends and patterns of Phanerozoic ichnofabrics. Annual Review of
Earth and Planetary Sciences, 21, 205–225.
Dumas, S., Arnott, R.W.C., 2006. Origin of hummocky and swaley cross-stratification - The controlling
influence of unidirectional current strength and aggradation rate. Geology, 34, 1073-1076.
Dunn, S.K. 2020. Middle Ordovician ironstone of North Wales, United Kingdom: sedimentologic and
oceanographic evidence for a ferruginous Rheic Ocean, Msc Thesis, Acadia University, Wolfville,
Canada, 170pp.
Emsbo, P., McLaughlin, P.I., Breit, G.N., Du Bray, E.A., Koenig, A.E., 2014. Rare Earth Elements in
Sedimentary Phosphate Deposits: Solution to the Global REE Crisis? Gondwana Research, 27, 2,
776-785.
Felitsyn, S., Morad. S., 2002. REE patterns in latest Neoproterozoic–early Cambrian phosphate
concretions and associated organic matter. Chemical Geology 187, 257– 265.
Fischer, W.W., Knoll, A.H., 2009. An iron shuttle for deepwater silica in Late Archean and early
Paleoproterozoic iron formation. Geological Society of America Bulletin, 121, 222-235.
Follmi, K., 1996. The phosphorus cycle, phosphogenesis and marine phosphate-rich deposits. Earth-
Science Reviews, 40, 55–124.
Follmi, K.B., Badertscher, C., Kaenel, E., Stille, P., John, C.M., Adatte, T., Steinmann, P., 2005.
Phosphogenesis and organic-carbon preservation in the Miocene Monterey Formation at Naples
Beach, California -The Monterey hypothesis revisited. GSA Bulletin 117, 589–619.
Follmi, K.B., Garrison, R.E., Grimm. K.A., 1991. Stratification in phosphatic sediments: illustrations from
the Neogene of California. In: Einsele, G., Ricken, W, Seilacher, A. (Eds.), Cycles and Events in
Stratigraphy. Springer-Verlag, Berlin, 492- 507.
Froelich, P.H., Klinkhammer, G.P., Bender, M.L., Luedtke, N.A., Heath, G.R., Cullen, D., Dauphin, P.,
Hammond, D., Hartman, B., Maynard, V. 1979. Early oxidation of organic matter in pelagic
140

sediments of the eastern equatorial Atlantic: suboxic diagenesis. Geochimica et Cosmochimica


Acta, 43, 1075-1090.
Gao, G. 1993. The temperatures and oxygen-isotope composition of early Devonian oceans. Nature, 361,
712–714.
Garnit, H., Bouhlel, S., 2017. Petrography, mineralogy and geochemistry of the Late Eocene oolitic
ironstones of the Jebel Ank, Southern Tunisian Atlas. Ore Geology Reviews, 84, 134-153.
Girard, J.-P., Flicoteaux, R., Walter, A.-V., Savin, S.M., and Nahon, D., 1993. Oxygen and carbon stable
isotope composition of structural carbonate in weathering apatites from laterites, southern Brazil
and western Senegal. Applied Geochemistry, 8, 617–632.
Glenn, C.R., Arthur, M.A., 1988. Petrology and major element geochemistry of Peru margin phosphorites
and associated diagenetic minerals: Authigenesis in modern organic-rich sediments. Marine
Geology, 80, 231–267.
Glenn, C.R., Garrison, R.E., 2003. Phosphorites. In: Encylopedia of Sediments and Sedimentary Rocks,
G. Middleton (Ed.). Kluwer Academic, p.519-526.
Glenn, C.R., Ollmj, K.F., Gr, K.A., Trappe, J.O., Abed, A.M., Galli, C., Garr, R.E., 1994. Phosphorus and
phosphorites: Sedimentology and environments of formation. Eclogae Geologicae Helvetiae, 87,
747-788.
Góes, A.M.O., Feijó, F.J., 1994. Bacia do Parnaíba. Rio de Janeiro. Boletim de Geociências da Petrobras,
8, 1, 57–67.
Goevert, D., Conrad, R., 2008. Carbon Isotope Fractionation by Sulfate-Reducing Bacteria Using
Different Pathways for the Oxidation of Acetate. Environental Science Technology, 42, 7813-
7817.
Goldberg, K., Humayun M., 2016.Geochemical paleoredox indicators in organic-rich shales of the Irati
Formation, Permian of the Paraná Basin, southern Brazil. Brazilian Journal of Geology, 46, 3,
377-393.
Grahn, Y., Young, C., Borghi, L., 2008. Middle Devonian chitinozoan biostratigraphy and sedimentology
in the eastern outcrop belt of the Parnaíba Basin, northeastern Brazil. Revista Brasileira de
Paleontologia, 11, 137–146.
Grandjean-Lécuyer, P., Feist, R., Lbarbde, F. 1993. Rare earth elements in old biogenic apatites.
Geochimica et Cosmochimica Acta, 57, 2507-2514.
Gromet, L.P., Dymek, R.F., Haskin, L.A., Korotev, R.L. 1994. The “North American shale composite”:
Its compilation, major and trace element characteristics. Geochimica et Cosmochimica Acta, 48,
2469-2482.
Guerrak, S., 1988. Geology of the Early Devonian oolitic iron ore of the Gara Djebilet field, Saharan
Platform, Algeria. Ore Geology Reviews, 3, 333-358.
Haley, B.A., Klinkhammer, G.P., Mcmanus, J., 2004. Rare earth elements in pore waters of marine
sediments. Geochimica et Cosmochimica Acta, 68, 6, 1265–1279.
Harding, S.C., Nash, B.P., Petersen, E.U., Ekdale, A.A., Bradbury, C.D., Dyar, M.D., 2014. Mineralogy
and geochemistry of the main glauconite bed in the Middle Eocene of Texas: Paleoenvironmental
implications for the Verdine Facies. PLoS ONE 9, e87656, doi:10.1371/journal.pone.0087656.
Hatch, J.R., Leventhal, J.S. 1992. Relationship between inferred redox potential of the depositional
environment and geochemistry of the Upper Pennsylvanian (Missourian) Stark Shale Member of
the Dennis Limestone, Wabaunsee County, Kansas, U.S.A. Chemical Geology, 99, 65-82.
141

Heggie, D.T. Skyring, G.W., O’brien, G.W., Reimers, C., Herczeg, A., Moriarty, D.J.W., Burnett, W.C.,
Milnes, A.R., 1990. Organic carbon cycling and modern phosphorite formation on the East
Australian continental margin: an overview, In: Notholt, A.J.G., Jarvis, I. (Eds.). Phosphorite
Research and Development. London: Geological Society, Special Publication, 52:5287-117.
Hein, J.R., Koschinsky, A., Mikesell, M., Mizell, K., Glenn, C.R., Wood, R., 2016. Marine phosphorites
as potential resources for heavy rare earth elements and yttrium. Minerals, 6, 88.
Doi:10.3390/Min6030088.
Hiatt, E.E., Budd, D.A. 2003. Extreme paleooceanographic conditions in a Paleozoic oceanic upwelling
system: organic productivity and widespread phosphogenesis in the Permian Phosphoria Sea. In:
Chan, M.A., Archer, A.W. (Eds.). Extreme depositional environments: Mega end members in
Geologic Time: Boulder., Colorado. Geological Society of America, Special Paper, 370, 245-264.
Holser, W.T., 1997. Evaluation of the application of rare-earth elements to paleoceanography.
Palaeogeography, Palaeoclimatology, Palaeoecology, 132, 309-323.
Humphreys, B., Smith, S.A., 1989. The distribution and significance of sedimentary apatite in Lower to
Middle Devonian sediments east of Plymouth Sound. Proceedings of the Ussher Society, 7, 118-
124.
Hutchins, D.A., Hare, C. E., Weaver, R. S., Zhang, Y., Firme, G.F., G. R. DiTullio, G.R., Alm, M.B.,
Riseman, S.F., Maucher, J.M., Geesey, M.E., Trick, C.G., Smith, G. J. Rue, E. L. Conn, J.,
Bruland, K.W. Limnology and Oceanography, 47, 4, 997–1011.
Iacoviello, F., Giorgetti, G., Nieto, F., Memmi, I.T., 2012. Evolution with depth from detrital to authigenic
smectites in sediments from AND-2A drill core (McMurdo Sound, Antarctica). Clay Minerals, 47,
481–498.
Irwin, H., Curtis, C., Coleman, M. 1977. Isotopic evidence for source of diagenetic carbonates formed
during burial of organic-rich sediments. Nature, 269, 209–213.
Jackson, J. E., 1991. A user´s guide to Principal Components. New York, Willey, 592pp.
Jarvis, I., Burnett, W.C., Nathan, Y., Almbaydin, F.S.M., Attia, A.K.M., Castro, L.N., Flicoteaux, R.,
Hilmy, M.E., Husain, V., Qutawnah, A.A., Serjani, A., Zanin, Y.N., 1994. Phosphorite
geochemistry - state-of-the-art and environmental concerns. Eclogae Geologicae Helveticae, 87,
643–700.
Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochemistry, Geophysics, Geosystems, 11,
1-30.
Johnson, J.E., Muhling, J.R., Cosmidis, J., Rasmussen, B., Templeton, A.S., 2018. Low‐Fe (III) Greenalite
was a primary mineral from Neoarchean oceans. Geophysical Research Letters, 45, 3182–3192.
Johnson, J.G., Klapper, G., Sandberg, C.A., 1985. Devonian eustatic fluctuations in Euramerica:
Geological Society of America Bulletin 96, 567–587.
Jones, B., Manning, D.A.C., 1994. Comparison of geochemical indices used for the interpretation of
palaeoredox conditions in ancient mudstones. Chemical Geology, 111, 111-129.
Kaiho, K., Katabuchi, M., Oba, M., Lamolda, M., 2014. Repeated anoxia-extinction episodes progressing
from slope to shelf during the latest Cenomanian. Gondwana Research, 25, 1357-1368.
Kendall, C., Caldwell, E.A., 1998. Fundamentals of isotope geochemistry. In: Kendall, C., McDonnel, J.J.
(Eds), Isotope tracers in catchment hydrology. Elsevier Science B.V., Amsterdam, 51-86.
Kidder, D.L., Krishnaswamy, R., Mapes, R.H., 2003. Elemental mobility in phosphatic shales during
concretion growth and implications for provenance analysis. Chemical Geology, 198, 335– 353.
142

Kronen, J.D.Jr., Glenn, C.R., 2000. Pristine to reworked minerals of the verdine facies: Keys to
interpretating sequence stratigraphy and sequence condensation in mixed carbonate-siliciclastic
forereef sediments (Great Barrier Reef). In: Glenn, C.R., Prévôt-Lucas, L., Lucas, J. (Eds.),
Marine Authigenesis: From Global to Microbial, SEPM Special Publication, 66, 387-403.
Kuzyk, Z.Z.A., Macdonald, R.W., Tremblay, J-E, Stern, G. A. 2010. Elemental and stable isotopic
constraints on river influence and patterns of nitrogen cycling and biological productivity in
Hudson Bay. Continental Shelf Research, 30, 163–176.
Lawrence, M.G., Greig, A., Collerson, K.D., Kamber, B.S., 2006. Rare earth element and yttrium
variability in South East Queensland waterways. Aquatic Geochemistry, 12, 39-72.
Lécuyer, C., Reynard, B., Grandjean, P. 2004. Rare earth element evolution of Phanerozoic seawater
recorded in biogenic apatites. Chemical Geology 204, 63-102. Piper, D.Z., Baedecker, P.A.,
Crock, J.G., Burnett, W.C., Loebner, B.J., 1988. Rare Earth elements in the phosphatic-enriched
sediment of the Peru Shelf. Marine Geology, 80, 269-285.
Lewan, M.D., Maynard, J.B., 1992. Factors controlling enrichment of vanadium and nickel in the bitumen
of organic sedimentary rocks. Geochimica et Cosmochimica Acta, 46, 2547- 2560.
Li, Y., Schieber, J., 2015. On the origin of a phosphate enriched interval in the Chattanooga Shale (Upper
Devonian) of Tennessee - A combined sedimentologic, petrographic, and geochemical study.
Sedimentary Geology, 329, 40–61.
Liao, W; Wang, Y., Kershaw, S., Weng, Z., Yang, H., 2010. Shallow-marine dysoxia across the Permian–
Triassic boundary: Evidence from pyrite framboids in the microbialite in South China.
Sedimentary Geology, 232, 77–83.
Loboziak, S., Caputo, M.V., Melo, J.H.G., 2000. Middle Devonian-Tournaisian miospore biostratigraphy
in the southwestern outcrop belt of the Parnaíba Basin, North-Central Brazil. Revue de
Micropaléontologie, 43, 301–318.
MacRae, N.D., Nesbitt, H.W., Kronberg, B.I., 1992. Development of a positive Eu anomaly during
diagenesis. Earth and Planetary Science Letters, 109, 585-591.
Majewski, W., 2000. Middle Jurassic concretions from Częstochowa (Poland) as indicators of
sedimentation rates. Acta Geologica Polonica, 50, 431-439.
Maynard, J.B., 1982. Extension of Berner's "New Geochemical Classification of Sedimentary
Environments" to ancient sediments. Journal of Sedimentary Petrology, 52, 4, 1325—1331.
Markello, J.R., Koepnick, R.B., Waite, L.E., Collins, J.F., 2008. The Carbonate Analogs Through Time
(CATT) hypothesis and the global atlas of carbonate fields - A systematic and predictive look at
Phanerozoic carbonate systems. In: Lukasik, J., Simo, J.A.T. (Eds.), Controls on carbonate
platform and reef development. SEPM Special Publication, 89, 15–45.

Martins, L.P., 2019. Estratigrafia química e potencial gerador da Formação Pimenteiras, Bacia do
Parnaíba. Tese de Doutorado. Universidade Estadual do Rio de Janeiro, Rio de Janeiro, 170 pp.

Masterson, W.D., Paris, C.E., 1987. Depositional history and reservoir description of the Kuparuk River
oil field, North Slope, Alaska. In: Alaskan North Slope Geology, Volumes I and II, 012 Pacific
Section, Society for Sedimentary Geology (SEPM), 95-106.
Mazzullo, S.J., 2000. Organogenic Dolomitization in Peritidal to Deep-Sea Sediments. Journal of
Sedimentary Research, 70, 10 - 23.
McArthur, J.M., Benmore, R.A., Coleman, M.L., Soldi, C., Yeh, H.M., O'Brien, G.W. 1986. Stable
isotopic characterisation of francolite formation. Earth and Planetary Science Letters, 77, 20-34.
143

McArthur, J.M., Walsh, J.N. 1984. Rare-earth geochemistry of phosphorites. Chemical Geology, 47,
191—220.
McEachern, J.A., Pemberton, S.G., Gingras, M.K., Bann, K.L., 2010. Ichnology and facies models. In:
James, N.P., Dalrymple, R.W. (Eds.), Facies Models 4. Geological Association of Canada, St.
John’s, Geotext, 6, 19–58.

Meyer, K.M.; Kump, L.R., 2008. Oceanic Euxinia in Earth History: Causes and Consequences. Annual
Review Earth Planetary Sciences 36, 251–288.
Melo, J.H.G., 2002. Revisão da biocronoestratigrafia de miósporos do Devoniano – Carbonífero Inferior
da bacia do Amazonas e correlação com outras bacias paleozoicas brasileiras. Tese de Doutorado,
IGEO-UFRJ, Rio de Janeiro. 103p.

