Download PDF
ads:
UNIVERSIDADE FEDERAL DO PARÁ
CENTRO DE GEOCIÊNCIAS
CURSO DE PÓS-GRADUAÇÃO EM GEOLOGIA E GEOQUÍMICA
FÁCIES, PETROGRAFIA E GEOQUÍMICA DA
FORMAÇÃO CODÓ, NEO-APTIANO, BACIA DE SÃO
LUÍS-GRAJAÚ
TESE APRESENTADA POR
JACKSON DOUGLAS SILVA DA PAZ
Como requisito parcial à obtenção do Grau de Doutor em Ciências
na área de GEOLOGIA
DATA DA APROVAÇÃO
:
COMITÊ DE TESE:
_____________________________________
DILCE DE FÁTIMA ROSSETTI
(ORIENTADORA)
_____________________________________
MOACIR JOSÉ BUENANO MACAMBIRA
(CO-ORIENTADOR)
_____________________________________
ALCIDES NÓBREGA SIAL
_____________________________________
ANA MARIA GÓES
_____________________________________
VIRGÍNIO NEUMANN
_____________________________________
WERNER TRUCKENBRODT
Belém
-
2005
ads:
Livros Grátis
http://www.livrosgratis.com.br
Milhares de livros grátis para download.
À Nazaré
ii
ads:
AGRADECIMENTOS
À Dilce de Fátima Rossetti, para quem eu não tenho palavras suficientes para expressar toda a minha gratidão,
afeto e admiração;
Ao Prof. Moacir Macambira, pela amizade e pela co-orientação em todas as etapas dentro do Laboratório de
Geocronologia da UFPA;
Ao CNPq pela concessão da bolsa de estudo;
Ao Grupo Itapicuru Agro-Industrial S/A pela colaboração e pela liberdade que nos foi concedida de acessar as
minas exploradas por esta companhia nos arredores do Município de Codó;
Ao Grupo Mineradora Vale do Grajaú S/A pela atenção e pela liberdade concedidas para observar as minas
exploradas por esta empresa na região do Município de Grajaú;
À hospitalidade do povo tanto de Codó quanto de Grajaú, onde sempre fomos bem recebidos. Em especial, às
Sras. D. Gonçala e D. Edite, donas das pousadas que nos receberam muito cordialmente neste trabalho durante as etapas
de campo, e à D. Fátima, cuja comida foi um alento durante estas etapas;
Ao Museu Paraense Emílio Goeldi pela infra-estrutura concedida durante os vários anos deste trabalho;
À Universidade Federal do Pará, ao Centro de Geociências e ao Curso de Pós-Graduação em Geologia e
Geoquímica, pela infra-estrutura, pelo pessoal e pela organização que me auxiliaram muito neste trabalho;
À Coordenação do Curso de Pós-Graduação em Geologia e Geoquímica, nas pessoas do seu atual coordenador
Prof. Dr. Cândido Veloso, do antigo coordenador Prof. Dr. Paulo Gorayeb e das secretarias Gladys Pimentel e Cleidiane
Caldeira, pela atenção e competência que dispensaram a mim em todas as etapas deste trabalho;
À comissão que formou a banca examinadora desta tese: à Dra. Ana Maria Góes, amiga de todos, membro do
nosso grupo de pesquisa em geologia histórica e sedimentar (GSED) e por quem nossa admiração vai além da geologia;
ao Dr. Werner Truckenbrodt, pela paciência com que tentou melhorar cada aspecto dos mais variados textos relacionados
com este trabalho, durante os anos em que trabalhei ao seu lado; ao Dr. Alcides Nóbrega Sial, cujo apoio foi fundamental
na execução da parte deste trabalho relacionada aos isótopos estáveis; e ao Dr. Virgínio Neumann, que acreditou neste
trabalho de imediato, sabendo que seria difícil mas executável no prazo exigido;
Ao Grupo de Geologia Histórica e Sedimentar (GSED) do Museu Goeldi e Universidade Federal do Pará, em
especial aos estudantes do grupo que sempre estiveram um ao lado do outro a fim de ajudar da melhor forma possível:
Emídio, Carol Miranda, Daniele, Marivaldo, Anderson, Leandro, Hebérton, Cristovam, Leonardo, Denis, Sue Anne,
Samanta e Paulinha;
Ao Laboratório de Isótopos Estáveis do Núcleo de Estudos de Granitos da Universidade Federal de Pernambuco
(LABISE/NEG/UFPE), nas pessoas da Profa. Dra. Valderez Ferreira, que além disso foi extremamente atenciosa com este
trabalho quando da minha estadia no LABISE, auxiliando-me totalmente na parte de análise elementar com fluorescência
de raios-X; à Silvana Diene Barros, pela amizade, pelo amor e pelo carinho mútuos; Roberta Brasilino, pela amizade e
carinho recebidos na sua casa e de Xuxxu; à Gilsa, pela experiência compartilhada na separação química de C e O; e, de
novo, ao Prof. Sial, pela paciência e atenção com que me recebeu em seu laboratório.
Ao Laboratório de Geocronologia e Isótopos Estáveis da Universidade Federal do Pará (Pará-Iso/UFPA), nas
pessoas do Prof. Dr. Moacir Macambira, das Químicas Rose Brabo, Roberta e da estudante Gilmara, pela atenção
despendida com meu trabalho e pelo auxílio na execução das etapas mais difíceis;
Ao Laboratório de Geoquímica Orgânica da Universidade Estadual do Rio de Janeiro, nas pessoas do Prof. Dr.
René Rodrigues, do Químico Luis Freixo e da secretária Sra. Rosalva, pela ajuda dispensada durante a minha estadia
naquele laboratório;
Ao pessoal técnico do Laboratório de Laminação do Centro de Geociências, cujos técnicos Sra. Shirley Tavares,
Sr. Eduardo Soares e Sr. Israel Tavares, eu sou muito grato pela franca amizade, e ao Laboratório de Sedimentologia do
Centro de Geociências, na pessoa do Sr. Antônio Lopes, pela amizade e paciência. Todos com atenção e cuidado no trato
deste trabalho;
Ao pessoal da Coordenação de Ciências da Terra do Museu Paraense Emílio Goeldi, nas pessoas do Chico,
Cristina, Márcia, Cléia, Paulo, Hudson, Maria Tereza, Amílcar, Josué e D. Maria, pelo companheirismo compartilhado
nestes anos;
Ao pessoal de transporte do Museu Goeldi, pela amizade nas duras etapas de campo:
Aos meus amigos Braga e Paulo Benevides e suas famílias pela amizade e cordialidade em me receber durante a
minha estadia na sua cidade, realizando etapas deste trabalho;
Aos amigos mais de perto, daqui de Belém, Leonilde, D. Siroca e Seu Joca, Leonete, Leonora, Vante, Paulinho,
Veruska, Angelina e a toda a família Aviz, pelo bem-querer como a de um membro da família;
Ao amigos próximos Erimar, Vanderlei, Cláudia e Heloísa, que são amigos de todas as horas;
E aos amigos distantes Eisner, Elias, Izaías, Naná, Roseane, Douglas e Ana Maria, que são amigos para sempre.
À minha família, mais perto, Naíde, Élson, Everton, Felipe, Chico e todos os outros mais longe;
À minha mãe...
Aos meus irmãos Jailton, Derimar e Janderson ...
À Leninha...
MUITO OBRIGADO!
iii
O Lago
Eu não vi o mar.
Não sei se o mar é bonito,
não sei se ele é bravo.
O mar não me importa.
Eu vi o lago.
O lago, sim.
O lago é grande
E calmo também.
Na chuva de cores
da tarde que explode
o lago brilh
a
o lago se pint
a
de todas as cores.
Eu não vi o mar. Eu vi o lago...
(Corrompido a partir de Carlos Drummond de
Andrade... no original “A lagoa”).
iv
SUMÁRIO
DEDICATÓRIA .…………………………………………………………………........ ii
AGRADECIMENTOS .................................................................................................... iii
EPÍGRAFE ...................................................................................................................... iv
LISTA DE FIGURAS ..................................................................................................... vii
RESUMO ......................................................................................................................... 1
ABSTRACT .................................................................................................................... 5
APRESENTAÇÃO .......................................................................................................... 9
1. INTRODUÇÃO ......................................................................................................... 10
1.1. ARCABOUÇO GEOLÓGICO E PROBLEMÁTICA ............................................. 10
1.2. OBJETIVOS ............................................................................................................. 14
1.3. ÁREA DE ESTUDO ................................................................................................ 14
1.4. MÉTODOS ............................................................................................................... 15
1.5. GEOQUÍMICA ISOTÓPICA ................................................................................... 17
1.5.1. Fundamentação teórica ...................................................................................... 17
1.5.2. Aplicação em depósitos lacustres ....................................................................... 19
REFERÊNCIAS .............................................................................................................. 24
2. LINKING LACUSTRINE CYCLES WITH SYN-SEDIMENTARY
TECTONIC EPISODES: AN EXAMPLE FROM THE CODÓ FORMATION
(LATE APTIAN), NORTHEASTERN BRAZIL……………………………………
29
2.1. ABSTRACT .....................................................................................………….…... 29
2.1. INTRODUCTION ..................................................................................………….. 30
2.3. GEOLOGICAL SETTING ………………………………………………………... 31
2.4. DEPOSITIONAL ENVIRONMENT ……………………………………………... 35
2.4.1. Central lake facies association ………………………………………………… 36
2.4.2. Transitional lake facies association …………………………………………… 38
2.4.3. Marginal lake facies association …………………………………………….… 39
2.4.4. Saline pan/sabkha facies association …………………………………………. 40
2.4.5. Interpretation …………………………………………………………………. 41
2.5. SHALLOWING-UPWARD CYCLES AND DEPOSITIONAL UNITS …….…... 43
2.5.1. Description .....................................................................................………….…. 43
2.5.2. Interpretation .....................................................................................…………. 48
2.6. ORIGIN OF THE SHALLOWING UPWARD CYCLES ……………………….. 49
2.7. CONCLUSION .....................................................................................………….. 54
REFERENCES .........................................................................................................…... 55
3. PETROGRAPHY OF GYPSUM-BEARING FACIES OF THE APTIAN
CODÓ FORMATION .....................................................................................………..
60
3.1. ABSTRACT .....................................................................................………….…... 60
3.2. INTRODUCTION ..................................................................................………….. 61
3.3. GEOLOGICAL SETTING ………………………………………………………... 62
3.4. PETROGRAPHY OF THE EVAPORITES ………………………………………. 65
3.4.1. Chevron gypsum (selenite) .................................................................................. 65
3.4.2. Nodular/lensoidal gypsum/anhydrite ................................................................. 70
3.4.3. Acicular gypsum .................................................................................................. 70
3.4.4. Mosaic gypsum ..................................................................................………....... 70
3.4.5. Brecciated gypsum/gypsarenite ……………………………………………….. 70
v
3.4.6. Pseudo-nodular anhydrite/gypsum .................................................................... 72
3.4.7. Rosettes of gypsum .............................................................................................. 72
3.5. PARAGENESIS ................................................................................................…... 74
3.6. CONCLUSIONS ..................................................................................……….…... 79
REFERENCES .........................................................................................................…... 80
4. GENESIS AND PALEOHYDROLOGY OF A SALINE PAN/LAKE SYSTEM
(LATE APTIAN) FROM THE BRAZILIAN EQUATORIAL MARGIN:
INTEGRATION OF FACIES, SR AND S ISOTOPES …………………………….
85
4.1. ABSTRACT ..................................................................................………………... 85
4.2. INTRODUCTION ..................................................................................………….. 86
4.3. MATERIAL AND METHODS .............................................................…………... 87
4.4. GEOLOGICAL FRAMEWORK AND PALEOENVIRONMENTAL CONTEXT 89
4.5. CHARACTERIZATION OF THE EVAPORITES ………………………………. 92
4.5.1. Description ..................................................................................………………. 92
4.5.2. Depositional setting …………………………………………………………….. 94
4.6. Sr AND S ISOTOPES FROM THE EVAPORITES ………………………….…... 96
4.7. BURIAL OVERPRINT ……………………………………………………….…... 100
4.8. DISCUSSION ..................................................................................………………. 102
4.9. CONCLUSIONS ..................................................................................……………. 108
REFERENCES ................................................................................................................ 109
5. PALEOHYDROLOGY OF A LATE APTIAN LACUSTRINE SYSTEM
FROM NORTHEASTERN BRAZIL WITH BASIS ON THE INTEGRATION
OF FACIES AND ISOTOPIC GEOCHEMISTRY ………………………………...
113
5.1. ABSTRACT ..................................................................................………………... 113
5.2. INTRODUCTION ..................................................................................………….. 114
5.3. GEOLOGICAL SETTING ................................................................................…... 116
5.4. FACIES ARCHITECTURE AND DEPOSITIONAL SYSTEM ...………………. 118
5.4.1. Description ..................................................................................………………. 118
5.4.2. Interpretation ..................................................................................………….… 122
5.5. GEOCHEMICAL TREATMENT …………………………………………….…... 123
5.6. EVALUATION OF DIAGENETIC IMPRINT ……………………………….….. 124
5.7. STABLE ISOTOPES ..................................................................................………. 128
5.8. ISOTOPIC CHARACTERIZATION OF THE LACUSTRINE SYSTEM ………. 129
5.9. FINAL REMARKS ……………………………………………………………….. 138
REFERENCES ................................................................................................................ 140
6. CONCLUSõES GERAIS ..................................................................................…… 147
vi
LISTA DE FIGURAS
Fig. 1-1: A) Corte esquemático do Sistema de Gráben Gurupi, incluindo as sub-
bacias do Grajaú, Sao Luís e Ceaté; B) estratigrafia, preenchimento e
tectônica da Bacia de Sao Luís- Grajaú; C) localização da área de estudo
nas porções leste e sul da Bacia do Grajaú; D) detalhe da figura anterior,
indicando as localidades com minas a céu aberto estudadas neste
trabalho (ver texto).
...................... 12
Fig. 1-2: Respiração do lago, consumo de matéria orgânica, soterramento de
carbono orgânico, influxo de águas fluviais com bicarbonato mais
“leve”, baixa troca de gás carbônico com a atmosfera e participação do
gás carbônico metanogênico alternam-se em importância durante fases
de lago alto e baixo e influenciam indiretamente no valor do d13C do
carbonato lacustre (Modificado de Guzzo, 1997).
...................... 20
Fig. 1-3: A) processos de enriquecimento em
18
O de águas lacustres ligados ao
ciclo hidrológico (cf. Lister et al., 1991); B) modelo de sinal
paleohidrológico de isótopos de oxigênio em carbonatos lacustres,
considerando bacias hidrologicamente fechadas (Lister et al., 1991;
Bellanca et al., 1992).
...................... 21
Fig. 1-4: Ciclo resumido do enxofre, enfatizando a passagem do sulfato para
sulfeto de hidrogênio por respiração anaeróbica de bactérias sulfato-
redutoras (modificado de Hoefs, 1980).
...................... 22
Fig. 1-5: Fatores influentes na composição da razão
87
Sr/
86
Sr de depósitos
lacustres e de bacias oceânicas (cf. Holser et al., 1996).
...................... 23
Fig .2-1: Location map of the study area in the eastern and southern of São Luís-
Grajaú Basin, northern Brazil, with indication. The map to the right
indicates the studied localities where the Codó Formation is well
exposed in limestone and evaporite quarries. (A=Santo Amaro; B-
C=Itapicuru Evaporite Quarry; D-E=Itapicuru D-6 Limestone Quarry;
F= Transmarenhense Road-Km 22; G=Olho D´água Quarry; H=Chorado
Quarry; I=Transmaranhense Road-Km 16.
...................... 31
Fig .2-2: a) Map displaying the structural lineaments of the São Luís-Grajaú
Basin. Note the main NW-SE, and NE-SW oriented fault traces, and the
subordinate E-W lineaments, the later representing the record of the
strike-slip phase of the basin. b) A geological section interpreted from a
seismic line and the well logs indicated in figure a, with the plot of the
main fault traces. Note that these are vertical to nearly vertical and
represent reactivations of faults derived from the crystalline and
Palaeozoic basement. The faults cut throughtout the entire basin,
particularly disrupting the Cretaceous successions, attesting the rift
development (Modified after Góes & Rossetti, 2001).
...................... 33
Fig .2-3: Stratigraphy of the São Luís-Grajaú Basin.
...................... 34
Fig .2-4: Exposure of the Codó Formation in a limestone quarry, near the town of
Codó, eastern São Luís-Grajaú Basin, with indication of the syn-
sedimentary soft sediment deformation zones formed by seismic activity
as described in the text.
...................... 34
vii
Fig .2-5: A summary of sedimentary facies described in the Codó Formation,
with their distribution arranged according to facies associations Fa1 to
Fa4, attributed to marginal, intermediate and central lake environments,
as well as saline pan/sabkha complex, respectively.
...................... 36
Fig .2-6: Sedimentary facies representative of the Codó Formation in the study
area. a, b) Central lake deposits, with massive macronodular evaporites
(Gy) and bituminous black shales (Bsh) within evaporites. The hatched
line in a, indicates the top of the evaporite beds. The deposits overlying
this surface consists of limestone and grey/green shale interbeddings,
attributed to transitional lake settings. (person in a for scale=1.6 m tall;
rod subdivision in b=10 cm). c) Transitional lake deposits, consisting of
limestone (Lm) and grey/green shale (Sh) interbeddings. (pen for
scale=15 cm long) d-g) Marginal lake deposits, illustrating: rhythmites
of ooidal/peloidal packstone (p) and ostracodal wackestone-grainstone
(g) with cryptomicrobial mats (am) in d; a detail of the ooidal/peloidal
packstone in e; fractured massive pelite in f; and photomicrography of
intraclastic grainstone in g. (pen for scale in e and f=15 cm long). h-i)
Saline pan/sabkha deposits, illustrating the layered gypsum. (arrows
indicate laminae of gypsum formed by vertical aligned chevron
crystals). Note in h the cycles formed by darker/lighter couplets, which
consist of microgranular gypsum/vertical aligned gypsum, respectively,
attributed to seasonal fluctuations.
...................... 37
Fig .2-7: Lithostratigraphic profiles representative of the Codó Formation
exposed in the study area, with the main facies characteristics and the
two ranks of shallowing-upward cycles described in the text. (See figure
1 for location of the profiles A-I). Datum=discontinuous surface with
evidences for maximum subaereal exposure within the Codó Formation.
...................... 45
Fig .2-8: Diagrams illustrating the four types of lower-rank, shallowing-upward
cycles of the Codó Formation. Thickness of individual cycles range from
0.3-5.6 m. See figure 2-7 for legend.
...................... 46
Fig .2-9: Types of shallowing-upward units present in the Codó Formation,
illustrating: a) Unit 2, represented by one complete cycle type 1 with
evaporites and shales. b) Unit 3, presented from base to top by two
incomplete cycles type 1 and one complete cycle type 2. c) The
uppermost portion of unit 2 and the base of unit 3, with the latter
showing three incomplete cycles type 1 and an upper incomplete cycle
type 2.
...................... 47
Fig .2-10: A representative lithostratigraphic profile of the Codó Formation
showing the good match between higher-rank shallowing-upward
cycles, represented by depositional units 2 to 4, and the syn-sedimentary
deformational zones described by Rossetti and Góes (2000). (See Fig. 2-
7 for legend).
...................... 53
Fig. 3-1: A) Location map of the study areas in the São Luís-Grajaú Basin,
northeastern Brazil. B) A close up map, indicating the location of the
studied sections in the Codó and Grajaú areas.
...................... 62
Fig. 3-2: A) Diagram with the sketched representation of the proposed structural
pattern of the São Luís-Grajaú Basin in the Gurupi Graben System, and
its relation with the Pará-Maranhão Basin. In this model, the Ferrer-
Urbano Santos Arch is considered as an intrabasin horst within an
...................... 63
viii
abandoned intracontinental rift system formed by combination of pure
shear stress and strike-slip deformation. The northeastward rifting
migration through time gave rise to the development of a deeper basin,
represented by the Caeté Sub-basin (cf. Góes & Rossetti, 2001). B)
Stratigraphic framework of the São Luís-Grajaú Basin (cf. Rossetti,
2001).
Fig. 3-3: Depositional model proposed for the Codó Formation in the Codó and
Grajaú areas (cf. Paz & Rossetti, 2001).
...................... 64
Fig. 3-4: Measured vertical sections representative of the evaporites studied in the
Codó Formation (see Fig. 3-1 for section location).
...................... 67
Fig. 3-5: The several phases of evaporites recognized in the Codó Formation,
with paragenesis indicated by lateral position of the boxes.
...................... 68
Fig. 3-6: Textures of the evaporites from the Codó Formation exposed in the
study areas. A) Photomicrography of chevron gypsum. B)
Photomicrography of the nodular gypsum (ng) within bituminous black
shales (sh). Note calcite relicts (arrow) within nodules (crossed nicols).
C) A close up in SEM view from the gypsum nodules shown in figure B,
illustrating their composition by micrometric, equant, lath-like crystals,
typical of anhydrite. D) Photomicrography of the acicular gypsum and,
at the base, mosaic gypsum. E) Photomicrography of relics of anhydrite
(arrow) within a large mosaic gypsum crystal.
...................... 69
Fig. 3-7: Textures of the evaporites from the Codó Formation exposed in the
study areas. A) Brecciated gypsum (upper part) evolved from mosaic
gypsum (lower part). The dark lines around the clasts in the upper part
of the photograph were formed by muds resulting from mechanical
infiltration (parallel nicols). B) A detailed of the brecciated gypsum
illustrating a crystal with several fractures (arrow). Note that the optical
continuity beyond fractures, attesting fracturing occurred within a single
large crystal (crossed nicols). C and D) Brecciated gypsum locally
displaying well rounde clasts resembling gypsarenite (C=parallel nicols;
D=crossed nicols). Note in D that several clasts are in optical continuity,
revealing they were most likely formed by fracturing of a same large
crystal.
...................... 71
Fig. 3-8: Textures of the evaporites from the Codó Formation exposed in the
study areas. A) Overlay field drawing illustrating a spot within the
pseudo-nodular anhydrite/gypsum with relics of a complex arrangement
formed by mosaic, acicular and fibrous gypsum, bounded by films of
calcimudstone. B) Pseudo-nodular gypsum, formed by fracturing. C) A
spot within the pseudo-nodular anhydrite/gypsum, with nodules of
anhydrite bound by films of fibrous gypsum (arrows). Note the
superimposed rosettes of gypsum of variable sizes (rg). D) Alabastrine
(gy) and fibrous (gf) gypsum in gradational contacts. The arrows
indicate places where the fibrous gypsum is partly replaced by the
alabastrine gypsum, recorded by numerous crystals of the later over the
fibrous gypsum, which in turn remains as diffuse relics. E)
Photomicrography of the nodules shown in C, illustrating their
composition of tiny, equant, lath-like crystals (an). Note that the edges
of the nodules were replaced by fibrous (gf) or mosaic (gy) gypsum.
...................... 73
Fig. 3-9: Textures of the evaporites from the Codó Formation exposed in the
d)ifhddlhdi/
...................... 74
ix
study areas. A) SEM view of the pseudo-nodular anhydrite/gypsum,
with lath-like anhydrite (an) interlaced with gypsum (gy). This material
comes from a nodule of anhydrite depicted in figure 8C. B) Calcite (ca)
cementing fractures in the pseudo-nodular anhydrite/gypsum (gy). C) A
detail of the pseudo-nodular anhydrite/gypsum showing several fractures
(arrows) filled by a mixture of calcite and muds. Note larger calcite
crystals (ca) that grew sidewards from the fractures trough replacement
of gypsum (gy). D) Relics of calcite (ca) within alabastrine gypsum (gy)
from the pseudo-nodular anhydrite/gypsum. E) Rhombs of calcite after
dolomite (arrows). (Except for the SEM micrography shown in A, all the
other figures were obtained under petrographic microscope with crossed
nicols).
Fig. 3-10: Model to explain the formation of primary and early diagenetic
evaporite phases recorded in the Codó Formation.
...................... 75
Fig. 4-1: A) Location and geological map of the study area in the north of Brazil.
B) Stratigraphic framework of the São Luís-Grajaú Basin, with the
depositional sequences discussed in the text.
...................... 88
Fig. 4-2: A) A view of the unconformity between the Codó Formation and the
overlying Albian deposits of the Itapecuru Group. (Gy= gypsum;
Lsh=interbedded limestone and shale). B) Lithostratigraphic profile
representative of the Codó Formation in the study area, with indication
of the shallowing upward cycles formed by central lake (1), intermediate
lake (2) and marginal lake (3) deposits. The letters to the left locate
figures A-F. C) Bituminous shale (black areas) interbedded with
evaporite (white areas) from central lake deposits. D) Interbedded
limestones (lm) and shales (darker areas from intermediate lake deposits.
E) Interbedded lime-mudstone (Ml) and pisoidal limestones (Pl) and F)
Rhythmite of lime mudstone (white beds) and microbial mats (black
beds) from marginal lake deposits.
...................... 90
Fig. 4-3: Evaporite deposits from the Codó Formation. A) Layered gypsum.
(Person for scale=1.68 m tall). B) Detail of the layered gypsum
illustrating the alternating darker and lighter beds consisting of gypsum
crystals or nodules within a matrix of black shale (Dg), and upward-
oriented chevron/acicular gypsum crystals (Lg), respectively. C) A plan
view of the gypsarenite facies (pen for scale=15 cm long). D)
Massive/macronodular gypsum (Gy) underlying interbedded limestones
and shales (Lsh). The white dotted line indicates the discontinuity
surface between the Codó Formation and the Itapecuru Group (Albian).
(Person for scale=1.68 m tall) E) A detail of the massive/macronodular
gypsum strongly affected by fractures, resulting in almond-shaped
anhydrite nodules. Note the abundant rosettes of gypsum (rg) that are
disperse in this facies. (Lens cap for scale=10 cm in diameter).
...................... 93
Fig. 4-4: Microscopy of the evaporites from the Codó Formation. A) Gypsum
with nodular texture of the dark layered facies (parallel polarizers). B)
Layered nodular gypsum with relics of anhydrite (=an; crossed
polarizers). C) Recrystallized nodular gypsum displaying mosaic of
gypsum crystals that grew beyond the nodule boundaries (crossed
polarizers). D) Scanning electron photomicrography of the light beds in
the layered gypsum, showing chevron gypsum locally replaced by
acicular gypsum. E) A petrographic view of the gypsarenite formed by
rounded gypsum grains after overgrowth. F) Alabastrine gypsum.
...................... 95
x
Fig. 4-5: A summary of facies distribution and depositional interpretation
proposed for the Codó Formation in the study areas.
...................... 97
Fig. 4-6: Lithostratigraphic profiles representative of the Codó Formation in the
study areas with the plot of
87
Sr/
86
Sr, Sr (ppm) and δ
34
S isotope values.
...................... 99
Fig. 4-7:
87
Sr/
86
Sr and δ
34
S values from the Codó Formation and comparison with
the values from the Aptian seawater. Note that the strontium values for
both of the study areas are much higher than those expected from Aptian
seawaters, which support a continental-sourced brine for the evaporites.
This interpretation is further supported by the higher values of S
isotopes, and the scatter nature of both isotopes. The lower S values
from the Grajaú area are interpreted to have a facies control, and
probably respond to increased evaporation (see text for further
discussion).
...................... 104
Fig. 4-8: Model to explain the distribution of
87
Sr/
86
Sr according to the low
frequency cycles of the Codó Formation. (Inspired on Rhodes et al.,
2002).
...................... 106
Fig. 5-1: A) Location of the study area in the Codó region, eastern margin of the
Grajaú Basin. B) Stratigraphy and main tectonic stages of the São Luís-
Grajaú Basin.
...................... 117
Fig. 5-2: Diagram illustrating the proposed lacustrine depositional model for the
Codó Formation, characterized by central to marginal lake deposits. A)
General view of marginal lake deposits between intermediate lake
deposits. B) A detail showing the upward gradation from interbedded
limestones (Lm) and laminated argillites (Al; intermediate lake) to
rhythmites (Rh; marginal lake). C) Fenestral calcarenite from marginal
lake deposits. D) Rhythmite of limestones (lighter color) and microbial
mats (darker color) from marginal lake deposits. E) General view of
central lake deposits (Ev=evaporite). F) Bituminous black shales (Sh)
interbedded with evaporites (Ev). G) Layered gypsum.
...................... 119
Fig. 5-3: Shallowing-upward cycles of the Codó Formation. A) Examples of first-
and second-order cycles. B-D) Third-order cycles formed by alternations
of bituminous black shale with streaks of lime-mudstone (Bsl) and
bituminous black shales with native sulphur (Bss) (B), ostracodal
grainstone (Gro) and vadose pisoidal packstone (Pp) with microbial mats
(M) (C), ostracodal grainstone (Go) and wackestone (Wo) (D).
...................... 120
Fig. 5-4: First- and second-order cycles of the Codó Formation with relation to
soft-sediment deformation zones attributed to syn-sedimentary seismic
activity. (See Fig . 5-7 for legend).
...................... 121
Fig. 5-5: Photomicrographies of biogenetically not affected facies utilized for the
isotopic analysis. A,B) Calcimudstone (A=crossed nichols; B=scanning
electron microscopy). C) Ostracodal grainstone (crossed nichols). D)
Electron photomicrography illustrating ostracode shells of ostracodal
grainstone, formed by densely packed, columnar calcite crystals
attributed to primary origin.
...................... 125
Fig. 5-6: Distribution of the trace elements Fe (A), Mn (B), Mg (C) and Sr (D)
from samples used in the isotopic analysis presented herein (I=meteoric
...................... 127
xi
calcite; II=marine calcite; III=diagenetic calcite). (After Tucker and
Wright, 1990).
Fig. 5-7: Lithostratigraphic profiles representative of the Codó Formation
exposed in the study area, with the stratigraphic distribution of facies
associations, first- and second-order cycles, and δ
18
O
(%o PDB)
and δ
13
C
(%o
PDB)
values.
...................... 130
Fig. 5-8: Plot and correlation of carbon and oxygen isotope data from the Codó
Formation in the eastern Grajaú Basin. The positive correlation in
profiles B and C is attributed to episodes of dominantly closed lake
system, while the negative correlation in profile A records lake opening.
...................... 131
Fig. 5-9: δ
13
C trend of marine-bearing Aptian carbonates (Modified from Jones
and Jenkins, 2001).
...................... 131
Fig. 5-10: Plots of carbon and oxygen stable isotope data from several marine and
lacustrine deposits throughout the world (Modified from Hendry and
Kalin, 1997), and their comparison with data obtained in the Codó
Formation. This diagram shows that the isotopic composition of the
Codó Formation is in conformity with isotope data from lacustrine
limestones.
...................... 133
Fig. 5-11: Schematic diagram illustrating the main mechanisms that contribute to
modify δ
13
C
(%o PDB)
and δ
18
O
(%o PDB)
of the carbonates in a lake system.
...................... 134
Fig. 5-12: Summary of the depositional model proposed for the origin of the
Codó lake system in the eastern of the Grajaú Basin, illustrating the
close relationship of facies development, and thus the distribution of
oxygen and carbon isotopes, with alternation between syn-sedimentary
fault displacement and uplift. A) Offset of few meters along faults
displaced along the basin margins resulted in the creation of
accommodation space along subsiding areas, where the lake system
developed, giving rise to first-order cycle represented by unit 2. B)
Uplift contributed to decrease the lake level with the consequent spread
out of marginal deposits at the top of unit 2, culminating with lake
dryness and formation of a discontinuity surface with paleosols. C) Fault
reactivation resulted in a renewed phase of lake deepening, with
deposition of central and intermediate facies deposits recorded by unit 3.
D) Renewed uplift promoted the fall in lake level and widespread
formation of marginal deposits represented by pisoidal packstone to
grainstone and rhythmite, which culminated with lake exposure and soil
development at the top of unit 3. E) Fault reactivation with renewed
deposition of laminated argillites, recorded by unit 4. F) Increasing
stability led to progressive decrease in water level resulting from the
abandonment of the lacustrine deposition in the study area, with
subaerial exposure and formation of an unconformity at the top of the
Codó Formation.
...................... 136
xii
RESUMO
A Formação Codó, objeto deste estudo, corresponde a uma unidade geológica
neoaptiana bem conhecida por ser o único registro exposto de rochas desta idade na margem
equatorial brasileira. Esta formação, constituída de folhelhos betuminosos, calcários e
evaporitos, é particularmente bem exposta nas bordas sul e leste da Bacia de São Luís-Grajaú,
MA, áreas aqui investigadas com o intuito: 1. de aprimorar o entendimento do sistema
deposicional, discutindo-se a hipótese de formação em ambientes lacustres; e 2. reconstituir as
condições paleohidrológicas com base na integração de dados faciológicos, estratigráficos,
petrográficos e isotópicos (C, O, Sr e S). Os dados de campo confirmaram sistema lacustre
para a área de Codó, onde se desenvolveram lagos salinos, estáveis, bem estratificados, e com
períodos de fechamento, quando prevaleceram condições anóxicas acompanhadas pela
precipitação de sais em subambientes de lago central. Na região de Grajaú, por outro lado,
prevaleceram condições mais efêmeras, com desenvolvimento de complexo de sabkha/saline
pan, e precipitação de evaporitos principalmente nas margens do sistema, sob condições de
salinas marginais e de planícies lamosas.
Os estudos faciológico e estratigráfico mostraram, também, que a Formação Codó em
ambas as áreas estudadas está organizada em ciclos de arrasamento ascendente, que registram
a progradação de depósitos de lago marginal sobre os de lago central. Três categorias de ciclos
foram distinguidos, designados aqui de inferior, intermediário, e superior. Os ciclos de ordem
inferior, de espessuras variando entre milímetros a poucos centímetros, são formados por
depósitos com acamamentos constituídos de um dos seguintes arranjos litológicos: a) folhelho
negro betuminoso e evaporito; b) folhelho negro betuminoso e calcimudstone; c) folhelho
negro betuminoso e packstone-wackestone peloidal; d) folhelho cinza-esverdeado e
calcimudstone; e) folhelho cinza-esverdeado e packstone-wackestone peloidal; f) folhelho
cinza-esverdeado e packstone-wackestone ostracodal; ou g) grainstone-wackestone ostracodal
e/ou calcimudstone com tapetes criptomicrobiais e packstone ooidal-pisoidal. Estes ciclos são
atribuídos a depósitos sazonais, tendo em vista as suas espessuras regulares na escala
milimétrica, típicas de depósitos climaticamente controlados.
Os ciclos de ordem intermediária têm, em média, 1,7 m de espessura e são
subdivididos por ciclos completos e incompletos. Ciclos completos são compostos de
2
depósitos de lago central, que gradam para acima a depósitos de lago intermediário e marginal,
sendo representados por dois tipos: ciclos com depósitos de lago central, constituídos por
folhelhos e evaporitos (C1); e ciclos com depósitos de lago central, constituídos por folhelho
cinza esverdeado (C2). Ciclos incompletos são formados por sucessões faciológicas onde pelo
menos uma das associações de fácies está ausente. São também de dois tipos: ciclos com
depósitos de lago central e intermediário (I1); e ciclos com depósitos de fácies de lago
intermediário e central (I2).
Os ciclos de ordem superior medem, em média, 5,2 m de espessura e consistem em
quatro unidades deposicionais, limitadas por superfícies de descontinuidade, sendo
internamente constituídas por ciclos intermediários, tanto completos quanto incompletos, e de
distribuição variável em direção ao topo das seções. A unidade 1, mais inferior, está apenas
parcialmente exposta, com 2,7 m de espessura em média, sendo formada por um intervalo
constituído por ciclos I1 delgados. A unidade 2 tem, em média, 5,2 m de espessura e contem
todos os tipos de ciclos, principalmente ciclos completos. A unidade 3, com 2,6 m de
espessura em média, é constituída por quase 80% de ciclos I2. A unidade 4 apresenta 2,2 m de
espessura média, inclui exclusivamente ciclos incompletos, embora a maior parte desta
unidade tenha sido destruída pela formação do limite da seqüência aptiana.
A caracterização sedimentar detalhada e o padrão de empilhamento dos ciclos de
ordens intermediária e superior suportam gênese ligada à atividade tectônica sin-sedimentar.
Isto é particularmente sugerido pela alta variabilidade de fácies, pela extensão lateral limitada,
e por mudanças aleatórias na espessura e freqüência dos ciclos de ordem intermediária. Além
disto, os quatro ciclos de ordem superior são correlacionáveis com zonas estratigráficas
apresentando diferentes estilos de estruturas de deformação sin-sedimentar, atribuídos em
trabalhos anteriores a atividades sísmicas sin-deposicionais. Portanto, os vários episódios de
arrasamento do lago, registrados na Formação Codó pelos ciclos de ordem intermediária e
superior, são atribuídos a flutuações no nível de água do lago promovidas por pulsos sísmicos
contemporâneos à sedimentação.
A análise petrográfica dos evaporitos da Formação Codó permitiu que se definissem
melhor as histórias tanto deposicional do sistema lago-sabkha-saline pan, quanto pós-
deposicional. Sete morfologias de evaporitos foram reconhecidas: 1. gipso en chevron; 2.
gipso/anidrita nodular/lenticular; 3. gipso acicular; 4. gipso em mosaico; 5. gipso
3
brechóide/gipsarenito; 6. gipso/anidrita pseudo-nodular; e 7. gipso em roseta. A despeito desta
ampla variedade de fases, a abundância de gipso en chevron, gipso/anidrita nodular/lenticular
e gipso brechóide/gipsarenito, registra a boa preservação de formas primárias. Esta
interpretação é suportada pela associação destas morfologias de gipso com depósitos
mostrando acamamento horizontal de natureza cíclica, que são atribuídos a flutuações do nível
de base do lago, eventualmente culminadas com períodos de exposição subaérea. Mesmo o
gipso acicular e o gipso em mosaico, interpretados como produtos de substituição do gipso en
chevron e do gipso brechóide/gipsarenito, mostram características de formação autigência
ainda sob influência do ambiente deposicional. Fases de formação de gipso sob condições
diagenéticas mais profundas são registradas somente no gipso/anidrita pseudo-nodular,
atribuídos a mobilizações durante halocinese. Além disto, gipso em rosetas, que interceptam
todas as outras fases evaporíticas, têm também origem diagenética ligada a processos tardios
por interação com água subterrânea e/ou intemperismo superficial.
A constatação de forte influência deposicional registrada em, pelo menos, grande parte
das morfologias dos evaporitos da Formação Codó (i.e., gipso primário ou eodiagenético),
além da constatação de microfácies carbonáticas com poucas modificações diagenéticas,
motivaram a aplicação de métodos isotópicos com propósitos de reconstituição
paleoambiental. Os resultados obtidos mostram que ciclos de expansão/contração do sistema
deposicional em ambas as áreas estudadas são acompanhadas por variações significativas nos
valores isotópicos. A ampla dispersão de valores dos isótopos de Sr e S dentro de cada ciclo
deposicional reforça a interpretação petrográfica de que a diagênese não modificou a
assinatura geoquímica deposicional dos evaporitos, confirmando seu valor como ferramenta
paleoambiental. Além disto, origem não marinha para os evaporitos é sugerida pelas razões
87
Sr/
86
Sr, que variaram de 0,70782 a 0,70928, consideradas mais altas do que aquelas
esperadas para evaporitos oriundos da água do mar no Neo-Aptiano (entre 0,70720 e 0,70735).
O δ
34
S variou nas amostras estudadas de 16.12‰ to 17.89 ‰
(V-CDT)
na região de Codó,
mostrando-se também em total desarmonia com valores marinhos do Neo-Aptiano (i.e., entre
13‰ e 16‰
(V-CDT)
). Tanto Sr quanto S foram influenciados pelas características das fácies
deposicionais, de tal forma que, durante a expansão do sistema deposicional, os valores de
87
Sr/
86
Sr decresceram devido à inibição do
87
Sr liberado a partir de argilominerais pela
4
drenagem interna de planícies lamosas. Nos picos de expansão, os valores de
87
Sr/
86
Sr eram os
mais baixos, o que é relacionado à submergência de planícies lamosas e introdução de águas
depletadas em
87
Sr oriundo do intemperismo de calcários e evaporitos marinhos permianos a
neocomianos, tanto quanto basaltos triássicos a neocomianos.
Enquanto o estudo dos isótopos de Sr e S observou o comportamento destes nos
evaporitos da Formação Codó, análises de isótopos de C e O foram realizadas nos carbonatos
e também revelaram uma ampla distribuição isotópica, com valores exclusivamente baixos de
δ
13
C e δ
18
O, ou seja , entre –5.69‰ e –13.02‰
(PDB)
e –2.71‰ e –10.80‰
(PDB)
,
respectivamente. Adicionalmente, estas razões variam de acordo com ciclos de arrasamento
considerados neste trabalho como de origem tectônica e que, em geral, mostram razões de
δ
13
C e δ
18
O mais leves na base, onde predominam depósitos de lago central, e
progressivamente mais pesados em direção ao topo, onde depósitos de lago marginal são mais
expressivos. Também confirmando a assinatura deposicional, este comportamento leva a
propor um modelo isotópico controlado por eventos de sismicidade sin-sedimentar. Assim,
razões isotópicas com valores mais leves parecem estar relacionados com eventos de
inundação promovidos por subsidência, que resultou no desenvolvimento de um sistema de
lago perene. Razões isotópicas com valores mais pesados estariam relacionados a fases de lago
efêmero e seriam favorecidas pelo soerguimento e/ou aumento da estabilidade tectônica. Além
disso, os resultados mostram que sistemas de lagos fechados predominaram em pelo menos
parte do tempo de evolução desses depósitos, o que é indicado pela boa covariância positiva
(i.e., +0,42 e +0,43) entre o carbono e o oxigênio, embora fases de lago aberto também sejam
registradas pelos valores de covariância negativa (i.e., –0,36).
5
ABSTRACT
The Codó Formation is an important geological unit in Brazil, representing the only
record of Neoaptian rocks exposed along the Brazilian equatorial margin. This unit consists of
bituminous black shales, limestones and evaporites, which are particularly well represented in
the south and east margins of the São Luís-Grajaú Basin, around the towns of Codó and
Grajaú, State of Maranhão. These areas were investigated in order to: 1. improve the
depositional system, discussing the hypothesis that the Codó Formation was produced in a
lacustrine setting; and 2. reconstruct the paleohydrological conditions with basis on the
integration of facies, stratigraphy, petrography and isotope (C, O,Sr and S) data. Hence, the
field data presented herein confirmed a lacustrine system for the Codó area, where prevailed
stable, well-stratified, saline lakes characterized by periods of closure, anoxia and salt
precipitation in the central saline lakes. On the other hand, ephemeral conditions with
development of a sabkha/saline pan complex prevailed in the Grajaú area, where salts
precipitated mostly in the marginal portions of the system (i.e., marginal saline pans and
mudflats).
Studies focusing facies and stratigraphy also revealed that in both areas the Codó
Formation is arranged into several shallowing-upward cycles formed by progradation of
marginal into central lake deposits. Three types of cycles were distinguished, referred to here
as lower, intermediate and higher rank cycles. The lower rank cycles correspond to millimetric
interbeddings of: a) bituminous black shale and evaporite; b) bituminous black shale and
calcimudstone; c) bituminous black shale and peloidal wackestone-packstone; d) grey/green
shale and calcimudstone; e) grey/green shale and peloidal wackestone-packstone; f)
grey/green shale and ostracodal wackestone/grainstone; h) ostracodal wackestone/grainstone
and/or calcimudstone with cryptomicrobial mats and ooidal/pisoidal packstone. These are
attributed to seasonal deposition with basis on their regular nature forming very thin cycles
resembling varves.
The intermediate rank cycles average 1.7 m thick and are formed by complete and
incomplete cycles. Complete cycles show an upward transition from central to intermediate
and then marginal facies associations, and include two types: C1 cycles with central lake
deposits consisting of evaporites and black shales; and C2 cycles with central lake deposits
6
formed by gray/green shale. Incomplete cycles are those formed by successions lacking at
least one of the facies associations, consisting of either central and intermediate lake deposits
(cycles I1) or intermediate and marginal lake deposits (cycles I2).
The higher rank cycles average 5.2 m thick and consist of four depositional units,
which display shallowing-upward successions formed by both complete and incomplete,
intermediate rank cycles that vary their distribution upward in the section, and are bounded by
sharp surfaces. Unit 1, the lowermost one, averages 2.7 m in thickness, being entirely
composed by thin I1 cycles. Unit 2 averages 5.2 m thick, and displays all of the
aforementioned intermediate cycles, especially complete ones. Unit 3, averaging 2.6 m thick,
consists of 80% of cycles I2. Finally, unit 4, which averages 2.2 m in thickness, displays only
incomplete cycles, though its uppermost part was not preserved due to erosion during the
development of the Aptian sequence boundary.
The detailed sedimentological characterization and the stratal stacking patterns of the
intermediate and higher rank cycles support a genesis linked to syn-sedimentary tectonic
activity, particularly suggested by high facies variability, limited lateral extension, as well as
frequent and random thickness changes of the intermediate-rank cycles. Additionally, the
matching between the four higher rank cycles with four stratigraphic zones having different
styles of soft-sediment deformation structures previously described in the literature as
resulting from seismic activities, is a further argument to corroborate this interpretation.
Therefore, the several episodes of lake shallowing recorded in the intermediate and higher
rank cycles of the Codó Formation are attributed to fluctuations in the lake water level,
triggered by seismic pulses alternating with sediment deposition.
The petrographic analysis of the evaporites from the Codó Formation allowed to better
defining both the lake-sabkha-saline pan depositional system and the post-depostional
histories. Seven evaporite morphologies were recognized: 1. chevron (selenite) gypsum; 2.
nodular/lensoidal gypsum/anhydrite; 3. acicular gypsum; 4. mosaic gypsum; 5. brecciated
gypsum/gypsarenite; 6. pseudo-nodular anhydrite/gypsum; and 7. rosettes of gypsum. Despite
of this large variety of evaporite phases, the chevron gypsum, the nodular/lensoidal
gypsum/anhydrite and the brecciated gypsum/gypsarenite record the preservation of primary
features. The association of these morphologies with deposits displaying cyclic horizontal
bedding, attributed to lake level fluctuations eventually culminated with subaerial exposure,
7
reinforces this interpretation. Even acicular gypsum and mosaic gypsum, which replaced the
chevron and brecciated gypsum/gypsarenite, respectively, formed under the influence of the
depositional surface. Burial phases of gypsum are only recorded in the pseudo-nodular
anhydrite/gypsum, attributed to salt mobilization induced by halokinesis. In addition, rosettes
of gypsum, which crosscut the other evaporite morphologies, diagenetic in origin, have
probably formed as the latest evaporite phase of the study area, under the influence
groundwater and/or surface weathering.
In the present research, isotope studies aiming paleoenvironmental purposes were
motivated by both confirmation of strong depositional influence for at least great part of the
evaporites from the Codó Formation (i.e., primary and eodiagenetic gypsum), and the low
diagenetic modification recorded for the limestones. Results of these approaches show that
expansion/contraction cycles in both studied areas were accompanied by significant changes
in isotope values. The wide dispersion of Sr and S isotope data within individual depositional
cycles reinforces the lack of significant diagenetic modification as suggested by the
petrographic analysis, and confirms the utility of these isotopes as environmental tools.
Additionally, a non-marine brine source is suggested by
87
Sr/
86
Sr ratios ranging from 0.707824
to 0.709280, which are higher than those from late Aptian seawater (i.e., between 0.70720 and
0.70735). The δ
34
S varies from 16.12 to 17.89 %
o(
V-CDT)
in the Codó area, which is also in
disagreement with late Aptian marine values (ranging from 13 to 16 %
o
(V-CDT)
). Both
geochemical tracers were influenced by facies characteristics, and thus a model is provided
where expansion of saline pan/lake systems led to decreasing
87
Sr/
86
Sr values due to the
inhibition of the
87
Sr from clay minerals originated during the internal draining of mudflats.
During expansion peaks, the
87
Sr/
86
Sr values were lower due to submergence of mud flats and
introduction of external
87
Sr-depleted waters related to weathering of Permian to Neocomian
marine limestones and evaporites, as well as Triassic to Neocomian basaltic rocks.
Furthermore, the sulphur isotope values decrease in the southern margin of the basin from
14.79 to 15.60 %
o
(V-CDT)
probably due to increased evaporation in shallower water settings.
While the studies of Sr and S isotopes emphasized the evaporites of the Codó
Formation, the analysis of C and O isotopes were carried out on the carbonates. The data
revealed a wide distribution of dominantly low δ
13
C and δ
18
O values, ranging from –5.69‰ to
8
–13.02‰ and from –2.71‰ to –10.80‰, respectively. It was also observed that these ratios
vary according to seismically-induced shallowing-upward cycles, in general becoming lighter
in their bases, where central lake deposits dominate, and progressively heavier upward, where
marginal lake deposits are more widespread. In addition to confirm a depositional signature
for the analysed samples, this behavior led to introduce a seismic-induced isotope model.
Hence, lighter isotope ratios appear to be related with flooding events promoted by
subsidence, which resulted in the development of a perennial lake system, while heavier
isotope values are related to ephemeral lake phases favored through uplift and/or increased
stability. Furthermore, the results show that a closed lake system dominated, as indicated by
the overall good positive covariance (i.e., +0.42 to +0.43) between the carbon and oxygen
isotopes, though open phases are also recorded by negative covariance values of –0.36.
9
APRESENTAÇÃO
Esta tese acha-se organizada sob a forma de quatro artigos, já submetidos ou aprovados
em periódicos científicos indexados na categoria internacional A ou B. Adicionam-se a estes,
um capítulo introdutório de integração e um capítulo final com uma síntese das principais
conclusões advindas desta pesquisa.
O capítulo 1 tem por meta fazer uma introdução, com a apresentação da área de estudo,
problemática que motivou a pesquisa, objetivos, métodos, bem como um breve sumário dos
princípios teóricos relacionados com a aplicabilidade da geoquímica isotópica na reconstrução
paleohidrológica de sistemas lacustres, tema de abordagem de dois dos capítulos desta tese. O
capítulo 2 trata da descrição litológica e caracterização das diferentes categorias de ciclos
sedimentares reconhecidos na Formação Codó, com discussão dos possíveis fatores
causadores. O capítulo 3 contém a descrição petrográfica dos evaporitos da unidade
investigada, com o objetivo de: 1. reconstituir sua paragênese; e 2. distingüir entre fases
evaporíticas formadas sob influência do ambiente deposicional e aquelas geradas por
processos diagenéticos, possibilitando a escolha de amostras com melhor potencial para os
estudos isotópicos de Sr e S. O capítulo 4 apresenta os resultados da análise isotópica de Sr e S
dos evaporitos da Formação Codó, com o objetivo de fornecer elementos adicionais que
contribuam na determinação da origem da salmora. O capítulo 5 contém os resultados dos
estudos isotópicos de C e O das rochas calcáreas intercaladas aos evaporitos e folhelhos da
unidade estudada, tendo por objetivo de contribuir para no melhor entendimento do regime
hidrológico vigente durante a deposição. Por fim, o capítulo 6 compreende as principais
conclusões sintetizadas neste estudo.
A grande contribuição desta tese foi a consolidação do conhecimento a respeito da
Formação Codó, fazendo uso do estudo integrado combinando análise faciológica com estudos
petrográficos e métodos isotópicas enfocando Sr, S, C e O, como ferramenta chave para a
caracterização mais precisa de sistemas deposicionais lacustres e entendimento de sua
paleohidrologia.
10
1. INTRODUÇÃO
1.1. ARCABOUÇO GEOLÓGICO E PROBLEMÁTICA
A configuração atual da costa brasileira e as principais feições estruturais que
controlam a dinâmica geológica da margem continental brasileira começaram a ser
estabelecidos durante o Mesozóico, ao mesmo tempo em que as principais reservas de petróleo
do mundo se formavam. No Eocretáceo, a região nordeste do Brasil formava o último elo de
ligação entre a América do Sul e África, o que se estendeu até o período Aptiano, o qual
marca, assim, a separação definitiva entre estes dois continentes e a livre circulação das águas
entre o Atlântico Norte e Sul (Szatmari et al., 1987; Zanotto & Szatmari, 1987; Feijó, 1996).
Este processo de separação foi desencadeado por um complexo sistema de rifteamento,
marcado inicialmente pela implantação de diversos sistemas deposicionais lacustres ao longo
de toda margem continental brasileira (inclusive continente adentro), e que culminou com
transgressão marinha. A implantação dos lagos foi assíncrona ao longo da margem continental
brasileira, ocorrendo genericamente de sul para norte. Desta forma, as bacias das regiões sul e
sudeste brasileiras começaram a apresentar lagos representativos da fase pré-rifte ainda no
Neocomiano, enquanto que nas bacias da região equatorial, os lagos são registrados apenas a
partir do Aptiano.
Na Bacia de São Luís-Grajaú, localizada em grande parte do Estado do Maranhão, a
sedimentação aptiana é bastante expressiva, merecendo destaque por ser a única bacia da
11
margem equatorial onde a sucessão aptiana pode ser investigada na escala de afloramentos.
Esta bacia foi formada durante a separação dos continentes sul-americano e africano, nos
últimos estágios da fragmentação do Gondwana, no Cretáceo Inferior, por processos
transtensivos associados com o estabelecimento do Sistema de Gráben Gurupi (Góes &
Rossetti, 2001). Este sistema interliga as sub-bacias de Grajaú, São Luís e Caetés numa única
feição estrutural tipo rift (Fig. 1-1A). A Bacia do Grajaú, por si só, abrange uma área de
aproximadamente 130.000 km
2
e atinge espessuras de até 900 m, e, como continuidade ao sul
da Bacia de São Luís, tem sido redenominada por Bacia de São Luís-Grajaú (Góes & Rossetti
2001).
O embasamento da Bacia de São Luís-Grajaú é composto de rochas da Bacia do
Parnaíba, de cinturão de dobramentos Gurupi e Tocantins-Araguaia e do Cráton São Luís
(Góes, 1995). O preenchimento sedimentar é representado por três megasseqüências
deposicionais, designadas S1, S2 e S3 (cf. Rossetti, 2001; Fig. 1-1B). A seqüência S1 tem
idade aptiana superior a albiana inferior e inclui as formações Grajaú e Codó em sua porção
basal. A Formação Grajaú consiste de arenitos esbranquiçados, quartzosos, com granulação
fina, e conglomerados quartzosos de origem flúvio-deltaica com contribuição eólica (e.g.,
Rezende, 1997). A Formação Codó constitui-se em folhelhos betuminosos, anidritas, calcários
e arenitos lacustres (Campbell et al., 1949; Aranha et al., 1991). A seqüência S2 depositou-se
durante o Eo/Mesoalbiano, correspondendo à Unidade Indiferenciada descrita na Bacia de São
Luís (cf. Rossetti & Truckenbrodt, 1997), cujas características faciológicas são inferidas
principalmente a partir de dados de subsuperfície e que levam a interpretá-la como registro de
ambientes fluvio-deltaico a marinho marginal (Rossetti, 2001). A seqüência S3 depositou-se
entre o Neoalbiano e um tempo incerto no Cretáceo Superior/Terciário, sendo representada
por sucessões sedimentares atribuídas a sistemas estuarinos de vales incisos, cujas duas
sucessões mais superiores desta seqüência acham-se bem expostas na Bacia de São Luís e
correspondem às formações Alcântara e Cujupe (Rossetti, 2001). A Formação Alcântara
(Albiano superior a Cenomaniano) constitui-se de arenitos bem estruturados, de coloração
característica marrom-chocolate, granulometria fina a média e cimentação de calcita.
Subordinadamente, ocorrem pelitos, conglomerados e calcários. Estes depósitos foram gerados
principalmente em ambiente de ilha-barreira sujeito à ação de ondas de tempestade (Rossetti,
1998). A Formação Cujupe (Neocretáceo ?) é constituída por arenitos cauliníticos, de
12
granulometria fina a muito fina, bem selecionados, cores branca, rósea ou amarelada, com
estruturas geradas dominantemente por correntes de maré (Rossetti, 1998). Sobrepostos a estas
unidades, ocorrem depósitos terciários atribuídos às formações Pirabas, Barreiras e depósitos
Pós-Barreiras (Rossetti & Truckenbrodt, 1997).
A Formação Codó, objeto deste estudo, representa os depósitos neoaptianos da Bacia
de São Luís-Grajaú. Resgata-se a definição original de Lisboa (1914) e Campbell et al. (1949)
13
para o termo Formação Codó que aqui se emprega para referenciar folhelhos betuminosos
intercalados a evaporitos e calcários, aflorantes nas regiões em torno das cidades de Codó e
Grajaú, no Estado do Maranhão.
A Formação Codó desperta os mais diversos interesses econômicos e científicos.
Economicamente, esta unidade pode representar a geradora da bacia, dado o seu alto teor de
carbono orgânico associado a folhelhos betuminosos (Fernandes & Della Piazza, 1978). Além
disto, calcário, gesso e argilas nobres têm sido exaustivamente explorados, devido à alta
qualidade para uso industrial (Rezende, 1997). Cientificamente, o posicionamento
estratigráfico da Formação Codó no período Aptiano a coloca como uma chave para o
entendimento da paleogeografia, paleoclima e paleohidrologia do Cretáceo e ajuda a compor
um quadro mais completo a respeito da implantação dos sistemas lacustres ao longo da
margem continental brasileira. O conteúdo fóssil desta unidade, rico em elementos delicados e
raros do Aptiano como, por exemplo, insetos, plantas e peixes, bem preservados em ritmitos
de calcário e folhelho, ainda são atrativos importantes da Formação Codó (Pinto & Ornellas,
1974; Duarte & Silva-Santos, 1993; Popov & Pinto, 2000).
Dado o interesse despertado sob os mais diversos aspectos, vários estudos teceram
considerações sobre os ambientes deposicionais da Formação Codó, com a proposição de
modelos mais ou menos concordantes entre si (e.g., Aranha et al., 1991; Batista, 1992; Paz,
2000; Paz & Rossetti, 2001). Porém, a caracterização precisa do sistema deposicional (p.e.,
Batista, 1992; Paz, 2000) e, principalmente, a origem dos evaporitos (p.e., Rodrigues, 1995;
Rossetti et al., 2000) são temas que permaneceram por ser mais sistematicamente abordados.
Além disto, estudos petrográficos detalhados dos evaporitos nunca foram abordados
anteriormente, tendo merecido apenas uma caracterização geral em dissertação de mestrado
(i.e., Paz, 2000). Da mesma forma, apesar da abundância de calcários e evaporitos em
afloramentos, a Formação Codó havia sido investigada em estudos isotópicos somente com
base em dados de testemunho (Rodrigues & Takaki, 1993; Rodrigues, 1995). Por fim, a
organização estratigráfica desta unidade, e a caracterização e entendimento da origem de sua
natureza cíclica, são temas que foram somente superficialmente abordados em trabalhos
prévios (p.e., Paz 2000; Gonçalves, 2004).
Assim, a situação existente no início desta tese era de um panorama com informações
ainda inadequadas para sustentar solidamente o modelo deposicional dos depósitos aptianos da
14
Bacia de São Luís-Grajaú. Por outro lado, os estudos preliminares disponíveis sugeriam que
estes estratos tinham grande potencial para servir como modelo de integração de dados
faciológicos, petrográficos e isotópicos, com o intuito de reconstituição de seu sistema
deposicional e entendimento das condições hidrológicas dominantes durante a deposição.
1.2. OBJETIVOS
O objetivo desta tese foi aprimorar (ou redefinir) o modelo paleoambiental da
Formação Codó nas regiões de Codó e Grajaú, bordas leste e sul da Bacia de São Luís-Grajaú,
respectivamente, integrando-se informações faciológicas, petrográficas e isotópicas (C, O, Sr e
S), como ferramenta de análise do sistema deposicional, bem como das condições
hidrodinâmicas dominantes durante a sedimentação. O surgimento de novas exposições em
minas a céu aberto na região em torno da cidade de Grajaú, ainda não documentada na
literatura, mostravam grande potencial de se incrementar as informações de campo. Uma vez
superado este obstáculo (Rossetti et al., 2004), esta região tornou-se promissora para: 1) testar
o modelo deposicional lacustre anteriormente adotado para a região de Codó, confrontando-o
diante de novas evidências; 2) reconstituir os processos de formação dos evaporitos da região
de Grajaú, comparando-os com a área de Codó; e 3) auxiliar no entendimento do significado
das variações do sinal isotópico dos depósitos das áreas de estudo comparando-se ambientes
deposicionais em duas áreas distintas da bacia.
1.3. ÁREA DE ESTUDO
As áreas estudadas neste trabalho localizam-se nas proximidades das cidades de Grajaú
e Codó, Estado do Maranhão (Fig. 1-1C), onde ocorrem os melhores registros aflorantes da
Formação Codó na Bacia de São Luís-Grajaú. Na área de Codó, esta unidade aflora em 3
minas a céu aberto, uma na BR 316, ao lado da Fábrica da Itapicuru Agroindustrial S/A, a 5
Km da entrada para a cidade de Codó, uma na estrada que leva à sede do município de Codó, a
15 Km após o cruzamento com a BR 316, e outra na localidade de Santo Amaro (Fig. 1-1D).
A área de Grajaú possui inúmeras minas rudimentares, sendo que para este estudo foram
utilizados 7 perfis, correspondentes às localidades de Chorado, Barreirinho, Olho D´Água,
Olaria (ou Cerâmica), Faz. São João Oneida, Faz. Pau Ferrado e Faz. Fortaleza, além de um
15
ponto à beira da estrada, no Km 22 da Rodovia Transmaranhense (MA-026). Estas sessões
foram escolhidas com base na boa preservação dos estratos. Em geral, os perfis apresentam
espessuras em torno de 10 m, com continuidade lateral de até 400 m, como no caso da mina de
gesso a 5 Km da entrada para a cidade de Codó da Fábrica Itapicuru Agroindustrial S/A.
O acesso às áreas de estudo é relativamente fácil, sendo feito pelas rodovias BR-316,
BR-226 e MA-026. Esta última, entretanto, é intransitável em certos pontos como, por
exemplo, entre os municípios de Arame e Entroncamento. Embora a infraestutura seja
precária, é possível encontrar boa hospedagem e até ótima uma alimentação tanto em Codó
quanto em Grajaú.
1.4. MÉTODOS
A fim de alcançar o objetivo proposto nesta tese foram adotados os seguintes métodos:
análise faciológica/estratigráfica, análise petrográfica e análise geoquímica (elementos
principais, traços e isótopos).
A análise faciológica e estratigráfica foi feita com base em dados de afloramentos,
tendo abrangindo mapeamento vertical e horizontal das fácies deposicionais, observando-se as
seguintes feições: litologia, cor, textura e estruturas deposicionais, geometria da fácies e
conteúdo paleontológico (cf. Tucker, 2003). Os dados foram registrados pela confecção de
perfis litoestratigráficos, mapeamento de superfícies-chaves e documentação fotográfica de
seções panorâmicas. Adicionalmente, procedeu-se com a coleta orientada de amostras para
estudos de laboratório. Assim, tornou-se possível detalhar a geometria das fácies e reconhecer
superfícies de importância estratigráfica. Esta metodologia tem sido adotada com sucesso em
diversos trabalhos realizados na Bacia de São Luís-Grajaú (p.e., Anaisse, 1999; Lima &
Rossetti, 1999; Paz, 2000; Rossetti & Góes, 2000; Rossetti et al., 2000; Paz & Rossetti, 2001).
A análise petrográfica consistiu no estudo de 220 seções delgadas de carbonatos e
evaporitos. A caracterização petrográfica das lâminas foi feita pela: 1) determinação da
composição dos carbonatos, através do seu tingimento com alizarina vermelha-S, ferricianeto
de potássio e ácido clorídrico diluído em água destilada; 2) identificação e estimativa dos
constituintes da seção delgada; e 3) determinação do relacionamento entre as diversas fases
mineralógicas. Como ferramenta auxiliar à microscopia óptica convencional (i.e., microscópio
16
petrográfico), aplicação de microscópio eletrônico de varredura (MEV) e espectrometria de
energia dispersiva de raios-X (EDS), do Laboratório de Microscopia do Museu Goeldi,
ajudaram a solucionar problemas de definição de aspectos morfológicos, composicionais e
mineralógicos observados nas amostras e seções delgadas estudadas.
A análise geoquímica elementar consistiu na determinação de Ca, Mg, Fe, Mn, Sr e Na
das amostras de rocha total de carbonato, a fim de investigar possíveis modificações
diagenéticas. Estes elementos são os mais importantes na estrutura da calcita tanto marinha
quanto não marinha e, considerando que os teores destes elementos mantiveram-se mais ou
menos constantes através do Fanerozóico (cf. Holland, 1978), uma comparação adicional com
valores comumente esperados em águas de formação poderia detectar potenciais influências
diagenéticas. O procedimento consistiu na secagem de 1,5 g de 70 amostras a 1000
o
C por
duas horas, sendo fundidas com tetraborato de lítio e fluoreto de lítio, fornecendo pastilhas
fundidas que foram analisadas pelo espectrômetro de fluorescência de Raios-X do Laboratório
de Isótopos Estáveis da Universidade Federal de Pernambuco (LABISE/UFPE), aos cuidados
da Profa. Dra. Valderez Ferreira.
A análise geoquímica isotópica consistiu na determinação das razões isotópicas de C e
O em carbonatos e de S e Sr em evaporitos da Formação Codó, visando testar variações no
regime hidrológico do ambiente deposicional, que ajudassem a entender e compor um quadro
mais refinado do sistema deposicional. Esta caracterização tem se mostrado bem sucedida em
diversos trabalhos visando-se estabelecer arcabouços hidrológicos e climatológicos, tanto em
ambientes modernos, quanto em depósitos de ambientes antigos (p.e., Chivas et al., 1991; Bird
et al., 1991; Bellanca et al., 1992; Camoin et al., 1997; Chaves & Sial, 1998; Li & Ku, 1997;
Neumann, 1999; Playà et al., 2000; Ortí e Rosell, 2000; Lu et al., 2001; Wignall & Tretchett,
2002; Krull et al., 2004).
Maiores detalhamento sobre os procedimentos utilizados durante as análises
geoquímicas encontram-se inseridos nos capítulos respectivos.
17
1.5. GEOQUÍMICA ISOTÓPICA
1.5.1. Fundamentação teórica
A geoquímica isotópica como ferramenta para auxiliar na interpretação paleoambiental
de depósitos lacustres é uma abordagem relativamente nova (Stuiver, 1970). Este tema
representa o cerne dos capítulos 4 e 5 desta tese e prima pelo ineditismo deste tipo de
abordagem nos depósitos da área de estudo. Merece, portanto, que seus fundamentos teóricos
sejam revisados em mais detalhes ao longo deste item.