Menzies, L.A., Carter, A., MacDonald, D.I.M., 2018. Evolution of a cratonic basin: insights from the
stratal architecture and provenance history of the Parnaíba Basin. In: Daly, M.C., Fuck, R.A.,
Julià, J., Macdonald, D.I.M., Watts, A.B. (Eds.), Cratonic Basin Formation: A Case Study of the
Parnaíba Basin of Brazil. Geological Society, London, Special Publications, 472, 157–179.
Merrot, P., Juillot, F., Noël, V., Lefebvre, P., Brest, J., Menguy, N., Guigner, J-M., Blondeau, M., Viollier,
E., Fernandez, J-M., Moreton, B., Bargar, J.R., Morin, G., 2019. Nickel and iron partitioning
between clay minerals, Fe-oxides and Fe sulfides in lagoon sediments from New Caledonia.
Science of the Total Environment, 689, 1212–1227.
Morad, S., Ketzer, J.M., De Ros, L.F., 2012. Linking Diagenesis to Sequence Stratigraphy: An Integrated
Tool for Understanding and Predicting Reservoir Quality Distribution. In: Morad, S., Ketzer,
J.M., De Ros, L.F. (Eds.), Linking Diagenesis to Sequence Stratigraphy. IAS Special Publications
45, 1–36.
Mortimer, R. J. G., Coleman M. L, Era J. E., 1997. Effect of bacteria on the elemental composition of
early diagenetic siderite: implications for palaeoenvironmental interpretations. Sedimentology, 44,
759-765.
Mozley, P.S., 1989. Relation between depositional environment and the elemental composition of early
diagenetic siderite. Geology, 17, 704-706.
Mozley, P.S., Carothers, W.W., 1992. Elemental and isotopic composition of siderite in the Kuparuk
Formation, Alaska: effect of microbial activity and water/sediment interaction on early pore-water
chemistry. Journal of Sedimentary Petrology, 62, 681–692.
Mozley, P.S., Wersin, P., 1992. Isotopie composition of siderite as an indicator of depositional
environment. Geology, 20, 817-820.
Murphy, A.E., Sageman, B.B., Hollander, D.J., Lyons, T.W., Brett, C.E., 2000. Black shale deposition and
faunal overturn in the Devonian Appalachian Basin: Clastic starvation, seasonal water-column
mixing, and efficient biolimiting nutrient recycling. Paleoceanography, 15, 280–291.
Nathan, Y., 1984. The mineralogy and geochemistry of phosphorite. In: (Nriagu, J. O. and Moore, P. B.,
(Eds.), Phosphate Minerals, (Chapter 8), Spriger-Verlag, Berlin, 257–291.
Nozaki, Y., Zhang, Amakawa, J.H., 1997. The fractionation between Y and Ho in the marine
environment. Earth and Planetary Science Letters, 148, 329-340.
Odin, G.S., 1988. Green Marine Clays, Developments in Sedimentology 45. Elsevier, Amsterdam.
Odin, G.S., Sen Gupta, B.K., 1988. Geological significance of the verdine facies. In: Odin, G.S. (Ed.),
Green Marine Clays. Developments in Sedimentology 45, Elsevier, Amsterdam, 205–219.
144

Oliveira, J.C., Barros, F.L., 1976. Projeto Fosfato de São Miguel do Tapuio. Recife: CPRM, 1976. 4v.
Convênio DNPM/CPRM. Relatório interno.
Oliveira, L.C., Pereira, E., 2011. Ocorrência de Ironstones no Devoniano da Bacia do Paraná. Revista
Brasileira de Geociências, 4, 447-462.

Orris, G.J., Dunlap, P., Wallis, J.C., 2015, Phosphate occurrence and potential in the region of
Afghanistan, including parts of China, Iran, Pakistan, Tajikistan, Turkmenistan, and Uzbekistan,
with a section on geophysics by Jeff Wynn: U.S. Geological Survey Open-File Report, 1121, 70
p. http://dx.doi.org/10.3133/ofr20151121.

Parrish, J.T., 1982. Upwelling and petroleum source beds, with reference to Paleozoic. American
Association of Petroleum Geologists Bulletin, 66, 750-774.
Pearce, T.J., Jarvis, I., 1992. Composition and provenance of turbidite sands: Late Quaternary, Madeira
Abyssal Plain. Marine Geology, 109, 21-51.
Pecoits, E., Gingras, M.K., Barley, M.E., Kappler, A., Posth, N.R., Konhauser, K.O., 2009. Petrography
and geochemistry of the Dales Gorge banded iron formation: Paragenetic sequence, source and
implications for palaeo-ocean chemistry. Precambrian Research, 172, 163–187.
Pemberton, S.G., MacEachern, J.A., Dashtgard, S.E., Bann, K.L., Gingras, M.K., Zonneveld, J.P., 2012.
Shorefaces. In: Knaust, D., Bromley, R.G. (Eds.), Trace fossils as indicators of sedimentary
environments. Elsevier, Developments in Sedimentology, 64, 563–603.
Percival, L.M.E., Selby, D., Bond, D.P.G., Rakociński, M., Racki, G., Marynowski, L., Adatte, T.,
Spangenberg, J.E., Föllmi, K.B., 2019. Pulses of enhanced continental weathering associated with
multiple Late Devonian climate perturbations: Evidence from osmium-isotope compositions.
Palaeogeography, Palaeoclimatology, Palaeoecology, 524, 240–249.
Pereira, E., Rodrigues, R., Souza, M.S.P., Bergamaschi, S., 2009. Chemostratigraphy Applied to Paleozoic
Black-Shale Intervals. AAPG International Conference and Exhibition, Rio de Janeiro, Brazil, 15-
18.
Petsch, S.T., Berner, R.A., Eglinton, T.I., 2000. A field study of the chemical weathering of ancient
sedimentary organic matter. Organic Geochemistry, 31, 475-487.
Phan, T.T., Hakala, J.A., Lopano, C.L., Sharma, S. 2019. Rare earth elements and radiogenic strontium
isotopes in carbonate minerals reveal diagenetic influence in shales and limestones in the
Appalachian Basin. Chemical Geology, 509, 194–212.
Ponciano, L.C.M. de O., da Fonseca, V.M.M., Machado, D.M. da C., 2012. Taphofacies analysis of late
early Givetian fossil assemblages of the Parnaíba Basin (State of Piauí, northeast Brazil).
Palaeogeography, Palaeoclimatology, Palaeoecology 326–328, 95–108.
Potter, P., Maynard, J., Depetris, P., 2005. Mud and mudstones: introduction and overview. Berlin:
Springer.
Poulton, S.W., Canfield, D.E., 2011. Ferruginous Conditions: A dominant feature of the ocean through
Earth’s history. Elements, 7, 107–112.
Pufahl, P.K., 2010. Bioelemental Sediments. In: James, N.P.; Dalrymple, R.W. (Eds.) Facies Models 4.
Canada, Geological Association of Canada, 477-504.
Pufahl, P.K., Grimm, K.A., 2003. Coated phosphate grains: proxy for physical, chemical and ecological
changes in seawater. Geology, 31, 801-804.
145

Pufahl, P.K., Grimm, K.A., Abed, A.M., Sadaqah, R.M., 2003. Upper Cretaceous phosphorites in Jordan:
implications for the formation of a south Tethyan phosphorite giant. Sedimentary Geology, 161,
175-205.
Pufahl, P.K., Hiatt, E.E., 2012. Oxygenation of the Earth’s atmosphere–ocean system: a review of
physical and chemical sedimentologic responses. Marine and Petroleum Geology, 32, 1–20.
Pufahl, P.K., Squires, A.D., Murphy, B., Quesada, C., Lokier, S.W., Álvaro, J.J., Hatch, J., 2020.
Ordovician ironstone of the Iberian margin: coastal upwelling, ocean anoxia and Palaeozoic
biodiversity. The Depositional Record, 00, 1–24.
Pufahl, P.K.; Groat, L.A., 2017. Sedimentary and Igneous Phosphate Deposits: Formation and
Exploration: An Invited Paper. Economic Geology, 112, 3, 483-516.
Pye, K., 1981. Marshrock formed by iron sulphide and siderite cementation in saltmarsh sediments,
K. Nature, 294, 650–652.
Pye, K., Dickson, J.A.D., Schiavon, N., Coleman, M.L., Cox, M., 1990. Formation of siderite-Mg-calcite-
i-iron sulphide concretions in intertidal marsh and sandflat sediments, north Norfolk, England.
Sedimentology, 31, 325-343.
Redfield, A.C., Friedman, I., 1965. Factors affecting the distribution of deuterium in the ocean. In
Symposium on Marine Geochemistry. Rhode Island University Narragansett Marine Laboratory
Occasional Publication, 3, 149-168.
Redfield, A.C., Ketchum, B.H., Richards, F.A., 1963. The influence of organisms on the composition of
seawater. In: Hill, M.N. (Ed.), The Sea, 2. Wiley, New York, 26–77.
Ribeiro C.C. & Dardenne M.A., 1978. O minério de ferro da Formação Pimenteiras na borda SW da Bacia
do Maranhão (Goiás). In: SBG, Congresso Brasileiro de Geologia, 30, Anais 4, 1583-1595.
Ribeiro C.C., 1984. Caractérisation sédimentologique et géochimique d’un milieu sédimentaire. Cas du
Dévonien Moyen et Supérieur de la Région de Paraíso do Norte – Miranorte (Bassin de
Maranhão, Goiás, Brésil). Thèse Docteur de 3ème cycle. I’Institut National Polytechinique de
Lorraine. 252p.
Riggs S. R., Sheldon R.P., 1990. Paleoceanographic and paleoclimatic controls of the temporal and
geographic distribution of Upper Cenozoic continental margin sediments.
In: Burnett W. C., Riggs S. R. (Eds), Phosphate Deposits of the
World. Neogene to Modern Phosphorites. Cambridge University Press, 3, 53–61.
Rimmer, S.R., Thompson, J.A., Goodnight, S.A., Robl, T.L., 2004. Multiple controls on the preservation
of organic matter in Devonian–Mississippian marine black shales: geochemical and petrographic
evidence. Palaeogeography, Palaeoclimatology, Palaeoecology, 215, 125– 154.
Rimstidt, J.D., Chermak, J.A., Schreiber, M.E., 2017. Processes that control mineral and element
abundances in shales. Earth-Science Reviews, 171, 383–399.
Rodrigues, R., 1995. A geoquímica orgânica na Bacia do Parnaíba. Tese (Doutorado), Universidade
Federal do Rio Grande do Sul, Porto Alegre, 226 pp.
Rodrigues, R; Pereira, E.; Bergamaschi, S.; Bastos, L.P.H. 2019. Stable isotopes as a tool for stratigraphic
studies: Insights from the Brazilian sedimentary record. In: Montenary, M. Case Studies in
Isotope Stratigraphy, Volume 4, AcademicPress, 134-164.
Royer, D.L., 2006. CO2-forced climate thresholds during the Phanerozoic. Geochimica et Cosmochimica
Acta, 70, 5665–5675.
146

Rudmin, M., Mazurov, A., Banerjee, S., 2019. Origin of ooidal ironstones in relation to warming events:
Cretaceous-Eocene Bakchar deposit, south-east Western Siberia. Marine and Petroleum Geology,
100, 309–325.
Ruttenberg, K. C., 2003. The Global Phosphorus Cycle. In: Schlesinger, W.H. (Ed.); Holland, H.D.;
Turekian, K.K. (Exec. Ed.). Treatise on Geochemistry 8. ISBN 0-08-043751-6. Elsevier, p.585-
643.
Ruttenberg, K.C., Goñi, M.A., 1997. Phosphorus distribution, C:N:P ratios, and δ13Coc in arctic,
temperate, and tropical coastal sediments: tools for characterizing bulk sedimentary organic
matter. Marine Geology, 139, 123–145.
Sageman, B.B., Murphy, A.E., Werne, J.P., Ver Straeten, C.A., Hollander, D.J., Lyons, T.W., 2003. A tale
of shales: the relative roles of production, decomposition, and dilution in the accumulation of
organic-rich strata, Middle–Upper Devonian, Appalachian Basin. Chemical Geology, 195, 229–
273.
Salama, W., Khirekesh Z., Amini, A., Bafti, B.S., 2018. Diagenetic evolution of the upper Devonian
phosphorites, Alborz Mountain Range, Northern Iran. Sedimentary Geology, 376, 90-112.
Scholle, P.A., Arthur, M.A., 1980. Carbon Isotope Fluctuations In Cretaceous Pelagic Limestones:
Potential Stratlgraphic and Petroleum Exploration Tool. The American Association of Petroleum
Geologists Bulletin, 64, 1, 6 7 - 87.
Schieber, J., 2011. Marcasite in Black Shales--a mineral proxy for oxygenated bottom waters and
intermittent oxidation of carbonaceous muds. Journal of Sedimentary Research 81, 447–458.
Schieber, J., Baird, G., 2001. On the origin and significance of pyrite spheres in Devonian Black Shales of
North America. Journal of Sedimentary Research, 71, 155–166.
Schobbenhaus, C., Gonçalves, J.H., Santos, J.O.S., Abram, M.B., Leão Neto, R., Matos, G.M.M., Vidotti,
R.M., Ramos, M.A.B., Jesus, J.D.A., 2004. Carta Geológica do Brasil ao Milionésimo, Sistema de
Informações Geográficas-SIG e 46 folhas na escala 1:1.1.000.000. CPRM, Brasilia.
Schovsbo, N.H. 2001. Why barren intervals? A taphonomic case study of the Scandinavian Alum Shale
and its faunas. Lethaia, 34, 271–285.
Scott, B.T., 2009. Variation in C/P ratios in Devonian-Mississippian marine shales: testing the
productivity-anoxia feedback model. University of Kentucky Master's Theses. 609.
https://uknowledge.uky.edu/gradschool_theses/609.
Selleck, B., Koff, D., 2008. Stable Isotope Signature of Middle Devonian Seawater from Hamilton Group
Brachiopods, Central New York State. Northeastern Geology & Environmental Sciences, 30, 4,
330-343.
Sharafi, M., Mahboubi, A., Moussavi-Harami, R., Mosaddegh, H., Gharaie, M.H.M., 2014. Trace fossils
analysis of fluvial to open marine transitional sediments: Example from the Upper Devonian
(Geirud Formation), Central Alborz, Iran. Palaeoworld, 23, 50–68.
Shields, G., Stille, P., 2001. Diagenetic constraints on the use of cerium anomalies as palaeoseawater
redox proxies: an isotopic and REE study of Cambrian phosphorites. Chemical Geology, 175, 29-
48.
Sholkovitz, E.R., Landing, W.M., Lewis, B.L., 1994. Ocean particle chemistry: The fractionation of rare
earth elements between suspended particles and seawater. Geochimica et Cosmochimica Acta, 58,
6, 1567-1579.
Silveira, D.A., Sachs, L.L.B., Moraes Filho, J.C.R., Cunha, I.A., Batista, I.H., Silva, R.C., 2016. A
pesquisa para fosfato na Bacia Parnaíba, área Picos, Piauí. In: Abram, M.B., Bahiense, I.C.,
147

Almeida, R.C. (Eds.), Projeto Fosfato Brasil: parte II, Informe de Recursos Minerais, Série
Insumos Minerais para Agricultura 17, Salvador, CPRM, 1034-1086.
Smith, K.S., 2007. Strategies to predict metal mobility in surficial mining environments. Reviews in
Engineering Geology, 17, 25-45.
Souza, I.V.A.F., 2007. Faciologia orgânica de seções Devonianas da bacia do Parnaíba (Formação
Pimenteira): Implicações para geração de petróleo. Dissertação de Mestrado, IGEO/UFRJ, Rio de
Janeiro, 164pp.
Squires, A., 2019. Ordovician ironstone of the Western Asturian-Leonese zone, Spain: coastal upwelling,
Ocean anoxia and paleozoic biodiversity. Msc Thesis. Wolfville, Nova Scotia, Canada.
https://scholar.acadiau.ca/islandora/object/theses:3230.
Storms, J.E.A., 2003. Event-based stratigraphic simulation of wave-dominated shallow-marine
environments. Marine Geology, 199, 83-100.
Taylor S.R., McLennan S.H.,1985. The Continental Crust: Its Composition and Evolution. Blackwell.
Oxford, 312 p.
Taylor, K.G., Curtis, C.D., 1995. Stability and facies association of early diagenetic mineral assemblages:
an example from a Jurassic ironstone-mudstone succession. Journal of Sedimentary Research, 65,
358-368.
Taylor, K.G., Macquaker, J.H.S., 2011. Iron minerals in marine sediments record chemical environments.
Elements, 7, 113–118.
Todd, S.E., Pufahl, P.K., Murphy, J.B. and Taylor, K.G., 2019. Sedimentology and oceanography of Early
Ordovician ironstone, Bell Island, Newfoundland: Ferruginous seawater and upwelling in the
Rheic Ocean. Sedimentary Geology, 379, 1–15.
Tosca, N.J., Guggenheim, S., Pufahl, P.K., 2015. An authigenic origin for Precambrian greenalite:
Implications for iron formation and the chemistry of ancient seawater. Geological Society of
America Bulletin, 128, 511–530.
Trabucho-Alexandre J., Dirkx R., Veld H., Klaver G., de Boer P.L. 2012. Toarcian black shales in the
Dutch Central Graben: Record of energetic, variable depositional conditions during an oceanic
anoxic event. Journal of Sedimentary Research, 82, 2, 104-120.
Trappe, J., 1998. Phanerozoic Phosphorite Depositional Systems: A Dynamic Model for a Sedimentary
Resource System. Lecture Notes Earth Science 74, 316pp.
Tribaldos, V.R., White, N., 2018. Implications of preliminary subsidence analyses for the Parnaíba
cratonic basin. In: Daly, M.C., Fuck, R.A., Julià, J., Macdonald, D.I.M., Watts, A.B. (Eds.),
Cratonic Basin Formation: A Case Study of the Parnaíba Basin of Brazil. Geological Society,
London, Special Publications 472, 147–156.
Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox and
paleoproductivity proxies: An update Chemical Geology, 232, 12–32.
Trindade, V.S.F., Carvalho, M. de A., 2018. Paleoenvironment reconstruction of Parnaíba Basin (north,
Brazil) using indicator species analysis (IndVal) of Devonian microphytoplankton. Marine
Micropaleontology, 140, 69–80.