Isótopos são átomos que possuem o mesmo número atômico, mas que apresentam
diferentes números de massa, conseqüentemente de quantidades diferentes de nêutrons em seu
núcleo. Na natureza, os elementos químicos são formados por uma mistura de, pelo menos,
dois isótopos, onde um deles em quantidade abundante e o outro, subordinada. Os isótopos
podem também ser considerados estáveis ou instáveis, sendo que a estabilidade é um fator
relativo ao tempo de decaimento radioativo muito longo, da ordem de 10
18
anos para os
isótopos mais leves (massa atômica 34). Como ocupam a mesma posição na tabela
periódica, formam um mesmo elemento químico e compartilham das mesmas propriedades
químicas, mas por causa do número de massa distinto, apresentam propriedades físico-
químicas (p.e., densidade, temperatura de congelamento e de ebulição, viscosidade, pressão de
vapor) sutilmente distintas entre si. Esta distinção nas propriedades físico-químicas dos
isótopos que constituem um mesmo elemento químico é chamada efeito isotópico.
Por causa do efeito isotópico, durante eventos físico-químicos de transformação ou
transição dos elementos, estes acabam por se tornar mais enriquecidos ou mais empobrecidos
(=depletados) em alguns dos isótopos que os constituem. Este efeito é conhecido por
fracionamento isotópico, que ocorre fundamentalmente por troca isotópica ou por efeito
cinético.
O fracionamento por troca isotópica diz respeito aos processos em que não ocorrem
mudanças ordinárias no sistema químico, mas na distribuição dos isótopos entre diferentes
substâncias químicas, fases, ou ainda, moléculas (cf. Hoefs, 1980). Simplificadamente, a troca
isotópica pode ser entendida como um caso especial de equilíbrio químico:
18
aA
1
+ bB
2
aA
2
+ bB
1
onde A e B representam os elementos que contêm as moléculas leve ou pesada 1 ou 2. O
fracionamento por efeito cinético diz respeito àquelas reações químicas unidirecionais, em que
as taxas de reação entre os reagentes e os produtos são sensíveis às diferenças de massas dos
átomos. Quando o fracionamento ocorre em equilíbrio isotópico (i.e., sem modificações no
sistema físico-químico durante a reação), os isótopos são redistribuídos proporcionalmente de
acordo com um fator de fracionamento da reação. O fator de fracionamento α entre duas
substâncias A e B é a razão R
A
do número de quaisquer dois isótopos na substância A dividida
pela razão R
B
deste dois isótopos na substância B, de tal maneira que:
α
A-B
= R
A
/R
B
O fator de fracionamento em equilíbrio isotópico é dependente da temperatura,
experimentalmente determinável e diretamente relacionado à constante de equilíbrio da
reação. Em termos absolutos, os valores das razões isotópicas são números muito pequenos e,
por convenção, são expressos em notação δ, que expressa o desvio (em partes por mil) da
razão isotópica medida em relação a uma amostra padrão, de acordo com a expressão:
δ = ((R
amostra estudada
– R
padrão
) / R
padrão
) * 1000
Existem diferentes padrões para cada isótopo estudado. Entre os principais, estão o
PDB (PeeDee Belemnite), usado para expressar razões isotópicas de C e/ou O em amostras de
carbonatos, o SMOW (Standard Mean Ocean Water), usado para expressar razões isotópicas
de H e/ou O em amostras de água, silicatos e, ainda, carbonatitos, e o CDT (Canyon Diablo
Troilite – um composto de FeS oriundo de meteorito), usado para expressar as razões
isotópicas do S em amostras dos mais diversos materiais. A principal fornecedora destes
padrões ao redor do mundo são a Agência Internacional de Energia Atômica (Áustria) e o
National Bureau Standards (EUA). Alguns destes padrões já se exauriram ou estão muito
perto disso. Por isso, tem sido comum a adoção de padrões alternativos àqueles originais,
19
reconhecidos pelo acréscimo da partícula V- (
Viena-)
ao nome do padrão (por exemplo, V-
PDB, V-SMOW e V-CDT).
1.5.2. Aplicação em depósitos lacustres
O princípio do estudo de isótopos estáveis em depósitos lacustres é que tanto
carbonatos quanto evaporitos precipitam-se em equilíbrio isotópico com a água do lago
(Stuiver, 1970; Talbor, 1990; Talbot & Kelts, 1990). Os isótopos estáveis leves mais utilizados
em estudos de reconstituição paleoambiental de depósitos lacustres são carbono, oxigênio e
enxofre.
O carbono possui três isótopos (i.e.,
12
C,
13
C e
14
C). O
12
C é o mais abundante (~99%),
sendo que em depósitos muito antigos (acima de 50.000 anos), todo
14
C já sofreu decaimento
radioativo para as formas estáveis. O carbono sofre fracionamento por troca isotópica durante
a transformação do gás carbônico da atmosfera em bicarbonato dissolvido e precipitação do
bicarbonato em carbonato dentro do lago. O fracionamento por efeito cinético ocorre durante a
fotossíntese, que favorece o acúmulo de
12
C mais “leve” na matéria orgânica do lago e torna a
razão isotópica δ
13
C residual na água do lago mais “pesada”. Em sistemas lacustres, a variação
da composição isotópica de C depende da produtividade do lago, da estratificação da lâmina
de água, do tipo de degradação bacteriana, da geologia da área fonte, do tipo de vegetação da
área de captação da bacia e da troca de gás carbônico na interface água-ar (Stuiver, 1970;
Hoefs, 1980; McKenzie, 1985; Herczeg, 1988; Li & Ku, 1997; Talbot & Lærdal, 2000;
Wignall & Twitchett, 2002). Durante períodos em que o nível do lago está mais alto (Fig. 1-2),
processos como maior respiração do lago, elevado consumo de matéria orgânica, alto
soterramento de carbono orgânico, maior influxo de águas fluviais teoricamente com
bicarbonato mais “leve”, baixa troca de gás carbônico com a atmosfera e menor participação
do gás carbônico metanogênico na formação do carbono orgânico total dissolvido na água do
lago, favorecem que o δ
13
C da água do lago seja mais “depletado” do que em períodos em que
o nível do lago está mais baixo (cf. Talbot, 1990; Talbot & Kelts, 1990; Guzzo, 1997).
20
O oxigênio também possui três isótopos (
16
O,
17
O e
18
O). O
16
O é o mais abundante,
respondendo por mais de 99% do oxigênio disponível na natureza. Em geral, a razão mais
usada é
18
O/
16
O por causa das maiores abundâncias em relação ao
17
O e da maior diferença de
21
massa entre eles. O oxigênio sofre fracionamento por troca isotópica, durante as fases do ciclo
hidrológico (Fig. 1-3A), quando se torna mais pesado nas fases residuais, e por efeito cinético,
durante a fotossíntese, quando provavelmente contribui significaivamente com valores
bastante negativos produzidos durante a respiração (cf. Hoefs, 1980, p.122). Embora a
temperatura seja o fator mais importante no fracionamento do oxigênio, a variação da
composição isotópica do O nas águas do lago refletem preferencialmente variações no balanço
hídrico da bacia (cf. Lister et al., 1991; Fig. 1-3B). Durante períodos de nível do lago mais
alto, o δ
18
O tende a se tornar mais negativo dado o aporte de águas depletadas em
18
O. O
período de rebaixamento do nível do lago que se segue favorece a evaporação de
16
O e torna a
água residual do lago mais enriquecida em
18
O. Depósitos cretáceos lacustres da região do
Recôncavo na Bahia que registram este comportamento isotópico são interpretados neste
sentido (p.e., Guzzo, 1997).
O enxofre possui quatro isótopos (i.e.,
32
S,
33
S,
34
S e
36
S), dos quais os mais abundantes
são o
32
S (~ 95%) e o
34
S (~ 4%). O fracionamento de enxofre em sedimentos ocorre quase
exclusivamente por redução bacteriana de sulfato para H
2
S (Fig. 1-4). A troca isotópica
durante processos de redox provavelmente desempenha um papel importante ainda não bem
definido no registro antigo (Lu et al., 2001). Desta maneira, períodos em que a estratificação
da coluna d’água no lago é mantida por um longo período favorecem a proliferação de
22
bactérias sulfato redutoras, que por sua vez promovem o enriquecimento em
34
S do sulfato
residual na água do lago.
Fig. 1-4: Ciclo resumido do enxofre, enfatizando a passagem do sulfato para sulfeto de hidrogênio por respiração
anaeróbica de bactérias sulfato-redutoras (modificado de Hoefs, 1980).
O
87
Sr não é um isótopo estável leve, uma vez que surge como produto de decaimento
do
87
Rb. Entretanto, a análise da razão
87
Sr/
86
Sr tem acompanhado diversos trabalhos
orientados por isótopos estáveis, uma vez que também apresenta um tempo de meia-vida
bastante longo, comparável ao dos isótopos estáveis leves (Holser et al., 1996). Ao contrário
dos isótopos leves, o Sr não sofre fracionamento no ambiente sedimentar e é facilmente
capturado pela cristalização de evaporitos, que reflete exatamente a composição isotópica da
água da qual se precipitou. Outro aspecto curioso, é que as razões dos isótopos de Sr são
escritas diretamente de um isótopo em relação ao outro, sem necessidade de padrões de
comparação. Os principais meios de alteração da razão
87
Sr/
86
Sr numa bacia deposicional são
através de maior influxo fluvial, que favorece o aporte de Sr oriundo do intemperismo de
rochas continentais, em geral produzindo valores superiores a 0,707, e a “troca” de Sr com
basalto de crosta juvenil, cujas razões
87
Sr/
86
Sr são iguais ou inferiores a 0,706, mas que só
ocorre em ambiente hidrotermal marinho abaixo das cordilheiras meso-oceânicas (Fig. 1-5;
Holser et al., 1996). Em sistemas continentais, portanto, a variação na razão
87
Sr/
86
Sr está
diretamente ligada a variações no aporte fluvial da bacia de captação do sistema (cf. Rhodes et
al. 2002).
Este quadro mostra que a deposição de carbonatos e evaporitos em lagos reflete
fortemente as condições paleohidrológicas atuantes durante a deposição. As variações na
composição isotópica destes sedimentos revelam mudanças no balanço físico- e/ou bioquímico
23
na água do lago e/ou na água vadosa da partes emersas do ambiente lacustre. Eventualmente,
estas variações isotópicas refletem indiretamente perturbações de ordem climática e, quiçá,
tectônica.
24
REFERÊNCIAS
ANAISSE, J., Jr. 1999. Fácies costeiras dos Depósitos Itapecuru (Cretáceo), região de
Açailância, Bacia do Grajaú. Belém, Univ. Fed. Pará, 86p. (Dissertação de Mestrado)
ARANHA, L.G.; LIMA, H.P.; MAKINO, R.K.; SOUZA, J.M. 1991. Origem e evolução das
bacias de Bragança-Viseu, São Luís e Ilha Nova. In: E.J. MILANI & G.P. RAJA
GABAGLIA (eds.) Origem e evolução das bacias sedimentares. Rio de Janeiro,
PETROBRÁS, p.221-234.
BATISTA, A. M. 1992. Caracterização paleoambiental dos sedimentos Codó-Grajaú, Bacia
de São Luís (MA). Belém, Univ. Fed. Pará, 102 p. (Dissertação de Mestrado)
BELLANCA, A., CALVO, J.P., CENSI, P., NERI, R. & POZO, M. 1992. Recongnition of
lake-level changes in Miocene lacustrine units, Madrid Basin, Spain. Evidence from facies
analysis, isotope geochemistry and clay mineralogy. Sediment. Geol., 76: 135-153.
BIRD, M.I., CHIVAS, A.R., RADNELL, C.J., BURTON, H.R. 1991. Sedimentological and
stable isotope evolution of lakes in the Vestfold Hills, Antarctica. Palaeogeog.,
Palaeoclim., Palaeoecol., 84: 109-130.
CAMOIN, G., CASANOVA, J., ROUCHY, J.M., BLANC-VALLERON, M.M,
DECONINCK, J.F. 1997. Environmental controls on perenial and ephemeral carbonate
lakes: the central palaeo-Andean Basin od Bolivia during Late Cretaceous to early Tertiary
times. Sedimentary Geology, 113: 1-26.
CAMPBELL, D.F., ALMEIDA, L.A., SILVA, S.O. 1949. Relatório preliminar sobre a
geologia da Bacia do Maranhão. Boletim do Conselho Nacional do Petróleo, 1: 1-160.
CHAVES, N.S. & SIAL, A.N. 1998. Mixed oceanic and freshwater depositional conditions
for beachrocks of Northeast Brazil: evidence from carbon and oxygen isotopes. Int. Geol.
Rev., 40: 748-754
CHIVAS, A.R., ANDREW, A..S., LYONS, W.B., BIRD, M.I., DONELLY, T.H. 1991.
Isotopic constraints on the origin of salts in Australian playas. 1. Sulphur. Palaeogeog.,
Palaeoclim., Palaeoecol., 84: 309-332.
DUARTE, L.S. & SILVA-SANTOS, R. 1993. Plant and megafossils of the Codó Formation,
Parnaiba Basin, NE Brazil. Cretaceous Research, 14: 735-746.
25
EHLERINGER, J.R. & OSMOND, C.B. 1989. Stable isotopes. In: R.W. PEARCY, J.
EHLERINGER, H.A. MOONEY, P.W. RUNDD (Eds) Plant Physiological Ecology: Field
Methods and Instrumentation. London, Chapman & Hall, p. 281-300.
FEIJÓ, F.J. 1996. O início da livre circulação das águas do Oceano Atlântico. Bol. Geoc.
Petrobrás, 10: 157-164
FERNANDES, G. & DELLA PIAZZA, H. 1978. O potencial oleogenítico da Formação Codó.
Bol. Técn. PETROBRAS, 21: 3-16.
GÓES, A.M. 1995. A Formação Poti (Carbonífero inferior) da Bacia do Parnaíba.
Universidade de São Paulo, São Paulo, 171 p. (Tese de Doutorado)
GÓES, A.M. & ROSSETTI, D.F. 2001. Gênese da Bacia de São Luís-Grajaú, Meio-Norte do
Brasil. In: D.F. ROSSETTI, A.M. GÓES, W. TRUCKENBRODT (Eds) O Cretáceo na
Bacia de São Luís-Grajaú. Belém, Coleção Friedrich Katzer, Museu Paraense Emílio
Goeldi, p. 15-29.
GONÇALVES, D. F. 2004. Argilominerais da Formação Codó (Aptiano superior) – Bacia do
Grajaú: implicações climáticas e ambientais. Belém, Univ. Fed. Pará, 100 pp.
(Dissertação de Mestrado)
GUZZO, J.V.P. 1997. Estratigrafia integrada e paleolimnologia de uma seção de idade Aratu
(Eocretáceo) da Bacia do Recôncavo, NE do Brasil. Porto Alegre, Univ. Fed. do Rio
Grande do Sul, 210 p. (Dissertação de Mestrado)
HERCZEG, A.L. 1988. Early diagenesis of organic matter in lakes sediments: a stable carbon
isotope study of pore waters. Chemical Geology, 72: 199-209. (Isotope Geoscience
Section)
HOEFS, J. 1980. Stable isotope geochemistry. 2
a
ed. Berlim, Springer-Verlag, 208 p.
HOLLAND, H.D., 1978. The Chemistry of the Atmosphere and Oceans. Wiley-Interscience,
New York, 351 p.
HOLSER, W.T.; MAGARITZ, M.; RIPPERDAN, R.L. 1996. Global isotopic events. In: H.
WALLISER (Ed.) Global events and event stratigraphy in the Phanerozoic. Berlim,
Springer-Verlag, p. 63-88.
LI, H.-C. & KU, T.-L. 1997. δ
13
C and δ
18
O covariance as a paleohydrological indicator for
closed-basin lakes. Palaeogeog., Palaeoclim., Palaeoecol., 133: 69-80.
26
LIMA, R.D. & ROSSETTI, D. F. 1999. Depositional facies in Late Cretaceous-?Lower
Tertiary deposits from northwestern Maranhão State, Brazil. Rev. Bras. Geoc., 29: 237-
244.
LISBOA, M.A.R. 1914. The Permian geology of northern Brazil. Am. J. Sci., 177: 425-442.
LISTER, G.S., KELTS, K., CHEN, K.Z., YU, J.-Q., NIESSEN, F., 1991. Lake Qingai, China:
closed-lake basin levels and the oxigen isotope record for ostracoda since the latest
Pleistocene. Palaeogeogr., Palaeoclimatol., Palaeoecol., 84: 141-162.
LU, F.H.; MEYERS, W.J.; SCHOONEN, M.A. 2001. S and O (SO
4
) isotopes, simultaneous
modeling, and environmental significance of the Nijar messinian gypsum, Spain. Geochim.
Cosmochim. Acta, 65: 3081-3092.
KRULL, E.S.; LEHRMANN, D. J.; DRUKE, D.; KESSEL, B.; YU, Y.Y.; LI, R. 2004. Stable
carbon isotope stratigraphy across the Permian-Triassic boundary in shallow marine
carbonate platforms, Nanpanjiang Basin, south China. Palaeogeogr., Palaeoclimatol.,
Palaeoecol., 204: 297-315.
MCKENZIE, J. A. 1985. Carbon Isotopes and Productivity in the Lacustrine and Marine
Environment. In: STUMM, W. (Ed.) Chemical Processes in Lakes. New York, Wiley, 99-
118.
NEUMANN, V.H.M.L. 1999. Estratigrafía, sedimentología, geoquímica y diagénesis de los
sistemas lacustres aptienses-albienses de la Cuenca de Araripe (Noreste de Brasil).
Barcelona, Universitat de Barcelona, 234 pp. (Tese de doutorado)
ORTÍ, F. & ROSELL, L. 2000. Evaporative systems and diagenetic patterns in the Calatayud
Basin (Miocene, Central Spain). Sedimentology, 47: 665-685.
PAZ, J.D.S. 2000. Análise faciológica da Formação Codó (Aptiano superior), na região de
Codó (MA), borda leste da Bacia do Grajaú. Belém, Univ. Fed. Pará, Belém, 117p.
(Disertação de Mestrado)
PAZ, J.D.S & ROSSETTI, D.F. 2001. Reconstrução paleoambiental da Formação Codó
(Aptiano), borda leste da Bacia do Grajaú, MA. In: D.F. ROSSETTI, A.M. GÓES, W.
TRUCKENBRODT (Eds) O Cretáceo na Bacia de São Luís-Grajaú. Belém, Coleção
Friedrich Katzer, Museu Paraense Emílio Goeldi, p. 77-100.
27
PINTO, I.D. & ORNELLAS, L.P. 1974. New cretaceous hemiptera (insects) from Codó
Formation – Northern Brazil. In: CONGRESSO BRASILEIRO DE GEOLOGIA, 28.
Porto Alegre, Anais... SBG, p. 289-304.
PLAYÀ, E.; ORTÍ, F.; ROSELL, L. 2000. Marine to non-marine sedimentation in the upper
Miocene evaporites of the Eastern Betics, SE Spain: sedimentological and geochemical
evidence. Sediment. Geol., 133: 135-166.
POPOV, Y.A. & PINTO, I.D. 2000. On some Mesozoic burrower bugs (Heteroptera;
Cydnidae). Paleontological Journal, 34: 298-302 (Suppl. 3)
REZENDE, N.G.A.M. 1997. Argilas nobres e zeólitas na Bacia do Parnaíba. Belém, CPRM,
50 p. (Relatório final de projeto)
RHODES, M.K.; CAROLL, A.R.; PIETRAS, J.T.; BEARD, B.L.; JOHNSON, C.M. 2002.
Strontium isotope record of paleohydrology and continental weathering, Eocene Green
River Formation, Wyoming. Geology, 30: 167-170.
RODRIGUES, R. 1995. A geoquímica orgânica na Bacia do Parnaíba. Porto Alegre, Univ.
Fed. Rio Grande do Sul, 225 p. (Tese de Doutorado).
RODRIGUES, R. & TAKAKI, T. 1993. Estratigrafia química da Formação Codó, Cretáceo
inferior da Bacia do Parnaíba. In: SIMPOSIO SOBRE O CRETACEO DO BRASIL, 3.
Rio Claro, Boletim... Rio Claro, SBG. v.2 , p. 115-117.
ROSSETTI, D.F. 1998. Facies architecture and sequential evolution of incised valley estuarine
fills: the Upper Itapecuru Formation (São Luís Basin), northern Brazil. J. Sediment. Res.,
68: 299-310.
ROSSETTI, D.F.
2001. Arquitetura deposicional da Bacia de São Luís-Grajaú. In: D.F.
ROSSETTI, A.M. GÓES, W. TRUCKENBRODT (Eds) O Cretáceo na Bacia de São
Luís-Grajaú. Belém, Museu Paraense Emílio Goeldi, Belém, Museu Paraense Emílio
Goeldi, pp. 31-46. (Coleção Friedrich Katzer)
ROSSETTI, D.F. & GÓES, A.M. 2000. Deciphering the sedimentological imprint of
paleoseismic events: an example from the Aptian Codó Formation, northern Brazil.
Sediment. Geol., 135: 137-156.
ROSSETTI, D.F. & TRUCKENBRODT, W. 1997. Classificação estratigráfica para o
Albiano-Terciário Inferior (?) na Bacia de São Luís, MA. Boletim do Museu Paraense
Emílio Goeldi - Série Ciências da Terra, 9: 31-43.
28
ROSSETTI, D.F.; PAZ, J.D.S.; GÓES, A.M. 2004. Facies analysis of the Codó Formation
(Late Aptian) in the Grajaú Area, Southern São Luís-Grajaú Basin. An. Acad. Bras. Ciênc.,
76: 791-806.
ROSSETTI, D.F.; PAZ, J.D.; GÓES, A.M.; MACAMBIRA, M. 2000. A marine versus non-
marine origin for the Aptian-Albian evaporites of the São Luís and Grajaú basins,
Maranhão state (Brazil) based on sequential analysis. Rev. Bras. Geoc., 30: 642-645.
STUIVER, M. 1970. Oxygen and carbon isotope ratios of freshwater carbonates as climatic
indicators. Journal of Geophysical Research, 75: 5247-5257.
SZATMARI, P., FRANÇOLIN, J.B.L., ZANNOTTO, O.; WOLFF, S. 1987. Evolução
tectônica da margem equatorial brasileira. Rev. Bras. Geoc., 17: 180-188.
TALBOT, M.R., 1990. A review of the paleohydrological interpreatation of carbon and
oxygen isotopic ratios in primary lacustrine carbonates. Chem. Geol., 80: 261-279.
TALBOT, M.R. & ALLEN, P.A. 1996. Lakes. In: H.G. READING (Ed.) Sedimentary
Environments: Process, Facies and Stratigraphy. Oxford, Blackwell, pp. 83-124.
TALBOT, M.R. & KELTS, K., 1990. Paleolimnological signatures from carbon and oxygen
isotopic ratios in carbonates from organic carbon-rich lacustrine sediments. In: B.J. KATZ
(Ed.) Lacustrine Basin Exploration: Case Studies and Modern Analogs. Tulsa, AAPG, pp.
99-112. (Memoir, 50)
TALBOT, M.R. & LÆRDAL, T. 2000. The late Pleistocene-Holocene palaeolimnology of
Lake Victoria, East Africa, based upon elemental and isotopic analyses of sedimentary
organic matter. Journal of Paleolimnology, 23: 141-164.
TUCKER, M.E. 2003. Sedimentary rocks in the field (3
rd
Ed.). John Wiley & Sons, 244 p.
WIGNALL, P.B. & TWITCHETT, R.J. 2002. Extent, duration, and nature of the Permian-
Triassic superanoxia event. Geol. Soc. Am. Spec. Publ., 104: 171-176.
ZANNOTTO, O. & SZATMARI, P. 1987. Mecanismo de rifteamento da porção ocidental da
margem norte brasileira, Bacia do Pará-Maranhão. Rev. Bras. Geoc., 17: 189-197.
29
2. LINKING LACUSTRINE CYCLES WITH SYN-SEDIMENTARY
TECTONIC EPISODES: AN EXAMPLE FROM THE CODÓ
FORMATION (LATE APTIAN), NORTHEASTERN BRAZIL
*
2.1. ABSTRACT
The Codó Formation exposed in the eastern Grajaú Basin consists mostly of black
shales, limestones and evaporites arranged into several shallowing-upward cycles formed by
progadation of lake deposits. Three ranks of cycles are distinguished. The lower rank cycles
correspond to milimetric interbeddings of: bituminous black shales with evaporites,
calcimudstones or peloidal wackestone-packstone; grey/green shale with calcimudstone,
peloidal wackestone-packstone or ostracodal wackestone/grainstone; and ostracodal
wackestone/grainstone and/or calcimudstones with cryptomicrobial mats and ooidal/pisoidal
packstones. The intermediate rank cycles average 1.7 m thick and are formed by complete and
incomplete cycles. Complete cycles show a transition from central to intermediate and then to
marginal facies associations and include two types: C1 cycle with central lake deposits
consisting of evaporites and black shales; and C2 cycle with central lake deposits formed by
grey/green shale. Complete cycles are attributed to the upward gradation from central to
marginal environments of the lake or saline pan-sabkha system. Incomplete cycles are either
successions with central and intermediate facies associations (I1) or successions with
*
Authors: J.D.S. Paz & D.F. Rossetti. Geological Magazine, in press.
30
intermediate and marginal facies associations (I2), which represent shallowing-upward
successions where at least one facies association is lacking. The higher rank cycles average 5.2
m thick and consist of four depositional units that display shallowing-upward successions
formed by complete and incomplete intermediate rank cycles that vary their distribution
upward in the section, and are bounded by sharp surfaces.
While the lower rank cycles display characteristics that reveal their seasonal signature,
detailed sedimentological characterization and understanding of stratal stacking patterns
related to the intermediate and higher rank cycles supports a genesis linked to syn-sedimentary
tectonic activity. This is particularly suggested by the high facies variability, limited lateral
extension, and frequent and random thickness changes of the intermediate-rank cycles.
Additionally, the four higher rank cycles recognized in the Codó Formation match with
stratigraphic zones having different styles of soft-sediment deformation structures attributed to
seismic activities. Therefore, the several episodes of lake shallowing recorded in the Codó
Formation are linked to seismic pulses that alternated with sediment deposition. This process
would have created significant changes in the lake water level, and resulted in sharply
bounded successions internally displaying deeper to relatively shallower facies associations in
the upward direction.
Key-words: lake system, cyclic sedimentation, evaporites, syn-sedimentary tectonics,
Late Aptian, Brazil.
2.1. INTRODUCTION
Shallowing-upward cycles are one of the most common characteristics of lacustrine
successions, and many authors have related them to orbital forcing (e.g., Olsen, 1986; Glenn
& Kelts, 1991; Olsen & Kent, 1996, 1999; Juhász et al., 1997; Vugt et al., 1998; Steenbrick et
al., 2000; Hofmann et al., 2000; Aziz et al., 2000). Developments in the understanding of
facies relationship and thickness consistency have also allowed lacustrine cycles to be related
to tectonic pulses. However, only a few studies have clearly demonstrated the linkage between
shallowing-upward cycles and tectonics (e.g., Martel & Gibling, 1991; Anadon et al., 1991).
The Codó Formation (Upper Aptian) exposed in several quarries along the eastern
Grajaú Basin, northeastern Brazil (Fig. 2-1), is a saline lacustrine to sabkha complex
31
characterized by different ranks of shallowing-upward cycles. This unit, particularly well
exposed in the Codó and Grajaú areas, consists chiefly of bituminous shales, evaporites and
limestones, forming a succession up to 25 m thick. Internally, the Codó Formation displays
several depositional cycles formed by episodes of upward shallowing. Although smaller scale
cycles are probably attributed to seasonal fluctuations, intermediate and higher rank cycles
show features that do not match with climate influence. In this work we present a detailed
sedimentological and stratigraphic analysis of the shallowing upward cycles recognized in the
Codó Formation, in order to demonstrate that their genesis are, at least in part, related to syn-
sedimentary tectonic pulses.
2.3. GEOLOGICAL SETTING
The split up of Africa and South America continents led to the establishment of several
rift basins along the equatorial Brazilian margin. The São Luís and Grajaú basins are
expressive examples of this tectonic phase, occupying together more than 150,000 km
2
. These
basins are interpreted to represent a unique structural feature formed by combination of pure
32
shear stress and strike-slip deformation associated with an intracontinental rift (Góes &
Rossetti, 2001). Fault displacement (Fig. 2-2a,b) initiated during the Aptian, resulting in a
shallow basin where the Codó Formation was deposited. During the main rifting in the Albian,
fault offsets reached up to 400 m, culminating with the amplification of the rift system.
Vertical to sub-vertical, normal and reverse faults cut through the entire sedimentary package,
with some continuing downward into the crystalline and Palaeozoic basement. Three main
fault trends can be distinguished (Fig. 2-2a): NE/SW, NW/SE, and less commonly E-W
(Rezende & Pamplona, 1970; Azevedo, 1991
).
The latter is closely associated with the E-W-
oriented Sobradinho Fault, which represents the continuity of the Romanche Transcurrent
Zone (Pindell, 1985). E-W oriented thrust faults with offsets of a few meters are present in
outcrops located in the southern and eastern margins of the basin (Góes & Rossetti, 2001).
The Codó Formation records the Upper Aptian deposition of the São Luís-Grajaú
Basin. Gamma-ray log correlation in this basin shows that the sedimentary record consists,
from bottom to top, of three major, probably second-order, depositional sequences bounded by
regional discontinuity surfaces, interpreted as sequence boundaries (Rossetti, 2001; Fig. 2-3).
The two lowermost sequences (S1 and S2) are Aptian to Middle Albian in age, and show an
internal tripartite organization into lowstand, transgressive and highstand systems tracts
(Rossetti, 2001), as expected in a complete depositional sequence. The lowstand systems tracts
of sequences S1 and S2 comprise continental (i.e., fluvial, deltaic, eolian and lacustrine)
deposits in the southern margin of the Grajaú Basin, which interfinger with shallow
marine/estuarine deposits to the north in the São Luís Basin. In both sequences, the
transgressive systems tracts are characterized by muddy lithologies with marine fossils that
sharply cover the underlying sandier deposits, resulting in wedges that pinch out southward.
Discontinuity surfaces marked by sharp lithological contrast define the base and top of these
successions and are interpreted as transgressive and maximum marine flooding surfaces,
respectively (Rossetti, 2001). The highstand systems tracts are characterized by deposits
displaying aggradational stratal patterns that grade upward into intervals with an overall
prograding character, formed when the relative sea level started to decline at the end of the
highstand stage. In contrast to sequences S1 and S2, sequence S3 does not show a tripartite
internal organization, but consists of several sharply bounded, fining- and then coarsening-
33
upward successions. At least the two uppermost ones are partly exposed and record estuarine
systems formed during the Late Cretaceous (Rossetti, 2001).
34
35
upward successions. At least the two uppermost ones are partly exposed and record estuarine
systems formed during the Late Cretaceous (Rossetti, 2001).
The Codó Formation correlates with the lowstand deposits of sequence S1 described
above (Fig. 2-3). It is interesting to note that, in the southern margin of the basin, the Codó
Formation contains a horizon marked by soft-sediment deformation sandwiched within
entirely undisturbed deposits, which are attributed to syn-sedimentary seismic activity. Four
zones of deformation were recognized (Rossetti & Góes, 2000), which are briefly summarized
here (Fig. 2-4) due to their close relationship with the shallowing-upward cycles discussed in
this paper. They include: (1) zone Z1, with cracks filled by fine-grained calcite crystals, small-
scale faults, fissures and stylolites inclined at a high angle to bedding; (2) zone Z2, represented
by complex convolute folds associated with thrust faults, pseudonodules, and mound-and-sag
structures, the later requiring alternating periods of deposition and sediment deformation; (3)
zone Z3, which is associated with intraformational boulders up to 2.5 m long, consists of
normal faults and fissures that are vertical to near vertical, present ragged morphologies, with
small, delicate edges, and taper both downward and upward after a few centimeters; and (4)
zone Z4, characterized by shales with irregular convolute folds. This vertical succession of
deformation events were attributed to syn-sedimentary shear stresses associated with the early
rifting that gave rise to the São Luís-Grajaú Basin. These syn-sedimentary seismic pulses seem
to have had a strong influence on the evolution of the Codó lake and saline pan/sabkha
complex and on the origin of its shallowing-upward cycles, as proposed in this paper.
2.4. DEPOSITIONAL ENVIRONMENT
A detailed facies analysis of the Codó Formation was presented elsewhere (Paz &
Rossetti, 2001; Rossetti ert al., 2004). However, given its importance to provide a general
overview on the depositional setting and define the shallowing-upward cycles discussed here,
the main descriptions and interpretations will be summarized in this section, together with
some new information that helps to support the proposed paleoenvironmental model. The
Codó Formation can be described in terms of four facies associations in the study areas (Fig.
2-5), which are attributed to central lake, intermediate lake, marginal lake, and saline
pan/sabkha depositional environments. Ostracods, including two genera, Harbinia and
36
Candona, are abundant throughout these deposits, occurring either dispersed or as thin (up to
10 cm thick) beds of shell mittens. Disarticulated and articulated shells, including both young
and adult individuals, are present, as are freshwater Charophytes algae.