Trindade, V.S.F., Carvalho, M. de A., Borghi, L., 2015. Palynofacies patterns of the Devonian of the
Parnaíba Basin, Brazil: paleoenvironmental implications. Journal of South American Earth
Sciences, 62, 164–175.
148

Trucco-Pignata, P.N., Hernández-Ayón, J.M., Santamaria-del-Angel E., Beier, E., Sánchez-Velasco, L.,
Godínez, V.M., Norzagaray, O., 2019. Ventilation of the Upper Oxygen Minimum Zone in the
Coastal Region Off Mexico: Implications of El Niño 2015–2016. Frontiers in Marine Science, 6,
459. doi: 10.3389/fmars.2019.00459.

Vaz, P.T., Rezende, N.G.A.M, Wanderley Filho, J.R, Travassos, W.A.S., 2007. Bacia do Parnaíba.
Boletim de Geociências da Petrobras 15, 289-297.

Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.K.F., Diener, A., Ebneth, S.,
Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., Strauss, H. 1999. 87Sr/86Sr, δ
13
C and δ 18O evolution of Phanerozoic seawater. Chemical Geology, 161, 59–88.
Veizer, J., Prokoph, A. 2015. Temperatures and oxygen isotopic composition of Phanerozoic oceans.
Earth-Science Reviews, 146, 92-104.
Velde, B., 2005. Green clay minerals. In: Mackensie, F.T. (Ed.), Sediments, Diagenesis, and Sedimentary
Rocks 7, Treatise on Geochemistry (Eds. Holland, H.D., Turekian, K.K.). Elsevier, Pergamon,
Oxford, 309-324.
Villeneuve, M., 2005. Paleozoic basins in West Africa and the Mauritanide thrust belt. Journal of African
Earth Sciences, 43, 166–195.
Walker, R.G., Plint, A.G., 1992. Wave and storm shallow marine systems. In: Walker, R. G., and
James, N. P., (Eds.), Facies models: response to sea level change. Geological Association of
Canada, Geotext, 1, 219-238.
Watkins, R.T., Nathan, Y., Bremner, J.M., 1995. Rare earth elements in phosphorite and associated
sediment from the Namibian and South African continental shelves. Marine Geology, 129, 111-
128.
Weissert, H., Joachimski, M., Sarnthein, M., 2008. Chemostratigraphy. Newsletter in Stratigraphy, 42 (3),
145–179.
Wildman, R.A., Berner, R.A., Petsch, S.T., Bolton, E.W., Eckert, J.O., Mok, U., Evans,, B.J., 2004. The
weathering of sedimentary organic matter as a control on atmospheric O2: I. Analysis of a black
shale. American Journal of Science 304, 234–249.
Witkowska, M., 2012. Palaeoenvironmental significance of iron carbonate concretions from the Bathonian
(Middle Jurassic) ore-bearing clays at Gnaszyn, Kraków-Silesia Homocline, Poland. Acta
Geologica Polonica, 62, 307–324.
Yilmaz, İ.Ö., Göncüoğlu, M.C., Demiray, D.G., Gedik, I., 2015. An approach to paleoclimatic conditions
for Devonian (upper Lochkovian and middle Givetian) ironstone formation, NW Anatolian
carbonate platform. Turkish Journal of Earth Sciences, 24, 21-38.
Young, C.G.K., 2006. Estratigrafia de alta resolução da Formação Pimenteira (Devoniano, Bacia do
Parnaíba). Programa de Pós-graduação em Geologia, Universidade Federal do Rio de Janeiro,
Dissertação de Mestrado, 174 pp.
Young, T.P., 1989. Phanerozoic ironstones: an introduction and review. In: Young, T.P., Taylor, W.E.G.
(Eds.), Phanerozoic Ironstones. Geological Society, London, Special Publications 46, 9-25.
Zang, J., Liu, Q., He, Q., 2019. Rare Earth Elements and their isotope in the ocean. In: Cochran, J.K.,
Bokuniewicz, H.J., Yager, P.L. Encyclopedia of Ocean Sciences. Elsevier, London.
Zarasvandi, A., Fereydouni, Z., Pourkaseb, H., Sadeghi, M., Mokhtari, B., Alizadeh, B., 2019.
Geochemistry of trace elements and their relations with organic matter in Kuh-e-Sefid phosphorite
mineralization, Zagros Mountain, Iran. Ore Geology Reviews, 104, 72–87.
149

Zhang, C.L., Horita, J., Cole, D.R., Zhou, J., Lovley, D.R., Phelps, T.J., 2001. Temperature-dependent
oxygen and carbon isotope fractionations of biogenic siderite. Geochimica et Cosmochimica Acta,
65, 14, 2257–2271.
Zymela, S., 1996. Carbon, oxygen and strontium isotopic composition of diagenetic calcite and siderite
from the Upper Cretaceous Cardium Formation of Western Alberta, PhD thesis, McMaster
University, Canada, 156pp.
150

CAPÍTULO 4
CONCLUSÕES

Este trabalho permitiu a definição de um arcabouço estratigráfico para o intervalo entre o

Praguiano e o Givetiano da Bacia do Parnaíba e interpretações paleoambientais e diagenéticas que deram

suporte à caracterização das acumulações fosfáticas neste período, cruciais para o entendimento dos

processos de fosfogênese, suas características, restrições, bem como o estabelecimento de modelos

preditivos.

Observamos que uma série de eventos ocorreu do Devoniano Inferior ao Devoniano Médio que

propiciaram momentos episódicos de acúmulo de sedimentos moderamente enriquecidos em matéria

orgânica, fosforitos e rochas fosfáticas e ironstones muitas vezes fosfáticos. Estes eventos ocorreram no

norte do paleocontinente Gondwana, na margem sul do oceano Rheic, ao longo de latitudes que definem

zonas temperadas, mas numa época de condições climáticas de alto carbono atmosférico, baixo oxigênio e

efeito estufa. Este período coincide com momentos, onde a flora sofreu uma grande revolução devido ao

engrandecimento dos corpos das plantas e grande enraizamento que, por sua vez, devem ter facilitado os

processos de intemperismo e a biodisponibilidade de nutrientes para o oceano.

O padrão de empilhamento do Grupo Canindé descrito para este intervalo de tempo indica que a

deposição dos sedimentos ocorreu por transgressões marinhas, com pequenas flutuações sobrepostas

gerando discordâncias regionais e também locais, especialmente nas zonas mais marginais da bacia, sendo

estas controladas por fenômenos eustáticos. O intervalo do Praguiano ao Givetiano foi dividido em quatro

seqüências de terceira ordem (B, C, D e E) controladas principalmente por transgressões e regressões

normais. As litofácies e as icnofácies associadas representam sucessões hidrodinâmicas em um ambiente


151

marinho raso, que representam ambientes do tipo influenciado por tempestades, também plataforma

dominada por tempestades e ambientes flúvio-deltaicos que caracterizam principalmente tendências

progradacionais em tratos de sistema de nível alto e também em regressão forçada. Os fosforitos e os

ironstones fosfáticos estão relacionados a depósitos de offshore (outer shelf), da zona de transição

offshore-lower shoreface (inner shelf) até depósitos de lower shoreface que foram reconhecidos em

sucessões transgressivas e de regressão normal.

Verificou-se que as paraseqüências ricas em fosforito são compostas por folhelhos bioturbados

e rochas heterolíticas bioturbadas depositadas na zona plataformal distal, próxima à base de ondas de

tempestade e em zonas condensadas (TST e porção inicial do TSNA). Os ironstones fosfáticos e também

ironstones (com teores menores que 1% de P2O5 e considerados não fosfáticos) ocorrem em zonas mais

proximais, na zona de transição lower shoreface-offshore. Estes associam-se a sequências paragenéticas

caracterizadas por arenitos muito finos e rochas heterolíticas com estrutura wavy, ambos bioturbados,

ichnofacies Cruziana e arenitos finos a muito finos com estratificação cruzada tipo hummocky. Foram

depositados perto do limite da base de ondas de tempo bom (fairweather wave base - FWWB) e estão

associados a porções iniciais do Trato de Sistemas de Nível Alto (TSNA) e mais finais do Trato de

Sistemas de Nível Baixo (TSNB), refletindo um leve aumento na influência de sedimentos terrestres. Isto

significa que a mudança relativa do nível do mar e o aumento da influência marinha, bem como o aporte

de nutrientes devido a descragas fluviais, regulou a natureza, extensão e a deposição da materia orgânica

ao longo da bacia e consequentemente dos fosfatos e dos ferrosilicatos e carbonatos de ferro.

Para as seqüências estudadas, foram definidos quatro estágios paragenéticos principais para o

conjuto de litofácies portadoras de sucessões bioelementais: detrítico; autigênico/retrabalhado

(eodiagenético); soterramento raso (mesodiagenético raso) e; intemperismo. Os fosforitos e os ironstones

estão relacionados ao estágio autigênico.


152

De uma maneira geral a região costeira foi marcada por águas predominantemente óxicas,

entretanto marcada por momentos episódicos marcados por condições subóxicas e anóxicas. Portanto,

consideramos que a ressurgência (upwelling) episódica manteve o suprimento de Fe, Si e P, nas porções

mais distais e com fontes adicionais de ferro associadas às descargas dos rios diretamente para a zona da

plataforma mais mediana.

Dois períodos principais de deposição de fósforo foram associados às superfícies de inundação

máxima do Eifeliano tardio e do Givetiano, favorecida pelos momentos de transgressão e diminuição das

taxas de sedimentação, formando zonas condensadas. Nestes ambientes a acumulação de matéria orgânica

e degradação microbiana favoreceu a formação de fosforitos . As taxas de produtividade primária foram

suficientes para esgotar a água do fundo de oxigênio dissolvido, criando condições mais anóxicas. Estes

fosforitos foram depositados em um oceano estratificado, subóxica na zona de inner shelf e mais anóxica

na zona de outer shelf. Em condições de interface subóxica/anóxicas, o fósforo foi liberado por

degradação microbiana em uma zona sulfato redutora. A reciclagem de entrega de nutrientes alimentou a

produtividade orgânica em posições ligeiramente mais rasas, gerando associações ricas em glauconita em

zonas mais subóxicas (interface óxica/subóxica).

Os ironstones fosfáticos são representados por “coated grains” formados por odinita, chamosita

/ clinocloro e francolita principalmente híbridos, em uma matriz com chamosita autigênica (ou também

clinocloro), Mg siderita, siderita, framboides de pirita e matéria orgânica parcialmente degradada. Os

“coated grains” de odinita e francolita, foram precipitados nas primeiras fases eodiagenéticas associados a

zonas condensadas e ambientes subóxicos. A degradação progressiva da matéria orgânica alterou o limite

de oxi-redução, gerando condições mais anóxicas, sulfato redutoras, sob condições de consumo total de

H2S por formação de pirita e excesso de ferro, com precipitação de Mg siderita, devido a liberaçao

progressiva de HCO3-. A atuação de tempestades levaram ao retrabalhamento destes “coated grains”

formando grãos híbridos com alternância de feiçoes “redox-agraded” e “unconformity bounded”,


153

formando níveis de chamosita e ou/clinocloro sob condições mais anóxicas. Em seguida, com a retomada

de condições mais óxicas e expansão da bioturbação, e com a influência de soluções aquosas mais

alcalinas, ricas em ferro, em contato com zonas com HCO3-- em excesso, sem sulfato foram formadas as

ferro sideritas mais tardias, possivelmente relacionadas às zonas ferro redutoras. Possivelmente este

processo inibiu a contínua precipitação de francolita. Os processos dinâmicos induzidos pela tempestade

resultaram no acúmulo de fosforitos e ironstones retrabalhados.

Os padrões de TRY obtidos são marcados por enriquecimentos nas TR mais medianas (incluindo os

fosforitos, as rochas siliciclásticas fosfáticas, os francolita-glauconita nódulos, braquiópodes fosfáticos, os ironstones

fosfáticos, alguns ironstones não fosfáticos, sideritas e francolitas) relacionados ao processo de eodiagênese. Embora

estes padrões possam representar condições subóxicas ou anóxicas, este é um padrão destacado em fosforitos

autigênicos e conchas de braquiopódes observados no Devoniano em outras partes do mundo. Isto sugere uma

possibilidade de que estes padrões também possam de alguma forma representar um padrão de assinatura global que

reflita condições mais comuns de anoxia para a água do mar no Devoniano. Esta condição também é sugerida por

outros marcadores geoquímicos, como por exemplo pelas razões Corg/P globais (Algeo e Ingall, 2007).

As assinaturas isotópicas de δC13 e δO18 obtidas aqui também sugerem que as sideritas relacionadas aos

ironstones fosfáticos precipitaram sob condições de atividade microbiana sulfato redutoras relacionadas a uma fonte

enriquecida em ferro e sulfato dissolvido na água do mar. Entretanto também indica que parte das sideritas sofreram

a influência de outras zonas, aqui interpretada como ferro redutora, e de possíveis variações de

salinidade/temperatura. A matéria orgânica parcialmente degradada e substituída por cristais framboidais de pirita

confirma que a diagênese evoluiu em uma zona sulfato redutora. Dados geoquímicos em rocha total, através de

indicadores como as razões V/Cr, V/(V+Ni) e Mo confirmam que os intervalos com ironstones fosfáticos e

ironstones indicam condições subóxicas e anóxicas variáveis, com mínima quantidade de H2S dissolvido na água

interporos. Variações em δO18 ressaltam as possíveis misturas em água do mar e água meteórica nas fases mais

tardias de precipitação.
154

Este trabalho também sugere que eventos globais de acumulação de matéria orgânica de terceira

ordem podem possivelmente ser rastreados por seções na América do Sul e possivelmente se correlacionar

com as mudanças globais do nível do mar. Como exemplo citamos a correspondência de uma zona

diacrônica e estreita, porém presente em toda a bacia após o final do Emsiano que poderia se correlacionar

aos eventos de extinção em massa do Eifeliano/Givetiano. Este nível estratigráfico registra um pico de

acúmulo de matéria orgânica terrígena e marinha, e altos valores em elementos que indicam anoxia, que

poderia representar um maior fluxo de nutrientes derivados do continente e e eutrofização, após um

estágio de queda do nível do mar.

O mecanismo de ressurgência pode ter sido relacionado a uma série de fatores como

estratificação do oceano, batimetria e circulação dirigida pelos ventos na camada superior dos oceanos

relacionados aos giros de pressão atmosférica. A mudança no nível de base e as mudanças no nível do mar

podem ter agido para mudar o local da ressurgência. A correlação com outras acumulações de ironstones

no mundo indica que processos episódicos de anoxia tanto relacionados à ressurgência e aos processos de

eutrofização devido ao aumento de nutrientes terrígenos não eram incomum em condições de greenhouse.

A precipitação global destas rochas deve também ter contribuído para o aumento da oxigenação através do

soterramento de carbono orgânico seja pela matéria orgânica soterrada ou pela precipitação de carbonatos.

Este trabalho evidencia que a explicação das causas globais de anoxia e soterramento de CO2 deve incluir

também o papel das mudanças no nível do mar, circulação oceânica, clima, disponibilidade de nutrientes e

redes alimentares microbianas eficientes.

As baixas taxas de sedimentação e as condições de produtividade foram suficientes para gerar

os fosforitos e as rochas fosfáticas. Entretanto, as descargas fluviais provavelmente afetaram as

características físico-químicas da água do mar produzindo uma grande variabilidade na coluna d’água

mais superior, influenciando as condições de temperatura, salinidade e estratificação do oceano gerando

diferentes gradientes e concentrações de nutrientes.


155

Os dados obtidos com este trabalho mostram que as contribuições terrígenas de matéria

orgânica, aparentemente frequentes na bacia nos intervalos mais progradacionais, acentuaram as

condições de anoxia nos ambientes mais rasos, diminuindo a quantidade de oxigênio na água de fundo e

interporos disponível devido a demanda de respiração microbiana. Possivelmente isto resultou numa

redução na eficiência dos processos de fosfogênese e num aumento da preservação da matéria orgânica.

Assim, conclui-se que o ciclo do fósforo foi provavelmente afetado pelas condições globais de

baixo oxigênio, alto CO2 e anoxia, frequentes no Devoniano. Embora a reciclagem de nutrientes de volta à

coluna d’água sob condições de baixa oxigenação manteve os níveis mais elevados em produtividade, o

soterramento do fósforo foi provavelmente afetado ao longo da plataforma nas zonas de costa. Portanto,

embora estejam associadas a zonas de upwelling, este fato provavelmente influenciou a economicidade

dos depósitos de fosfato devonianos. Por outro lado, as associações dos fosforitos com teores econômicos

em TR abrem uma nova perspectiva de potencial atrelado a este tipo de depósito.

Referências

Algeo, T.J., Ingall, E., 2007. Sedimentary Corg/P ratios, paleocean ventilation, and Phanerozoic

atmospheric pO2. Palaeogeography, Palaeoclimatology, Palaeoecology 256, 130–155.


156

APÊNDICE A – JUSTIFICATIVA DA PARTICIPAÇÃO DOS CO-


AUTORES

A participação do Co-autor Michael Holz (UFBA) se justifica pela participação do mesma como co-
orientador desta tese de Doutorado, pela sua participação na discussão especialmente nas etapas de
estratigrafia de sequências.

Prof. Claudio Porto foi convidado pela primeira autora para participar do segundo artigo pela sua
experiência e especialidade em geoquímica. Sua participação foi importante para consolidar o resultado
deste trabalho.
157

APÊNDICE B – MAPA COM A LOCALIZAÇÃO DOS FUROS DE


SONDAGEM E AFLORAMENTOS DESCRITOS.
158

Figura B.1 - Mapa com a localização dos furos e afloramentos estudados na parte leste da Bacia do
Parnaíba.
159

Figura B.2 - Mapa com a localização dos furos e afloramentos estudados na parte oeste da Bacia do
Parnaíba. Os furos 2_NA-1-TO e 2_LT-1_TO, apesar de descritos e interpretados, fazem parte da
seção Frasniana (Devoniano Superior) e não foram utilizados na definição do arcabouço
estratigráfico do Devoniano Inferior a Médio.
160

APÊNDICE C – PERFIS DOS FUROS DE SONDAGEM


DESCRITOS
161

1-VL-03-PI

0,0m

Cenozoic
Sandy cover

Ps10c
transition zone
Offshore/Lower shoreface
He(w)

10,0m

FS

weathering
Sr/Hew

Sr/Hew

ii2-3

transition zone
Offshore/Lower shoreface
20,0m

Ps9c
ii2-3

HST
He(w)

ii/(?)
30,0m
ii3-4

Sb/lp
Sb(org)/lp
Sh/Ip
Ip/S (org) - black
ii 2
FS

Sb/lp
Lower/Middle Shoreface

ii4-5

Ss2 ii2
Shcs ii1

Hew/Sr/lp ii4
40,0m
Sb/lp
Shcs ii1
Shcs ii1

Ss2 ii2

Shcs ii1
Early Givetian

HST

Shcs ii1

Hew (b) ii3-4


Ps8c

He sty
He(b) ii3-4
He sty (f,b)
50,0m ii4-5
Fm ii1

He sty (f,b)
ii4-5 Ii1-3 (He?, Pa?)
MFS-C
He sty (f,b)
ii3-4 (h, v, t)
Fb/Fm

Sedimentary structures and related features


ii1-2
Offshore

Cross bedded Ripples

Hummocky Swaley
TST

Wavy
Fm/Fb ii3 Flat parallel lamination
ii1-2
60,0m Trough Cross Bedded
Ps7c

Phosphatic Ironstone

He silty (w,b) M Massive


ii3-4
Intraclasts

Normal grading

Fm (black shale)ii1-2
Bioturbation
ii - Bioturbation index
Offshore/Lower shoreface
transition zone

He sty (w,b) Clasts


ii3-4 (pa, und) Pyrite

linsen
Late eifelian

ii4-5
MRS-C
70,0m
Shcs/Sb/lp

ii1
Shcs/Ss2/lp (green spots
- glauconite?)
transition zone
Lower Shoreface/offshore

Ps6c

Sb/Ip ii2-3

He sty (b) / Ip
LST

Ip/ Hew/ Sb(org)

He (w)/Ip (b)
C
S

ii2-3
C
VFS

CS
G
MS
C
S
RAIOS
ST-18-PI 0 GAMA 100
0,0m 162
CPS

xxxxxxxxx
xxxxxxxxx
Lower shoreface

xxxxxxxxx
xxxxxxxxx

xxxxxxxxx
xxxxxxxxx
xxxxxxxxx

Ss2

xxxxxxxxx
?weathered He? or FB or Fm?
Offshore/ Lower shoreface

TST
Ps12d
Transition zone

10,0m

xxxxxx
xxxxxx
Inferred interval

2% P2O5

2% P2O5
Sb-Sborg)/Ss2(green)

Sp
20,0m
M Ss2
Sp

Sb
ii 4
Ps11d

xxxxxxxxx Sb
xxxxxxxx Ss2M - 0
xxxxxxxxx
Sb (org)
xxxxxxxxx Sb ii 5
Lower to Middle Shoreface

xxxxxxxxx M

Shcs
xxxxxxxxx

Shcs
LST

30,0m M
Offshore/ Lower shoreface

Shcs

M
Transition zone

xxxxx

ii 2-3
xxxxx

I
Shoreface/distal delta?

Sp/Sc (? or structured )
ii 1
Lower to Upper

He silty/Sb
ii 4-5

ii 1
Ss1?

40,0m xxxxx M

xxxxx
Offshore/Lower shoreface transition zone

He silty/Ip
xxxxx

xxxxx ii 1-3
xxxxx

xxxxx Obs.: This well presents evidence of a structured


xxxxx part of the basin as the lamination can
He silty/(Ip)
be dipping around 25 degrees. So the stratigraphy
xxxxx
can have been affected by the meteor impact of
São Miguel do Tapuio.
xxxxx ii 1-3

xxxxx

xxxxx

xxxxx

xxxxx
He silty/ Ip

50,0m xxxxxxx
ii 4-5 Sb
Lower shoreface/offshore transition zone
Inferred interval

Sb
Giventian

ii 1-2
?
SB-D?
He (b)
ii 4-5
Lithology
60,0m Shcs ii 1-2
Sandstones

Heterolithe with microhummocky,


xxxxxx wavy or rarely linsen structures,

He (b) Bioturbated Shale - laminated


xxxxxx

ii 4-5
Pebbly sandstone

Contacts between bedding plans


xxxxxx Gradational
Smooth
Giventian?

He (b) Irregular

xxxxxx
Sedimentary structures and related features
Offshore/Lower shoreface transition zone

xxxxxx Cross bedded Ripples


70,0m
Hummocky Swaley
ii 5
HST

Wavy
Ps10c

Flat parallel lamination


Shcs/Sr/Hew
Trough Cross Bedded

Phosphatic Ironstone

M Massive

Intraclasts
N.D.
Normal grading
Sr? Hew?

Bioturbation
ii - Bioturbation index
Clasts
Pyrite
Shcs/Sr/Hew
linsen

80,0m
C
S
VFS
FS
MS
CS
163

ST-17-PI RAIOS
0 GAMA 100
0,0m
CPS

Ss2

xxxxxx
Shcs

Ss2

10,0m

Shcs ii1

xxxx
Weathered - mainly kaulinization

HST
Ps8c

Fb (silty)

ii 2-3 (Pa)
C sequence (Pimenteira Formation)

Fb (silty)

ii 2-3 (Pa)
Offshore

Fb (silty)

ii 2-3 (Pa)

MFS-C
xxxx
Fm

ii 1-2
20,0m

He (b, w)
Offshore/Lower shoreface transition zone

xxxx ii 2-3 (Pa)

xxx Fm
ii 1-2
TST
Ps7c

He (w, b)

ii 3-4

xxxxx

Fm

ii 1-2
30,0m xxxx

Lithology
He (b, w) Sandstones
Late Eifelian/ Early Giventian

ii 2-3
MRS-C
Heterolithe with microhummocky,
wavy or rarely linsen structures,
xxxxxx Hew/Sr
Lower shoreface /

Bioturbated Shale - laminated


Lower shoreface-

Ps6c

ii 3-4
LST

[org]
xxxxxx Sb/Ip
xxxxxx
Shcs
Pebbly sandstone
Offshore

He (b, w) Contacts between bedding plans


xxxxx ii 3-4 (Pa, Zoo)
Sr/He(w) Gradational
ii 3-4 (und, h)
xxxxxxx
Smooth
Sb [org]/Ip
xxxxxxx Irregular
SB-C
xxxxxxx ii 3-4

Sedimentary structures and related features


40,0m
Cross bedded
B sequence (Itaim Formation)

Hummocky
Praguian/Emsian?
Upper Shoreface

Ss1
Wavy
HST
Ps5b

Flat parallel lamination

Trough Cross Bedded

Phosphatic Ironstone

M Massive
Ssw
Intraclasts
Sm
M Normal grading

Bioturbation
ii - Bioturbation index
Clasts
Pyrite

linsen
50,00m
C
S
VFS
FS
MS
CS
00,0m
ST-15-Pi
164

distal?
Wave influenced delta
Ss2?/Sp/ (Shcs)

Transition zone
Offshore/Lower shoreface

Ps14d
Shcs/Ip/He silty
ii2-3 (v, As)
10,0m

FS

HST
Sb

Lower shoreface
M Ss/Sp

Ps13d
Transition zone
Lower shoreface/Offshore
Shcs/Ip

M Ss2
MFS-D
Ip (b)/He silty
20,0m

Fm

Offshore

TST

Ps12d
Hew (b)

Fm

Sb
ii5-6 (v,h,t)
muddy
Shcs

Hew (b)
30,0m

Ps11d
Transition zone
Lower shoreface/ Offshore
Shcs

Hew (b)

SB-D
40,0m Shcs/Sb (lag)
Hew (b)
Lower shoreface/Offshore transition zone

Sh
Hew (b)

Shcs

Hew (b)
ii2-3 (Sk, Spy)
Ps10c
Offshore

FS Fm

Hew ( b)
ii3-4 (sk, hor)
Offshore

FS Fm
Sb/Sr
ii2-3

60,0m
Transition zone
Offshore/Lower shoreface

He sty

He sty
Sedimentary structures and related features

Cross bedded Ripples

Hummocky Swaley

Wavy
70,0m
Inferred Interval

Flat parallel lamination

Trough Cross Bedded


HST

Ps9c

Phosphatic Ironstone

M Massive

Intraclasts

Normal grading

Bioturbation
ii - Bioturbation index
He sty Clasts
ii1-2 (As?) Pyrite

linsen

He sty

80,0m

FS
82,0m
165

RAIOS
ST-12-PI 0 GAMA 100
0,0m CPS

Sb
ii 3

ii 1

Sp

10,0m
Sc
Sp
Fm2

Quasi planar

Sp

M
Deformação sin-sedimentar
20,0m
ii 1

Ssg
Ss1
Ssg Asg quase planar
sigmoidal
M Ss1
M Ss1
Am
Hew2

M
fluvial?

Am M
ii 2
LST

Hew2 ii 1
M
Ss1

Argila vermelha
30,0m Vermelho
Fm2
Argila
Ps17e

Ss1

Hew2
~ (Outra cor)

M Quasi planar
Ss1/Sc

Argilito vermelho (Fm2)

símbolo

M ii 1

Argila vermelha
SB-E
M Sgb congl (verm, branco, arg)

Ssg

40,0m
M Ss1
Lithology
Foreshore/Delta/Fluvial

Ssg

Sandstones
Quasi planar
Am ii 1 Heterolithe with microhummocky,
Ps16d
FSST

Sp
M Red mud
wavy or rarely linsen structures,
Sp
ii 1
Ssg quasi planar Bioturbated Shale - laminated
sigmoidal
ii 1
Sc
Pebbly sandstone
BSFR Sc ii 1

Shcs/Ip Contacts between bedding plans


ii
Gradational
Fb ii 3-4 (h)
Smooth
Irregular

50,0m Sedimentary structures and related features


He silty/Fm
Cross bedded Ripples
He silty/Fm (Sk, Pa, Trans)
Hummocky Swaley
Sh ii3
(4%, P2O5)
~ Wavy
He silty
Flat parallel lamination
ii 3 (Sk, Pa)
Trough Cross Bedded

Phosphatic Ironstone
Early Giventian

Ps15d

M Massive

Intraclasts
HST

Normal grading

Bioturbation
ii - Bioturbation index

FS Clasts
He silty
60,0m Pyrite

linsen
166
ST-08-PI RAIOS
0 GAMA 100
0,0m
CPS

Shcs
Shoreface
Lower to Middle

Shcs
xxxxxxxx
xxxxxxxx
xxxxxxxx
xxxxxxxx
xxxxxxxx
xxxxxxxx
xxxxxxxx Ip/Shcs
xxxxxxxx
xxxxxxxx

HST
Weathering

Fm

Ps8c
Sh

Fm
ii 1-2

10,0m ii 2-3 (und)

Hew (b)

MFS-C ii 4-5 (Pa, h)


xxxx

Fm
Early Givetian

ii 1-2
Offshore

Fm
ii 1-2

20,0m
A sequence (Pimenteira Formation)

ii 2-3 (Pa, Ru)


Lithology
Fb

xxxxxx Sandstones

Fm
Heterolithe with microhummocky,
wavy or rarely linsen structures,
TST

Fm
Ps7c

ii 1-2 (Pa)

Bioturbated Shale - laminated


Lower shoreface/offshore transition zone

Pebbly sandstone
Hew (b)

ii 4-5(Pa,Sk) Contacts between bedding plans


xxxxxx Gradational
Fm Smooth
ii2
MRS-C Irregular
Hew (b)
ii2
Late Eifelian?