2.4.1. Central lake facies association:
The central lake deposits (Fig. 2-6a,b) occur at the base of the shallowing-upward
cycles, and consist mostly of bituminous black shale and evaporite (mostly gypsum). The
bituminous black shale is the most frequent facies in this association, occurring as packages up
to 3 m thick. It consists of bituminous shales (Fig. 2-6b), with total organic content up to 30%,
which contain plant remains, pyrite crystals that are as large as 0.5 cm, and lenses of native
sulphur. Silt grain sizes include quartz and feldspar grains, while clays are composed mostly
of detrital (irregular, laminated flakes) smectite and, secondarily, kaolinite and illite.
Bioturbation does not occur in this facies association.
37
38
The evaporite facies reaches up to 5 m in thickness and includes massive/macronodular
(Fig. 2-6a) and, subordinately, layered gypsum. Packages of gypsum up to 1 m thick showing
well-developed horizontal lamination are locally present. The massive gypsum forms
unstructured bodies intergraded with macronodular gypsum, the latter consisting of
centimeter-sized nodules with or without a shale matrix. The gypsum nodules display an
almond-like form, defined by a web of undulating, horizontal to oblique fractures filled by
bladed crystals up to 0.5 cm thick and with compromise contacts that grow perpendicularly to
the fracture walls. Rosettes of dark gypsum up to 5 cm in diameter are common in this facies.
The massive/macrogranular gypsum is in sharp contact with the layered gypsum, locally
forming diapirs several meters long. The latter comprises laterally continuous, horizontal
alternating dark/light beds that vary from few mm up to 10 cm thick. The dark beds are
formed by either crystals or micronodules of gypsum less than 0.5 cm long distributed within a
matrix of black shale, while the light beds consist of upward-oriented chevron gypsum
crystals.
2.4.2. Transitional lake facies association:
The transitional facies association (Fig. 2-6c) consists of interbedded grey/green shale
and limestones. Bioturbation is locally present, as are symmetrical ripple marks at the top of
some limestone beds. The grey/green shale is the dominant facies in this association, occurring
particularly well developed in the Grajaú area, where it reaches up to 4 m thick. The total
organic content of this facies is much lower than in the black shale, reaching only up to 1 %.
Films of gypsum or calcite as acicular or fibrous, vertically aligned crystals, are locally present
on bed planes.
The limestones can be described in terms of three facies: calcimudstone, peloidal
wackestone-packstone, and sparstone. The calcimudstone is the dominant limestone facies in
this association. It consists of microcrystalline calcite, locally silicified to chalcedony, which
occurs as laterally continuous, either laminated or massive beds up to 15 cm thick. Grains of
quartz and mica (mostly muscovite) occur disperse or parallel to bed planes. The peloidal
wackestone-packstone facies forms beds up to 20 cm thick, characterized by well sorted,
rounded to sub-rounded, spherical to oblate peloids within a micrite matrix. Disarticulated
ostracod shells and microspherules are frequently present, reaching up to 15 % of the
39
allochemical grains. This facies is commonly parallel laminated and, less commonly, massive.
Films of microbial mats might be present paralleling bed planes. The sparstone facies (cf.
Wright, 1992) consists exclusively of calcite crystals ranging in size from 250-500 µm and
displaying sutured contacts, arranged as a blocky mosaic.
2.4.3. Marginal lake facies association:
The marginal lake facies association (Fig. 2-6d-g) occurs at the top of the shallowing-
upward cycles and comprises a variety of intergrading lithofacies, including: pelite,
intraclastic grainstone, ostracodal wackestone to grainstone, ooidal/pisoidal packstone,
rhythmite (Fig. 2-6D,E), tufa, and sandstone. Wave-ripple marks are commonly observed
within this facies association, and gypsum occurs locally as discontinuous laminae.
Bioturbation is, in general, absent, though a few undetermined tiny traces were locally
observed.
The pelite facies is endurated, massive, and display a blocky texture (Fig. 2-6f). Its
colour varies upward from olive-green to brownish-red. The intraclastic grainstone (Fig. 2-6g)
consists of well sorted and well rounded, fine- to coarse-grained calcite grains. This facies
forms structureless and, less commonly, tabular and sigmoidal cross-stratified beds that are up
to 20 cm thick. Mm-cm long lensoid cavities (i.e., fenestrae), locally filled by mosaics of
sparite, are typical of this facies, as are microkarstic structures, and vadose meniscate calcite
cement. Ostracodal wackestone to grainstone occurs as beds or concretions up to 15 cm thick,
the latter typically displaying articulated shells with a mixture of young and adult individuals.
Ooidal/pisoidal packstone consists of coalescing, well-rounded or elongated ooids and pisoids
(diameter<5 mm) displaying internal botryoidal or fibrous radial fabric. This facies forms
layers up to 5 cm thick, but which are laterally continuous for the entire extension of the
exposures (at least hundreds of meters). Rounded to slightly elongated fenestrae are abundant
in this facies. The tufa facies is a highly porous limestone with sparry calcite filaments
arranged into a dendritic web. Rounded calcite grains less than 200 µm of diameter occur
encrusting the branches. Rhythmite occurs as packages 10-20 cm thick of thinly (millimeter-
scale) laminated, ostracodal wackestone-grainstone, shale, and less commonly, siltstone beds.
Fish are abundant in this facies. Cryptomicrobial mats are abundant in association with
40
ooidal/pisoidal packstone and rhythmites, being recognized in the field as black, corrugated
films parallel to bed planes.
The sandstone facies is structureless and occurs only locally as lenses up to 0.5 m
thick, being represented by moderately sorted, well rounded, fine-grained, quartz grains.
2.4.4. Saline pan/sabkha facies association
This facies association is up to 6 m thick, being mostly present in the Grajaú area,
where it occurs between lacustrine deposits. This is an evaporite-dominated association
constituted by layered gypsum, gypsarenite, and secondarily tufa, grey/green shale, and
calcimudstone. The three latter facies will not be described here, as they have the same
characteristics described in previous facies associations. Tufa is much more expressive in this
facies association than in the marginal lake deposits, occurring as beds up to 5 cm thick that
disappear laterally within few meters. A typical feature of this facies association is the
preservation of primary horizontal layering.
The layered gypsum (Fig. 2-6h,i) consists of horizontal beds typically formed by
alternating darker and lighter couplets that become progressively thicker upward, varying from
few mm to 30 cm. The darker beds consist of either crystals or micronodules of gypsum, in
general less than 5 mm long, which occur within a matrix of dark mudstone. The lighter
gypsum consist of upward-oriented crystals displaying aligned twin-planes and superimposed
growth faces with acute angles arranged in zig-zag, perpendicularly to the crystal long axe
(i.e., chevron gypsum); the chevron gypsum intergrades with acicular gypsum.
The gypsarenite is interbedded with the layered gypsum, and forms layers few cm thick
of moderate to poorly sorted, rounded to sub-rounded gypsum grains with sizes varying from
very coarse to pebbly. Calcite cement is common. This facies increases in frequency upward
in this association. Shallow (i.e., <2 mm deep) potholes averaging 3 cm in diameter and with
ragged concave up shapes completely filled up by layers of growth-aligned gypsum crystals,
are present at the top of some gypsarenite beds. A film of yellow to light brownish calciferous
clay is settled out on the bottom of these structures.
41
2.4.5. Interpretation:
Although displaying similar lithologies, a comparison of the sedimentary features and
facies architecture between the Grajaú and Codó areas revealed some basic differences in their
depositional systems. Hence, the Codó area displays deposits formed in comparatively more
stable, well-stratified lakes with significant periods of anoxia and closure, when evaporites
where formed almost exclusively in deeper, central areas. Facies analysis, added to the
abundance of continental ostracods and Charophyte algae, as well as the absence of any
marine fauna, led us to attribute these deposits to a dominantly lacustrine system. The
abundance of shales in the central lake deposits shows a low energy depositional setting.
Pyrite, bitumen, well-preserved fossils (i.e., ostracods and fishes) and the almost complete
lack of bioturbation support anoxic conditions. Under such circumstance, the preservation of
large amounts of organic matter is favoured, being derived from organisms living in upper,
more oxygenated water layers (Trewin, 1986; Martill, 1988).
The transitional lake facies association is characterized by its position between central
and marginal lake facies associations. The abundance of fine-grained deposits, recorded by
calcimudstones and grey/green shales consistent with a low energy environment with
abundant mud deposition below wave base (e.g., Specht & Brenner, 1979; Shinn et al., 1989).
In this sense, this association resembles central lake deposits, but the lighter colour, suggesting
less preservation of organic matter and absence of bitumen, as well as the local bioturbated
horizons, symmetrical ripple marks, and disarticulated ostracod shells, support a shallower
environment relative to the central lake association, with more oxygenated waters and wave
action.
The marginal lake facies association was recognized by its position at the top of the
shallowing-upward cycles and by the abundance of sedimentary features attributed to
subaerial and/or meteoric exposure, such as palaeosoil (recorded by the massive pelite),
microkarstic surface, fenestrae and meteoric cement. The presence of coalescing pisoids with
variable sizes, elongated shapes that are associated with microbial mats suggests stagnant,
shallow water condition and/or even subaerially exposed environments (e.g., Risacher &
Eugster, 1979; Chafetz & Butler, 1980; Schreiber et al., 1981; Tucker & Wright, 1990). Under
such conditions, microbes lead to in situ formation of pisoids (cf. Ferguson et al., 1978), an
42
interpretation favoured in the study area due to the close association of pisoids and microbial
mats. Fenestrae support formation close to the vadose zone. The concretions of ostracodal
wackestone-packstone associated with these deposits might have resulted from a microbial
influence (Raiswell, 1976) combined with anisotropic permeability, since the concretions
occur within more permeable limestone/shale rhythmite that overlies less permeable shales.
The local occurrence of sandstone facies in this association is probably related to episodic
influx of sands during raining episodes through small deltas.
As opposed to the Codó area, much more ephemeral conditions prevailed in the Grajaú
area, which presented better-oxygenated water pans having evaporite precipitation only in
their margins and along the surrounding mudflats, configuring a saline pan/sabkha complex.
This depositional setting displays widespread pans and evaporitic flats, the first being
restricted to more depressed local areas of the system. The distinguished depositional setting
recorded in the Grajaú area might have been crucial to control the distribution of the evaporite
deposits. A saline pan/sabkha facies association is indicated by evaporite precipitation in a
very shallow subaqueous environment that was interrupted by periodic exposure. The thin
beds of grey/green shales and calcimudstones are attributed as the record of deeper waters,
probably representing central saline pans. Shallower areas of the pans and surrounding flats
were dominated by evaporite deposition. The dominance of gypsum with well-preserved
horizontal lamination reflects original bedding on flatter-lying environments surrounding the
saline pans. The chevron growth-aligned crystals of the lighter beds attest to primary gypsum
deposition on the floor of brine pools. Precipitation of similar features in many modern and
ancient environments occurs when water depths is less than only 2 m (e.g., Logan, 1987;
Hovorka, 1987; Handford, 1991; Smoot & Lowenstein, 1991). Shallow waters favour a high
degree of supersaturation, as well as a less dense and unstable brines, which are conditions
required to promote the upward growth of crystals. The darker beds in the couplets record
displacive intrasediment growth of crystals beneath the brine from supersaturated pore fluids
in the capillary and/or upper phreatic zone, thus requesting periods of descending ground
waters and eventual exposure (e.g., Kerr & Thomson, 1963, Warren, 1999). These deposits
formed slightly post-depositionally, but still under a strong influence of the depositional
setting as precipitation took place within only a few mm of the depositional surface. The
alternating dark and light gypsum couplets are attributed to pulsating episodes of water table.
43
The growth-aligned crystals grew up in the interface sediment-brine during times when
saturated brines were flushed into the basin. The formation of such evaporite layers requires a
subaqueous environment with stable phases to allow the aggradation of the upper euhedral
surface of the crystals (Warren, 1999). As the water level falls, precipitation of this type of
crystals is precluded. During these episodes, evaporite precipitation in the study area occurred
only below the sediment interface.
The gypsarenite records moments when evaporite crystals were reworked. Because
these deposits are mostly associated with the layered gypsum, it is possible that its formation
is due to reworking of growth-aligned crystals as the water level slightly decreased, but with
the process occurring still subaqueously. The increased occurrence of this facies upward in the
evaporite section attests to progressive shallowing, an interpretation that is supported by the
presence of numerous potholes attributed to partial evaporite dissolution due to subaerial
exposure. As the evaporite basin was dissected, the deposits were at least momentously kept
above the water level and dissolution took place forming these small depressions. Water
accumulated in these depressions, bringing suspended sediments, that were settle down
forming the clay films. A later submergence of the dissolved planes led to the infill of the
potholes by thin layers of growth-aligned crystals.
2.5. SHALLOWING-UPWARD CYCLES AND DEPOSITIONAL UNITS
2.5.1. Description
Three ranks of shallowing-upward cycles are distinguished in the Codó Formation. The
lower rank cycles display regular thickness ranging from 5 to 10 cm, and consist of facies that
vary according to the position in the lake setting. Hence, the central lake deposits show
interbeddings either of bituminous black shales and evaporites, or bituminous black shales
with streaks of calcimudstone and bituminous black shales with native sulphur. The
intermediate lake deposits display bituminous black shale interbedded with peloidal
wackestone-packstone or grey/green shale interbedded either with calcimudstone or peloidal
wackestone-packstone. The marginal lake deposits show either grey/green shale and
ostracodal wackestone/grainstone, as well as alternations of ostracodal wackestone/grainstone
and/or calcimudstones with cryptomicrobial mats and ooidal/pisoidal packstones. The lower
44
rank cycles in the saline pan/sabkha deposits consist of alternating dark and lighter gypsum,
consisting of microgranular gypsum grown within shales and vertical aligned gypsum crystals,
respectively. These are locally arranged into packages containing 3 to 7 couplets (Fig. 2-6H); a
succession of 4/7/5/6/6/7/6 couplets was counted at one place.
The intermediate rank cycles vary from a few 0.3 m up to 5.6 m thick (averaging 1.7 m
thick), and contain the lower rank cycles. They can be either complete or incomplete
according to the relative proportion of facies associations representing the several lacustrine
and saline pan/sabkha sub-environments (Fig. 2-7). Hence, complete cycles display facies
associations that record the upward gradation from central to marginal environments of the
lake or saline pan-sabkha system, attesting to the complete preservation of one shallowing
episode. Three different types of complete cycles were recognized (Fig. 2-8): complete cycles
type 1 (C1), where central lacustrine deposits are entirely formed by evaporites and
bituminous black shales; and complete cycle type 2 (C2), where central lake deposits consist
of black and/or grey/green shales and limestones; and complete cycle type 3 (C3), where
grey/green shale and calcimudstone record central saline pan environments, and evaporites
(layered gypsum, gipsarenite) and tufa record shallower saline pan and surrounding evaporitic
flats. Incomplete cycles are defined by shallowing-upward successions where at least one
facies association is lacking. There are also three types of incomplete cycles (Fig. 2-8):
incomplete cycle type 1 (I1), where central and intermediate lake associations dominate;
incomplete cycle type 2 (I2), composed mostly of intermediate and marginal lake facies
associations; incomplete cycle type 3 (I3), consisting of only shallower saline pan and
evaporitic flat deposits.
45
46
Fig .2-8: Diagrams illustrating the four types of lower-rank, shallowing-upward cycles of the Codó Formation.
Thickness of individual cycles range from 0.3-5.6 m. See figure 2-7 for legend.
The higher rank cycles define four laterally continuous depositional units, referred as 1
to 4 from bottom to top (Figs. 2-7, 9a,b). Unit 1 is only partly exposed at the base of the
sections and form an interval that reaches up to 3.6 m thick (averaging 2.7 m thick) composed
of thin I1 cycles. Its top is either marked by a horizon of breccia containing clasts of
grey/green shale up to 5 cm long, or by lenses of medium- to coarse-grained sandstones,
which occur along a sharp surface with erosional relief up to 4 m. Unit 2 (Fig. 2-7) reaches up
to 8 m thick (averaging 5.2 m) and contains all types of cycles. Although in number there is a
prevalence of incomplete cycles, this unit displays the highest volume of complete cycles of
the whole Codó Formation exposed in the study area. Up to five successive shallowing-
upward cycles were observed in this unit, with C1 cycles being the thickest and dominant
ones. These occur mainly in profiles 2 and 3 of the Codó area and profiles 7 and 8 of the
Grajaú area (Fig. 2-7). It is noteworthy that the marginal facies of the shallowing-upward
cycles located closer to the top of unit 2 are characterized by the abundance of gypsarenite,
intraclastic grainstone with fenestrae, karstic features and nodular chert (silcrete). The top of
unit 2 is marked by a sharp bounding surface that is either planar or displays an erosional
relief up to 1 m at the outcrop scale.
47
Unit 3 (Figs. 9b and c), confined to the eastern portion of the study area, is up to 3.8 m
thick (averaging 2.6 m thick) and is comprised mostly (i.e., nearly 80%) of I2 cycles, with the
remaining 20% being represented by C2 cycles. A remarkable and exclusive feature of this
unit is the presence of ooids/pisoids and concretions of ostracodal wackestone-grainstone,
which constitute important stratigraphic markers throughout the study area. The latter occur
invariably at the transition from intermediate to marginal lacustrine deposits, while the
ooids/pisoids are present in marginal lake deposits located close to the top of unit 3. The
concretions bearing articulated ostracod shells include a mixture of juvenile and adult
individuals. These lithologies are interbedded with rhythmites bearing abundant articulated
and disarticulated fish bones, and which also contain a high volume of former microbial mats.
Unit 3 is also bounded at the top by a sharp surface with erosional relief up to 2,5 m.
Fig .2-9: Types of shallowing-upward units present in the Codó Formation, illustrating: a) Unit 2, represented by
one complete cycle type 1 with evaporites and shales. b) Unit 3, presented from base to top by two incomplete
cycles type 1 and one complete cycle type 2. c) The uppermost portion of unit 2 and the base of unit 3, with the
latter showing three incomplete cycles type 1 and an upper incomplete cycle type 2.
48
The uppermost unit 4 (Fig. 2-9b) is up to 4.6 m thick (averaging 2.2 m thick) and
typically starts at the base with grey/green shales containing only thin (< 1 mm thick) laminae
of gypsite and/or fibrous calcite, both formed by vertically aligned crystals. Upward, the
shales are interbedded with few and thin (mm to few cm) layers of calcimudstones. I1 cycles
are dominant in this interval, but no black shales are present. Secondarily, this unit also shows
I2 cycles. Unit 4 is truncated by a discontinuity surface showing erosional relief up to 5 m,
marked by a red-coloured palaeosol horizon overlain by Albian sandstones and argillites of the
Itapecuru Group (Fig. 2-7).
2.5.2. Interpretation:
The lower rank cycles recognized in the Codó Formation record minor changes in
depositional conditions, attesting to alternations between mud settling and chemical
precipitation of evaporites or limestones. This characteristic, added to the regular thickness
variation, is consistent with seasonal fluctuations, with mud deposition and chemical
precipitation taking place during less and more arid phases, respectively.
The intermediate rank, shallowing-upward cycles reflect periods of progradation of the
lake shoreline resulting from superposition of marginal lake deposits upon intermediate and/or
central lake deposits. The higher rank cycles record several episodes of upward-shallowing
due to lake desiccation, bounded by subsequent floodings. In this sense, they are good
continental analogs for parasequences described in marginal marine settings (Vandervoort,
1997). The maximum shallowing was reached at the top of each unit, and it is marked by
better-developed marginal lacustrine facies associations, evidenced by the features recording
wave reworking and subaerial exposure such as palaeosol, karstic features, fenestrae, and chert
(silcrete) nodules.
Units 1 and 2 reflect the prevalence of anoxic conditions and a water column with
sufficient depth to favour stratification. The abundance of C1 cycles in unit 2, rich in bitumen
and evaporites, is the most representative record of this phase, and is attributed to a phase
when the lake had the maximum relative depth, which is consistent with the fact that in this
unit cycles are thicker than in the other units.
Unit 3 records a time with the greatest development of marginal lake deposits,
suggesting prevalence of shallower water conditions due to lake desiccation. The abundance
49
of ostracodal wackestone-grainstone displaying a mixture of juvenile and adult, articulated
ostracod shells, as well as both articulated and disarticulated fish bones, attests to episodes of
mass mortality. This event was probably associated to the reduction in the lake area due to
shallowing, which is suggested by the fact that only marginal lake deposits display evidence
for ostracod and fish mortality. Fish mass mortality recorded in lacustrine settings of the
Achanarras Middle Old Red Sandstone of Middle Devonian age from the Orcadian Basin in
Scotlant has also been interpreted in a similar way (Trewin, 1986). The reasons that led to this
rapid drought in the study area will be discussed below.
Unit 4 records a return to dominantly deeper water conditions, but without significant
evaporite precipitation and no black shale formation, as indicated by the prevalence of I1
cycles, characterized by deepest-water deposits with only grey/green shales. This unit is
interpreted as deposited during a time when the lake was oxygenated throughout the water
column and no water stratification was present. The thickness of the shallowing-upward cycles
suggests that the lake depth was again as deep as in unit 2. However, a better understanding of
the conditions leading to the formation of unit 2 is precluded due to the strong erosion
associated with the development of the late Aptian/Albian unconformity (Fig. 2-2).
2.6. ORIGIN OF THE SHALLOWING-UPWARD CYCLES
Analysis of the vertical stratal stacking patterns represented by the shallowing-upward
cycles recognized in the Codó Formation is the key to understand their nature. These cycles
were formed as the lake or saline pan/sabkha base level episodically decreased through time.
In this depositional system, drop in base level, with resulting facies progradation, is caused
chiefly from one or the interaction of the following factors: increase in sediment supply either
from a fluvial drainage or a marine inflow, progressively increased aridity, or increase in
subsidence due to syn-sedimentary tectonics.
In the particular case of the Codó Formation, the influence of sediment supply might
be considered negligible, since siliciclastic sands supplied into the lake system was reduced, as
indicated by their scarcity even in marginal deposits. Deciphering whether climate or
subsidence linked with syn-sedimentary seismic activity, was the main cause for progradation
50
is not so straightforward, particularly because both factors might be combined in order to
produce shallowing-upward units (e.g., Anadon et al., 1991).
Climate has been claimed to explain many shallowing-upward lacustrine deposits
recorded in ancient and modern environments (e.g., Smoot & Olsen, 1994; Olsen & Kent,
1996; Hofmann et al., 2000; Aziz et al., 2000; Steenbrink et al., 2000). The small, lower rank
cycles, particularly recognized in the saline pan/sabkha deposits of the Codó Formation,
record minor changes in depositional conditions, attesting to alternations between mud settling
and chemical precipitation of either evaporites or limestones. This characteristic, added to the
rhythmic and regular thickness variation, is consistent with seasonal fluctuations, with
alternating mud deposition and chemical precipitation taking place during less and more arid
phases, respectively. This is specially suggested by the succession of evaporite couplets,
which are attributed to pulsating episodes of water table. Although locally observed, packages
displaying regularly distributed bundles varying from 4 to 6 are probably resulting from
seasonal base level changes.
While seasonal fluctuations may be applied to explain the lower rank cycles recognized
in the Codó Formation, the origin of the intermediate and higher rank cycles are probably
related to syn-sedimentary tectonics. A climatic cause seems to be unlikely in these instances
because, if so, then one would expect an increased precipitation of evaporite minerals during
phases of maximum shallowing, when the lake and saline pan level was at the minimum
(Carroll & Bohacs, 1999). This did not happen in the Codó Formation, neither in the
intermediate cycles, nor in the overall higher rank cycles that form the depositional prograding
units described herein.
The intermediate and higher rank cycles recognized in the Codó Formation show
characteristics that are better explained by a tectonic signature. A line of evidence might be
claimed to suggest tectonics as the main cause for the intermeditate rank cycles: 1. the high
variability of facies within individual cycles; 2. the limited lateral distribution, usually on the
order of less than a few tens of meters; and 3. the frequent and random change in cycle
thickness in the upward direction, ranging from few centimetres up to several meters. These
characteristics match better with those typically recorded from tectonically-influenced,
shallowing-upward cycles (Martel & Gibling, 1991; Benvenuti, 2003), as climatically related
51
cycles are expected to have more regular thicknesses and more monotonous facies distribution
throughout large distances (Hofmann et al., 2000; Harvey, 2003).
The higher rank cycles of the Codó Formation exposed in the Codó area are
particularly suggestive of formation under the influence of syn-sedimentary tectonics. These
units have been correlatable to Grajaú area based on distinctive features of the surfaces
overlying each unit, as formerly discussed in the item 2.5.1. The four higher rank cycles
reflect lower frequency episodes of lake shallowing. The two lowermost units record
progressive episodes of lake shallowing in deep-water, anoxic conditions, when lake
stratification prevailed. Following this phase, progradation of lake shoreline proceeded under
shallower water conditions, as recorded by unit 3. The fossil assemblage preserved in this unit
indicates mass mortality, suggesting that the reduction in the lake area might have been subtle.
Stagnant waters favoured a widespread distribution of microbial mats. A renewed episode of
significant lake deepening took place during deposition of unit 4, which at this time occurred
under dominantly oxidizing conditions.
A tectonic influence is also proposed for the origin of the higher-rank cycles recorded
by deformational structures in the depositional units 1 to 4. This is suggested because these
units have good correspondence with stratigraphic intervals representing different styles of
deformation zones that characterize the Codó Formation in the Codó area. As summarized in a
previous section (see geological setting), these deformation zones (Fig. 2-4) were attributed to
alternating periods of sediment accumulation and shear stress associated with syn-sedimentary
seismic activity linked to the rifting of the São Luís-Grajaú Basin during the Late Aptian
(Rossetti & Góes, 2000). Hence, depositional units 1 and 2 are stratigraphically correlatable
with deformational zones Z1 and Z2 (Fig. 2-10), respectively. In other words, during the two
first episodes of shallowing recorded in the study area, sedimentation in the lake system was
strongly affected first by extensional and, then by compressional forces, as shown by the
upward gradation from small scale cracks and normal faults to complex convolute folds
associated with thrust faults and vertical to sub-vertical stylolites. Deformation would have
initially increased the accommodation space, promoting the development of a thicker higher-
rank shallowing-upward cycle (i.e., unit 1) with increased deeper water facies, as recorded by
a higher volume of C1 and I1 cycles. As the sedimentary succession evolved, the area might
52
have experienced some compression, with consequent local uplift giving rise to the
development of shallower-water facies at the top of depositional unit 2.
Uplift or more stable tectonic conditions would have produced even shallower-water
environments in the Grajaú area, represented by the saline pan/sabkha deposits that are
represented by complete and incomplete cycles of types C3 and I3, respectively. After this
time, a relative stability seems to have prevailed, with the lake basin becoming progressively
shallower due to low accommodation rate and resulting in deposition of unit 3, which is
characterized by the abundance of I2 cycles. A period of extension led to development of
several normal faults and slumpings at the top of unit 3, which corresponds to deformation
zone Z3. As a result of this extension, new accommodation space was created and the lake
system became relatively deeper, favouring deposition of unit 4, characterized by thicker C2
and I1 cycles. However, anoxic conditions no longer existed during this time, as revealed by
the scarcity of organic matter in the shales of this unit. Tectonic activity seems to have
continued, resulting in convolution of these deposits and producing deformation zone Z4.
53
Fig .2-10: A representative lithostratigraphic profile of the Codó Formation showing the good match between
higher-rank shallowing-upward cycles, represented by depositional units 2 to 4, and the syn-sedimentary
deformational zones described by Rossetti and Góes (2000). (See Fig. 2-7 for legend).
The seismic interpretation of the shallowing-upward cycles provided here is in
agreement with the structural framework of the São Luís-Grajaú Basin. As previously
presented, the main rift stage of this basin took place during the Albian, but reactivation of
ancient fault systems started earlier in the Aptian, with the establishment of a shallow, but
widespread basin where the Codó Formation was deposited. The presence of deposits with
evidence for soft sediment deformation in this unit is taken as an indication for syn-
sedimentary tectonics (Rossetti & Góes, 2000). At the basin margins, where the study areas
are located, the displacement of faults with small offsets would have favoured the
54
development of subsiding areas, promoting the establishment of lake systems. The prevalence
of fine- grained and chemical deposits in the depositional setting is not inconsistent with this
model. Tectonically-influenced settings are usually considered to be recorded by a dominance
of coarse-grained siliciclastic deposits. However, modern examples have shown that coarse-
grained deposits will not occur immediately following tectonics. Instead, it has been suggested
that the first response to tectonics in several settings from marine to lacustrine is represented
by fine-grained deposition, as a depositional setting usually takes a time to re-equilibrate and
respond to tectonics (Blair & Bilodeau, 1988). In areas dominated by mild tectonics with fine-
grained and chemical sedimentation, as occur in the Codó Formation, higher rates of
accommodation will give rise to deposition of evaporites and preservation of shales with high
organic content in central lake areas (Carroll & Bohacs, 1999). On the other hand, periods of
quiescence will favour deposition of shallower-water limestones and evaporites as the
accommodation space is reduced. It is possible that the increased sediment reworking,
recorded by the occurrence of gypsarenites and intraclastic grainstones in marginal deposits of
depositional unit 2, might also be related to this tectonic phase. The presence of breccia and
coarse-grained sandstones only at the top of the lowermost higher-rank cycle, represented by
depositional unit 1, is probably an indication for a progressive decrease in the intensity of the
tectonic process through time.
2.7. CONCLUSION
Although more often attributed to climate fluctuations, many shallowing-upward
lacustrine cycles might be resulting from pulsating tectonism taking place contemporaneously
with sediment deposition. The Codó Formation, exposed in the eastern Grajaú Basin, seems to
be an unequivocal example of an ancient lacustrine system displaying two ranks of
shallowing-upward cycles reflecting prograding episodes driven by syn-sedimentary seismic
activity. Despite the seasonal signature recognized in the lower rank cycles, a tectonic cause is
proposed for the intermediate and higher rank cycles described in this unit. In the particular
case of the intermediate rank cycles, a tectonic origin is revealed on the basis of: (1) high
facies variability; (2) limited lateral extension; and (3) frequent and random thickness change.
The higher-rank cycles were also formed as a result of tectonic episodes that alternated with
55
sediment deposition, a conclusion supported by their matching with stratigraphic zones
characterized by different styles of soft sediment deformation that are attributed to
contemporaneous seismic activity. Based on the observations made in the study area, one can
state that extension affecting lake deposits, with subsequent creation of accommodation space,
promotes the development of prograding successions internally formed by thicker (deeper)
water cycles. Bed shortening by compression and/or stability reduces the water depth and lead
to the development of thinner and shallower-water cycles. Therefore, different styles or/and
intensities of seismic pulses alternating with sediment deposition might cause substantial
changes in lake level, promoting alternating deeper and shallower water phases, and ultimately
resulting in cyclic deposition. Deciphering the genesis of such prograding episodes in ancient
lacustrine successions is a task that requires detailed facies analysis and precise mapping of
the stratal stacking patterns, as well as their association with tectonically-driven structures.
Acknowlegement. The Itapicuru Agroindustrial S/A is acknowledged for the permission to
access the quarries with the exposures of the Codó Formation. This work was supported by the
Conselho Nacional de Desenvolvimento Científico e Tecnológico-CNPq (Grant #460252/01).
The authors want to gratefully thank Paulo Milhomem (Petrobras) for the ostracod
identification.
REFERENCES
A
NADON, P., CABRERA, Ll., JULIA, R. & MARZO, M. 1991. Sequential arrangement and
asymmetrical filling the Miocene Rubielos de Mora Basin (northeast Spain). In Lacustrine
Facies Analysis (eds P. Anadón, Ll. Cabrera and K. Kelts), pp. 257-275, International
Association of Sedimentologists Special Publication no. 13. Oxford: Blackwell.
A
ZEVEDO R.P., 1991. Tectonic evolution of Brazilian Equatorial Continental Margin Basins.
Doctoral Thesis, University of London, London, 455 p.
A
ZIZ, H.A., HILGEN, F., KRIJGSMAN, W., SANZ, E. & CALVO, J.P. 2000. Astronomical forcing
of sedimentary cycles in the middle to late Miocene continental Catalayud Basin (NE
Spain). Earth and Planetary Science Letters 177, 9-22.
56
B
ENVENUTI, M. 2003. Facies analysis and tectonic significance of lacustrine fan-deltaic
successions in the Pliocene-Pleistocene Mugello Basin, Central Italy. Sedimentary
Geology 157, 197-203.
B
LAIR
, T.C. & B
ILODEAU
, W.L. 1988. Development of tectonic cyclothems in rift, pull-apart,
and foreland basins: sedimentary response to episodic tectonism. Geology 16, 517-520.
C
ARROLL, A.R. & BOHACS, K. M. 1999. Stratigraphic classification of ancient lakes:
balancing tectonic and climatic controls. Geology 27, 99-102.
C
HAFETZ
, H.S. & B
UTLER
, J.C. 1980. Petrology of recent caliche pisoliths, spherulites (after
Microcodium) and speleothem deposits. Sedimentology 27, 497-518.
F
ERGUSON, J., BUBELA, B. & DAVIES, P.J. 1978. Synthesis and possible mechanism of
formation of radial carbonate ooids. Chemical Geology 22, 285-308.
G
LENN
, C.R. & K
ELTS
, K. 1991. Sedimentary rhythms inlake deposits. In Cycles and events in
stratigraphy (eds G. Einsele, W. Ricken, A. Seilacher), pp. 188-221. Berlin: Spring-Verlag.
G
ÓES
, A.M. & R
OSSETTI
, D.F. 2001. Gênese da Bacia de São Luís-Grajaú, Meio-Norte do
Brasil. In O Cretáceo na Bacia de São Luís-Grajaú (eds D.F. Rossetti, A.M. Góes, W.
Truckenbrodt), pp. 15-29, Coleção Friedrich Katzer. Belém: Museu Paraense Emílio
Goeldi.
H
ANDFORD, C.R. 1991. Marginal marine halite: sabhkas and salinas. In Evaporites, petroleum
and mineral resources (ed J.D. Melvin), pp. 1-66, Developments in Sedimentology 50,
Amsterdam: Elsevier.
H
ARVEY, A.M. 2003. The role of base-level change in the dissection of alluvial fans: case
studies from southeast Spain and Nevada. Geomorphology 45, 67-87.
H
OFMANN, A., TOURANI, A. & GAUPP, R. 2000. Cyclicity of Triassic to lower Jurassic
continental red beds of the Argana Valley, Morocco: implications for paleoclimate and
basin evolution. Palaeogeogeography, Palaeoclimatology, Palaeoecology 161, 229-266.
H
OVORKA, S.D. 1987. Depositional environments of marine-dominated bedded halite,
Permian SanAndres Formation, Texas. Sedimentology 34: 1029-1054.
J
UHÁSZ, E., KOVÁCS, L.Ó., MÜLLER, P., TOTH-MAKK, Á., PHILLIPS, L. & LANTOS, M. 1997.
Climatically driven sedimentary cycles in the Late Miocene sediments of the Pannonian
Basin, Hungary. Tectonophysics 282, 257-276.
57
K
ERR, S.D. & THOMSON, A.1963. Origin of nodular and bedded anhydrite in Permian shelf
sediments, Texas and New Mexico. AAPG Bulletin 47, 1726-1732.
L
OGAN
, B.W. 1987. The MacLeod Evaporite Basin, Western Australia: Tulsa: American
Association of Petroleum Geologists Memoir 44, 140 pp.
M
ARTEL
, A.T. & G
IBLING
, M.R. 1991. Wave-dominated lacustrine facies and tectonically
controlled cyclicity in the Lower Carboniferous Horton Bluff Formation, Nova Scotia
Canada. In Lacustrine Facies Analysis (eds P. Anadón, Ll. Cabrera and K. Kelts), pp. 223-
244, International Association of Sedimentologists Special Publication no. 13. Oxford:
Blackwell.
M
ARTILL, D. 1988. Preservation of fish in the Cretaceous Santana Formation. Paleontology
31, 1-18.
O
LSEN
, P.E. 1986. A 40-million-year lake record of Early Mesozoic orbital climatic forcing.
Science 234, 842-848.
O
LSEN
, P.E. & K
ENT
, D.V. 1996. Milankovitch climate forcing in the tropics of Pangea during
the Late Triassic. Palaeogeogeography, Palaeoclimatology, Palaeoecology 122, 1-26.
O
LSEN, P. E. & D.V. KENT. 1999. Long-period Milankovitch cycles from the Late Triassic
and Early Jurassic of eastern North America and their implications for the calibration of
the early Mesozoic time scale and the long-term behavior of the planets. Trans. Roy. Soc.
Lond. A, 357, 1761-1786.
P
AZ, J.D.S & ROSSETTI, D.F. 2001. Reconstrução paleoambiental da Formação Codó
(Aptiano), borda leste da Bacia do Grajaú, MA. In O Cretáceo na Bacia de São Luís-
Grajaú (eds D.F. Rossetti, A.M. Góes, W. Truckenbrodt), pp. 77-100, Coleção Friedrich
Katzer. Belém: Museu Paraense Emílio Goeldi.
P
INDELL, J.L. 1985. Alleghenian reconstruction and the subsequent evolution of the Gulf of
Mexico, Bahamas and proto-Caribbean. Tectonics 4, 1-39.
R
AISWELL, R. 1976. The microbiaological formation of carbonate concretions in the Upper
Lias of NE England. Chemical Geology 18, 227-244.
R
EZENDE, W.M. & PAMPLONA, A.H.R.P., 1970. Estudo do desenvolvimento do Arco Ferrer-
Urbano Santos. PETROBRAS, Boletim Técnico 13, 5-14.
R
ISACHER, F. & EUGSTER, H.P. 1979. Holocene pisoliths and encrustations associated with
spring-fed surface pools, Pastos Grandes, Bolivia. Sedimentology 26, 253-270.
58
R
OSSETTI, D.F. 2001. Arquitetura deposicional da Bacia de São Luís-Grajaú. In O Cretáceo na
Bacia de São Luís-Grajaú (eds D.F. Rossetti, A.M. Góes, W. Truckenbrodt), pp. 31-46,
Coleção Friedrich Katzer. Belém: Museu Paraense Emílio Goeldi.
R
OSSETTI
, D.F. & G
ÓES
, A.M. 2000. Deciphering the sedimentological imprint of
paleoseismic events: an example from the Aptian Codó Formation, northern Brazil.
Sedimentary Geology 135, 137-156.
R
OSSETTI, D.F., PAZ, J.D.S. & GÓES, A.M. 2004. Facies analysis of the Codó Formation (Late
Aptian) in the Grajaú area, southern Sao Luís-Grajaú Basin. Anais da Academia Brasileira
Ciências, in press.
S
CHREIBER, B.C., SMITH, D. & SCHREIBER, E. 1981. Spring peas from New York State:
nucleation and growth of fresh water hollow ooliths and pisoliths. Journal of Sedimentary
Petrology 51, 1341-1346.
S
HINN, E.A., STEINEN, R.P., LIDZ, B.H. & SWART, R.K. 1989. Whitings, a sedimentologic
dilemma. Journal of Sedimentary Petrology 59, 147-161.
S
MOOT, J.P. & LOWENSTEIN, T.K. 1991. Depositional environments of non-marine evaporites.
In Evaporites, petroleum and mineral resources (ed J.D. Melvin), pp. 189-347,
Developments in Sedimentology 50, Amsterdam: Elsevier.
S
MOOT, J.P. & OLSEN, P.E. 1994. Climatic cycles as sedimentary controls of rift-basin
lacustrine deposits in the Early Mesozoic Newark Basin based on continuous core. In
Lacustrine Reservoirs and Depositional Systems (eds A.J. Lomando, B. Charlotte
Schreibber & P.M. Harris), pp. 239-295, SEPM Core Workshop no. 19, Tulsa: Society for
Sedimentary Geology.
S
PECHT, R.W. & BRENNER, R.L. 1979. Storm-wave genesis of bioclastic carbonates in Upper
Jurassic epicontinental mudstones, East-Central Wyoming. Journal of Sedimentary
Petrology 49, 1307-1322.
S
TEENBRINK, J., VAN VUGT, N, KLOOSTERBOER-VAN HOEVE, M.L., HILGEN, F.J. 2000.
Refinement of the Messinian APTS from sedimentary cycle patterns in the lacustrine Lava
section (Servia Basin, NW Greece). Earth and Planetary Science Letters 181, 161-173.
T
REWIN
, N. H. 1986. Palaeoecology and sedimentology of the Achanarras fish beds of the
Middle Old Red Sandstone, Scotland. Transactions of the Royal Society of Edinburgh:
Earth Sciences 77, 21-46.
59
T
UCKER, M.E. & WRIGHT, V.P. 1990. Carbonate sedimentology. Oxford: Blackwell, 482 pp.
V
ANDERVOORT
, D.S. 1997. Stratigraphic response to saline lake-level fluctuation and the
origin of cyclic non-marine evaporite deposits: the Pleistocene Blanca Lila Formation,
northwest Argentina. Geological Society of America Bulletin 109, 210-224.
V
UGT
, N. V
AN
, S
TEENBRICK
, J., L
ANGEREIS
, C.G., H
ILGEN
, F.J. & M
EULENKAMP
, J.E. 1998.
Magnetostratigraphy-based astronomical tuning of the early Pliocene lacustrine sediments
of Ptolomais (NW Greece) and bed-to-bed correlation with the marine record. Earth and
Planetary Science Letters, Amsterdam 164, 535-551.
W
ARREN
, J. 1999. Evaporites: Their Evolution and Economics. Oxford, Blackwell Science,
438 pp.
W
RIGHT, V.P. 1992. A revised classification of limestones. Sedimentary Geology 76, 177-185.
60
3. PETROGRAPHY OF GYPSUM-BEARING FACIES OF THE
LATE APTIAN CODÓ FORMATION
*
3.1. ABSTRACT
An original and detailed study focusing the petrographic aspects of evaporites from the
Neoaptian Codó Formation exposed in the eastern and southern margins of the São Luís-
Grajaú is presented herein, with the aim to discriminate between evaporites with primary and
secondary origin and reconstruct their post-depositional evolution. Seven phases of evaporite
formation were recognized: 1. chevron (selenite) gypsum; 2. nodular/lensoidal
gypsum/anhydrite; 3. acicular gypsum; 4. mosaic gypsum; 5. brecciated gypsum/gypsarenite;
6. pseudo-nodular anhydrite/gypsum; and 7. rosettes of gypsum. The chevron gypsum, the
nodular/lensoidal gypsum/anhydrite and the brecciated gypsum/gypsarenite display
petrographic characteristics that conform to a primary nature. Their occurrence forming well-
layered, horizontal beds displaying a cyclic arrangement are consistent with this interpretation.
The acicular and mosaic gypsums were formed by replacement of these primary gypsums, but
their formation took place very early in the diagenetic history, being still influenced by the
depositional environment. Noteworthy is that these gypsum morphologies are closely related
to the layered evaporites, serving to demonstrate that their formation were related to
replacements that were mild, not enough to cause any significant change in the primary
*
Authors: J.D.S. Paz & D.F. Rossetti. Submitted to Anais da Academia Brasileira de Ciências
61
sedimentary structures. The pseudo-nodular anhydrite/gypsum seems to have originated due to
intrastratal fluids during burial, probably been related to halokinesis. The rosettes of gypsum,
which intercept all the other variety of gypsum, represent the latest phase of evaporite
formation in the study area, resulting from either intrastral waters or surficial waters during
weathering.
Key-words: evaporite, petrography, paleolake, sabkha, late Aptian, São Luís-Grajaú
Basin
3.2. INTRODUCTION.
Ancient evaporites are economically important as sources for salts (e.g., halite and
potash) and metals (e.g., Cu, Zn, Au), and as indicators for structural traps of oil and gas
(Warren, 1999). These deposits are abundant in sedimentary basins located along the Brazilian
continental passive margin, being particularly developed in the Aptian-Albian transition
(Hashimoto et al., 1987; Uesugui, 1987; Silva-Telles, 1996). The evaporites of the Codó
Formation in the São Luís-Grajaú Basin (Fig. 3-1) are the only ones available for surface
studies in the north equatorial Brazilian margin. Detailed facies analysis of fresh exposures in
several open quarries located in the southern and eastern margins of this basin has provided
elements to support a depositional system representing a lacustrine/sabkha complex (Paz &
Rossetti, 2001; Rossetti et al., 2004). In addition, the deposits exposed in these localities
represent an excellent opportunity to better understand the post-depositional processes that
took place after primary evaporite formation. Such approach concerning to the Aptian
Brazilian evaporites from the north equatorial Brazilian margin has not been conducted yet.
A preliminary field investigation showed that the evaporites from the Codó Formation
include several morphologies, suggesting they might not be all primary in origin. Petrographic
studies of evaporites are useful to help deciphering their origin (e.g., Ogniben, 1955; Kerr &
Thomson, 1963; Holliday, 1970; Arakel, 1980; El-Tabakh et al., 1997; Aref, 1998), thus this
type of study applied to the Codó Formation might contribute to better define and/or
reconsider the proposed lacustrine/sabkha system. Furthermore, petrographic studies can serve
as the basis to determine the sequence of events that took place following deposition (e.g.,
Arakel, 1980; Morad et al., 2000; Pérez et al., 2002). Determining the post-depositional
62
history of these evaporites might help to decide on their potential for future strontium and
sulphur isotopic studies attempting to test the hypothesis of a possible marine influence during
sedimentation, as previously proposed elsewhere (e.g., Batista, 1992; Rodrigues, 1995;
Rossetti et al., 2000).
3.3. GEOLOGICAL SETTING
The Gondwana split up took place through several steps, culminating in the Aptian
with the break up of African and South American continents (Feijó, 1996; Góes & Rossetti,
2001). This process led to final establishment of several rift basins along the equatorial
Brazilian margin, where the São Luís-Grajaú Basin is one of the largest, occupying 150,000
km
2
. This basin has been interpreted as a unique structural feature formed by combination of
pure shear stress and strike-slip deformation (Góes & Rossetti, 2001). The main rifting
developed in the Albian, when fault offsets reached up to 400 m and (Fig. 3-2A,B), but fault
displacement initiated in the Aptian, giving rise to a shallow and widespread basin where the
Codó Formation was deposited in the late Aptian. This unit is defined at the top by a
regionally correlatable unconformity marked by erosion and paleosols (Rossetti et al., 2001).
63
64
Exposures of the Codó Formation include shallowing-upward deposits attributed to
lacustrine or saline pan/sabkha depositional environments (Paz & Rossetti, 2001; Rossetti et
al., 2004; Fig. 3-3). In the Codó area, where a lacustrine system dominates, the shallowing-
upward cycles consist of bituminous black shales, evaporites and, subordinately,
calcimudstones attributed to central lake depositional environments. These deposits grade
upward into gray/green shale and limestones (i.e., calcimudstone, laminated to massive
peloidal wackestone to grainstone, and sparstone), attributed to intermediate lake
environments. The top of the shallowing-upward cycles comprises massive pelite, shale,
limestone (e.g., intraclastic grainstone, ostracodal wackestone/grainstone, ooidal/pisoidal
packstone, tufa) and rhythmite of limestone and shale. These lithologies display with a variety
of sedimentary features consistent with marginal lake deposition, such as paleosol, karstic
surface, fenestrae, meteoric calcite cement, vadose pisoid. In the Grajáu area, a saline
pan/sabkha depositional environment was proposed (Rossetti et al., 2004), being represented
by evaporite and locally tufa, gray-green shale, and calcimudstone.
Fig. 3-3: Depositional model proposed for the Codó Formation in the Codó and Grajaú areas (cf. Paz & Rossetti,
2001).
65
The evaporite facies consists of layered gypsum (formed by alternating darker and
lighter couplets that are up to 15 cm thick), gypsarenite, and massive-macronodular gypsum,
which record deposition in flat lying areas, intraformational reworking of previously deposited
evaporites, and remobilization of salts during burial, respectively. Layered gypsum occurs in
both of the study areas, being dominant in the Grajaú area, while gypsarenite was observed
only in the latter. Massive/macronodular gypsum is widespread in the Codó area, occurring
mostly in the nucleus of evaporite lenses that are up to 5 m thick, which enclose chunks of
black bituminous shales. In the Grajaú area, these deposits form diapiric features that are up to
6-8 m high at the outcrop scale. The contact between the massive-macrogranular gypsum and
the layered gypsum is gradational, except locally where the diapirs are well defined by sharp
boundaries.
3.4. PETROGRAPHY OF THE EVAPORITES
The evaporite facies from the Codó Formation were characterized petrographycally
through the analysis of 86 thin sections distributed along four vertical sections (Fig. 3-4).
Based on morphology and crystal relationships, eight phases of evaporite formation were
recognized, which include (Fig. 3-5): 1. chevron (selenite) gypsum; 2. nodular/lensoidal
gypsum/anhydrite; 3..acicular gypsum; 4. mosaic gypsum; 5. brecciated gypsum/gypsarenite;
6. pseudo-nodular anhydrite/gypsum; and 7 rosettes of gypsum. In addition to these evaporite
phases, an episode of carbonate (calcite/dolomite) cementation/replacement took place in
these deposits.
3.4.1. Chevron gypsum (selenite)
This type of gypsum was recorded in both areas, however it is much more widespread
in sections located in the Grajaú area. The chevron gypsum, which represents the lighter
component of the layered gypsum facies, was described in a previous work (e.g., Rossetti et
al., 2004), and it forms horizontal beds that are up to 10 cm thick of vertically aligned crystals.
Under the microscope, the selenite crystals form twin planes and superimpose growth faces
with acute angles arranged as a zig-zag, perpendicularly to the crystal long axis (Fig. 3-6A).
The selenite crystals usually grew up or are mantled by thin discontinuous layers of shales
66
(mostly smectites). Some thicker layers of chevron gypsum might display packages of crystals
that slumped down into the muds, resulting in a series of segments with superimposed lower
concave-up shapes.
67
68
Fig. 3-5: The several phases of evaporites recognized in the Codó Formation, with paragenesis indicated by
lateral position of the boxes.
69
70
3.4.2. Nodular/lensoidal gypsum/anhydrite
This evaporite was recorded only in sub-surface, occurring in association with shales.
It consists of either nodules of gypsum or isolated lensoidal gypsum crystals that are parallel
to bedding planes (Fig. 3-6B,C). The nodules are up to 5 mm long, and internally composed of
a dark mixture of anhydrite and gypsum that form a massive cryptocrystalline framework.
3.4.3. Acicular gypsum
The acicular gypsum occurs invariably in association with the chevron gypsum,
likewise forming vertically oriented crystals similar to needles (Fig. 3-6D). Although areas
with dominance of acicular gypsum were observed, in general these types of gypsum are
intergraded, resulting in a closely interlaced framework. Noteworthy it is the frequent presence
of relics and/or ghosts of the chevron gypsum within the acicular crystals.
3.4.4. Mosaic gypsum
This type of evaporite occurs particularly within the dark bundles of the layered
gypsum. It consists of inequidimensional crystals of gypsum averaging 300 µm in length that
are arranged into a mosaic framework. The contact between crystals is usually ragged and
sutured. The mosaic gypsum typically displays an abundance of relics of floating anhydrite,
forming a poikilitic texture (Fig. 3-6E).
3.4.5. Brecciated gypsum/gypsarenite
This type of evaporite occurs most frequently intergrading with the mosaic gypsum
within the darker bundles of the layered gypsum facies. It consists of gypsum that occur as
clasts ranging from µm to 1 cm in diameter (Fig. 3-7A,B). The clasts might be either angular
to subangular, when a fitted texture is recognized, or rounded, forming locally a texture that is
similar to gypsarenite (Fig. 3-7C). The clasts are surrounded by either a continuous or
discontinuous dark cutan formed by a mixture of clay, organic matter and iron oxides. Under
crossed nichols, several clasts might display optical continuity, forming large crystals that can
reach up to 3 mm of diameter (Fig. 3-7D). These contain abundant relics of anhydrite.
71
72
3.4.6. Pseudo-nodular anhydrite/gypsum
At the outcrop scale, this evaporite phase consists of either slightly elongated or
spherical fragments up to 5 cm long of evaporites displaying a fitted texture (Fig. 3-8A-C).
Petrographically, the clasts are composed of alabastrine gypsum, fibrous gypsum, and
anhydrite. In general, alabastrine gypsum is the most widespread, occurring as a mosaic of
crystals less than 200 µm in diameter (Fig. 3-8D). The crystals are very limpid, i.e., typically
with no inclusions. The fibrous gypsum occurs secondarily as a series of parallel columns that
can reach up to 1 cm long, which grew within the fractures that define the clasts, being
perpendicular to fracture walls (Fig. 3-8A-C). Unlike to the alabastrine, the fibrous gypsum
might display relics of anhydrite. In addition, ghosts of fibrous crystals may occur within the
alabastrine gypsum (Fig. 3-8D). In places, some elongated clasts are composed of
microcrystalline lath-like anhydrite (Fig. 3-8E). Where these clasts are in contact with mosaics
of large gypsum crystals similar to the ones described above, the latter form concave
embayments toward the anhydrites, leaving a “dust” of anhydrite relics behind (Fig. 3-8E).
Inasmuch, both the anhydrite clasts and the mosaic gypsum might display etched margins as
they grade into fibrous and alabastrine gypsum. Electronic scanning microscopic analysis of
the anhydrite clasts reveal a mixture of anhydrite and limpid gypsum crystals that are similar
to those observed in the nodular/lensoidal gypsum/anhydrite (Fig. 3-9A, compare with Fig. 3-
6C). Noteworthy is that the anhydrite clasts occur as spots that intergrade with alabastrine and
fibrous gypsum.
3.4.7. Rosettes of gypsum
Rosettes of amber-colored gypsum occur disperse within the evaporites of the Codó
area, being restricted to the massive/macronodular gypsum (Fig. 3-8C). The rosettes reach up
to 5 cm of diameter, and consist of fibrous crystals arranged into a radial pattern. The rosettes
unconformably cut into all the other gypsum phases described above.
The carbonates associated with the gypsum/anhydrite consist of calcite cement filling
fractures in the brecciated gypsum/gypsarenite (Fig. 3-9B), even where this facies is
obliterated by pseudo-nodular anhydrite/gypsum. The calcite also occurs as enlarged crystals
that grew upon the gypsum from both sides of the fractures (Fig. 3-9C), in which case it forms
73
caries in the mosaic gypsum. On the other hand, the alabastrine gypsum enters into relics of
calcite, forming caries toward them (Fig. 3-9D). Rhombs of calcite after dolomite are locally
present (Fig. 3-9E). Celestite might be also present in association with the carbonates,
occurring as pyramidal crystals with etched edges.
Fig. 3-8: Textures of the evaporites from the Codó Formation exposed in the study areas. A) Overlay field
drawing illustrating a spot within the pseudo-nodular anhydrite/gypsum with relics of a complex arrangement
formed by mosaic, acicular and fibrous gypsum, bounded by films of calcimudstone. B) Pseudo-nodular gypsum,
formed by fracturing. C) A spot within the pseudo-nodular anhydrite/gypsum, with nodules of anhydrite bound by
films of fibrous gypsum (arrows). Note the superimposed rosettes of gypsum of variable sizes (rg). D) Alabastrine
(gy) and fibrous (gf) gypsum in gradational contacts. The arrows indicate places where the fibrous gypsum is
partly replaced by the alabastrine gypsum, recorded by numerous crystals of the later over the fibrous gypsum,
which in turn remains as diffuse relics. E) Photomicrography of the nodules shown in C, illustrating their
composition of tiny, equant, lath-like crystals (an). Note that the edges of the nodules were replaced by fibrous
(gf) or mosaic (gy) gypsum.
74
Fig. 3-9: Textures of the evaporites from the Codó Formation exposed in the study areas. A) SEM view of the
pseudo-nodular anhydrite/gypsum, with lath-like anhydrite (an) interlaced with gypsum (gy). This material comes
from a nodule of anhydrite depicted in figure 8C. B) Calcite (ca) cementing fractures in the pseudo-nodular
anhydrite/gypsum (gy). C) A detail of the pseudo-nodular anhydrite/gypsum showing several fractures (arrows)
filled by a mixture of calcite and muds. Note larger calcite crystals (ca) that grew sidewards from the fractures
trough replacement of gypsum (gy). D) Relics of calcite (ca) within alabastrine gypsum (gy) from the pseudo-
nodular anhydrite/gypsum. E) Rhombs of calcite after dolomite (arrows). (Except for the SEM micrography
shown in A, all the other figures were obtained under petrographic microscope with crossed nicols).
3.5.PARAGENESIS
The combination of facies and petrographic characteristics revealed that the foregoing
described evaporites show features supporting both primary and secondary formations (Fig. 3-
10). In addition, the petrographic study also showed that at least great part of the secondary
gypsum might have been formed under influence of the depositional surface, thus reflecting
75
Fig. 3-10: Model to explain the formation of primary and early diagenetic evaporite phases recorded in the Codó
Formation.
the original brine characteristics. Evidences for primary evaporite precipitation is particularly
recorded in the layered gypsum. First, this is suggested by the preservation of sedimentary
structures forming laterally continuous, horizontal beds. Second, the primary nature is
confirmed by the presence of chevron gypsum in the lighter bundles of this facies. Such
feature is attributed to the progressive upward precipitation of salts on the floor of shallow
(usually less than 2 m deep), supersatured brine pools (e.g., Logan, 1987; Hovorka, 1987;
Handford, 1991; Smoot & Lowenstein, 1991), as suggested in a previous work (i.e, Rossetti et
al., 2004).
76
recorded in the layered gypsum. First, this is suggested by the preservation of sedimentary
structures forming laterally continuous, horizontal beds. Second, the primary nature is
confirmed by the presence of chevron gypsum in the lighter bundles of this facies. Such
feature is attributed to the progressive upward precipitation of salts on the floor of shallow
(usually less than 2 m deep), supersatured brine pools (e.g., Logan, 1987; Hovorka, 1987;
Handford, 1991; Smoot & Lowenstein, 1991), as suggested in a previous work (i.e, Rossetti et
al., 2004).
The cyclic alternation of selenite beds with dark gypsum containing nodular/lensoidal
gypsum/anhydrite is also a reflex of changes in the depositional conditions. The formation of
these morphologies is consistent with accumulation taking place a few mm beneath the
depositional surface by displacive intrasediment growth of crystals from supersaturated pore
fluids in the capillary and/or upper phreatic zone (e.g., Kerr and Thomson 1963, Warren
1999). As the brine level decreased, nodular/lensoidal gypsum grew displacively within the
underlying sediments, here represented by shales. This is the most common habit of gypsum
precipitated within sediment, either in mudflats or other environments subjected to palustrine
conditions (Magee, 1991).
The presence of nodules composed of a dark mixture of anhydrite and gypsum forming
a massive cryptocrystalline framework is consistent with this interpretation, being related to
increased evaporation, probably due to climatic changes (Rossetti et al., 2004). Similar masses
of mixing gypsum have been considered as the record of sulphate replacements during
seasonal variations next to the depositional surface (p.e., Arakel, 1980; Mees, 1998; Mees &
Stoops, 2003).
The mosaic gypsum was formed by replacement of anhydrite, as indicated by the fact
the abundant remains of this mineral within it. Mosaics of gypsum crystals with sutured
contacts have been interpreted as a non-equilibrium texture of grain interpenetration at low
temperature (cf. Voll, 1960), probably reflecting formation under early diagenesis (e.g.,
Spencer & Lowenstein, 1990). The occurrence of mosaic gypsum in association with the dark
bundles alternated with selenite/acicular gypsum within the layered gypsum might be taken as
an evidence for replacement soon after deposition.
Extreme low brine level with subaereal exposure resulted in gypsum fracturing due to
desiccation, a process that gave rise to in situ brecciated gypsum and ultimately gypsarenite,
77
the latter recording local reworking at the surface as recorded in other evaporite deposits (e.g.,
Sanz-Rubio et al., 1999; Schreiber & El Tabakh, 2000). The frequent upward gradation from
the brecciated to the gysarenite is consistent with this interpretation. The presence of mud
cutans surrounding the clasts is attributed either to mechanical infiltration from downward
flows or to the adhesion of residues on clast surfaces during reworking.
Following brecciation, there was a phase of widespread replacement, which gave rise
to mosaics of large gypsum crystals affecting the dark beds of the layered gypsum facies. This
sequence of event is proposed based on the observation that large gypsum crystals encompass
several clasts, attesting to pervasive cementation and/or replacement. The presence of
abundant relics of microcrystalline lath-like anhydrite within the mosaics attests to their
formation following a period of anhydritization. Noteworthy is the fact that the mosaics were
developed only in the darker bundles of the layered gypsum, not affecting the lighter
components. Instead, the chevron gypsum was replaced by acicular gypsum, as indicated by
the presence of ghosts of selenite within the latter. This morphology reflects crystal growth
from very pure, supersaturated fluids, which favors extreme elongation parallel to the c-axis
(Magee, 1991). The close association of the acicular gypsum with the selenites leads to
suggest that formation fluids were driven from remobilization of sulphates from the selenites,
whose formation naturally required high brine saturation. It is important to mention that
neither of these authigenic processes were enough to destroy the primary lamination. This is
taken as evidence to suggest that the formation of mosaic and acicular gypsum took place
cyclically close to the depositional surface just shortly after formation of the individual
bundles.
Fracturing at the surface created a porosity that was cemented by calcite. Calcite also
replaced gypsum near the fracture sides. Under subaerial conditions (i.e., vadose to freshwater
phreatic), sulphate-undersaturated pore fluids dissolves gypsum and/or anhydrite and release
Ca
2+
for precipitation of calcite as the CO
3
2-
has more affinity with calcium than with SO
4
2-
(cf. Back et al., 1983). Rare dolomite rhombs might have been formed either by replacement
of calcite or simultaneous dolomite precipitation. The dolomite was in turn replaced by calcite.
Dedolomitization is a process that might be closely linked to karstification (Cañaveras et al.,
1996). This process might have contributed to release Sr
2+
, which combined with SO
4
2-
,
promoted precipitation of celestite (cf. Olaussen, 1981; Taberner et al., 2002). The celestite
78
formation in turn also releases Ca
2+
that increases the Ca/Mg ratio, and could dissolve
dolomite and renew precipitation of calcite (p.e., Back et al., 1983; Kushnir, 1985). The close
association of calcite, dolomite and celestite in the study evaporites of the Codó Formation
suggest these processes as the most likely.
The pseudonodular anhydrite/gypsum is interpreted to represent a later phase of
gypsum formation. First, this is suggested by its occurrence restricted to the
massive/macronodular gypsum, where primary sedimentary structures were almost entirely
lost. The presence of massive gypsum grading into the layered gypsum, and the diapiric
geometry enclosing chunks of black bituminous shales is related to salt remobilization during
halokinesis. Second, the pseudonodular anhydrite/gypsum contains limpid alabastrine and
fibrous crystals, not observed in association with the layered gypsum, which is consistent with
the proposed late formation. Fine crystalline gypsum have been considered as secondary in
origin (Holliday, 1970), forming in consequence of hydration, especially of anhydrite, induced
by diapirism or other mechanism that allow percolation of water through evaporite rocks
(Holliday, 1970; Price & Cosgrove, 1990; Warren, 1999). Fracturing seems to have been the
cause of the pseudonodular aspect of this gypsum. Under stress, probably related to salt
remobilization, part of the gypsum might have had a ruptile behavior, ultimately braking apart
to form individual fragments (e.g., Price & Cosgrove, 1990; Marco et al., 2002). Saturated
fluids percolating along the secondary porosity created by this process would have promoted
precipitation of the fibrous gypsum. In the following, there might have had the enlargement of
the fractures due to forcing caused by crystal growth. While this process took place, most of
the primary sedimentary features, as well as the previous phases of evaporites, became
obliterated. As the salt became mobile, there was a pervasive development of alabastrine
gypsum. The later nature of this phase of gypsum is revealed by its limpid aspect free of
anhydrite inclusions and by the fact that it contains ghosts of fibrous gypsum.
The spots with microcrystalline lath-like anhydrite that occur within the macronodular
gypsum is probably a relic of one of the earliest phase of evaporite precipitation in the study
area The paragenesis reconstituted with basis on petrographic relationships shows that these
anhydrites form nodules that were in part replaced by the mosaic gypsum, as revealed by the
fact that the later form concave embayments that enter into the anhrydrite. The margins of this
gypsum is in turn ragged due to reaction with fibrous and alabastrine gypsum. These
79
relationships support that the anhydrite nodules had an early development relative to all other
gypsum phases present in the pseudonodular gypsum. If so, then it is possible that the
anhydrite is temporally related to the anhydrite formed in the dark bundles of the layered
gypsum, which represents one of the earliest evaporite phases developed in the Codó
Formation.
The final phase of evaporite formation is recorded by the rosettes of gypsum, as
confirmed by the fact that these truncate all the other evaporite morphologies. Aggregates of
large fibrous gypsum crystals forming rosettes similar to the ones of the study area have been
attributed to the action of either intrastratal waters during burial or surficial waters during
weathering (e.g., Shearman, 1966; Holliday, 1970; Warren, 1999).
3.6.CONCLUSIONS
Despite the many varieties of evaporites recorded in the Codó Formation exposed in
the eastern and southern margin of the Grajaú Basin, these deposits display many petrographic
and faciological attributes that are consistent with an early formation either by primary
precipitation or early replacements when the sediments were still under the influence of the
depositional surface. These types of evaporites prevail particularly in the Grajaú area, where
layered gypsum is the dominant facies. In both areas, burial phases of gypsum seem to have
developed only where gypsum was remobilized during halokinesis.
The lack of significant deep diagenetic modification of the Codó Formation is recorded
also by studies focusing the limestones and shales interbedded with the evaporites. The
limestones are dominated by lithologies consisting of calcimudstones and peloidal
wackestones/packstones with only local evidences of cementation or replacement. These
lithologies usually display primary features, as normal grading and horizontal crenulated
lamination. Inasmuch, the shales associated with the evaporites consist almost entirely by
smectite, with only subordinate kaolinite and illite. Among these clay minerals, smectite is far
the dominant one, being represented by detrital flakes, while the kaolinite and illite are mostly
authigenic but related to pedogenetic horizons (Gonçalves, 2004). These data are consistent
with the proposition that burial did not cause significant textural or mineralogical modification
of the Codó Formation.