Sedimentary structures and related features


xxxxxx
xxxxxx
xxxxxx Cross bedded Ripples
Ps6c
LST

31,0m
Hummocky Swaley
xxxxxx
xxxxxx He sty (b, w)/
Ip (b) Wavy
transition zone
Offshore
Lower shoreface/

xxxxxx
Flat parallel lamination
xxxxxx
ii 3-4 (Pa, Sk)
Trough Cross Bedded
xxxxxx
xxxxxx
xxxxxxxxxxxx
xxxxxxxxxxxx Phosphatic Ironstone
xxxxxxxxxxxx
xxxxxxxxxxxx
xxxxxxxxxxxx Ip
xxxxxxxxxxxx
xxxxxxxxxxxx M Massive
SB-C xxxxxxxxxxxx
xxxxxxxxxxxx
Am ii 1 Ss1
Sp Intraclasts
Ss1 (conglomeratic)
mud films
Floodplain
Channel/
Fluvial

Normal grading
B sequence (Itaim Formation)

Stc

ii 1 Ss1
Af
M Ss1 Bioturbation
White mud ii - Bioturbation index
Ps5b

intraclasts
HST

Fm2 (white mud)


M Ss1 Clasts
Delta front

Sp
Pyrite
M Ss1
40,00m linsen
C
S
VFS
FS
MS
CS
167

ST-04-PI RAIOS
0 GAMA 100
0,0m
CPS

ii 4
Sb
Lower-Middle Shoreface

Shcs

Sp
Givetian?

10,0m

Shcs
transition zone
Offshore/Lower shoreface

Ps8c

Fm/Fb
ii 1-3 [org]
(Pa)

Sh
HST

Sh

Fm ii 1-2

Hew (b) ii 4 (Pa) [org]


C sequence (Pimenteira Formation)

MFS

xxxx
ii 4-5 (Pa) [org]
Fb

20,0m
Offshore

ii 2-3
Fm

Fb
ii 4-5 (Pa, v)
TST

Fm

ii 2-3
ii 4-5 (Pa, Spy) Fb
ii 2-3 Fm

Fb ii 4-5 (Pa)
ii 2-3 Fm

Hew(b)
ii 3-4 (fossilífero?)
Conchas lamelibranquios
transition zone
Offshore/Lower shoreface

ii 2-3 (Pa?)
Fb/Fm
30,0m

Lithology
xxxxxx

Sandstones
Hew(b)
ii 4-5 Heterolithe with microhummocky,
wavy or rarely linsen structures,

Bioturbated Shale - laminated


Ps7c

xxxxxx
He silty (b)
Late Eifelian

xxxxxx
Pebbly sandstone
xxxxxx
He silty/Sb [org]
xxxxxx ii3-4
xxxxxx Contacts between bedding plans
Gradational
xxxxxx Smooth
Irregular

Hew(b) ii3-4
MRS-C ii 1-2
40,0m
Fm Sedimentary structures and related features

xxxxxxxx [org] Cross bedded


LST
Shoreface

xxxxxx
Ripples
Lower

Ps6c

xxxxxx
xxxxxx
Sb [org] /Shcs Hummocky Swaley

Wavy
B Sequence (Itaim Formation)

SB-C Sc (?)
(?)
Flat parallel lamination
Fracture zone

Trough Cross Bedded


Praguian-Emsian?
Proximal delta
Foreshore/

Phosphatic Ironstone
Ps5b

Ss1
HST

M
M Massive

Intraclasts
Sr/Sb ii3

Normal grading

Bioturbation
ii - Bioturbation index
Clasts
Pyrite
50,00m
C
S
VFS
FS
MS
CS

linsen
168

SM-ST-04-PI RAIOS
0 GAMA 100
0,0m
CPS

Ss2
shoreface?
Lower to Middle
Early Giventian

Ss2

Ss2

Ss2

Fm
HST

Shcs/Sr
Ps8c

Fm
shoreface
Lower to Middle

drill core it was not possible to better

Ss2
Shcs
Sm
The absence of gamma ray and

M
10,0m

Shcs
define the MFS

?
MFS-C?
Offshore

Fm

iii 3
Missing data
TST
Lower shoreface/Offshore transition zone

Ps7c

xxxxx
xxxxx He (b)
xxxxx
iii-4 (Pa)
20,0m
Sborg F(b)ii
Hew (b) ii3
Sborg (Sk, Pa)
xxxxxx

Hew (b)

II 4-5
(Pa, Sk,und)

Hew (b) ii 4-5

xxxxx ii 4-5 Hewb (Pa, Sk)


Sborg ii2
C sequence (Pimenteira Fm.)

xxxxx ii Fm
Hewb ii4-5 (Pa, Sk)

F(b) ii 3-4
xxxx
Offshore

F(b)
ii 3-4
Sb [org] ii 2
Late Eifelian

ii Fm

Heb
ii 1 ii 2-3

MRS-C Hew (b) Lithology


30,0m ii 1-2
Sandstones
Hew(b)
Heterolithe with microhummocky,
ii 2-3
wavy or rarely linsen structures,
xxxxx
offshore transition zone
Lower shoreface/

Sborg/He sty/
xxxxx Ip Bioturbated Shale - laminated

ii 3-4
Ps6c

Pebbly sandstone
LST

xxxxxxxx
Sborg/He sty

ii 3
Contacts between bedding plans
Ip
Gradational
Smooth
SB-C Irregular

Sp

ii 1
Sedimentary structures and related features

M Cross bedded
40,0m
Hummocky
Ss1
Foreshore/delta

Wavy
B sequence (Itaim Formation)

Flat parallel lamination


Praguian/Emsian

Ps5b

Trough Cross Bedded


HST

Ss1
Phosphatic Ironstone

ii 1 M Massive
?
Intraclasts

M Normal grading

Bioturbation
ii - Bioturbation index
Intraclasts (organic phosphate?)
Pyrite
linsen

50,00m
C
S
VFS
MS
CS
G
RAIOS
3-PM-21-PI 0 GAMA 100
0,0m 169
Shoreface transition zone CPS
Offshore/Lower

He (w)

10,0m
Offshore

Ps10c

Fm ii 1-2

FS
ii 3 (As)
Shoreface transition zone

Sf
Offshore/Lower

ii 1-2
Ps9c

20,0m
Offshore

Fm

ii 1-2
TST

FS
He,w (org)ii 4-5 (Pa/Sk)

Sb ii 4-5
ii 4-5

Sb Op?/Ar? (Nearshore trace fossil


/Sklolithos ichnofacies)

Sp

30,0m
Sb ii 4-5
Middle/Upper Shoreface

(SK,Ar)

Ferrous Ip M

Sb ii 5-6
Giventian?

Ferrous Ip M

M
Sp/Sm
HST

ii 1

Sc
Ps8c

ii -
Offshore/Lower shoreface transition zone

Sh

ii 2-3 (Pa,h)
40,0m
He (b, w) ii 3

Fb ii 2-3 (Pa)

He (b, w)

He (b, w) ii 4-5

MFS-C
Fm/(Fb)

(Ph;h)
ii 2-3

50,0m

Fm/(Fb)
ii 2-3 (Ph;h)
Ps7c

ii 2-3 He (silty)

Fm ii 2

He (silty)
TST

Fb Lithology
ii 3-4
Sandstones
Fm/Fb ii 2
Heterolithe with microhummocky,
wavy or rarely linsen structures.
He (silty)

ii 2-3 Bioturbated Shale - laminated

60,0m Pebbly sandstone

He (silty)
ii 3
He (silty)/Ip
Contacts between bedding plans
MRS-C
Gradational
He (silty) ii 5
Smooth
Shcs
Irregular

He (silty)/Ip
Sedimentary structures and related features
b, w
Late Eifelian?

Ip (h,v)ii 4
Cross bedded Ripples

Ip M ii 2-3 Hummocky Swaley


LST

Ps6c

Wavy
M ii 3 Sb(org)/Ip
Lower shoreface

Flat parallel lamination


M
Sborg/Ip ii 3 Trough Cross Bedded
M Ip ii 3-4
Phosphatic Ironstone
Sb/Ip M Massive
70,0m ii 2-3
Intraclasts
SB-C
Shcs
Normal grading
plan?
Delta

ii 4 Fm2
Shcs (glauconitic?)
ii 1 Ss1 (glauconitic?)
Bioturbation
Praguian/Emsian?

HST

Ps5b

M
ii - Bioturbation index
Delta front

Ss1 Clasts
Pyrite
ii 1

linsen
75,0m
C
S
VFS
FS
MS
CS
170

RAIOS
1-PM-11-PI 0 GAMA 100
0,0m CPS

xxxxxxxxxxx

Shcs

Ps10c
HST
Offshore transition Zone
Lower Shoreface/

Heb(?)/Sr(?)
Inferred Interval

FS
Shcs
Sb
Lower Shoreface/Middle Shoreface

Givetian

10,0m
Shcs or Sscs?

ii 1-2

Black mud clasts

xxxxxxxx Sb (muddy)/Ip (ii 3-4)


xxxxxxxx
Sb (ii 3-4)
xxxxxxxxx Shcs
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx Sb/Ip
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx ii 4-5 (As, others und..)
xxxxxxxxxx

He sty /Shcs
Offshore transition Zone
Lower Shoreface/

Ip/He sty ii-3


~ He sty ii-3
Ps9c

xxxxx Ip/He sty/ii 3


xxxxx
Fossil?

~
HST

Silicified

ii 3 (v,h,t)

20,0m
xxxxx
xxxxx Ip/He sty ii 3 (v,h,t)

~
Offshore

Fm ii 1-2
FS

Sb (muddy)
ii 5 (v,h,t)

xxxxxxx
He sty/Shcs
xxxxxxxxx ii 3 (v)
xxxxxxxxx
xxxxxxx
He sty (ii 3)
(v, h)

xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxxx

xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx Sb/Shcs
30,0m xxxxxxxxxx

xxxxxxxxxx
Middle/Lower Shoreface

Sb/Shcs/Ip (muddy)

ii-1

ii5 (undist. diversified.)

He (w,b)/Sr

ii 4-5 (Pa>>>Sk)

Shcs
Ps8c

Shcs
Hew(b)
Shcs
~ ii 4-5 (Pa, Sk,...)

M ii-2 (Fos)

Sb/Ip
xxxx
ii 4-5 (Ar, Pa and others)
HST

xxxx
Fm
xxxx

40,0m
xxxxx
xxxxx
xxxxx Lithology
xxxxx
MFS-C
xxx Fm M ii
Sandstones

Heterolithe with microhummocky,


Offshore

Fb ii 2-3 (Pa, v) wavy or rarely linsen structures,

Fb ii 2-3 (Pa, v) Bioturbated Shale - laminated

Pebbly sandstone
Fm ii 2 (Pa, h)
Offshore transition Zone
Lower Shoreface/

Contacts between bedding plans


Gradational
Smooth
ii 3-9 (Sr, Pa)
Irregular

He (b,w)
ii 3-4 (Sk, Pa)
Sedimentary structures and related features

Cross bedded
50,0m
Ps7c

Hew (b) Hummocky


TST

Wavy
ii 3-4 (Sk, Pa)
Flat parallel lamination

Trough Cross Bedded

Phosphatic Ironstone

M Massive
Late Eifelian

Intraclasts
Offshore

ii 2 Fm Normal grading
(h, v)

Bioturbation
ii - Bioturbation index
1-weak
2-medium
ii 2 Fm (h)
3-strong

60,0m
C
S
VFS
FS
MS
CS
171

RAIOS
PM-10-PI 0 GAMA 100
0,0m CPS

Fm

Sb/Sr/Hewb

TST
Ps12d
ii 4-5
xxxxxxxx

Sr/Sb
ii 4-5

ii 4-5 Fb
FS
xxxxxxx
Ip
Sr/Hewb
ii 4-5

10,0m

xxxxxx
xxxxxx
TST
Inferred interval

xxxxxx
xxxxxx 2%P2O5
xxxxxx

xxxxxx

xxxxxx
20,0m xxxxxx
xxxxxx
xxxxxx

xxxxxx

xxxxxx

xxxxxx

Sb/Ss2

He silty

ii3-4

30,0m

Sb
Ps11d
Lower to Middle Shoreface

Shcs

ii 1-2

40,0m
He silty
Pimenteiras Formation

Sh
SB-D
ii 3-5

Sr ii 3

xxxxxx Heb silty


xxxxxx ii 1-2

xxxxxx
xxxxxx
ii 2-3
Silty (org)

Heb silty

He silty

ii 2-3

50,0m
ii 5 (Sk, Pa)

He (w,b)

ii 5
Lower shoreface/offshore transition zone (transgressive)

He (b)
ii 5 (Pa, Sk)
Giventian

xxxxxx

He (w,b)
xxxxxx ii 2-3 (h)

60,0m
ii 2-3 (h)
Ps10c

He (w, b)

ii 1-2

Sr
Givetian

Inferred
interval
Hewb/Sr

70,0m
FS
I

ii 3
Lower shoreface /Ofsshore transition zone
Pimenteira Formation

Sr/Hewb
(regressive)

HST
Ps9c

Hew
Lithology

80,0m Sandstones
Offshore

Heterolithe with microhummocky,


FS Fm wavy or rarely linsen structures,

Bioturbated Shale - laminated

Laminated shale
Lower shoreface

Sb

ii 4-5
Contacts between bedding plans
Gradational
Smooth
Irregular
Interval
Inferred

Ss2
Sedimentary structures and related features
N.D.

Cross bedded Ripples


90,0m
Hummocky
Ps8c

Wavy
shoreface-offshore
Transition zone lower

Hb ii 2 PS
Flat parallel lamination
Sh Trough Cross Bedded

Sb ii 3-4 WS Phosphatic Ironstone


XXXXXX
M Massive
xxxxxx [org] PS
XXXXXX
Intraclasts
xxxxxxx
SIM-C Sh Dark gray mud clasts
Normal grading
Offshore

Late Eifelian
TST

Fm Bioturbation
Ps7c

[org]
ii - Bioturbation index
Clasts
Pyrite

linsen

100,0m
C
S
VFS
FS
MS
CS
RAIOS
1-PM-09-Pi 0 GAMA 100
0,0m CPS 172

10,0m
Lower Shoreface/Offshore
transition Zone

Fm ii 1-2
Ps10c

He(w)
HST

ii 1-2

Fs
Sr
Offshore transition zone
Lower Shoreface/

20,0m He(w)

ii 1-2
Offshore

Fm
Ps9c

Fm
ii 1-2

Fm ii 1-2
Offshore transition
Lower Shoreface/
zone

Sb ii 4-5
He(w,b) (Pa?)