80
The petrographic studies presented here strongly motivate to undertake isotopic studies
in the evaporites of the Codó Formation. The analysis should be carried out using samples
from the layered facies only, which preserves evaporites formed both primarily and shortly
after deposition. For these reasons, these deposits should provide information of the original
brine characteristics. On the other hand, the pseudo-nodular anhydrite/gypsum and the rosettes
of gypsum should not be considered in this type of studies, as they represent later stages of
gypsum formation due to deeper salt remobilization. Inasmuch, the spots of anhydrite within
the pseudo-nodular anhydrite/gypsum should be also discarded in these analyses. Despite the
interpretation that these evaporites might have formed contemporaneously to the early-formed
nodular/lensoidal anhydrite, the microscopic studies revealed they were strongly replaced by
limpid, alabastrine gypsum, being inappropriate for isotopic analysis that can be used for
paleoenvironmental purposes.
Acknowledgements. The Itapicuru Agroindustrial S/A is acknowledged for the
permission to access the quarries with the exposures of the Codó Formation. This work was
financed by the Brazilian Council for Research–CNPq (Project #460252/01).
REFERENCES
Arakel, A.V.1980. Genesis and diagenesis of Holocene evaporitic sediments in Hutt and
Leeman lagoons, western Australia. J. Sedim. Petrol. 50: 1305-1326.
Aref, M.A.M. 1998. Holocene stromatolites and microbial laminites associated with lenticular
gypsum in a marine-dominated environment, Ras El Shetan Ara, Gulf of Aqaba, Egypt.
Sedimentology 45: 245-262.
Back, W., Hanshaw, B. B., Plummer, L. N., Rahn, P. H., Rightmire, C. T. and Rubin, M.,
1983, Process and rate of dedolomitization: mass transfer and
14
C dating in a regional
carbonate aquifer: G.S.A. Bull. 94: 1414-1429.
Batista, A.M.N. 1992. Caracterização Paleoambiental dos sedimentos Codó-Grajau, Bacia de
São Luís (MA). Belém, 102 p. (Tese de Mestrado, UFPA)
Cañaveras, J.C., Sánchez Moral, S., Calvo, J.P., Hoyos, M. & Ordoñez, S. (1996):
Dedolomites associated with karstification: An example of early dedolomitization in
81
lacustrine sequences from the Tertiary Madrid Basin, Central Spain. Carbonates and
Evaporites 11: 85-103.
El-Tabakh, M., Riccioni, R., Schreiber, B.C., 1997. Evolution of Late Triassic rift basin
evaporites (Passaic Formation); Newark Basin, eastern North America. Sedimentology 44:
767-790.
Feijó, F.J. 1996. O início da livre circulação das águas do Oceano Atlântico. Bol. Geoc.
PETROBRAS 10: 157-164.
Góes, A.M. & Rossetti, D.F. 2001. Gênese da Bacia de São Luís-Grajaú, Meio-Norte do
Brasil. In D.F. Rossetti, A.M. Góes, W. Truckenbrodt (Eds) O Cretáceo na Bacia de São
Luís-Grajaú. Belém: Coleção Friedrich Katzer. Belém: Museu Paraense Emílio Goeldi. p.
15-29.
Gonçalves, D. F. 2004. Argilominerais da Formação Codó (Aptiano superior) – Bacia do
Grajaú: implicações climáticas e ambientais. Belém, 100 pp. (Tese de Mestrado, UFPA)
Handford, C.R. 1991. Marginal marine halite: sabkhas and Salinas. In: J.L. Melvin (Ed.)
Evaporites, petroleum and mineral resources. Amsterdan: Elsevier, Developments in
Sedimentology 50: 1-66.
Hashimoto, A.T., Appi, C.J., Soldan, A.L. & Cerqueira, J.R. 1987. O Neo-Alagoas nas bacias
do Ceará, Araripe e Potiguar (Brasil): caracterização estratigráfica e paleoambinetal.
Revista Brasileira de Geociências 17: 118-122.
Holliday, D.W. 1970. The petrology of secondary gypsum rocks: a review. J. Sedim. Petrol.
40: 734-744.
Hovorka, S.D. 1987. Depositional environments of marine-dominated bedded halite, Permian
SanAndres Formation, Texas. Sedimentology 34: 1029-1054.
Kasprzyk, A. & Ortí, F. 1998. Paleogeographic and burial controls on anhydrite genesis: tha
Badenian Basin in the Carpathian Foredeep (southern Poland, western Ukraine).
Sedimentology 45: 889-907.
Kerr, S.D. & Thomson, A. 1963. Origin of nodular and bedded anhydrite in Permian shelf
sediments, Texas and New Mexico. AAPG Bull 47: 1726-1732.
Kushnir, S.V. 1985. The epigenetic celestite formation mechanism for rocks containing
CaSO
4
. Geokhimiya 10: 1455-1463. (Translation from Scripta Technica, Inc., 1986 ©)
82
Logan, B.W. 1987. The MacLeod evaporite basin, western Australia. A.A.P.G. Memoir 44,
140p.
Magee, J.W. 1991. Late Quaternary lacustrine, groundwater, Aeolian and pedogenic gypsum
in the Prungle Lakes, southeastern Australia. Palaeogeography, Palaeoclimatology,
Palaeoecology 84: 3-42.
Marco, S.,Weinberger, R, & Agnon, A. (2002). Radial fractures formed by a salt stock in the
Dead Sea Rift, Israel. Terra Nova 14:288-294.
Mees, F. 1998. The alteration of glauberite in lacustrine deposits of the Taoudenni-Agorgott
basin, northern Mali. Sedimentary Geology 117: 193-205.
Mees, F. & Stoops, G. 2003. Circumgranular bassanite in a gypsum crust from eastern Algeria
– a potential palaeosurface indicator. Sedimentology 50: 1139-1145.
Morad, S., Ketzer, J.M. & De Ros, L.F. 2000. Spatial and temporal distribution of diagenetic
alterations in siliclastic rocks: implications for mass transfer in sedimentary basins.
Sedimentology, Supple.1, 47: 95-120.
Ogniben, L. 1955. Inverse graded bedding in primary gypsum of chemical deposition. J. Sed.
Petrol. 25: 273-281.
Olaussen, S. 1981. Formation of celestite in the Wenlock, Oslo region Norway – evidence for
evaporitic depositional environments. J. Sed. Petrol. 51: 37-46.
Paz, J.D.S & Rossetti, D.F. 2001. Reconstrução paleoambiental da Formação Codó (Aptiano),
borda leste da Bacia do Grajaú, MA. In: D.F. Rossetti, A.M. Góes, W. Truckenbrodt (Eds)
O Cretáceo na Bacia de São Luís-Grajaú. Belém: Coleção Friedrich Katzer, Museu
Paraense Emílio Goeldi, p. 77-100.
Price, N. J. & Cosgrove, J. W. 1990. Analysis of Geological Structures. Cambridge, 502 p.
(pp. 89-122)
Pérez, A., Luzón, A., Roc, A.C., Soria, A.R., Mayayo, M.J., Sánchez, J.A. 2002. Sedimentary
facies distribution and gênesis of recent carbonate-rich saline lake: Gallocanta Lake,
Iberian Chain, NE Spain. Sedimentary Geology 148: 185-202.
Rodrigues, R. 1995. A geoquímica Orgânica na Bacia do Parnaíba. Porto Alegre, 225 pp.
(Tese de Doutorado, UFRGS).
83
Rossetti, D.F. 2001. Arquitetura deposicional da Bacia de São Luís-Grajaú. In: D.F. Rossetti,
A.M. Góes, W. Truckenbrodt (Eds) O Cretáceo na Bacia de São Luís-Grajaú. Belém:
Coleção Friedrich Katzer, Museu Paraense Emílio Goeldi, p. 31-46.
Rossetti, D.F., Paz, J.D.S., Góes, A.M. 2004. Facies analysis of the Codó Formation (Late
Aptian) in the Grajaú Area, Southern São Luís-Grajaú Basin. An. Acad. Brás. Ciên. 76:
791-806.
Rossetti, D.F., Paz, J.D.S., Góes, A. M. & Macambira, M. 2000. A marine versus non-marine
origin for the Aptian-Albian evaporites of the São Luís and Grajaú basins, Maranhão State
(Brazil) based on sequential analysis. Rev. Bras. Geocienc. 30: 642-645.
Sanz-Rubio, E., Hoyos, M., Calvo, J.P. & Rouchy, J.M. 1999. Nodular anhydrite growth
controlled by pedogenic structures in evaporite lake formations. Sedimentary Geology
125: 195-203.
Schreiber, B.C. & El Tabakh, M. 2000. Deposition and early alteration of evaporites.
Sedimentology, Suppl. 1, 47: 215-238.
Shearman, D.J. 1966. Origin of marine evaporites by diagenesis. Inst. Min. Metal. Trans.,
Section B, 75: 208-215.Smoot, J.P. & Lowenstein, T.K. 1991, Depositional environments
of non-marine evaporites. In: JL. Melvin (Ed.) Evaporites, petroleum and mineral
resources, Amsterdam: Elsevier, Developments in Sedimentology 50: 189-347.
Silva-Telles Jr., A.G. 1996. Estratigrafia de Seqüências de Alta Resolução do Membro
Coqueiros da Formação Lagoa Feia (Barremiano?/Aptiano da Bacia de Campos-Brasil)
Porto Alegre, 2v., 268 pp. (Dissertação de Mestrado, UFRGS).
Smoot, J.P. & Lowenstein, T.K. 1991. Depositional environments of non-marine evaporites.
In: JL. Melvin (Ed.) Evaporites, petroleum and mineral resources, Amsterdam: Elsevier,
Developments in Sedimentology 50: 189-347.
Spencer, R.J. & Lowenstein, T.K. 1990. Evaporites. In I.A. Macilreath & D.W. Morrow (eds.)
Diagenesis. St. Johns, New Foundlands, Geological Association of Canada, Geosciences
Canada Reprint Series, 4, p.141-163.
Taberner; C., Marshall, J.D., Hendry, J.P., Pierre, C.; Thirlwall, M.F. 2002. Celestite
formation, bacterial sulphate reduction and carbonate cementation of Eocene reefs and
basinal sediments (Igualada, NE Spain). Sedimentology 49: 171-190.
84
Uesugui, N. 1987. Posição estratigráfica dos evaporitos da Bacia de Sergipe-Alagoas. Rev.
Bras. Geociênc. 17: 131-134.
Voll, G. 1960. New work on petrofabrics. Liverpool and Manchester Geological Journal 2:
503-567.
Warren, J. 1999. Evaporites: Their Evolution and Economics. Oxford: Blackwell Science,
438p.
85
4. GENESIS AND PALEOHYDROLOGY OF A SALINE PAN/LAKE
SYSTEM (LATE APTIAN) FROM THE BRAZILIAN EQUATORIAL
MARGIN: INTEGRATION OF FACIES, SR AND S ISOTOPES
*
4.1. ABSTRACT
Facies analysis was combined with Sr and S isotope data to unravel the brine source of
Late Aptian evaporites from the Codó Formation exposed in the southern and eastern margins
of the São Luís-Grajaú Basin, northern Brazil. Comparisons of facies distribution between the
two investigated areas show: 1. stable, well-stratified and hypersaline lakes with periods of
anoxia and closure prevailing in the eastern margin of the basin, with salt precipitation only in
more saturated, central basin environments; and 2. prevalence of relatively more ephemeral
conditions to the south, where a saline pan complex developed and evaporite precipitation
took place mainly in marginal salinas and surrounding mudflats. In both areas,
expansion/contraction cycles were formed as sedimentation took place, which was followed
by decrease, and then increase, in isotope values. This, combined with the wide dispersion of
Sr and S isotope data within individual depositional cycles, as well as petrographic and
scanning electronic microscopic data, led to the conclusion that diagenesis in some of the
examined facies (i.e., laminated argillite and gypsarenite) was not enough to modify the
primary texture or the geochemical signature. This is because the diagnetic processes took
*
Authors: J.D.S. Paz, D.F. Rossetti & M. Macambira. Submitted to Sedimentology
86
place shortly after deposition or even penecontemporaneously, thus the newly formed minerals
still keep the primary signature, thus serving for paleoenvironmental purposes. A non-marine
brine source is suggested by
87
Sr/
86
Sr ratios ranging from 0.707824 to 0.709280, which are
higher than those from Late Aptian seawaters (i.e., between 0.70720 and 0.70735). The δ
34
S
varies from 16.12 to 17.89 %
o(
V-CDT)
in the eastern margin of the basin, which is in
disagreement with Late Aptian marine values (13 to 16 %
o(
V-CDT)
)
Both geochemical tracers
were influenced by facies characteristics, and thus a model is provided, where expansion of
saline pan/lake systems led to decreasing
87
Sr/
86
Sr values due to the release of
87
Sr from clay
minerals by internal draining of mud flats. During expansion peaks, the
87
Sr/
86
Sr values were
lower because this process is cut off due to submergence of mud flats, added to the
introduction of external
87
Sr-depleted waters resulting from weathering of Permian to
Neocomian marine limestones and evaporites, as well as Triassic to Neocomian basaltic rocks.
Furthermore, the sulphur isotope values decrease in the southern margin of the basin to a range
that varies from 14.79 to 15.60 %
o(
V-CDT)
probably due to increased evaporation in shallower
water settings.
Keywords: Sr and S isotopes; facies; mineralogy; Late Aptian; evaporites; continental
paleoenvironment; Northern Brazil.
4.2. INTRODUCTION
Distinguishing between marine and nonmarine evaporites has been always challenging
(Hardie, 1984; Brookins, 1988; Hovorka et al., 1993; Plàya et al., 2000), especially in places
where the associated deposits lack diagnostic features to help reconstructing the depositional
environments. This problem increases when evaporites evolve from hybrid brines, which is
probably the most likely in many cases (Denison et al., 1998). Lack of sedimentological
parameters has motivated the use of Sr and S isotopes to help investigating the genesis of
many evaporite deposits throughout the world (e.g., Denison et al., 1998; Hovorka et al.,
1993; Playà et al., 2000;
Schreiber & El Tabakh, 2000). Both elements are abundant and
uniformly distributed in seawater, displaying isotope ratios that vary through time in a known
manner (e.g., Burke et al., 1982; Bralower et al., 1997). Once age is well established, these
isotope tracers might provide reliable information on brine sources. Excursions from the
87
established seawater curve are taken as indicators of continental or hybrid brines (Claypool et
al., 1980; Burke et al., 1982), as long as other influences such as diagenesis can be discarded.
Late Aptian Codó deposits from northern Brazil (Fig. 4-1A) contains evaporites
formed throughout the last stages of the Gondwana break-up, which ultimately led to opening
of the South Atlantic Ocean. Determining whether the evaporitic brine evolved in a marine or
non-marine setting has important implications for the paleogeographic modeling of the initial
Brazilian Equatorial Margin. Hence, facies analysis and isotope geochemistry based on
87
Sr/
86
Sr and
34
S/
32
S were applied to help ascertain the genesis of evaporites in the Codó
Formation exposed in the eastern and southern margins of the Grajaú Basin, near the towns of
Codó and Grajaú. Based on sequence stratigraphic modelling (Rossetti, 2001), the first
basinward marine transgression in this basin took place during the Albian when rifting fully
developed. However, the occurrence of evaporites up to the basin margins has led to argue a
widespread marine incursion as early as the Late Aptian (Batista, 1992; Rodrigues, 1995). Sr
and S isotope analysis might help to better define the origin of the evaporitic brine source, and
hence determine when the inland areas of the basin experienced the first marine incursion,
narrowing the estimate of the time when the Brazilian Equatorial Margin established an ocean
connection. In addition, previous facies analysis has attributed the deposits studied herein to a
cyclic record formed by episodes of expansion and contraction of saline pan/lake systems (Paz
& Rossetti, 2001). This study might also contribute to analyse the behavior of Sr and S
isotopes relative to facies changes, an approach not extensively focused in the literature yet,
but which is of great significance for reconstructing past hydrologic patterns.
4.3. MATERIAL AND METHODS
Sedimentary facies information available from previous field studies was combined
with isotopic analysis of 36 samples from the Codó Formation collected along fresh quarry
exposures. Isotopic analyses were carried out on whole-rock gypsum (eventually anhydrite)
from selected facies. The samples were subdivided into three aliquots. The first one was
disaggregated and the sulphate minerals picked from dried material. About 10 mg of powered
sample was dissolved in 2.5 ml HCl (2.5 N) and heated at 70
o
C during 12 hours to allow
complete evaporation. Then, 1 ml HNO
3
(3.5 N) was added in order to separate the Sr cation.
88
The
87
Sr/
86
Sr isotopic analyses were performed on mass spectrometers from the
Geochronology Laboratory of the Universidade Federal do Para (Brazil). The ratios were
normalized to
88
Sr/
86
=0.1194 in order to correct any mass discrimination, with results
indicating mean values of 0.704120+-30x10
-6
for the NBS 987 standard.
89
The second sample aliquot (circa 1.5 g) was used to quantify Rb by X-ray fluorescence
(XRF), to analyse the contribution of
87
Rb to the
87
Sr/
86
Sr ratio. Rb/Sr ratios vary between
0.01 and 0.02, indicating no need for corrections due to decay contribution (cf. Clauer, 1976).
The third sample aliquot (0.3 mg) was used for δ
34
S determination carried out in the
Iso-Analytical Ltd Laboratory, England. Vanadium pentoxide catalyst was added to the
sample in a tin capsule, placed in an automatic sampler and combusted at up to 1700° C. The
combusted gases were then swept in a helium stream over combustion catalysts (tungstic
oxide/zirconium oxide) and through a reduction stage of high purity copper wires to produce
SO
2
, N
2
, CO
2
, and water. Water was removed using a Nafion™ membrane. Sulphur dioxide
was resolved from N
2
and CO
2
on a packed GC column at a temperature of 30 °C. NBS-127
(barium sulphate, δ
34
S
V-CDT
= +20.3 ‰) distributed by the IAEA, Vienna. NBS-127, IAEA-S-
1 (silver sulfide, δ
34
S
V-CDT
= -0.3 ‰) and Iso-Analytical IA-BaSO
4
(barium sulphate, δ
34
S
V-
CDT
= +11.0 ‰) were used for calibration and correction.
Petrographic analysis of all evaporite samples and, whenever necessary, scanning
electron microscopy were performed to provide the best constraints in the evaluation of
possible diagenetic changes of the Sr and S values during burial, and thus decide on their
potential use as paleohydrological indicators.
4.4. GEOLOGICAL FRAMEWORK AND PALEOENVIRONMENTAL CONTEXT
The sedimentary fill of the São Luís-Grajaú Basin consists of three 2
nd
-order
sedimentary sequences (Fig. 4-1B) deposited from the Late Aptian to Quanternary (Rossetti,
2001). The Codó Formation developed during the lowstand stage of the oldest sequence. The
prevalence of the polen Sergipea variverrucata in the deposits of the study area confirms their
Late Aptian age (Batista, 1992; Antonioli, 2001; Rossetti et al., 2001). This unit records the
first deposits accumulated in a broad and shallow depression formed by mild tectonic
stretching before the main Albian rifting stage. These Late Aptian deposits are represented
throughout the basin by deltaic, eolian, lacustrine, as well as transitional to shallow marine
deposits. A sharp discontinuity surface with a paleosol horizon occurs between the Late
Aptian and the Albian deposits (Rossetti et al., 2001; Fig. 4-2A).
90
Fig . 4-2: A) A view of the unconformity between the Codó Formation and the overlying Albian deposits of the
Itapecuru Group. (Gy= gypsum; Lsh=interbedded limestone and shale). B) Lithostratigraphic profile
representative of the Codó Formation in the study area, with indication of the shallowing upward cycles formed
by central lake (1), intermediate lake (2) and marginal lake (3) deposits. The letters to the left locate figures A-F.
C) Bituminous shale (black areas) interbedded with evaporite (white areas) from central lake deposits. D)
Interbedded limestones (lm) and shales (darker areas from intermediate lake deposits. E) Interbedded lime-
mudstone (Ml) and pisoidal limestones (Pl) and F) Rhythmite of lime mudstone (white beds) and microbial mats
(black beds) from marginal lake deposits.
91
In the study area, the Late Aptian Codó Formation consist ofs a 25 m-thick lacustrine
prograding succession typically formed by shallowing-upward cycles (Fig. 4-2B). The base of
these cycles includes evaporites and bituminous black shales attributed to central lake deposits
(Fig. 4-2C). Pyrite and native sulphur are locally abundant in the shales. The central lake
deposits progressively grade upward into laminated argillite interbedded with limestones
(calcimudstone, laminated to massive peloidal wackestone to packstone and sparstone),
attributed to intermediate lake environments (Fig. 4-2D). Bioturbation is locally present, as are
symmetrical ripple marks. Ostracods represent circa 15-25% of the grains in the peloidal
limestones, which may contain microbial mats. The top of the shoaling-upward cycles
comprises a variety of intergrading lithofacies, including: 1) massive blocky pelite that varies
upward from olive-green to brownish-red colors; 2) fenestral calcarenite formed by calcite
grains; 3) ostracodal limestone (wackestone to grainstone); 4) pisoidal packstone (Fig. 4-2E)
with elongated and agglutinating pisoids of various sizes interlaminated with microbial mats;
5) tufa; and 6) ostracodal limestone/shale rhythmite with abundant microbial mats (Fig. 4-2F).
These deposits show an abundance of sedimentary features (i.e., paleosols, karstic surfaces,
fenestrae, meteoric cements, vadose pisoids) typical of subaerial and/or meteoric exposure,
consistent with their deposition in marginal lake environments. Most of the described
sedimentary facies are also observed in the Grajaú area. However, in this locality evaporites
dominate over the other facies, being represented by layered gypsum and gypsarenite.
The shallowing-upward cycles of the Codó Formation are arranged into a higher
category of lower frequency cycles that are bounded by sharp discontinuity surfaces (Fig. 4-
2B). At the base of these lower frequency cycles the superimposed shallowing-upward cycles
are characterized by a higher volume of central lake deposits, while in its top shallowing-
upward cycles with increased volume of marginal lake deposits dominate. It is noteworthy the
presence of lags of intraformational calcimudstone and shales, clastic gypsum, as well as
rhythmites with abundant microbial mats, at the base of the lower frequency cycles (Fig. 4-
2B). The origin of these cycles is attributed to expansion and contraction of the lake system,
probably related to syn-sedimentary seismic activity (Paz, 2000).
92
4.5. CHARACTERIZATION OF THE EVAPORITES
4.5.1. Description
The evaporites of the Codó Formation, represented by gypsum and locally anhydrite,
occur as discontinuous lenses that are up to 4.5 m thick and circa 300 m long at the outcrop
scale. In the Codó area, these deposits are interbedded with black shales, and exclusively
occur in deeper central lake deposits, while in the Grajaú area they are linked to facies
associations typical of shallower waters with evidence for exposure to meteoric and/or vadose
conditions and pedogenesis.
The evaporites consist of three facies: layered gypsum, gypsarenite, and
massive/macronodular gypsum. Layered gypsum dominates, and is comprised by laterally
continuous, alternating dark/light horizontal beds that vary from few mm up to 10 cm thick
(Fig. 4-3A-C). The dark beds are formed by either crystals or micronodules of gypsum less
than 0.5 cm long distributed within a matrix of black shale. The light beds consist of upward-
oriented fibrous gypsum crystals. The gypsarenite (Fig. 4-3C) is interbedded with the layered
gypsum, forming layers few cm thick of moderate to poorly sorted, rounded to sub-rounded
very coarse to pebbly gypsum grains. The massive/macronodular gypsum forms unstructured
bodies (Fig. 4-3D) that are closely intergraded with macronodular gypsum consisting of
centimeter-sized gypsum nodules, with or without a shale matrix. The gypsum nodules display
an almond-like form defined by a web of undulating, horizontal to oblique fractures. Gypsum
crystals up to 0.5 cm thick that grew perpendicularly to the fracture walls fill these fractures.
Where fractures are several cm long, gypsum nodules replaced by anhydrite up to 3 cm in
length are present (Fig. 4-3E). Rosettes of dark gypsum up to 5 cm in diameter are common in
this facies. The massive/macronodular gypsum is in sharp contact with the layered gypsum,
locally forming diapirs several meters long.
93
94
Petrographic analysis of the evaporite facies showed that the darker beds are formed by
nodular gypsum displaying laths up to 20 µm long of gypsum, and secondarily, by anhydrite,
as well as by relics of calcite (Fig. 4-4A and B). The nodules were pervasively recrystallized
to gypsum crystals averaging 200 µm long, internally displaying laths of gypsum and/or
anhydrite, as well as relics of calcite. The volume of matrix is low in these nodules. The
authigenic gypsum is in optical continuity beyond the nodule boundaries (Fig. 4-4C). The
crystals from the lighter beds of the layered gypsum occur as aligned twin planes and
superimposed growth faces with acute angles zig-zagging perpendicularly to the crystal long
axes, and characterizes the chevron gypsum (Fig. 4-4D). These crystals are intergrown by
acicular gypsum containing relics and/or ghosts of the chevron gypsum. In the gypsarenite
facies (Fig. 4-4E), clasts were replaced by a mosaic of gypsum crystals varying from 100 µm
to 2 mm in length. These display optical continuity beyond their boundaries and present
abundant inclusions, similarly to the gypsum in the darker beds. The massive/macronodular
gypsum is pervasively replaced by mosaics of either alabastrine gypsym (i.e., crystals up to
200 µm in diameter), or coarser crystals reaching up to 2 mm in diameter (Fig. 4-4F). As
opposed to the previously described evaporite facies, the gypsum crystals in this facies lack
inclusions, resulting in limpid mosaics. The anhydrite nodules are fine-grained to
cryptocrystalline and internally display few tabular anhydrite crystals up to 250 µm. The
nodules replace large (circa 8 mm) crystals of gypsum with well-preserved cleavage; they are,
in turn, locally replaced by fine-grained gypsum.
4.5.2. Depositional setting
As previously stated, detailed facies analysis of the Codó Formation led to attribute it
to a dominantly continental setting. This interpretation will be further supported by the isotope
data presented in this paper. However, before going into that issue, it is important to detail the
depositional setting of the evaporites as extracted from facies data, because both vertical and
lateral variations in paleoenvironmental conditions seem to have affected the distribution of
the isotope values in this instance.
95
Stable and well-stratified hypersaline lakes with significant periods of anoxia and
closure prevailed in the Codó area, which led to the preferential precipitation of salts in more
saturated deeper central areas (Paz & Rossetti, 2001). Much more ephemeral conditions seem
96
to have occurred in the Grajaú area, where more oxygenated water pans with evaporite
precipitation in their margins and along surrounding mudflats prevailed (Fig. 4-5). This is
demonstrated by the dominance of layered gypsum and gypsarenite in association with
marginal and intermediate lake environments. In these facies, the chevron gypsum crystals of
the light-layered gypsum attest to primary deposition on the floor of brine pools. Precipitation
of similar features in many modern and ancient environments occurs when water depth is less
than 2 m (e.g., Hovorka, 1987; Logan 1987; Handford, 1991; Smoot & Loweinstein, 1991).
The darker layered gypsum is attributed to displacive intrasediment growth of crystals beneath
the brine from groundwater brines in the capillary and/or upper phreatic zone, thus implying
periods of eventual exposure (e.g., Kerr & Thomson, 1963; Southgate, 1982; Warren, 1999).
The gypsarenite records periods when evaporite crystals were reworked. Therefore, the
evaporites in the Grajaú area were dominantly formed in subaqueous, but shallower
environments than in the Codó area, with better defined alternating periods of stable and
ephemeral waters, a condition that led to envisage formation in a saline pan complex with
widespread evaporitic mudflats.
Beside facies characteristics, a continental setting is also suggested by the abundance
of ostracods mostly including the genus Harbinia and Candona, presence of charophytes, and
absence of any marine tolerant fauna.
4.6. SR AND S ISOTOPES FROM THE EVAPORITES
Thirty-six samples were collected for Sr and S isotope analysis. However, only the
results obtained from the layered gypsum and gypsarenite facies were used for
paleoenvironmental purposes, because these lithologies showed the least degree of textural
and/or mineralogical modification, while the massive/macronodular gypsum was strongly
affected during burial, as discussed in the following section.
The results (Table 4-I; Fig. 4-6) for the layered gypsum and gypsarenite indicate
variable
87
Sr/
86
Sr ratios ranging from 0.708481 to 0.709280 in the Codó area and 0.708272 to
0.708935 in the Grajaú area, with only one sample displaying 0.708104 in the later. The
massive/macronodular gypsum displayed the lowest values, ranging from 0.707824 to
0.708190. Discarding these values from our results, it is noteworthy that the
87
Sr/
86
Sr ratios
97
vary from higher to lower and then higher again upward within low frequency cycles. The Sr
content ranges from 212 to 2051 ppm (mean value of 435 ppm). These values vary inversely
to the Sr isotopes within individual cycles.
98
Table 4-I: Results of the strontium and sulphur isotope analyses from the evaporites of the Codó Formation in the
study areas.
Sample #
87
Sr/
86
Sr
2σ δ
34
S (v-CDT)
Evaporite Facies
1 0.709280 18 16.59
2 0.708901 15 17.09
3 0.708860 9 17.89
4 0.708558 15 17.47
5 0.708652 12 16.99
6 0.708481 14 -2.23
7 0.708675 53 16.12
8 0.708992 48 16.50
Layered gypsum
9 0.708116 17 15.85
10 0.708153 14 15.66
11 0.708134 14 15.29
12 0.707824 46 15.21
13 0.707989 30 -
C
O
D
Ó
14 0.707960 34 -
Massive/
macronodular
gypsum
15 - -5.57
16 0.708821 25 15.04
17 0.708831 24 15.15
18 0.708822 19 14.79
19 0.708754 29 15.34
20 0.708416 20 15.10
21 0.708561 15 15.06
22 0.708441 27 15.60
23 0.708413 59 15.35
24 0.708272 49 15.12
25 0.708442 27 15.30
26 0.708555 23 15.54
27 0.708935 28 15.29
28 0.708842 22 14.56
29 0.708758 26 14.60
30 0.708477 27 14.69
31 0.708443 25 14.75
Layered gypsum
32 0.708190 32 15.52
33 0.708090 30 15.14
34 0.708138 55 15.24
35 0.708104 17 15.07
36 0.708104 11 14.60
G
R
A
J
A
Ú
Massive/
macronodular
gypsum
99
100
The
34
S/
32
S ratios from layered gypsum and gypsarenite in the Codó area varied from
16.12 to 17.89 %
o (
V-CDT)
, with only one sample having an anomalous value of –2.23 %
o (
V-CDT)
.
The Grajaú area, in turn, showed distinctively lower values ranging from 14.79 to 15.60 %
o (
V-
CDT)
, with one sample displaying a negative value of –5.57 %
o (
V-CDT)
. The distribution of S
isotope within individual low frequency cycles changes following the same pattern shown by
the Sr isotope data, except for local inverse covariance where marginal lake deposits
dominate. The massive/macronodular gypsum in both areas showed S isotope values between
15.07 and 15.85 %
o (
V-CDT)
.
4.7. BURIAL OVERPRINT
An important step to consider, before analysing the significance of the Sr and S isotope
data as potential paleoenvironmental indicators, is to eliminate diagenetic influences. A line of
evidence suggests that, if the deposits were modified after deposition, diagenesis was not
enough to cause any significant changes in the isotope values, which seems to rather reflect
the primary water composition.
First, absence of significant diagenetic imprint is suggested by the overall wide
dispersion of Sr and S isotope data within individual depositional cycles. Second, there is a
consistent change in isotope values along the deposits formed during the expansion and
contraction phases of the saline pan/lake systems, revealing a depositional, rather than a
diagenetic signature. Third, petrographic and scanning electronic microscopic studies helped
to select samples suitable for isotope investigation. In fact, most of the analyzed samples
revealed at least a certain degree of replacement, recrystallization, cementation, neomorfism or
dissolution. However, this study also led to conclude that the mineralogical modifications
observed in the layered gypsum and gypsarenite, as explained in the following, were not
enough to cause any significant change in the isotope signal, as opposed to the
massive/macronodular gypsum that was pervasively modified during burial.
As previously discussed, the crevron crystals in the lighter portion of the layered
gypsum record primary deposition on the floor of brine pools. The horizontal alternations of
lighter and darker gypsum in this facies represent a product of sedimentary processes. The
darker layered gypsum formed slightly after deposition, but still under strong influence of the
101
depositional setting as precipitation took place within only a few milimeters below the
depositional surface (cf. Warren, 1999). The chevron gypsum was in part replaced by acicular
gypsum, as indicated by relic and ghost features of chevron gypsum. But even in this case the
effects of diagenesis were rather mild, not disturbing bedding planes. Although ultimate
evidence is lacking, the absence of other types of secondary gypsum in these beds suggests
that such replacement took place also shortly after deposition, thus still under a strong
influence of the primary brine. A second phase of gypsum would have affected this facies, as
well as the gypsarenite, which locally resulted in gypsum crystals that grew beyond nodule or
clast boundaries. However, this process was not extensive enough to replace the internal laths
of anhydrite/gypsum and/or relics of calcite. This fact, added to isotope values similar to those
from the layered gypsum, as will be shown in the following sections, led us to attribute an
eodiagenetic origin for this second phase of gypsum formation. Gypsum crystals formed
shortly after deposition will keep a primary isotope signature, because the precipitating fluids
derive from dissolution of evaporites formed in the same depositional site (e.g., Denison et al.,
1998).