Fs Sb ii 4-5 (diverssos)
He(w,b) ii 3-4 (Spy, Pa)

Sb ii 4-5 (diverssos)

Shcs
30,0m

Sb ii 3-5 (diverssos)
Lower to Middle Shoreface

Giventian

Sorg/Ip ii 3-4
Sb (As)
Sb ii 3-4

Sb ii 3-4(?)
(Muddy)

Ssw/Sm

Shcs/Ip

Shcs
M
Ss2

Lag?
M Ss2
40,0m Sb/Ss2

Shcs

He sty
Offshore transition zone
Lower Shoreface/

ii 3
SE

He sty

(Selicif)
HST
Ps8c

Fb ii 3 (Pa)
He(w,b)

Fm
MFS-C ii 1-2

He(w,b) ii 3-4
(Pa, Spy)

50,0m
Lithology
Offshore

Fm

ii 1-2 (Pa)
Sandstones

Heterolithe with microhummocky


ii 3 (Pa, Sk) wavy or rarely linsen structures
He(w,b)
TST

Fm ii 1-2 (Pa) Bioturbated Shale - laminated


He(w,b)

Pebbly sandstone
Fb
ii 3 (Pa)

Contacts between bedding plans


ii 3 He(w,b) (Spy, Pa, Sk)
Gradational
Fb ii 3 (Pa)
Smooth
C sequence (Pimenteira Formation)

ii 3 He(w,b) Irregular

Fb ii 3
Sedimentary structures and related features
60,0m

Cross bedded Ripples


He(w,b)
Hummocky Swaley
ii 4 (Pa, Sk, As?)
Ps7c

Fm ii 1-2 (Pa) Wavy


Hewb/SrIi4
Late Eifelian

(Pa, Sk) Flat parallel lamination

Trough Cross Bedded

Phosphatic Ironstone
ii 2 Fm/Ip

M Massive
Sr ii 4 Pa?
Shcs Intraclasts
He(w,b)/Ip
ii 4 (Pa?)
Normal grading
Lower shoreface

Ps6c
LST

Sb/Ip Bioturbation
ii 4-5
ii - Bioturbation index
Clasts
Shcs
Pyrite
70,0m Sb ii 4-5
Shcs/Ip linsen
He(w,b)/Ip ii 3-4
Shcs
?
C
S
VFS
FS
MS
CS
RAIOS 173
PM-07-PI 0 GAMA 100
0,0m
CPS
Inferred interval

Ss2
Givetian?

Sb

10,0m
Shcs
Lower to middle shoreface

xxxxxxxx 2%P2O5
Sb/Ip ii 4-5 (ind)

Shcs

xxxxxxx
HST

xxxxxxx
Ps8c
transition zone
Lower shoreface-offshore

20,0m
Pimenteiras Formation

xxxx

MTS or
Offshore

MFS

ii
Fb/
Fm

xxxx
Hew (b) ii 4-5
TST

Fb ii 3-4
Ps7c

Hew (b) ii 4-5


30,0m
Fb ii 3-4

ii 4-5 (Pa)
Hew (b)
Fb ii 3-4 (Pa)
xxxx

Hew (b) ii 4-5


Offshore/Lower shoreface
transition zone

Fb ii 3-4

Hew (b) ii 5-6

xxxxxx
xxxxxx 6%P2O5
Fb ii 3-4 (ind)

ii 5-6
Lithology
Sandstones
ii 5-6 (Pa,S15, v,h)
Heterolithe with microhummocky,
wavy or rarely linsen structures,
xx Shcs
xx black
phosphate
xxx Sb/Ixx 1 a 2%P2O5
40,0m Bioturbated Shale - laminated
xxxxx
Inferred Interval

xxxxxx
xxxxxx
Pebbly sandstone
transition zone
Lower shoreface-offshore

xxxxxx 1 a 2%P2O5
xxxxxx
LST

Ss/Sb?
Contacts between bedding plans
xxxxxx
xxxxxx Gradational
Ps6c

Smooth
xxxxxx
Late Eifelian

xxxxxx Irregular
xxxxxx
xxxxxxx Ss?/Sb?
xxxxx
Ip
xxxxxxxxx
Sedimentary structures and related features
xx Sb(fl)xx ii 4-5
SB-C xxxxxxxxx (v, h) Cross bedded Ripples
Scg M
Hummocky Swaley
M
Fluvial? Hew2 (Delta related)
Wavy

Flat parallel lamination


Ps5b
HST

50,0m
Trough Cross Bedded

Phosphatic Ironstone
Delta front

Ss1 M Massive
Itaim Formation

Intraclasts
M
Normal grading
Praguian/Emsian?

Bioturbation
ii - Bioturbation index
Clasts
Pyrite

linsen

58,00m
C
S
VFS
MS
CS
G
174

PM-06-PI RAIOS
0 GAMMA 100
0,0m
CPS
Inferred interval

Ps9c
ii 4 - 3 (h,7)
Sb

Shcs

ii 4 - 3 (h,t)

Sb

ii 4

xxxxxxx
x
transition zone
Lower-shoreface to Lower-shoreface/offshore

Sb ii 3
xxxxxxx
Sb
xxxxxx ii 3
HST

Sp

Sb
ii 4
10,0m

Sp

Sb/Ip
xxxxxxxx
ii 5-6 [org]
Sb/Ip
ii 5-6

Sp
Ps8c

Shcs
Sp
Shcs
Sp
Shcs

20,0m

MFS-C
C sequence (Pimenteira Formation)

Offshore

ii 1-2

Fm
30,0m

Hew(b)
ii 2-3
TST
Giventian

ii 1-2 (h)
shoreface transition zone

Ps7c
Offshore/Lower

ii 3 (Pa,h)

ii 3
ii 3
ii 3
Hew(b) Hew(b) Lithology
ii 3
MRS-C Fb Sandstones
ii 4

Heterolithe with microhummocky,


He silty wavy or rarely linsen structures,
ii 4

XXXX Bioturbated Shale - laminated


Sb ii 4 [org]
Shcs
Lower shoreface/offshore transition zone

Sp/Shcs (?) M? [org]


ii 2 (h) Pebbly sandstone
xxxxxxxxxx
40,0m xxxxxxxxxx
xxxxxxxxxx
Lower shoreface to

xxxxxxxxxx
xxxxxxxxxx
Contacts between bedding plans
xxxxxxxxxx Gradational
Ip/(Sb) ii 4 (v,h)
xxxxxxxxxx
Smooth
[org]
xxxxxxxxxx Irregular
xxxxxxxxxx
Ps6c

xxxxxxxxxx
Late Eifelian?

xxxxxxxxxx
xxxxxxxxxx
Sedimentary structures and related features
LST

xxx Sb/(Ip)
xx ii 5 - 6 (v,h)

xxx Sb(Ip)xx ii 5 - 6 (v,h) [org]


Cross bedded Ripples
xxx Sb(Ip)xx ii 5 - 6 (At, v, h) Hummocky Swaley
SB-C Sp/Sm? M?
Sb (I) Wavy
xxxxxxxxxx ii 4

Sp
Flat parallel lamination

Trough Cross Bedded


Ss1 M
Phosphatic Ironstone
xxxxxxxxxx
xxxxxxxxxx M Massive
50,0m
B sequence (Itaim

Intraclasts
Delta front?

Ps5b
Formation)

Praguian-Emsian?
HST

M Normal grading
Ss1
WS
Bioturbation
ii - Bioturbation index
Clasts
54,00m Pyrite
C
S
VFS
FS
MS
CS

linsen
175

RAIOS
1-PM-05-Pi 0 GAMA 100
0,0m
CPS
Soil- C horizon

Sr/He

ii 2
Lower shoreface-Offshore
Transition zone

ii 3 (Tr,v)

10,0m

ii 2
Ps9c
Offshore

ii 1
FS
HST

Fm ii 1-2
20,0m
He(w) ii 1-2

Sb
ii 4-5 (?)

Shcs

ii 3 (Minor sized -not connected)

Sb
Lower shoreface - Midlle shoreface

Ps8c

ii 5 (As,?)

M
30,0m Ss2

ii 2

ii 4-5 (v,Ar, h)
Sb

ii 3
Yellow nodules
Ssw
M
Sm

Fm
Shcs

ii 2 Fm

MFS-C

40,0m Fb

ii 3 (Pa?, Sk?)
Offshore

Fb

2%P2O5

Fb
Ps7c

Hew(b)
50,0m ii 3-4
Lithology
TST

Fb Sandstones
ii 3
Heterolithe with microhummocky,
ii 3-4 Hewb
wavy or rarely linsen structures,
Offshore/Lower shoreface
transition zone

(Sk, Pa)
Fb
Bioturbated Shale - laminated
Fb

Pebbly sandstone
ii 3-4 (Pa, Sk)
Giventian

Fb ii 3-4 (Pa, Sk)


Contacts between bedding plans
MRS-C ii 3-4 (Sk, Pa, v, h)
Fb Gradational
(calciferous) Smooth
ii 2-3 (?)Hew (org)
Irregular
ii 2

Sedimentary structures and related features


60,0m Sorg/Hew (org)
5% P2O5
Cross bedded Ripples
ii 3-4
Lower Shoreface

Hummocky Swaley
(Pa, hz)
Wavy
Ps6c

Sb(org)Hew (org)
Flat parallel lamination
LST

ii 3-4(h,t) Trough Cross Bedded


Sorg/Ip
(muddy) Phosphatic Ironstone
ii 2
SB-C
M Massive
Reelaborated Delta front?
Upper Shoreface/

M
Praguian/Emsian

Intraclasts

Normal grading
Ps5b
HST

Ss1 (glauconitic?)

Bioturbation
ii - Bioturbation index
Clasts
Pyrite
70,0m
linsen
C
S
VFS
MS
CS
G
176

RAIOS
PM-02-PI 0 GAMA 100
0,0m CPS
Weathering - Caulinization

Acg

Recent sediments
Acg?

Acg?

10,0m
Ps9c

Ss2/Sb

Shcs

20,0m
Ss2/Sb
ii 2-3
Lower to Middle Shoreface

Sb ii 2
Ss2

Shcs

Ss2
HST

Sb ii 4-5 [org]
(Ar)
Shcs

Ss2

30,0m
Shcs ?
ii 1
Ps8c
HST

He (silt)
b, w

He (silt)
b, w
Offshore

ii 1

Fm

40,0m
TST

ii 2 (Pa) - Fm/He silty

ii 3-4 (Pa, Sk)

ii 2-3
transition zone
Offshore/Lower shoreface

Hew (b)

iilt co
ii 3 - 4
(Sk, Pa)

F(b) ii 2-3

F(b) ii 2-3
(Pa, Sk)
xxxxx
Ps7c

Lithology
50,0m ii 2 Fm
He silty Ironstone/phosphatic ironstone
ii 4-5
xxx
Sandstones
I ii 2 Fm
Heterolithe with microhummocky,
ii 2-3 He silty wavy or rarely linsen structures
xxxx (h)
ii 2-3 He silty
Bioturbated Shale - laminated
He silty
ii 2 (ind)
xxxx
xxxx
xxxx
xxxx Ip M Pebbly sandstone
Late Eifelian?

xxxx
xxxx
S[org]/Sb/Sh
Lag?

xxxx M
xxxx
xxxx
xxxxxxx
Ip M ii 4-5 Contacts between bedding plans
xxxxxxxx
LST

xxxxxxxx Ip/Sh
xxxxxxxx
Lower Shoreface

xxxxxxxx M Gradational
xxxxxxxx Smooth
Sb [org]/(Sb)/Ip

xxxxxxxx ii 3-4 Irregular


Ps6c

xxxxx
xxxxx Ip M
xxxxxxxxx Sb [org]
ii 3-4 Sedimentary structures and related features
xxxxx M Ip
60,0m ii 4-5
xxxxxxxxx S (b) org
Cross bedded Ripples
UN-C
Ss M ii 1
xxxxxxxxx Hummocky Swaley
Ss
Middle to Upper

Ssw
Shoreface

Wavy
Praguian/Emsian?

Ss
M
HST

Ssw
Ps5b

ii 1 Flat parallel lamination


Ss M Glauconitic?
Trough Cross Bedded

Phosphatic Ironstone

M Massive

Intraclasts

Normal grading

Bioturbation
ii - Bioturbation index
Clasts
Pyrite
70,0m
linsen
C
S
VFS
MS
CS
G
177

RAIOS
PM-08-PI 0 GAMA 100
0,0m CPS

Sb ii 4
ii 5

Shcs

x x
x x x M
x x
x x x
Weathered
Mottled

M
x Sb/ x
Ss2
ii 4,5,6
x x M
x x x
x x
x x x M

Shcs
xxxxxxxxxx
Sb ii 4
Shcs
Lower to Midlle Shoreface

Sb ii 4
xxxxxxxxxx
10,0m
Early Giventian

Shcs

xx Sb/Ipxx
xxxxxxxxxx
ii 4

Shcs

Shcs
Ss2/Sb
ii 2
Shcs
C sequence (Pimenteira Formation)

Ps8c

xxxx ii 4 (Sk, Pa)


Lower Shoreface

xxxx
HST

xxxx
xxxx
xxxx
xxxx

Shcs

20,0m Sr/Sb ii 4-5 (v, h)

ii 1-2

MFS-C

Fm
Offshore

[org]
xxxx
ii (1,2) (Pa)

Lithology
Sandstones
Inferred interval

Heterolithe with microhummocky,


wavy or rarely linsen structures,

Fm/Fb/He? Bioturbated Shale - laminated


30,0m

Pebbly sandstone
Late Eifelian?

Ps7c
TST

Contacts between bedding plans


Gradational
Sr Smooth
ii 3-4 (Sk, horiz)
He (b) Irregular
xxxxxxxxx Int Sh/Sb/Ip
ii 3a1
xxxxxxxxx
xxxx Fb
xxxxxxxxx Int Sh/Sb/Ip
xxxxxxxxx ii 3a4
xxxxx Sedimentary structures and related features
MRS-C Fb ii 4

xxxxxxxxx Int Sh/Sb/Ip ii


ii 3a4 Cross bedded Ripples
x x x
xxxxxxxxx
Sb/Ip ii 4-5 (cinza) Hummocky Swaley
xxxxxxxxx (Sk - cement burrowed)
x x x - hor/vert
Sh ii 1 Wavy

Flat parallel lamination


40,0m xxxxxxxxx Trough Cross Bedded
Hew?
xxxxxxxxx
xxxxxxxxx Phosphatic Ironstone
Inferred interval

2%P2O5

Ps6c

xxxxxxxxx
Lower shoreface?

xxxxxxxxx M Massive
LST

Sb? Muddy /Ip


xxxxxxxxx Intraclasts

xxxxxxxx Normal grading


xxxxxxxx
xxxxxxxx
xxxxxxxx
Bioturbation
ii - Bioturbation index
Praguian/Emsian?

SB-C Clasts
deltaic?
Fluvio-

HST

Pyrite
Ps5b

linsen

48,00m
G
C
S
VFS
FS
MS
CS
178

RAIOS
1-CT-02-Pi 0 GAMA 100
0,0m CPS

Soil

Fb?
FS?
Middle / Lower Shoreface

Shcs/Sb
D sequence (Pimenteira Fm.)

Ps14d
and Offshore transition zone
Lower Shoreface / Lower Shoreface

10,0m
Fb ii 3 (Pa)
HST

Hew(b)/Ip
FS

Sb ii 4

Sb ii 4
Lower Shoreface

Shcs

Shcs
He (w,b) org
ii 2-3 (?)
Ps13d

Shcs
Early Giventian

He (b,w)

20,0m Shcs/Ip

MFS-D
Offshore

Fm (org)
ii 1
Ps12d

He (w,b)/Ip
ii 4-5 (Sk, Pa, ?)
TST

30,0m
Offshore/ Lower Shoreface Transistion Zone

Sandy siltstone
Inferred - Missing data
He silty (w)
Ps11d

He sty

SB-D Sr, Ip
(conglomeratic?)

He sty (w,b)

ii 5-6 (Pa, Zoo, ?)

40,0m
He sty

ii 5-6 (Pa)
HST

Ps10c +Ps9c?
Missing data
interpreted mainly as Hesty from original data
Siltstones, sandstones (bioturbated)

Lithology
Sandstones

Heterolithe with microhummocky,


wavy or rarely linsen structures,

50,0m
Bioturbated Shale - laminated

Fm ii 1-2
Offshore

(Org) Pebbly sandstone


Early Giventian

FS He (w,b) org
ii 5-6 (Pa, Sic, ?) Contacts between bedding plans
ii 5-6 (Pa) Gradational
Offshore transition
shoreface/
Lower

HST

Smooth
Sr(b)
Irregular
Hew(org)
ii 4 (Pa, Sic)

Sedimentary structures and related features


Ps8c
Distal delta?
C sequence (Pimenteira Fm.)