On the other hand, the massive/macronodular gypsum records salt replacement and
recrystallization during deeper burial in association with halokinesis. As this process took
place, the original structures were greatly obliterated, forming a dominantly massive deposit.
The pervasive mosaics of limpid crystals typical of this facies might have required several
phases of replacement, during which any primary sedimentary feature was lost. Although
some of the macronodular texture might reflect relics of a primary nodulation, the majority of
the macronodules originated by salt displacement, giving rise to fracturing and precipitation of
bladed gypsum by introduction of sulphate-saturated intrastratal waters. The close association
of anhydrite nodules with fractures points to their formation by local dewatering due to
overpressure. In fact, these samples revealed anomalously the lowest isotope values, which is
consistent with its distinct post-depositional evolution. Furthermore, samples with clay
minerals, as revealed under the microscope, were not applied for analysis of geochemical
tracers in order to avoid Sr contamination.
102
4.8. DISCUSSION
Considering that a significant diagenetic imprint was discarded in samples of layered
gypsum and gypsarenite, the Sr and S isotopes could be used for paleoenviromental purposes.
On the other hand, as expected, the values obtained for the massive/macronodular gypsum
differed from the average strontium ratios. This is attributed to later growth of limpid gypsum
crystals and secondary nodules of anhydrite, as characterized by petrographic studies. As a
result, the isotope data from the massive/macronodular facies were discarded from our
analysis. In addition, the sample from the Grajaú area that also showed an anomalously low Sr
isotope value was also excluded, as this sample comes from a red-stained gypsum horizon
only 0.2 m below a discontinuity surface probably related to prolonged subaerial exposure at
the top of a shallowing-upward succession.
The
87
Sr/
86
Sr and
34
S/
32
S analyses have been successfully used to distinguish marine
and non-marine brines throughout the world (Table 4-II). These methods have been
increasingly applied in evaporites because: 1. evaporites are rich in strontium and sulphur; 2.
evaporites are widespread in both marine and non-marine settings, and their distinction is very
problematic where the associated deposits have no other diagnostic paleoenvironmental
features, requiring more sophisticated geochemical procedures; 3. other geochemical tracers
commonly used in paleoenvironmental reconstructions show many conflicting results (e.g.,
Hovorka et al., 1993); 4. distribution of Sr and S ratios in seawater is well known for the
Phanerozoic; 5. low permeability in evaporites difficults the introduction of external water
carrying Sr and/or S that may modify their isotopic composition; and 6. Sr isotope is very
sensible to detect the introduction of marine waters in continental settings.
Previous studies focusing on Sr isotopic composition of carbonates and evaporites,
originated from marine-based brines, have established values between 0.70720 and 0.70735
for the Aptian (Bralower et al., 1997; Jones & Jenkins, 2001; Fig. 4-7). The data generated
here are much higher than expected for Late Aptian marine deposits, suggesting a continental
brine source. The difference of the Sr isotope values from the Codó Formation relative to the
Aptian seawater is at least of 0.00075. Similar variations of strontium isotope data relative to
the seawater have also been measured from many other continental deposits recorded in the
literature (e.g., Faure et al., 1963; Jones & Faure, 1967, 1972; Faure & Barret, 1973; Denison
103
et al., 1998; Vonhof et al., 1998; Playà et al., 2000). Modern settings record even lower
differences (i.e, 0, 00021) when continental and marine brines are compared, as measured for
instance in the Abu Dhabi Sabhka (Müller et al., 1990). Therefore, the high Sr isotope data
from the layered gypsum and gypsarenite are related to a continental brine source, which is
consistent with the proposed depositional model interpreted with basis on facies analysis.
Table 4-II:
87
Sr/
86
Sr values of non-marine source brines and rocks.
Age (
87
Sr/
86
Sr
seawater)
Origin Material
87
Sr/
86
Sr N
*
E. Great Salt Lake
2
Brine 0.7167 to 0.7179 9
W. Great Salt Lake
2
Brine 0.7125 to 0.7139 4
Lakes and rivers from
Canadian Shield
3
Brine 0.712 to 0.726
Lake Vanda (Antartida)
4
Brine 0.7149
Holocene
(0.709173)
Lake George
2
Brine 0.7184
Pebas Formation
6
Carbonate 0.708160 to
0.709051
10 Miocene
(0.7081 to 0.7089)
Eastern Betics
7
Evaporite 0.70804 to 0.70888
Jurassic
(~0.7069)
(~0.7073)
Todilto Formation
12
Carbonate
Gypsum
0.707273 to 0.708673
0.706873 to 0.708073
40
35
Beacon Supergroup
16
Carbonate 0.7150 to 0.7280 12 Permian
Salado-Tansill formations
9
Evaporite 0.7069 to 0.7076 5
Devonian
(~0.7077 to 0.7086)
Beacon Supergroup
16
Carbonate 0.7260 to 0.7291 3
N*= number of samples.
2
Jones and Faure (1972);
3
Faure et al. (1963);
4
Jones and Faure (1967);
6
Vonhof et al. (1998);
7
Playà et al. (2000);
12
Denison
et al. (1998);
16
Faure and Barret (1973).
104
In addition to the high
87
Sr/
86
Sr values, their wide distribution in a same section and
between sections, with a standard deviation of 0.000262, conforms to the proposition of
continental-derived brines (Playà et al., 2000). Oceans favor geochemical homogenization,
thus marine deposits should provide more consistent Sr isotope data during a given time
interval (Denison et al., 1998). For instance, studies in a modern sabkha from the Persian Gulf
indicate a standard deviation of
87
Sr/
86
Sr much higher (i.e., 0.000177) in continental brines
than in marine brines (i.e., 0,000042) (Müller et al., 1990). The standard deviation of the Sr
105
isotope values in the Codó Formation is much higher than measured in this modern analog,
reinforcing a continental origin for the brine.
The upward decrease, and then increase, in
87
Sr/
86
Sr values within the low frequency
depositional cycles in both of the study areas seem to be better explained by facies changes
related to expansion and contraction of saline pan/lake systems (Fig. 4-8). Analysis of facies
distribution within these cycles suggests either rapid or progressive expansion of the
depositional system. Rapid expansion is recorded where central lake black shales and
laminated argillites overlie directly the basal discontinuity surfaces at the base of the cycles.
On the other hand, progressive expansion is suggested where this surface is overlain by clastic
evaporites, rhythmites, and lags of intraformational calcimudstone/shales that grade upward
into laminated argillites and black shales. Contraction of the depositional system is recorded
by the increased dominance of marginal facies in the uppermost portions of the low frequency
cycles.
The distribution of
87
Sr/
86
Sr values according to the expansion and contraction phases
of the saline pan/lake systems reinforces their primary nature. Hence, the progressive
expansion of the depositional system gave rise to evaporitic precipitation. While evaporites
precipitated in central environments, low lying marginal areas dominated by mud flats were
exposed to substantial leaching due to weathering (Fig. 4-8). During this process, the draining
of clay minerals might have released
87
Sr, increasing
87
Sr/
86
Sr values of the evaporitic
precipitating brines (cf. Hovorka et al., 1993; Playà et al., 2000). Similar facies control on the
distribution of
87
Sr/
86
Sr ratios has been documented in contraction/expansion cycles in the
Eocene Green River Formation (cf. Rhodes et al., 2002).
Following initial expansion, increased contribution of external drainage resulted in a
period of maximum expansion of the saline pan/lake systems. Flooding promoted water
dilution, ending the evaporite precipitation. Increased inflow also resulted in the introduction
of large volumes of muds in suspension, which were eventually settled down, forming the
laminated argillites and black shales. Expansion caused flooding of widespread marginal mud
flats and, as a consequence, the release of
87
Sr from clay minerals was cut off, resulting in
relatively lower
87
Sr/
86
Sr values.
106
Additionally, the Sr isotope ratios might have decreased due to the contribution of Sr
released from basaltic source rocks. This is suggested due to the fact that the study areas are
located near Permian to Neocomian marine-derived limestones and evaporites from the
intracratonic Parnaíba Basin, as well as Triassic to Neocomian basaltic rocks associated with
the Xambioá-Teresina Arch (Fig. 4-1A). Weathering of these rocks might have contributed
with significant amounts of
87
Sr-depleted waters into the evaporitic basin, resulting in
relatively lower
87
Sr/
86
Sr values. However, it is noteworthy that, even considering the
influence of a basaltic sources, the
87
Sr/
86
Sr values are still much higher than expected to Late
107
Aptian seawaters. Mixing of waters from both basaltic and sedimentary rocks might have
provided continental brines with
87
Sr/
86
Sr ratios comparable to the ones obtained for the study
areas.
During a following stage, the inflow was interrupted and the saline pan/lake systems
contracted. As the introduction of mud progressively reduced and evaporation increased,
carbonates were formed and marginal facies became more widespread. As a consequence, the
87
Sr/
86
Sr data increased.
Likewise the strontium, sulphur is also abundant and uniformly distributed in
seawaters, and its isotope ratio changed through time, opening the possibility for its use to
determine brine sources (Claypool et al., 1980). During the Late Aptian, the δ
34
S values
ranged from 13 to 16 %
o(
V-CDT)
(Claypool et al., 1980; Fig. 4-7). Similarly to strontium, the
sulphur isotope values obtained in the Codó Formation are highly variable, as expected in
continental-derived brines. In addition, the values from the Codó area are higher than 16 %
o(
V-
CDT)
, thus also conforming to continental-derived brines. However, values from the Grajaú
area are lower, situating in the upper part of the range expected for marine waters. A marine
contribution to the saline pan complex could explain these values, but this is not supported by
the strontium nor by the facies data, as previously discussed. Therefore, if hybrid brine is to be
considered in this instance, then the marine contribution was not enough to cause a significant
geochemical imprint other than in the S isotope composition. Taking into account previous
studies discussing this method, this interpretation is disregarded here because in general
sulphur isotope is considered much less diagnostic of brine sources than strontium isotope
(Chivas et al., 1991; Denison et al., 1998; Playà et al., 2000). Considering the difference in
facies patterns between the study areas, an alternative hypothesis is that the higher rates of
evaporation in the Grajaú area, favored by the prevalence of shallower evaporitic
environments, might have contributed to the depletion in
34
S. In seawaters, this process might
account for about 2%o of
34
S depletion (Hoefs, 1980), though there is no information available
from lake systems, and the effect in the gypsum precipitate-brine relationship is an issue open
for further research.
Considering a non-marine brine interpretation, then the changes in S isotope values
between the Codó and Grajaú areas might have had a facies control. The increase in δ
34
S
108
values in the Codó area (Fig. 4-7) is attributed to a more restricted environment, resulting from
the prevalence of a hydrologically closed lake system, as revealed by facies analysis (Paz and
Rossetti, 2001). Under such conditions, the δ
34
S tends to be increased due to the activity of
sulphate-reducing bacteria, resulting in SO
4
depletion in the water and increasing the δ
34
S
(Hoefs, 1980). On the other hand, more oxygenated conditions would have prevailed in the
Grajaú area, which led to a relative enrichment of sulphates and a consequent decrease in δ
34
S.
The two anomalous negative sulphur isotope values obtained in the study area were derived
from evaporite sampled from central lake black shales recording the beginning of flooding
stages, and they may be related to recycling of underlying sulphides.
4.9. CONCLUSIONS
This study led to several conclusions: 1. when combined with facies analysis and
optical (petrography and scanning electron microscopy), Sr isotope is a powerful tool for
distinguishing evaporites derived from marine and non-marine brines. This effort led to the
confirmation that the deposition of the Codó Formation in the southern and eastern margins of
the Grajaú Basin, northern Brazil, took place dominantly in saline pan/lake depositional
systems; 2. Sr isotopes revealed to be very sensible to detect hydrological changes in this type
of environments; 3. however, fluctuations on the distribution of Sr values due to changes in
sedimentary patterns are not enough to preclude distinguishing between marine and non-
marine evaporites; 4. during initial expansion phases, the Sr isotopes in saline pan/lake
environments tend to decrease due to leaching of
87
Sr from clay minerals of mud flats exposed
to internal drainage; 5. during maximum expansion of this type of system, these values
progressively decrease due to the cut off of this process, as the mud flats become flooded; 6, in
the particular case of the study area, the lowering in Sr isotope values might have additionally
caused the introduction of
87
Sr-depleted waters derived from weathering of older marine
limestone and evaporites, as well as basaltic rocks; and 7. interruption of external drainage and
contraction of the system caused increased carbonate precipitation, with consequent slight
increase in Sr isotope values. The data from the study area allowed also to conclude that the
application of sulphur isotope as a parameter to decipher brine sources is limited. This is
because, even where strontium values and geochemical tracers indicate a non-marine brine
109
source, the sulphur values might coincide with those from seawater for a given time interval,
as observed in the Grajaú area. Hydrologically closed systems tend to develop anoxia, when
the activity of sulphate-reducing bacteria results in SO
4
depletion in the water, with the
consequent increase of δ
34
S. On the other hand, more oxygenated conditions lead to a relative
enrichment of sulphates and a consequent decrease in δ
34
S. Therefore, significant fluctuations
in sulphur isotope composition might take place due to changes in hydrological regime, and
thus the use of this parameter alone is not recommended to decipher between marine/non-
marine brines.
Acknowledgements. The Itapicuru Agroindustrial S/A is acknowledged for the permission to
access the quarries with the exposures of the Codó Formation. This work was financed by the
Brazilian Council for Research–CNPq (Project #460252/01).
REFERENCES
Antonioli, L. (2001) Estudo palino-estratigráfico da Formação Codó–Cretáceo Inferior do
Nordeste brasileiro. Ph.D. Thesis, Universidade Federal do Rio de Janeiro, Rio de Janeiro,
RJ, 265 pp.
Batista, A.M.N. (1992) Caracterização Paleoambiental dos sedimentos Codó-Grajau, Bacia de
São Luís (MA). M.Sc. Thesis, Universidade Federal do Para, Belém, PA, 102 pp.
Bralower, T.J., Fullagar, P.D., Paul, C.K., Dwyer, G.S. and Leckie, R.M. (1997) Mid-
Cretaceous strontium-isotope stratigraphy of deep-sea sections. Geol. Soc. Am. Bull. 109,
1421-1442.
Brookins, D.G. (1988) Seawater
87
Sr/
86
Sr for the Late Permian Delaware Basin evaporites
(New México, U.S.A.). Chem. Geol., 69, 209-214.
Burke, W.H., Denison, R.E., Hetherington, E.A., Koepnick, R.B., Nelson and H.F., Otto,
J.B. (1982) Variation of seawater
87
Sr/
86
Sr throughout Phanerozoic time. Geology, 10,
516-519.
Chivas, A.R., Andrew, A. S., Lyons, W.B., Bird, M.I. and Donelly, T.H. (1991) Isotopic
constraints on the origin of salts in Australian playas. 1. Sulphur. Palaeogeogr.,
Palaeoclimatol., Palaeoecol., 84, 309-332.
110
Clauer, N. (1976)
87
Sr/
86
Sr composition of evaporitic carbonates and sulphates from Miocene
sediment cores in the Mediterranean sea (D.S.D.P., Leg 13). Sedimentology, 23, 133-140.
Claypool, G.E., Holser, W.T., Kaplan, I.R., Sakai, H. and Zak, I. (1980) The age curves of
sulphur and oxygen isotopes in marine sulphates and their mutual interpretation. Chem.
Geol., 28, 199-260.
Denison, R.E., Kirkland, D.W. and Evans, R. (1998) Using strontium isotopes to determine
the age and origin of gypsum and anhydrite beds. J. Geol., 106, 1-17.
Faure, G. and Barrett, P.J. (1973) Strontium isotope composition of non-marine carbonate
rocks from the Beacon Supergroup of the Transantarctic Mountains. J. Sedim. Petrol., 43,
447-457.
Faure, G., Hurley, P.M. and Fairbairn, W.H. (1963) An estimate of the isotopic
composition of strontium in rocks of the Precambrian Shield of North America. J.
Geophys. Res., 68, 2323-2329.
Handford, C.R. (1991) Marginal marine halite: sabhkas and salinas. In: Evaporites,
Petroleum and Mineral Resources (Ed. J.D. Melvin), Developments in Sedimentology 50,
pp.1-66. Elsevier, Amsterdam.
Hardie, L.A. (1984) Evaporites: Marine or non-marine?: Am. J. Sci, 87, 193-240.
Hoefs, J. (1980) Stable isotope geochemistry. Springer-Verlag, Berlim, 208 pp.
Hovorka, S.D. (1987) Depositional environment of marine-dominated bedded halite, Permian
Sant Andreas Formation, Texas. Sedimentology, 34, 1029-1054.
Hovorka, S.D., Knauth, L.P., Fisher, R.S. and Gao, G. (1993) Marine to nonmarine facies
transition in Permian evaporites of the Palo Duro basin, Texas: geochemical response.
Geol. Soc. Am. Bull., 105, 1119-1134.
Jones, C.E. and Jenkins, H.C. (2001) Seawater strontium isotopes, oceanic anoxic events,
and seafloor hydrothermal activity in the Jurassic and Cretaceous. Am. J. Sci. 301, 112-
149.
Jones, L.M. and Faure, G. (1967) Origin of the salts in Lake Vanda, Wright Valley, southern
Victoria Land, Antarctica. Earth Planet. Sci. Lett., 3, 101-106.
Jones, L.M. and Faure, G. (1972) Strontium isotope geochemistry of Great Salt Lake, Utah.
Geol. Soc. Am. Bull., 83, 1875-1880.
111
Kerr, S.D. and Thompson, A. (1963) Origin of nodular and bedded anhydrite in Permian
shelf sdiments, Texas and New Mexico. AAPG Bull., 47, 1726-1732.
Logan, B.W. (1987) The Lake MacLeod Evaporite Basin, Western Australia–Holocene
Environments, Sediments, and Geological Evolution. AAPG Mem., 44, 140 pp.
Müller, D.W., McKenzie, J.A. and Mueller, P.A. (1990) Abu Dhabi Sabkha, Persian Gulf,
revisited: application of strontium isotopes to test an early dolomitization model. Geology,
18, 618-621.
Paz, J.D.S. (2000) Análise Faciológica da Formação Codó (Aptiano Superior), na Região de
Codó (MA), Borda Leste da Bacia do Grajaú. M.Sc. Thesis, Universidade Federal do Pará,
Belém, PA, 117 pp.
Paz, J.D.S. and Rossetti, D.F. (2001) Reconstrução paleoambiental da Formação Codó
(Aptiano), borda leste da Bacia do Grajaú, MA. In: O Cretáceo na Bacia de São Luís-
Grajaú (Eds. D.F. Rossetti., A.M. Góes and W. Truckenbrodt), pp. 77-101. Goeldi Press,
Belém.
Playà, E., Ortí, F. and Rosell, L. (2000) Marine to non-marine sedimentation in the upper
Miocene evaporites of the Eastern Betics, SE Spain: sedimentological and geochemical
evidence. Sed. Geol., 133, 135-166.
Rhodes, M.K., Caroll, A.R., Pietras, J.T., Beard, B.L., and Johnson, C.M. (2002)
Strontium isotope record of paleohydrology and continental weathering, Eocene Green
River Formation, Wyoming. Geology, 30, 167-170.
Rossetti, D. F. (2001) Depositional architecture of the São Luís-Grajaú Basin. In: O Cretáceo
na Bacia de São Luís-Grajaú (Eds. D.F. Rossetti., A.M. Góes and W. Truckenbrodt),
pp.31-46. Goeldi Press, Belém.
Rossetti, D.F., Góes, A.M. and Arai, M. (2001) A passagem Aptiano-Albiano na Bacia de
São Luís-Grajaú, MA. In: O Cretáceo na Bacia de São Luís-Grajaú (Eds. D.F. Rossetti.,
A.M. Góes and W. Truckenbrodt), pp. 101-117. Goeldi Press, Belém.
Schreiber, B.C. and El Tabakh, M. (2000) Deposition and early alteration of evaporites.
Sedimentol., 47, 215-238 (Supplement 1).
Smoot, J.P. and Lowenstein, T.K. (1991) Depositional environments of non-marine
evaporites. In: Evaporites, Petroleum and Mineral Resources (Ed. J.D. Melvin).
Developments in Sedimentology 50, 189-347. Elsevier, Amsterdam.
112
Southgate, P.N. (1982) Cambrial skeletal halite crystals and experimental analogues.
Sedimentol., 29, 391-407.
Vonhof, H.B., Wesselingh, F.P. and Ganssen, G.M. (1998) Reconstruction of the Miocene
western Amazonian aquatic system using molluscan isotopic signatures. Palaeogeogr.,
Palaeoclim., Palaeoecol., 141, 85-93.
Warren, J. (1999) Evaporites: Their Evolution and Economics. Blackwell Science, Oxford,
438 pp.
113
5. PALEOHYDROLOGY OF A LATE APTIAN LACUSTRINE
SYSTEM FROM NORTHEASTERN BRAZIL WITH BASIS ON THE
INTEGRATION OF FACIES AND ISOTOPIC GEOCHEMISTRY
*
5.1. ABSTRACT
The Codó Formation records the initial evolutionary stages of an intracontinental rift
system formed along the Brazilian equatorial margin in the Late Aptian. Deposits of this unit
exposed in the eastern margin of the Grajaú Basin include evaporites, bituminous black shales
and limestones. These lithologies were formed in a low energy, well stratified, predominantly
anoxic and hypersaline lake system developed in a dominantly arid/semi-arid climate. This
lacustrine succession is internally organized into three categories of shallowing-upward
cycles, with the first- and second-order cycles being related to seismic activity associated with
fault reactivations, and the third-order cycles recording seasonal fluctuations. Studies
emphasizing petrography and analysis of the minor and trace elements Fe, Mg, Sr, Mn, and Na
led to the evaluation that the Codó Formation was appropriate for isotopic investigations
having paleoenvironmental and paleohydrologic purposes. The results of this study revealed a
wide distribution of dominantly low δ
13
C and δ
18
O values, ranging from –5.69 ‰ to –13.02
(PDB)
and from –2.71‰ to –10.80‰
(PDB)
, respectively. This paper demonstrates that the
*
Authors: J.D.S. Paz & D.F. Rossetti. Submitted to Palaeogeography, Palaeoclimatology,
Palaeoecology
114
isotope ratios vary according to seismically-induced shallowing-upward cycles, in general
becoming lighter in their bases, where central lake deposits dominate, and progressively
heavier upward, where marginal lake deposits are more widespread. In addition, to confirm a
depositional signature for the analysed samples, this behavior led to introduce a seismic-
induced isotope model. Hence, lighter isotope ratios appear to be related with flooding events
promoted by subsidence, which resulted in the development of a perennial lake system, while
heavier isotope values are related to ephemeral lake phases favored through uplift and/or
increased stability. Furthermore, the results show that a closed lake system dominated, as
indicated by the overall good positive covariance (i.e., +0.42 to +0.43) between the carbon and
oxygen isotopes, though open phases are also recorded by negative covariance values of –
0.36. During closed phases, the δ
18
O displayed the highest interval of variation (i.e., –3.63‰
to –4.89‰) due to the increased residence time, while this variation was low (i.e., –0.09‰ to -
1.87 ‰) during open lake phases, when there was a balance in the water isotope composition
caused by the continuous basin inflow.
5.2. INTRODUCTION
δ
13
C and δ
18
O records have been successfully applied for reconstructing the evolution
and circulation patterns of oceanic basins throughout the geological time (e.g., Abell and
Williams, 1986; Charisi & Schmitz, 1995; Hendry & Kalin, 1997). These studies are mostly
based on the principle that the organic matter in marine sediments is characterized by
extremely uniform isotopic compositions, which vary according to climate, as well as oceanic
hydrology. The interpretation of these geochemical indicators in lacustrine settings is more
complex, mostly because lake environments are more diversified, as evidenced by a wider
distribution of carbon isotope ratios ranging from –25.9 to –10.5‰ (Bird et al., 1991). Since
the pioneer work of Stuiver (1970), several documentations on modern and Quaternary lake
systems have provided the basis for an increased discussion on the many parameters that
might influence the isotopic composition of the total inorganic carbon dissolved in lake waters
(e.g., Fritz et al., 1975; Katz et al., 1977; Anderson & Arthur, 1983; MacKenzie, 1985;
Hillaire-Marcel & Casanova, 1987; Bellanca et al., 1989; Gasse et al., 1989; Talbot & Kelts,
1990; Rosenmeier et al., 2002; Herczeg et al., 2003; Russell et al., 2003). Despite these
115
efforts, differentiation among the mechanisms that lead to variations in isotope composition of
lake waters is not straightforward, as local causes might be mistaken by externally forced
environmental changes (Talbot, 1990). In addition, in contrast to marine and modern
lacustrine systems, the record of chemical changes in ancient lake deposits is yet very limited
(Bird et al., 1991; Lister et al., 1991; Camoin et al., 1997; In Sung & Kim, 2003), which have
precluded a wider use of these geochemical tracers for paleoenvironmental purposes. Thereby,
geochemical analysis of carbonate lacustrine deposits from a larger volume of analogs, where
local causes can be distinguished from those of regional significance, are still needed in order
to provide a full understanding of the mechanisms controlling lacustrine carbonate
sedimentation.
In spite of the complex response, the available information concerning to carbon and
oxygen isotope variations has arrived to some important generalizations. The most significant
one for paleoenvironmental interpretation is the covariance of these geochemical tracers as an
indicator of hydrologically closed lake systems, as opposed to the non-covariance in inlet
lakes (e.g., Eicher & Siegnethaler, 1976; Gasse et al., 1987, 1989; Talbot, 1990; Talbot &
Kelts, 1990). Additionally, these isotopes have been applied for climate reconstructions of
lake systems (e.g., Talbot & Keltz, 1990; Lister et al., 1991; Valero-Garcés et al., 1995).
These applications are, however, highly dependent on the full understanding of facies
distribution and of the possible modifications occurred during burial.
The goal of this paper is to contribute for the documentation of δ
13
C and δ
18
O values in
ancient lacustrine systems, and discuss the causes of these variations analysing their
relationship with shallowing-upward cycles within a Late Aptian succession formed during the
early stages of a passive marginal rift. This unit, represented by the Codó Formation, is well
exposed along several quarries in the eastern margin of the Grajaú Basin, where detailed
studies focusing facies and facies architecture, stratigraphy, petrography, as well as Sr and S
isotopes, have provided the basis to support deposition in a dominantly lacustrine setting (e.g.,
Rossetti et al., 2000; Paz and Rossetti, 2001). An integrated approach combining facies and
isotope geochemistry provided the basis to analyze the distribution of carbon and oxygen
isotopes in this ancient lake system, as well as to investigate the main parameters controlling
its hydrology.
116
5.3. GEOLOGICAL SETTING
The Codó Formation records the first deposits accumulated within a broad and shallow
depression formed by mild tectonic streching before the main rifting stage that culminated
with the origin of the Equatorial South Atlantic Ocean during the Albian. These deposits are
well represented in the Grajaú Basin (Fig. 5-1A), a semi-graben formed by the
combination of
pure shear stress and strike-slip deformation (Azevedo, 1991; Góes & Rossetti, 2001).
This rift,
which is connected with the São Luís Basin to the north, became an aborted intracontinental
structure as the continental break up migrated northward.
The sedimentary fill of the São Luís-Grajaú Basin (Fig. 5-1B) reaches up to 4,000 m
thick in the depocenters, and consists chiefly of Cretaceous deposits organized into three
depositional sequences, i.e., S1, S2 and S3, formed during the Late Aptian/Early Albian,
Early-Middle Albian and Middle Albian/Late Cretaceous, respectively (Rossetti 2001). The
lowermost sequence S1 contains the Codó Formation, objective of this paper, and represents a
succession up to 450 m thick of sandstones, evaporites, shales and limestones. This sequence
displays a tripartite subdivision into systems tracts (Rossetti 2001), with the lowstand deposits
grading progressively southward from shallow marine to continental (i.e., fluvial, deltaic, and
lacustrine) in nature. The lowstand deposits are overlain by strata formed in the transgressive
systems tract, and consist of a wedge of richly fossiliferous (mostly bryozoans, equinoderms,
foraminifera and dinoflagellates) shales that pinches out to the basin margins. The highstand
systems tract consists of shallow marine to continental deposits typically displaying stratal
patterns varying upward from aggradational to progradational.
The maximum thickness of the Codó Formation in the Grajaú Basin is 150 m (Rezende
& Pamplona, 1970). Its paleontological content mostly includes pollens, continental ostracods,
insects, and fishes, which are all in agreement with a dominantly lacustrine interpretation for
the depositional system. Additionally, pollen has been recovered from these deposits and
allowed the establishment of a precise late Aptian age with basis on the presence of Sergipea
variverrucata (Batista, 1992; Lima, 1982; Rossetti et al., 2001). The Codó Formation either
grades downward into fluvial and deltaic deposits of the Grajaú Formation (e.g., Mesner &
Wooldridge, 1964), or sharply overlies an unconformity over older Paleozoic and Triassic
basement rocks. Its upper contact is an unconformity with Albian shallow marine, green to
117
brownish-red mudstones interbedded with fine- to very fine-grained, cross-stratified
sandstones of the Itapecuru Group (e.g., Rossetti & Truckenbrodt, 1997; Rossetti et al., 2001).
Fig . 5-1: A) Location of the study area in the Codó region, eastern margin of the Grajaú Basin. B) Stratigraphy
and main tectonic stages of the São Luís-Grajaú Basin.
118
5.4. FACIES ARCHITECTURE AND DEPOSITIONAL SYSTEM
5.4.1. Description
A detailed facies analysis of the Codó Formation exposed in the eastern margin of the
Grajaú Basin was previously reported elsewhere (e.g., Paz and Rossetti, 2001). There, the
Codó Formation consists of a lacustrine succession up to 25 m thick, attributed to three main
sub-environments (Fig. 5-2): 1. central lake deposits, consisting of evaporites and bituminous
black shales, locally with pyrite and native sulphur; 2. intermediate lake deposits, represented
by laminated argillites and limestones; and 3. marginal lake deposits, including massive
blocky pelites, fenestral calcarenites, ostracodal and pisoidal limestones, rhythmites of
limestones and microbial mats, as well tufas. Paleosol, karstic feature, meteoric cement and
vadose pisoid, typical of subaerial and/or meteoric exposures, are frequent features of the
marginal lake deposits.
Three categories of cycles were recognized in the Codó Formation (Fig. 5-3). A
detailed description of these cycles is presented in a separate paper (Paz and Rossetti,
submitted), but a summary is included here as they are critical to help understanding the
meaning of the isotope signals. Third-order cycles display regular thickness ranging from 5 to
10 cm, and consist of facies that vary according to the position in the lake setting. Hence, the
central lake deposits show interbeddings either of bituminous black shales and evaporites, or
bituminous black shales with streaks of calcimudstone and bituminous back shales with native
sulphur. The intermediate lake deposits display bituminous black shale interbedded with
peloidal limestone or green to gray laminated argillites interbedded either with calcimudstone
or peloidal wackestone-packstone. The marginal lake deposits show either green to gray
laminated argillites and ostracodal wackestone to grainstone, as well as alternations of
ostracodal and/or lime mudstones, microbial mats and vadose pisoidal packstones.
The second-order cycles consist of either complete or incomplete successions with
upward transitions from central to marginal lake deposits, with the latter displaying high
internal facies variability when comparing one cycle to another. These cycles are characterized
by limited lateral extension, as well as frequent and random thickness changes, which vary
from few cm up to 5 m.
119
120
Fig . 5-3: Shallowing-upward cycles of the Codó Formation. A) Examples of first- and second-order cycles. B-D)
Third-order cycles formed by alternations of bituminous black shale with streaks of lime-mudstone (Bsl) and
bituminous black shales with native sulphur (Bss) (B), ostracodal grainstone (Gro) and vadose pisoidal packstone
(Pp) with microbial mats (M) (C), ostracodal grainstone (Go) and wackestone (Wo) (D).
The first-order cycles define four episodes of shallowing (Fig. 5-4), organized from
bottom to top as units 1 to 4. Unit 1 is only partly exposed at the base of the sections,
consisting of bituminous black shales interbedded with calcimudstones, and are attributed to
central and intermediate lake settings. Unit 2 reaches up to 8 m thick and contains, at the base,
black bituminous shales interbedded with evaporites, which grade upward into limestones,
laminated argillites and massive block pelites displaying a variety of features related to
intermediate and marginal lake settings. Evaporites are, in general, absent or occur only as
milimetric lenses or isolated crystals of gypsum. Unit 3 reaches up to 4 m thick and is
constituted by intermediate and marginal lake deposits similarly to the underlying unit, but
constituted by intermediate and marginal lake deposits similarly to the underlying unit, bu
t
121
122
with an increased frequency relative to the latter. A remarkable and exclusive feature of this
unit 3 is the presence of oolites and calcareous (i.e., ostracodal packstone) concretions in its
upper portion, which constitute important stratigraphic markers. The uppermost unit 4 is up to
5 m thick, being represented by laminated argillites containing only thin (<1 mm thick)
laminae of gypsum or calcimudstone.