M
Cross bedded
Ss1?/Ss2?
ii 1 Hummocky
Mud films Wavy

Flat parallel lamination

60,0m Trough Cross Bedded


Offshore

Fm Phosphatic Ironstone
ii 1-2
M Massive
ii 3(Sk, ?)
He w Intraclasts
ii 2
Normal grading

Bioturbation
ii - Bioturbation index
Clasts
Pyrite

linsen

70,0m
C
S
VFS
MS
CS
G
179

CT-01-PI RAIOS
0 GAMA 100
0,0m
CPS
Sb

Shcs

Ss2

Sb
Middle shoreface
Lower to

Shcs?

Sb
ii 4 (Mc?)

Shcs?

Sb
ii 3 (indist)

Shcs?
HST
Upper Shoreface

Sp/Ss2

Fósseis? (1-2)?

10,0m Sscs

Sb ii 4-5 (Sk)

M
Lower to Middle shoreface

Sp/Ss2

Ss2 M

Shcs

Ss2 M

Shcs
Pimenteiras Formation

Ss2 M

Sp/Shcs

20,0m
ii 3
Hesty
Ps8c

(Silicif)

ii 3 He (b,w)
Sp
M Ss2
ii 1Shcs
Hew (b)
ii 4 (he?, V, esp, Sk, Pa)
Transition Zone
Lower shoreface/offshore

Shcs
Shcs

MFS-C

ii 4-5
(Pa, Tr, ind)

(Silicif)

30,0m
Fb
xxxx
He silt ii 4 (Pa, Sk)
ii 3 Fb
Lithology
transition zone
Offshore/lower shoreface

Late Eifelian

xxxx
Sandstones

He sty Org Heterolithe with microhummocky,


xxxx wavy or rarely linsen structures,
Ps7c
TST

xxxx

Bioturbated Shale - laminated

Fb ii 3

Pebbly sandstone
He sty
ii 4 (Pa, Sk)
Org
Fb
ii 3 (ind.) Org
MRS-C Contacts between bedding plans
Laminação
ondulada
Gradational
xxxx Hesty Smooth
Irregular
ii 4 (Pa, Sk)
Lower shoreface/offshore
transition zone

Linsen
xxxx
Org
40,0m Sedimentary structures and related features
(Silicif?)
LST
Ps6c

Cross bedded Ripples


M ii 2
Sborg Hummocky Swaley

1,2% P2O5 Wavy

Flat parallel lamination

Trough Cross Bedded

Phosphatic Ironstone

M Massive

Intraclasts

Normal grading

Bioturbation
ii - Bioturbation index
Clasts
50,00m Pyrite
C
S
VFS
MS
CS
G

linsen
180

RAIOS
ARN-1-TO GAMA
0 50 100 150
0,0m
CPS

10,0m

20,0m
xxxxxxxxxx
xxxxxxxxxx
Lower shoreface

ii 5
Sb (silty)

(Sk, He, Pa)


Cruziana

xxxxxxx
xxxxxxx
xxxxxxx Shcs
xxxxxxx
xxxxxxx
Sb ii5 (Pa, Sk, He?)

Hewb ii-3 muddy

Sp
xxxxxxx
Shcs
xxxxxxx
Shcs
xxxxxxx
Shcs

ii 5

Sb
Cruziana
Ps9c

30,0m
HST
Lower shoreface/offshore transition zone

Sb/Hew

ii 4

Shcs

Sb/Hew ii-4
Cruziana

Sb

xxxxxx Hewb
FS
xxxxxx

xxxxxxxx

Shcs
40,0m
Lower to Middle shoreface

Shcs
xxxxxxx
Shcs

xxxxxxx

Shcs

Shcs
xxxxxxx

Shcs

xxxxxxx

Sb
Shcs
Sp

50,00m
Upper shoreface

Sb
Sklolithos ichnofacies

ii 3-4

xxxxxxxx
xxxxxxxx Sb
Lower to Middle Cruziana ichnofacies
shoreface Sb
xxxxxxxx ii-5
Upper shoreface

xxxxxxxx
xxxxxxxx
xxxxxxx
Sb
Sklolithos ichnofacies
Sb ii-3 (A1)

Sb ii-4 (As?)
Lower to Middle ?
Sb
shoreface Cruziana ichnofacies

Black mud clasts and phosphatic ironstone


clasts at the base (15%)
60,00m
Hew (b)
Lower shoreface/Offshore

ii-4

ii-4 (Pa)
transition zone

Hew (b)

ii 2 Hewb
Pa
Hewb
ii-3 (Pa)

ii-4 Hewb
Ps8c

ii-3 Hewb

ii-4 (Pa, ?)

ii-3 Heb

ii-3 Hewb

ii-3 Hew(b) (Pa, Sk, ?)


70,00m
ii-2

Hew(b)

ii-3-4 (Pa, ?)

Hewb

ii-3-4 (Pa, ?)
MFS-C
F (b)
Offshore

ii-3-4 (Pa)
I
Lower shoreface/Offshore

80,00m

ii-5

(Pa, Zoo, ind)


transition zone

xxxxxxxxxxxxx Sb

Hew (b) (org)

ii-3

xxxxx
HST

xxxxxxxxxx
xxxxx
xxxxx M Hew(b)/Ip Sb ii-6 (Pa, Sk)
Ps7c

Hew(b) bituminous muddy


Offshore

ii-3 Fb

Hew(b) bituminous muddy


Givetian (Trindade et al., 2015)
C sequence (Pimenteira Fm.)

90,00m
ii-2-3 (Pa)

xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
Sb
xxxxxx
Lower shoreface/Offshore

xxxxxx Sb
Sb
xxxxxx
xxxxxx

xxxxxx
transition zone

xxxxxx

xxxxxxx ii-2-3 (Pa, ?)

xxxxxxx

xxxxxxx
xxxxxx

xxxx
xxxx
Offshore xxxx
xxxx
xxxx
MRS-C xxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx
Sb/Sr/Hew
ii-5
100,00m
(Pa, Ru?)
Sb/Sr
Ps6c
LST

xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx Ip/Sb
xxxxxxxxxxxxx
xxxxxxxxxxxxx
xxxxxxxxxxxxx
SB-C xxxxxxxxxxxxx
xxxxxxxxxxxxx M Ip

ii-3 (Ru?)
seixo < 1 cm
Middle-Upper shoreface/

Sb
[org]
Praguiam - Emeian

ii 5-6 (h,v, Sk, Pa)


Delta front

Sc
Sr
Sc
Praguian - Emsian
B sequence

Sc?

110,00m
Sr/Sb ii-3

xxxxxxxxxxxx Sr/Sb ii2-3


xxxxxxxxxxxx
Sc

xxxxxxxxxxx
Sb/Sr ii2-3
xxxxxxxxxxx

Sc

Sb

ii 6-5

xxxxxxxxxx Sc
xxxxxxx
xxxxxxxx
xxxxxxxxx Sb
HST

Sb

Sb/Sr ii-3
Ps5b

Sr/Sb

120,00m ii-3

ii-5 S(b)

Sb
Sr
xxxxxxxxx
xxxxxxxxxxxxxxxx Sc
Lower to Middle shoreface

xxxxxxxx Sb ii 3-4
Sr/Shcs/Sb

Sr/Shcs/Sb

ii4 (Pa, ind)

Sb

ii-4

130,00m

Sb ii-6 (ind)
[org]

xxxxxxxxxxxx
xxxxxxxxxxxx Ss2/Ip
xxxxxxxxxxxx
Lower to Middle shoreface

Sb
ii-6 (Pa, Sk, He?)

xxxxxxxxx
xxxxxxxxx

Sb/Ip

xxxxxxxx ii 5-6 (Pa, Sk, He?)

Sb
(Pa, Sk, He?)

xxxxxxxxx
xxxxxxxxx
xxxxxxxxx
140,00m Sb/Ip
Ps4b

ii 3-6
xxxxxxxxx
xxxxxxxxx
xxxxxxxxx
S(b) (Pa, Sk, He)

ii 3-6
xxxxxxxx
xxxxxxxx Sr/Sb/Ip
[org]
xxxxxxxx Sb/Ip
xxxxxxxx
xxxxxxxx ii 4-5
xxxxxxxx
xxxxxxxx Sr/Sb/Ip
xxxxxxxx
Sb
xxxxxxxx Sb/Ip
ii 3-4
xxxxxxxx
xxxxxxxx Sr(b)/Ip

Hew (b)
x ii 3
Lower shoreface-offshore transition zone

ii-5

xxxxx Ip/Hew(b) ii-3

xx x Hewb

ii-3
x xx
FS
150,00m
xxxxxxxxxx
xxxxxxxxxx
xxxxxxxxxx Sb/Ip
xxxxxxxxxx

Hew (b)(Sk, Pa)


ii-3
xxxxxxxxxx
xxxxxxxxxxx
Shcs/Ip concrecionary
xxxxxxxxxxx
xxxxxxxxxxx
xxxxxxxxxxx

ii-2
Muddy.
Hesty

Sb ii-5
xxxxxx
ii-4
xxxxxx
Shcs
xxxxxxx
ii-5 (Pa, He, ?)
Hew(b) (Pa, ?)
ii-4
Ps3b

Hew (b) ii-4

Hew (b)
ii-4

160,00m Hew (b)


HST

ii 3-4 (Pa, He, ?)

Lithology
x
ii 2 (Sk, Pa)
x Sandstones
Offshore

xx

Heterolithe with microhummocky,


wavy or rarely linsen structures,

MFS-C ii 2-3 (Sk, He?, Te?) Bioturbated Shale - laminated


F(b)
Ps2b
Offshore/Lower shoreface

Pebbly sandstone
event beds
TST

Contacts between bedding plans


transition zone

event beds Gradational


Ps1b

Smooth
x x
Irregular
xxx
170,00m
event beds
SB-B+
TSME-B
He (w,b)
ii-3 (Pa, He)
Sedimentary structures and related features
Crevasse splay

Ss1
Sp
Cross bedded Ripples
Flood plain

Ss1
Hummocky
Fluvial

Stc

Ss1 Wavy
Sp
M Ss1 Flat parallel lamination
Sgb Trough Cross Bedded
Channel

Ss1 Phosphatic Ironstone


Fluvial

HST?
Jaicós Formation

M Massive
M
Sgb
Intraclasts
Sc
Normal grading
Channel
Fluvial

Stc
Bioturbation
ii - Bioturbation index
180,00m Clasts
Stc
Pyrite
Stc
linsen
C
S
VFS
FS
MS
CS
VCS
181

ST-53-PI RAIOS
0 GAMA 100
0,0m
CPS

Ss1

Deep weathered
Delta front/Delta plan?

Ss1

5,0m

Sflaser
Tidal?

Sf/Hef

10,0m

Ss1
Pimenteira Formation (Passagem Member) or Cabeças Formation?

Ss1

Ss1

Ss1

Ss1
(muddy)
Ss1
15,0m

Ss1
Giventian or Frasnian?

(muddy)
Delta front

ii 2

20,0m
Ss1

ii 1
LST
Delta front

Ps17e

25,0m

Ss1
LST

Sp/Ssg

SB-E Stc Lithology


Sandstones
30,0m
Hesty Heterolithe with microhummocky,
(b?) wavy or rarely linsen structures,

Bioturbated Shale - laminated


Shcs
Lower soreface/offshore transition zone

Hesty (b)
Pebbly sandstone
ii 3-4 (Sk, Pa)
F(b) ii 3 (Pa, h)
Contacts between bedding plans
35,0m
Gradational
HST

Smooth
Hew (b) Irregular
Ps15d

ii 3-4 (Sk, Pa, he)


Giventian

Sedimentary structures and related features

FS Cross bedded
40,0m
Hummocky
D Sequence (Pimenteira Formation)

Shcs/Ip
xxxxxxxx
xxxxxxxx
Sb/Ip ii 4-5 Wavy
xxxxxxxx
Shcs ii
Flat parallel lamination
xxxxxxxx
xxxxxxxx
Offshore transition zone
Lower shoreface - Lower shoreface/

xxxxxxxx
Sb/Ip ii 4-5 (ind) Trough Cross Bedded
Shcs
Phosphatic Ironstone
xxxxxxxx
ii 3-4
xxxxxxxx M Massive
xxxxxxxx
Sb/Ip (ind)
xxxxxxxx
HST
Ps14d

xxxxxxxx He (b)
xxxxxxxx Intraclasts
xxxxxxxx
Sb
xxxxxxxx Normal grading
xxxxxxx
xxxxxxxx
xxxxxxxx Sb/(Ip)
Early Giventian?

xxxxxxxx
ii 3-4
Bioturbation
xxxxxxxx ii - Bioturbation index
Shcs ii 1
xxxxxxxx
Clasts
Pyrite
Offshore

ii 1-2

Fm [org] linsen

50,00m
C
S
VFS
MS
CS
G
182

3-PM-14-PI RAIOS
0 GAMA 100
0,0m
CPS
Interval without
register

He sty
I?
ii 1-2

M
10,0m

Ss2
Lower to Middle shoreface

Giventian?

Shcs

xxxxx

Shcs

He sty ii 2-3
C Sequence (Pimenteira Formation)

HST

Sh

Hew (b) ii 2
Shcs
Ps8c

He sty
ii 2-3
Offshore/Lower shoreface transition zone

Fm
ii 1-2 (Pa)
20,0m

He sty (b)

ii 3-4

Hew (b) ii 3-4 (Sk)

ii 2-3 Fb

M Ip

Hew (b)
ii 3-4 (Sk, and)
M Ip
xxxx F(b) ii 3-4 (Pa)
xxxx
MFS-C

Fm
Late Eifelian?
Offshore

ii 2 (Spy)

xxxx
xxxx
TST
Ps7c

30,0m
Interval without

Lithology
register

Sandstones

Heterolithe with microhummocky,


wavy or rarely linsen structures,

Hew ( b) Bioturbated Shale - laminated


ii 3-4 (Sk, h)

Pebbly sandstone

Contacts between bedding plans


He sty/Ip Gradational
Smooth
ii 1-3
Irregular

Shcs

xxxxxxxx
Shcs Sedimentary structures and related features
40,0m xxxxxxxxx
xxxxxxxxx ii 2-3
xxxxxxxxx Sborg M/Ip Cross bedded Ripples
zone?
Lower shoreface-Offshore Transition

He silty ii 2
LST
Ps6c

xxxxxxxxx
xxxxxxxxx Sborg M/Ip Hummocky
ii 2-3
Wavy
Shcs
ii 1 Flat parallel lamination

Sborg
Trough Cross Bedded
xxxxxxx
He sty/Ip Phosphatic Ironstone
xxxxxxx
Sh
xxxxxx He sty/Ip M Massive
xxxxxx ii 3 (ind)
Shcs
Intraclasts
xxxxxxxx Sborg
xxxxxx
xxxxxx M Ip
xxxxxx
Normal grading

Bioturbation
ii - Bioturbation index
Clasts
Pyrite
50,00m
linsen
C
S
VFS
FS
MS
CS
183

APÊNDICE D – RESULTADOS DAS ANÁLISES POR


DIFRATOMETRIA DE RAIOS X
184
185
186
187
188
189
190
191
192
193
194
195
196
197
198
199
200
201
202
203
204
205
206
207
208
209
210
211
212
213

APÊNDICE E – ESTUDOS POR MICROSSONDA –


COMPOSIÇÃO DAS SIDERITAS

Sideritas autigênicas formadas num estágio eodiagenético pode nos fornecer informações sobre a

geoquímica das águas inter-poros, uma vez que não sofrem facilmente reequilíbrios e processos mais

tardios de recristalização (Mozley, 1989). Em geral as feições petrográficas foram pobremente

distinguidas pela composição química das sideritas. As composições estão representadas nos diagramas e

tabela abaixo. Em geral as sideritas tipo 3 são mais puras (de: 87-99 (média 91) mol% FeCO3; 0-11

(média 6) mol% MgCO3; 0-9 (média 2) mol% CaCO3; 0-2 (média 1) mol% MnCO3). As mais precoces

são mais magnesianas – tipo 2 (80-90 (média 85) mol% FeCO3; 8-17 (média 12) mol% MgCO3; 0-10

(media 2) mol% CaCO3; 0-4 (média 1) mol% MnCO3; apresentando uma presença mais susbtancial de Ca

e Mg. Em ambos os casos o Mn é muito baixo, menor que 1.

Mozley, P.S., 1989. Relation between depositional environment and the elemental composition of

early diagenetic siderite. Geology, 17, 704-706


214

Amostra analisada No. MgO MnO CaO FeO


MA-T-115_Campo 1 _Pt 1 1 5,11 0,78 0,44 52,81
MA-T-115_Campo 1 _Pt 2 3 4,75 0,73 2,04 48,56
MA-T-115_Campo 1 _Pt 3 4 5,12 0,80 0,54 49,98
MA-T-115_Campo 1 _Pt 7 10 2,28 1,25 0,33 57,05
MA-T-115_Campo 1 _Pt 8 11 0,02 0,07 0,18 57,98
MA-T-115_Campo 2 _Pt 1 1 4,38 1,21 0,37 59,41
MA-T-115_Campo 2 _Pt 2 3 3,58 1,08 0,43 58,52
MA-T-115_Campo 2 _Pt 3 4 4,55 0,75 0,14 50,13
MA-T-115_Campo 2 _Pt 4 5 5,10 0,90 0,54 54,96
MA-T-115_Campo 2 _Pt 5 6 5,54 0,80 0,31 50,66
MA-T-115_Campo 2 _Pt 6 7 5,29 0,78 0,70 50,25
MA-T-146_Campo 1 _Pt 1 1 5,83 0,92 0,80 48,03
MA-T-146_Campo 1 _Pt 4 4 4,01 1,07 0,33 56,41
MA-T-146_Campo 1 _Pt 5 8 5,86 1,02 0,66 52,99
MA-T-146_Campo 1 _Pt 6 10 4,39 1,31 0,61 58,33
MA-T-146_Campo 1 _Pt 7 11 0,00 0,02 0,22 59,48
MA-T-146_Campo 1 _Pt 8 12 1,42 0,64 1,43 55,96
MA-T-146_Campo 1 _Pt 9 13 0,80 0,81 1,81 53,29
MA-T-146_Campo 1_Pt 11 15 4,13 0,68 0,13 52,01
MA-T-146_Campo 2 _Pt 1 3 6,73 0,87 0,65 47,16
MA-T-146_Campo 2 _Pt 2 4 5,33 0,75 0,72 50,02
MA-T-146_Campo 2 _Pt 3 5 3,13 0,73 0,68 54,72
MA-T-146_Campo 3 _Pt 2 2 4,68 0,64 0,16 52,60
MA-T-146_Campo 3 _Pt 3 3 3,60 0,76 0,17 53,77
MA-T-146_Campo 3 _Pt 4 4 0,71 0,84 1,26 54,15
MA-T-146_Campo 3 _Pt 5 5 1,65 0,82 0,37 59,59
MA-T-146_Campo 3 _Pt 6 6 2,75 0,71 0,32 57,26
MA-T-456_Campo 1 _Pt 1 2 5,27 0,07 0,85 51,67
MA-T-456_Campo 1 _Pt 3 4 5,51 0,25 0,80 51,39
MA-T-456_Campo 1 _Pt 9 11 0,08 0,02 0,13 60,94
MA-T-456_Campo 2 _Pt 1 1 3,65 0,16 0,54 55,38
MA-T-456_Campo 2 _Pt 2 2 5,65 0,44 0,92 51,39
MA-T-456_Campo 2 _Pt 3 3 7,03 0,67 0,74 49,57
MA-T-456_Campo 2 _Pt 4 4 7,19 0,29 1,09 48,52
MA-T-456_Campo 2 _Pt 5 6 5,32 0,03 0,79 50,88
MA-T-466_Campo 1_Pt 1 2 5,84 0,71 0,66 48,47
MA-T-466_Campo 1_Pt 5 11 3,90 0,21 3,14 44,91
MA-T-384_Campo 1_Pt 1 1 4,81 0,00 0,11 50,32
MA-T-384_Campo 1_Pt 2 2 5,25 0,04 0,44 49,24
MA-T-384_Campo 1_Pt 3 3 5,63 0,45 0,13 49,41
MA-T-384_Campo 1_Pt 4 4 6,19 0,74 0,18 50,05
MA-T-384_Campo 1_Pt 5 5 4,91 0,03 0,15 50,05
215

Amostra analisada No. MgO MnO CaO FeO


MA-T-384_Campo 1_Pt 9 12 3,59 2,19 0,32 49,20
MA-T-384_Campo 2_Pt 5 11 4,71 0,68 1,09 55,04
MA-T-384_Campo 3 _Pt 1 1 2,46 0,61 1,00 55,97
MA-T-384_Campo 3 _Pt 2 4 7,51 0,73 0,15 49,60
MA-T-384_Campo 3 _Pt 3 6 7,90 0,54 0,29 47,11
MA-T-384_Campo 3 _Pt 4 12 4,85 0,03 0,47 49,84
MA-T-384_Campo 3 _Pt 5 13 3,29 0,19 0,64 53,93
MA-T-384_Campo 3 _Pt 6 14 5,50 0,11 0,96 48,24
MA-T-384_Campo 3 _Pt 7 15 8,28 0,34 0,67 49,56

Amostra
mol% MgCO3 mol% MnCO3 mol% FeCO3 mol% CaCO3
MA-T-115_Campo 1 _Pt 1
10,90939961 1,295099226 86,99254479 0,802956365
MA-T-115_Campo 1 _Pt 2
10,44074764 1,235214349 82,20531334 6,118724675
MA-T-115_Campo 1 _Pt 3
11,38550557 1,369142877 85,61399975 1,631351807
MA-T-115_Campo 1 _Pt 7
4,776442663 2,037914408 92,25234642 0,933296512
MA-T-115_Campo 1 _Pt 8
0,044209465 0,123384028 99,29514416 0,537262349
MA-T-115_Campo 2 _Pt 1
8,476353694 1,819618898 88,73231235 0,971715061
MA-T-115_Campo 2 _Pt 2
7,148723971 1,674220811 90,01493857 1,162116647
MA-T-115_Campo 2 _Pt 3
10,34343484 1,319037059 87,91588222 0,421645881
MA-T-115_Campo 2 _Pt 4
10,43211371 1,427471952 86,64214463 1,498269712
MA-T-115_Campo 2 _Pt 5
12,13518408 1,361126034 85,56295859 0,940731291
MA-T-115_Campo 2 _Pt 6
11,59725441 1,317604676 85,00352044 2,081620475
MA-T-146_Campo 1 _Pt 1
13,0593956 1,589803261 82,89530656 2,455494578
MA-T-146_Campo 1 _Pt 4
8,219033342 1,693911712 89,17440479 0,912650156
MA-T-146_Campo 1 _Pt 5
12,10345759 1,632414518 84,4027999 1,861327993
MA-T-146_Campo 1 _Pt 6
8,580863532 1,985807891 87,8035897 1,629738873
MA-T-146_Campo 1 _Pt 7
0 0,027015326 99,31577083 0,657213839
MA-T-146_Campo 1 _Pt 8
3,029487964 1,047237194 91,779999 4,143275842
MA-T-146_Campo 1 _Pt 9
1,774883512 1,397385828 91,32768613 5,500044532
MA-T-146_Campo 1 _Pt 11
9,184107608 1,171011556 89,24900885 0,395871982
MA-T-146_Campo 2 _Pt 1
15,08274815 1,501726145 81,42750766 1,988018049
MA-T-146_Campo 2 _Pt 2
11,72149011 1,270327812 84,85648348 2,151698597
216

Amostra
mol% MgCO3 mol% MnCO3 mol% FeCO3 mol% CaCO3
MA-T-146_Campo 2 _Pt 3
6,688421867 1,208793976 90,1100311 1,99275306
MA-T-146_Campo 3 _Pt 2
10,19137058 1,085582002 88,24071585 0,482331573
MA-T-146_Campo 3 _Pt 3
7,848531459 1,289176068 90,34553602 0,516756456
MA-T-146_Campo 3 _Pt 4
1,584737715 1,447191959 93,12385708 3,844213241
MA-T-146_Campo 3 _Pt 5
3,379451646 1,309974939 94,28779108 1,022782331
MA-T-146_Campo 3 _Pt 6
5,752879088 1,148218493 92,17430583 0,92459659
MA-T-456_Campo 1 _Pt 1
11,37625639 0,121918066 86,00280113 2,499024408
MA-T-456_Campo 1 _Pt 3
11,87065629 0,42454695 85,3393261 2,365470663
MA-T-456_Campo 1 _Pt 9
0,167839666 0,036395861 99,42978347 0,365980999
MA-T-456_Campo 2 _Pt 1
7,733136642 0,269285713 90,44178891 1,555788733
MA-T-456_Campo 2 _Pt 2
12,06199254 0,723605386 84,54629277 2,668109303
MA-T-456_Campo 2 _Pt 3
15,02371488 1,105647293 81,71684167 2,153796158
MA-T-456_Campo 2 _Pt 4
15,52704686 0,479096141 80,77799715 3,215859854
MA-T-456_Campo 2 _Pt 5
11,65244504 0,057354177 85,91635911 2,373841676
MA-T-466_Campo 1_Pt 1
13,07096525 1,236217509 83,67149655 2,02132069
MA-T-466_Campo 1_Pt 5
9,078508926 0,383837595 80,57315555 9,964497929
MA-T-384_Campo 1_Pt 1
10,98510371 0 88,66547923 0,349417061
MA-T-384_Campo 1_Pt 2
11,97888089 0,065192179 86,57415242 1,381774515
MA-T-384_Campo 1_Pt 3
12,72311874 0,780715569 86,09019549 0,405970201
MA-T-384_Campo 1_Pt 4
13,57264664 1,25575194 84,63980666 0,531794758
MA-T-384_Campo 1_Pt 5
11,23707855 0,053319845 88,23738179 0,47221982
MA-T-384_Campo 1_Pt 9
8,230618534 3,885423977 86,88557922 0,99837827
MA-T-384_Campo 2_Pt 5
9,57706401 1,075234857 86,31084067 3,036860467
MA-T-384_Campo 3 _Pt 1
5,188420874 1,000330501 90,92879684 2,882451788
MA-T-384_Campo 3 _Pt 2
16,14015238 1,210026653 82,2172676 0,432553371
MA-T-384_Campo 3 _Pt 3
17,53873196 0,929415971 80,64776611 0,884085956
MA-T-384_Campo 3 _Pt 4
11,04921668 0,053423712 87,44319701 1,454162598
MA-T-384_Campo 3 _Pt 5
7,166470737 0,325882676 90,61673705 1,89090954
MA-T-384_Campo 3 _Pt 6
12,48697207 0,184497459 84,36174955 2,96678092
MA-T-384_Campo 3 _Pt 7
17,37470904 0,557559379 80,15195731 1,915774271
217

APÊNDICE F – TRATAMENTO DOS DADOS GEOQUÍMICOS E


OUTROS DIAGRAMAS GERADOS

1. P2O5 versus Terras Raras


218

2. Scatter diagrama

(a) Scatter diagrama de elementos traços para os fosforitos e ironstones fosfáticos estudados
(normalizado pelo UCC – Upper Continental Crust conforme Rudnick, R.L. e Gao, S., 2003). (b)
Scatter diagrama de elementos traços para o Kuh-e-Sefid e o Leste do Mediterrâneo e outros
fosforitos no mundo obtido em Zaravansdi et al. (2019) para comparar com o diagrama estudado.

a)

b)
Ba ppm Al2O3 %

2
4
6
8
10
12
14
16
18
20
22
24

1000
1500
2000
2500
3000

0
500
ironstone
ironstone

phosphatic ironstone phosphatic ironstone

phosphatic mudstone/heterolithe
phosphatic mudstone/heterolithe

phosphorite
phosphorite
3. Boxplots dos elementos segundo a classificação por litofácies

siderite concretion
siderite concretion

sideritic/phosphatic concretion (mudstone)

sideritic/phosphatic concretion (mudstone)


219
Cd ppm CaO%

10
15
20
25
30
35
40

0
5

0,0
0,2
0,4
0,6
0,8
1,0
1,2
ironstone ironstone

phosphatic ironstone phosphatic ironstone

phosphatic mudstone/heterolithe phosphatic mudstone/heterolithe

phosphorite phosphorite

siderite concretion siderite concretion

sideritic/phosphatic concretion (mudstone) sideritic/phosphatic concretion (mudstone)


220
Co

10
20
30
40
50
60
70
80

0
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


221
COT %

0,0
0,5
1,0
1,5
2,0
2,5
3,0

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


222
Cr ppm

20
40
60
80

0
100
120
140
160
180
200
220
240
260

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


223
Cu ppm

10
20
30
40
50
60
70
80

0
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


224
Fe2O3%

10
20
30
40
50
60
70

0
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


225
K2O%

0,0
0,5
1,0
1,5
2,0
2,5
3,0
3,5

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


226
LOI%

12
16
20
24
28

0
4
8
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


227
MgO%

0
1
2
3
4
5
6
7

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


228
MnO%

0,0
0,2
0,4
0,6
0,8
1,0
1,2
1,4
1,6

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


229
Mo ppm

10
20
30
40
50
60
70

0
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


230
Ni ppm

10
20
30
40
50
60
70
80

0
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


231
P2O5%

12
16
20
24
28

0
4
8
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


232
REE ppm

1000
1500
2000
2500
3000

0
500
ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


233
S%

0,0
0,4
0,8
1,2
1,6
2,0
2,4

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


234
SiO2%

10
20
30
40
50
60
70

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


235
Sr ppm
1000

0
200
400
600
800

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


236
Th ppm

50

0
100
150
200
250
300

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


237
U ppm

0
100
200
300
400
500
600
700

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


238
V ppm 1000

0
200
400
600
800

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


239
Y ppm
1000

0
200
400
600
800

ironstone

phosphatic ironstone

phosphatic mudstone/heterolithe

phosphorite

siderite concretion

sideritic/phosphatic concretion (mudstone)


240
241

1800
1600
1400
1200
Zr ppm

1000
800
600
400
200
0

sideritic/phosphatic concretion (mudstone)


phosphatic ironstone
ironstone

phosphatic mudstone/heterolithe

phosphorite

References siderite concretion

Rudnick, R.L., Gao, S., 2003. The composition of the continental crust. In: Rudnick, R.L. (Ed.), The
Crust. In: Holland, H.D., Turekian, K.K. (Eds.), Treatise on Geochemistry, vol. 3. Elsevier-Pergamon,
Oxford. pp. 1–64.

Zarasvandi, A., Fereydouni, Z., Pourkaseb, H., Sadeghi, M., Mokhtari, B., Alizadeh, B., 2019.
Geochemistry of trace elements and their relations with organic matter in Kuh-e-Sefid phosphorite
mineralization, Zagros Mountain, Iran. Ore Geology Reviews, 104, 72–87.
242

ANEXO A – REGRAS DE FORMATAÇÃO DA REVISTA

SEDIMENTARY GEOLOGY
243
244
245
246
247
248
249
250
251
252
253
254
255
256

ANEXO B – REGRAS DE FORMATAÇÃO DA REVISTA


JOURNAL OF SOUTH AMERICAN EARTH SCIENCES
257
258
259
260
261
262
263
264
265
266
267
268
269

ANEXO C – COMPROVANTE DE ACEITE DE ARTIGO 1


270

ANEXO D – COMPROVANTE DE SUBMISSÃO DE ARTIGO 2

You might also like