The first-order cycles closely match with stratigraphic horizons displaying syn-
sedimentary soft sediment deformation that occur between undeformed deposits (Fig. 5-4), an
observation that was crucial to reveal their genesis. Hence, units 1 and 2 correspond
respectively to undeformed strata, and the deformation zones 1 and 2 described in Rossetti and
Góes (2000). Deformation zone 1 consists of spar-filled cracks interconnected with small-
scale faults, fissures and stylolites inclined at a high angle to bedding. Deformation zone 2
consists of strata with complex convolute folds associated with thrust faults, pseudonodules,
and mound-and-sag structures, the latter correspond to synclines and anticlines mantled by
sigmoidal laminations inclined toward the sag centres. Unit 3 corresponds to a deformed
interval consisting of normal faults and fissures that are vertical to near vertical, present
ragged morphologies with small, delicate edges, and taper both downward and upward after a
few centimetres, being associated with intraformational boulders up to 2.5 m long. The
uppermost unit 4 consists of shales with irregular convolute folds.
5.4.2. Interpretation
The facies present in the Codó Formation exposed in the eastern margin of the Grajaú
Basin supports a low energy, well stratified, anoxic and hypersaline lake system developed
under a dominantly arid/semi-arid climate (Paz and Rossetti, 2001).
The third-order cycles record minor changes in depositional conditions, attesting to
alternations between mud settling and chemical precipitation of evaporites or limestones. This
characteristic, added to the regular thickness variation, is consistent with seasonal fluctuations,
with mud deposition and chemical precipitation taking place during less and more arid phases,
respectively.
The higher-order cycles seem to have a different origin. The second-order cycles
record successive episodes of upward gradation from deeper to relatively shallower lake
environments, which resulted in the superposition of marginal lake deposits upon intermediate
123
and/or central lake deposits. The high facies variability, when comparing one cycle to another,
the limited lateral extension and the frequent and random thickness variations are attributes
that match better with tectonically driven (e.g., Martel & Gibling, 1991; Benvenuti, 2003),
rather than more symmetrical climatic cycles (e.g., Olsen, 1986; Goldhammer et al., 1990;
Smoot & Olsen, 1994; Steenbrinck et al., 2000).
The first-order cycles also appear to have resulted from syn-sedimentary tectonics (Fig.
5-4), as suggested by their good correlation with the deformation zones attributed to
contemporaneous seismic activity related to fault reactivation (Rossetti and Góes, 2000).
Based on this fact, it has been proposed that the Codó lake system was affected by alternating
periods of distension and stability and even compression (Paz and Rossetti, chapter 2). The
prevalence of central lake deposits at the base of the first-order cycles would have formed
during higher subsidence, promoted by extension. On the other hand, the more widespread
distribution of marginal lake deposits in the top of these cycles would record periods of higher
stability or uplift. In addition to affect the development of the shallowing-upward cycles, these
processes appear to have had a strong control on the isotope evolution of this lake system, as
discussed in this paper.
5.5. GEOCHEMICAL TREATMENT
δ
13
C and δ
18
O data were analyzed from fresh samples in recently exposed quarries, to
guarantee they were free from the influence of modern weathering. In this study, the analyses
were performed using whole-rock limestone samples due to the fact that only ostracods are
present in the studied deposits, and their distribution is not uniform to guarantee a good record
of the individual cycles throughout the succession. Stable isotopic analysis has been
successfully performed in whole-rock carbonates with the advantage of minimizing possible
deviations related to vital effects, and even diagenesis (e.g., Urey, 1947; Camoin et al., 1997).
Twenty milligrams of powered sample reacted in vacuum with 100% of orthophosphoric acid
at 25
o
C during 12 hours. The released CO
2
and H
2
O were captured with liquid N
2
. The CO
2
was separated from the water with a solution of alcohol and liquid N
2
in an off-line gas
extraction line, and thereafter taken to the VG Isotech SIRA II mass spectrometer in the Stable
Isotope Laboratory at the Universidade Federal de Pernambuco (LABISE/UFPE). The results
124
are reported in δ notation, which is defined as the per mil deviation from a standard. The
Peedee Belemnite Standard (‰ PDB) were used to the notation of both isotopes. Replicate
analysis gave a standard deviation (2σ) lower than 0.02 %o for δ
13
C and 0.03 %o for δ
18
O.
The isotopic analyses were run using samples selected after their study under the
microscope to assure a primary signature. In addition, a double check on the diagenetic
influence was made through the analysis of Fe, Sr, Mg, Mn and Ca. The procedure consisted
in drying 1.5 g of sample at 1000
o
C for 2 hours. Samples were then fused using lithium
tetraborate and lithium fluorite, and analysed in the X-Ray Fluorescence Spectrometer of the
Stable Isotope Laboratory at the Universidade Federal de Pernambuco (LABISE/UFPE).
5.6. EVALUATION OF DIAGENETIC IMPRINT
The petrographic analysis of 83 limestone samples from the Codó Formation allowed
to evaluate its diagenetic signature, observing the amount of lime mud, recrystallization,
replacement, cementation, and fracturing. Several authigenic processes were observed, the
most important ones including recrystallization of calcite, cementation and filling of fractures
and secondary porosity by mosaics of calcite, replacement of micrite and ostracod shells by
chert and chalcedony, and pyrite formation within ostracod shells and dispersed in
calcimudstones. Despite these modifications, it was possible to select 53 samples consisting of
microfacies either not affected or only mildly affected by diagenesis, which enhanced their
potential to preserve the carbon and oxygen composition as a reflex of the depositional record.
The samples used in this study included mostly calcimudstone (36%) and ostracodal
wackestone to grainstone (45%; Fig. 5-5), and subordinately fenestral calcarenite (8%),
pisoidal packstone (6%), and peloidal packstone to grainstone (6%).
125
126
After the petrographic studies, samples were additionally investigated using
geochemical analysis of Fe, Sr, Mn, and Mg to help detecting diagenetic modification at a
higher degree of reliability. These elements are the principal trace consituents in the calcite
structure of both marine and non-marine limestones. Considering that their values remained
constant through time, which seems to have been the case at least for most of the Phanerozoic
(Holland, 1978), a comparison among values commonly expected from stratal waters provides
information to detect potentially significant diagenetic influences. The results (Fig. 5-6) show
that, in general, all the samples that appeared to be petrographically suitable for isotope
analysis contain minor and trace elements as expected for continental deposits not affected by
diagenesis. Exceptions are samples with high Fe content, which is due either to high volume
of organic matter or pyritization. However, these samples were also included in the isotope
analysis presented here, considering that: 1. the other geochemical tracers are within the range
expected for diagenetically non-affected rocks; 2. the isotope values do not show any
divergence with respect to the other samples; and 3. they derive from facies with pedogenetic
influence associated with marginal lake deposits, a situation that favours an increased Fe and
Mn circulation.
127
128
5.7. STABLE ISOTOPES
The δ
13
C isotope curves show an overall similar pattern in all the studied profiles, with
ratios ranging from –5.69 ‰ to –13.02 ‰ (Table 5-I). In general, two short intervals with
higher values at the base and top of the profiles are separated by a large interval with depleted
values (Fig. 5-7). The behavior of the isotope curves reflects the position of the first-order. For
instance, unit 2 displays increasing values at the bottom, followed by a short negative
excursion to ultimately oscillate toward heavier and, in more rarely, lighter values at the top.
The values in unit 3 are more regularly distributed, varying upward from lighter to heavier.
Table 5-I: δ
13
C
(
PDB)
and δ
18
O
(
PDB)
values observed in the study area. A, B and C mean the vertical profiles
showed in the Fig. 5-7.
Prof
Sample δ
18
O δ
13
C
Prof
Sample δ
18
O δ
13
C
01 pisoidal packstone -5,49 -10,10 31 ostracodal packstone -6,87 -10,96
0
2
ost
r
acoda
l
pac
k
sto
n
e
-7
,
24 -11
,80
3
2
ost
r
acoda
l
pac
k
sto
n
e
-
5,3
2 -
9,30
03
Ca
l
c
im
udsto
n
e
-
8,9
1 -1
0,03
33
ost
r
acoda
l
pac
k
sto
n
e
-7
,69
-1
0,95
0
4
Ca
l
c
im
udsto
n
e
-
9,
17 -
9,
1
3
3
4
p
i
so
i
da
l
pac
k
sto
n
e
-7
,6
2 -11
,5
1
05
Ca
l
c
im
udsto
n
e
-
9,
1
5
-
8,96
35
ost
r
acoda
l
pac
k
sto
n
e
-
8,38
-11
,00
06
f
e
n
est
r
a
l
ca
l
ca
r
e
ni
te
-
8,36
-
6,65
36
ost
r
acoda
l
pac
k
sto
n
e
-7
,83
-11
,8
7
0
7
ost
r
acoda
l
pac
k
sto
n
e
-
8,
4
5
-
9,9
2
3
7
Ca
l
c
im
udsto
n
e
-
8,69
-11
,
72
08
ost
r
acoda
l
pac
k
sto
n
e
-
8,36
-1
0,9
2
38
Ca
l
c
im
udsto
n
e
-
9,
41 -11
,03
09
Ca
l
c
im
udsto
n
e
-7
,
22 -
9,
2
6
39
pe
l
o
i
da
l
pac
k
sto
n
e
-
9,36
-1
0,89
1
0
Ca
l
c
im
udsto
n
e
-
8,9
2 -
5,
1
5
4
0
pe
l
o
i
da
l
pac
k
sto
n
e
-
9,
1
5
-12
,38
11
Ca
l
c
im
udsto
n
e
-
8,
1
6
-
6,09
41 f
e
n
est
r
a
l
ca
l
ca
r
e
ni
te
-
9,83
-1
0,
7
0
12
Ca
l
c
im
udsto
n
e
-
8,6
2 -
6,3
1 42 f
e
n
est
r
a
l
ca
l
ca
r
e
ni
te
-
8,00
-1
0,00
1
3
ost
r
acoda
l
pac
k
sto
n
e
-
8,
2
8
-
8,50
4
3
ost
r
acoda
l
pac
k
sto
n
e
-
8,6
4 -12
,
12
A
44 f
e
n
est
r
a
l
ca
l
ca
r
e
ni
te
-
9,9
4 -11
,80
4
5
Ca
l
c
im
udsto
n
e
-
8,83
-11
,
74
14
ost
r
acoda
l
pac
k
sto
n
e
-
6,05
-11
,99
4
6
ost
r
acoda
l
pac
k
sto
n
e
-
8,
1
8
-11
,5
7
1
5
ost
r
acoda
l
pac
k
sto
n
e
-7
,0
2 -12
,6
7 47
Ca
l
c
im
udsto
n
e
-
8,99
-12
,
12
1
6
ost
r
acoda
l
pac
k
sto
n
e
-
3,39
-7
,98
4
8
Ca
l
c
im
udsto
n
e
-
9,
24 -1
3,0
2
17
ost
r
acoda
l
pac
k
sto
n
e
-4
,
7
5
-
9,6
4
9
Ca
l
c
im
udsto
n
e
-
8,6
2 -11
,
2
3
1
8
p
i
so
i
da
l
pac
k
sto
n
e
-4
,
2
8
-1
0,8
50
Ca
l
c
im
udsto
n
e
-
9,
2
9
-11
,
1
8
1
9
ost
r
acoda
l
pac
k
sto
n
e
-
6,6
2 -1
0,
4
8
5
1
pe
l
o
i
da
l
pac
k
sto
n
e
-
8,06
-
6,
1
0
2
0
ost
r
acoda
l
pac
k
sto
n
e
-
6,5
1 -1
0,
71
5
2
ost
r
acoda
l
pac
k
sto
n
e
-
6,65
-1
0,
14
21
ost
r
acoda
l
pac
k
sto
n
e
-
5,
27 -12
,39
53
ost
r
acoda
l
pac
k
sto
n
e
-7
,
42 -
5,69
22
Ca
l
c
im
udsto
n
e
-1
0,8
-11
,
47
2
3
Ca
l
c
im
udsto
n
e
-1
0,36
-11
,
27
24
Ca
l
c
im
udsto
n
e
-1
0,
44 -11
,
14
2
5
Ca
l
c
im
udsto
n
e
-7
,58
-11
,89
2
6
ost
r
acoda
l
pac
k
sto
n
e
-7
,6
-12
,56
27
ost
r
acoda
l
pac
k
sto
n
e
-4
,
77 -12
,
1
6
2
8
ost
r
acoda
l
pac
k
sto
n
e
-
3,0
1 -
6,8
1
2
9
ost
r
acoda
l
pac
k
sto
n
e
-2
,
71 -7
,55
30
Ca
l
c
im
udsto
n
e
-4
,80
-
9,0
4
B
C
129
The fluctuation in carbon isotope ratios is fairly comparable within individual second-
order cycles, particularly in the intermediate and upper portions of the successions, where
values, in general, increase, or less commonly, remain constant. Downward in the cycles, the
carbon ratios oscillate slightly between lighter and heavier. A change in pattern occurs in the
lower portions of the profiles (Fig. 5-7), where there is an upward trend to lower values.
Similarly to the carbon, the oxygen isotope ratios obtained for the Codó Formation are
dominantly low, ranging from –2.71‰ to –10.80‰. The behavior of the curves along the
profiles, as well as within the first- and second-order cycles, also shows patterns that are very
similar to the ones described for the carbon, but with oscillations toward lighter values in the
lower and intermediate portions of the cycles. Exception is an increased value in the second-
order cycles located at the lower portions of profiles A and C, where the oxygen isotope
values increase.
Comparisons of the carbon and oxygen curves reveal a positive covariance in profiles
B and C, with values ranging from +0.42 to +0.43, while profile A shows a negative
covariance of –0.36 (Fig. 5-8). This behavior is reflected in both the first- and second-order
cycles. The only exception is the second-order cycle at the base of profile C, which displays an
inverse covariance.
5.8. ISOTOPIC CHARACTERIZATION OF THE LACUSTRINE SYSTEM
The carbon and oxygen data presented here have valuable application for further
support the lacustrine signature of the Codó Formation, as well as reconstruct its
paleohydrology and evolution through time. This procedure was made possible only
considering the primary signature of these data, as confirmed by the petrographic and minor
and trace elements discussed earlier in this paper. In addition, the wide distribution of the
carbon and oxygen values throughout the analysed profiles, their overall good covariance, and
pulsating oscillations according to the first- and second-order shallowing-upward cycles, as
described above, led to exclude a significant diagenetic influence.
130
131
132
The results of the isotope analysis are in agreement with a lacustrine interpretation for
the Codó Formation exposed in the eastern margin of the Grajaú Basin, as proposed in
previous publications (e.g., Campbell et al., 1949; Aranha et al., 1991; Rossetti et al., 2000;
Paz & Rossetti, 2001). Several observations related to the carbon and oxygen isotope data
support this interpretation. First, a non-marine setting is indicated by the exclusive occurrence
of values lighter than –5.15 ‰ for the carbon, which is well below the range of about –2‰
and +5‰ expected for marine limestones (e.g., Deines, 1980; Hoefs, 1980). Second, the
carbon values obtained in the study area are consistent with Late Aptian continental deposits,
since marine limestones of this age display values ranging from +2%o to +4%o (Jones &
Jenkins, 2001; Fig. 5-9). Third, the ratios –2.71‰ to -10.80 ‰ for the oxygen isotopes are also
consistent with a continental setting (Talbot, 1990; Bird et al., 1991). A marine-influenced
continental environment might show lighter values of up to –5‰ (e.g., Ingram et al., 1996;
Hendry & Kalin, 1997; Fig. 5-10), but this hypothesis is very unlikely in this instance because
91% of the analysed samples are below this value. Fourth, the overall wide range of both δ
13
C
and δ
18
O values, is typical of continental-derived waters, as considered in a number of works
(e.g., Talbot & Kelts, 1990; Casanova & Hillaire-Marcel, 1993; Camoin et al., 1997; Fig. 5-
10). Fifth, the good covariance between the oxygen and carbon isotope data (Fig. 5-8),
although not exclusive to, is more consistent with a non-marine setting (Turner et al., 1983;
Gasse et al., 1987; Talbot, 1990; Talbot & Kelts, 1990; Charisi & Schmitz, 1995).
133
Fig . 5-10: Plots of carbon and oxygen stable isotope data from several marine and lacustrine deposits throughout
the world (Modified from Hendry and Kalin, 1997), and their comparison with data obtained in the Codó
Formation. This diagram shows that the isotopic composition of the Codó Formation is in conformity with
isotope data from lacustrine limestones.
In addition to support a non-marine deposition, the carbon and oxygen isotope data
revealed to be valuable for reconstructing lake paleohydrology. Both of these isotopes have
been used directly or indirectly to interpret climate (Fig. 5-11). In fact, temperature and
hydrological balance are the main controllers of isotopic composition in lake systems (Kelts &
Talbot, 1990; Lister et al., 1991). It is well known from studies of modern settings that
temperature causes fractionation of the oxygen in a constant ratio of 0.26 ‰/
o
C in the
bicarbonate-water-carbonate system (cf. Craig, 1965; Friedman & O’Neill, 1977). However,
the wide range of δ
18
O values observed in the Codó Formation would require a temperature
gradient equivalent to about 40
o
C, not expected considering a low paleolatitudinal location
(>10
o
) of the study area during the Late Aptian (Scotese et al., 1989). On the other hand, the
balance between influx and evaporation causes drastic changes in lake isotopic composition
134
(Talbot, 1990; Lister et al., 1991). Evaporation leads to enriched
18
O, as lighter
16
O escape to
the atmosphere. Conversely, high water inflow results in the return of
16
O from the
atmosphere, causing depletion in δ
18
O values. Thus, low δ
18
O values have been related to
higher lake levels, while high δ
18
O are attributed to lower lake levels, parameters that have
been indirectly related to climate (e.g., Talbot, 1990; Caimon et al., 1997; Lister et al., 1991).
Carbon isotope ratios also have a direct relation to climate. Hence, higher δ
13
C values
have been associated with aridity, while lower δ
13
C values indicate relatively more humid
climates (e.g., Talbot & Keltz, 1990; Valero-Garcés et al., 1995). This interpretation is based
on the fact that dry climates favour evaporation, increased influence of C4-path vegetation
type, lower influx, and lake stratification, which ultimately lead to organic matter preservation
with the consequent output of
12
C from the lake water.
A close relationship between the carbon and oxygen isotope ratios and the first- and
second-order shallowing-upward cycles is recorded in the study area. These changes are
135
analysed in the following in terms of facies development, which is not necessarily related to
climate changes, as widely applied in the literature (e.g., Olsen, 1986; Smoot & Olsen, 1994;
Goldhammer et al., 1990; Steenbrick et al., 2000; Hofmann et al., 2000; Aziz et al., 2000). In
this instance, the variation in both isotopes from lighter to heavier values along individual
second-order cycles follows upward gradations from central to marginal facies deposits, which
are related to changes in subsidence rates, as previously discussed (Fig. 5-12). Such facies
stacking requires alternating episodes of lake deepening and shallowing, which is associated
with increased subsidence and the return to relative stability or even uplift, respectively.
Extreme shallowing might have culminated with periods of desiccation, as indicated by
deposits with paleosols. Heavier isotope values recorded during these situations could be
attributed to a significant enhancement of the isotopic exchange with the atmosphere. This is
because as the water became extremely shallow, evaporation increased significantly due to
heating (Fig. 5-12). Differences in carbon values according to location in the lake system, with
marginal areas displaying higher values, have been also noted by other authors (e.g., Camoin
et al., 1997; Casanova & Hillaire-Marcel, 1993). The loss of
12
C to the atmosphere as a cause
to enrich the
13
C in the dissolved inorganic carbonate appears to be an active process in the
epilimnion of lakes with low water inflow (Stiller et al., 1985; Talbot, 1990; Talbot & Kelts,
1990).
The upward lighter trends of carbon and oxygen isotope values observed in the lower
portions of the studied successions is related to lake stratification and bottom anoxia, as
suggested by the prevalence of central lake deposits consisting of evaporites and bituminous
black shales. Such situation leads to
12
C burial, increasing the amount of
13
C dissolved in the
central lake waters (Herczeg, 1988; Talbot & Kelts, 1990). The subtle upward decrease in
isotope values would record the transition to intermediate lake environments, where this effect
was less significant. In addition, as opposed to the shallowing-upward cycles located upward
in the studied succession, marginal lake deposits that could contribute to increase the isotope
values through atmosphere exchange, as proposed above, are in general lacking in the upper
portions of these cycles.
136
137
The overall upward high-low-high trend of the isotope values coincides with the
position of the first-order cycles. As presented earlier, these cycles record main episodes of
lake desiccation superposed upon second-order cycles. The lightest values in the lower
portions of the first-order cycles is related to periods of maximum flooding, when the lake was
established and deeper water levels prevailed, as reflected by the better development of central
lake deposits. As the lake evolved, increased evaporation led to a lower water level,
progressively enhancing the isotope values and ultimately culminating with lake desiccation,
when the isotope values reached the heaviest ratios. This trend, well documented in unit 3, is a
little modified in the lower portions of unit 2, where the heaviest ratios were recorded as a
result of lake stratification and anoxia, as this time coincides with the maximum development
of evaporites and/or bituminous black shales in this lacustrine basin. The abundance of these
lithologies in unit 2, where relatively deeper water facies prevail, lead to propose that, rather
than evaporation, lake stratification and anoxia were crucial to form the evaporites (e.g.,
Warren, 1999). As the lake evolved through time, shallower water conditions became
progressively better developed and evaporation increased, but with restrict or no precipitation
of evaporative minerals.
The overall moderate positive covariance between the carbon and oxygen isotope
curves shown in profiles B and C is attributed to the prevalence of a closed lake environment,
a pattern that have been noticed in many other ancient and modern lake systems throughout
the world (e.g., Eicher & Siegenthaler, 1976; Gat, 1981; McKenzie, 1985; Gasse et al., 1987;
Janaway & Parnell, 1989; Talbot, 1990; Talbot & Kelts, 1990; Lister et al., 1991). It is
interesting to note that unit 2 of profile C coincides with a period of high covariance. This is
consistent with the presence of evaporites in these beds, typical of closed basins (Warren,
1999). In spite the evidences for lake closure, a comparison between the isotope curves shows
alternating phases of higher and relatively lower covariance, attributed to episodes of lake
opening. This can be particularly seen in profile A (Fig. 5-7), where the overall covariance
between carbon and oxygen isotopes is negative.
It is interesting to observe that the δ
18
O displayed the highest interval of variation
during closed phases, which ranged from -3.63‰ to –4.89‰, against the interval of variation
of –0.09‰ to -1.87 ‰ that characterizes open phases. This is because closed lakes have better
138
chances to show oscillations in water levels as the residence time increases, which lead to
enhance the δ
18
O. Conversely, the isotopic composition of open lakes is more stable due to the
balance caused by the continuous basin inflow, as recorded in several lake systems, such as
Lake Henderson (Stuiver, 1970), Lake Huleh (Stiller & Hutchinson, 1980) and Greifensee
(McKenzie, 1985).
5.9. FINAL REMARKS
The carbon and oxygen isotopic composition of the Codó Formation in the eastern
Grajaú Basin can be directly related to facies changes, as revealed by the good correspondence
between the isotope values and the shallowing-upward cycles. Deciphering the causes of these
changes through time, whether related to climate fluctuation or to any other variation in the
basin, such as subsidence or sediment inflow, is not so straightforward. There has been an
agreement among the authors in relating carbon and oxygen isotopes with lake hydrology,
which is often used directly or indirectly to make inferences about climate (e.g., Smoot &
Olsen, 1994; Steenbrick et al., 2000; Hofmann et al., 2000; Aziz et al., 2000). However, this
study shows that significant changes in carbon and oxygen isotope composition of lake waters,
resembling climatic cycles, can be also related to fluctuations in subsidence rates caused by
syn-sedimentary seismic activity. In this case, combination of isotope and sedimentological
data provided the key for distinguishing which of these factors left the most significant imprint
in the sedimentary record.
We have shown in this paper the close relationship of both carbon and oxygen isotopes
with the first- and second-order shallowing-upward cycles that characterize the studied unit.
The dominant asymmetric nature of these cycles, inferred with basis on the high facies
variability when comparing one cycle to another, the limited lateral extension, as well as the
frequent and random thickness changes, led to propose that climate was not their primordial
controlling factor. On the other hand, sedimentological data favour to attribute these cycles to
syn-sedimentary seismic activity associated with the early tectonic evolution of the São Luís-
Grajaú Basin during the Late Aptian. The Codó Formation was deposited just prior to the main
rifting that culminated with the process of opening of the South Atlantic Equatorial Ocean.
During this initial time, there was the development of a shallow, but extensive subsiding
139
intracontinental basin (Azevedo, 1991; Góes & Rossetti, 2001). This characteristic matches
well with the presence of shallow lakes in marginal areas of the basin, where faults with
reduced offsets are expected to have prevailed. Subsidence gave rise to local water
accumulation, forming closed and perennial lake systems, but as compression took place, and
the area was uplifted, the development of ephemeral lakes appears to have been favoured. This
situation is recorded in the studied profiles by a change from shallowing-upward cycles with
dominance of central and intermediate lake deposits, to cycles with well-developed marginal
lake deposits, as occurs in the lower and upper portions of the first-order cycles, respectively.
Such facies arrangement records the upward transition from periods of maximum flooding to
periods when the lake level fell to lower levels. The carbon and oxygen isotopic composition
of such lake basins is expected to be characterized initially by light values, but as the residence
time increases due to shallowing, heavier values are reached due to the high evaporation-
inflow budget provided by the progressive decrease in water level.
Because arid climate prevailed along the Brazilian equatorial margin during the Late
Aptian (Lima et al., 1980; Lima, 1982; Batista, 1992; Rodrigues, 1995; Rossetti et al., 2001),
evaporite precipitation took place in central lake areas, induced by water stratification and
bottom anoxia (Kirkland & Evans, 1981). The highest value of carbon isotope coinciding with
the moment of maximum formation of evaporites and bituminous black shales is consistent
with this interpretation.
Therefore, different styles and/or intensities of seismic pulses alternating with sediment
deposition might cause changes in the lake level, promoting alternating periods of flooding
and fall in the lake level, and resulting in well-developed asymmetric shallowing-upward
cycles. Such scenario will ultimately affect the overall isotope composition of lake waters.
Acknowledgments. The Itapicuru Agroindustrial S/A is acknowledged for the
permission to access the quarries with the exposures of the Codó Formation. This work was
financed by the Brazilian Council for Research–CNPq (Project #460252/01). The authors are
grateful to Dr. Alcides Sial e Dr. Valderez Ferreira for helping with the geochemical analyses.
140
REFERENCES
Abell, P.I., Williams, M.A.J., 1986. Sedimentary carbonates as isotopic marker horizons at
Lake Turkana, Kenya. In: Frostick, L.E., Renaut, R.W. Reid, I., Tiercelin, J.-J. (Eds.),
Sedimentation in the African Rifts. Spec. Publ., Geol. Soc. Am. 25, 153-158.
Anderson, F.W., Arthur, M.A., 1983. Stable isotopes of oxygen and carbon and their
application to sedimentologic and paleoenvironmental problems. Short Course Notes,
SEPM, 10, 151 p.
Aranha, L.G., Lima, H.P., Makino, R.K., Souza, J.M., 1991. Origem e evolução das bacias de
Bragança-Viseu, São Luís e Ilha Nova. In: Milani, E.J., Raja Gabaglia, G.P. (Eds.),
Origem e Evolução das Bacias Sedimentares. PETROBRAS, Rio de Janeiro, pp. 221-234.
Azevedo R.P., 1991. Tectonic evolution of Brazilian Equatorial Continental Margin Basins.
Doctoral Thesis, University of London, London, 455 p.
Aziz, H.A., Hilgen, F., Krijgsman, W., Sanz, E., Calvo, J.P., 2000. Astronomical forcing of
sedimentary cycles in the middle to late Miocene continental Catalayud Basin (NE Spain).
Earth Planet. Sci. Lett. 177, 9-22.
Batista, A.M.N., 1992. Caracterização Paleoambiental dos sedimentos Codó-Grajaú, Bacia de
São Luís (MA). M.Sc.Thesis, Universidade Federal do Pará, Belém, 102 p.
Bellanca, A., Calvo, J.P., Censi, P., Elizaga, E., Neri, R., 1989. Evolution of lacustrine
diatomite carbonate cycles of Miocene age, southeastern Spain: petrology and isotope
geochemistry. J. Sediment. Petrol. 59, 45-52.
Benvenuti, M., 2003. Facies analysis and tectonic significance of lacustrine fan-deltaic
successions in the Pliocene-Pleistocene Mugello Basin, Central Italy. Sediment. Geol. 157,
197-203.
Bird, M.I., Chivas, A.R., Radnell, C.J., Burton, H.R., 1991. Sedimentological and stable
isotope evolution of lakes in the Vestfold Hills, Antarctica. Palaeogeog., Palaeoclim.,
Palaeoecol. 84, 109-130.
Botz, R., Stoffers, P., Faber, E., Tietze, K., 1988. Isotope geochemistry of carbonate sediments
from Lake Kivu (East Central Africa). Chem. Geol. 69, 299-308.
Camoin, G., Casanova, J., Rouchy, J.M., Blanc-Valleron, M.M, Deconinck, J.F. 1997.
Environmental controls on perennial and ephemeral carbonate lakes: the Central Palaeo-
141
Andean Basin of Bolivia during Late Cretaceous to early Tertiary times. Sediment. Geol.
113, 1-26.
Campbell, D.F., Almeida, L.A., Silva, S.O., 1949. Relatório preliminar sobre a geologia da
Bacia do Maranhão. Boletim do Conselho Nacional do Petróleo 1, 1-160.
Casanova, J., Hillaire-Marcell, C., 1993. Carbon and oxygen isotopes in African lacustrine
stromatolites: palaeohydrological interpretation. In: Swart, P.K., Lohmann, K.C.,
McKenzie, J., Savin, S. (Eds.), Climate Change in Continental Isotopic Record. Geophys.
Monogr., Am. Geophys. Union, 94, 123-133.
Craig, H., 1965. The measurement of oxygen isotope palaeotemperatures. In: Tongiorgi (Ed),
Stable Isotopes in Oceanographic Studies and Palaeotemperatures. Cons. Naz. Rich. Lab.
Geol. Nucl., Pisa, pp. 9-130.
Charisi, S.D., Schmitz, B., 1995. Stable (
13
C,
18
O) and strontium (
87
Sr/
86
Sr) isotopes through
the Paleocene et Gebel Aweina, eastern Tethyan region. Palaeogeog., Palaeoclim.,
Palaeoecol. 116, 103-129.
Deines, P., 1980. The isotopic composition of reduced organic carbon. In: Fritz, P.J., Fontes, J.
(Eds.), Handbook of Environmental Isotope Geochemistry: v.1, Terrestrial Environment.
Elsevier, Amsterdam, pp. 329-406.
Eicher, U., Siegenthaler, U., 1976. Palynological and oxygen isotope investigations on Late-
Glacial sediment cores from Switzerland. Boreas 5, 109-117.
Friedman, I., O’Neill, J.R., 1977. Compilation of stable isotope fractionation factors of
geochemical interest. In: Fleischer, M. (Ed.), Data of Geochemistry. Geological survey
Prof. Paper 440-KK, U.S. Gov. Print. Office, Washington, pp. 1-12.
Fritz, P., Anderson, T.W., Leqis, C.F.M., 1975. Late Quaternary trends and history of Lake
Erie from stable isotope studies. Science 190, 267-269.
Gasse, F., Fontes, J.C., Plaziat, J.C., Carbonel, P., Kaczmarska, I., De Decker, P., Soulié-
Marsche, I., Callot, Y., Depeuble, P.A., 1987. Biological remains, geochemistry and stable
isotopes for the reconstruction of environmental and hydrological changes in the Holocene
lakes from north Sahara. Palaeogeog., Palaeoclim., Palaeoecol. 60, 1-46.
Gasse, F., Lédée, V., Massault, M., Fontes, J.C., 1989. Water-level fluctuations of Lake
Tanganyika in phase with the oceanic changes during the last glaciation and deglaciation.
Nature 342, 57-59.
142
Gat, J.R., 1980. The isotopes of hydrogen and oxygen in precipitation. In: Fritz, P.J., Fontes, J.
(Eds.), Handbook of Environmental Isotope Geochemistry: v.1, Terrestrial Environment.
Elsevier, Amsterdam, pp. 21-47.
Gat, J.R., 1984. The stable isotope composition of Dead Sea waters. Earth Planet. Sci. Lett.
71, 361-376.
Góes, A.M., Rossetti, D.F., 2001. Gênese da Bacia de São Luís-Grajaú. In: Rossetti, D.F.,
Góes, A.M., Truckenbrodt, W. (Eds.), O Cretáceo na Bacia de São Luís Grajaú. Museu
Paraense Emilio Goeldi, Coleção Friedrich Katzer, Belém, pp. 15-29.
Goldhammer, R.K., Dunn, P.A., Hardie, L.A., 1990. Depositional cycles, composite sea-level
changes cycle stacking patterns, and hierarchy of stratigraphic forcing: examples from
Alpine Triassic platform carbonates. Geol. Soc. Am. Bull. 102, 535-562.
Hendry, J.P., Kalin, R.M., 1997. Are oxygen and carbon isotopes of mollusc shells reliable
palaeosalinity indicators in marginal marine environments? A case study from Middle
Jurassic of England. J. Geol. Soc. London 154, 321-333.
Herczeg, A.L., 1988. Early diagenesis of organic matter in lakes sediments: a stable carbon
isotope study of pore waters. Chem. Geol. 72, 199-209.
Herczeg, A.L, Leaney, F.W., Dighton, J.C., Lamontagne, S., Schiff, S.L., Telfer, A.L.,
English, M.C., 2003. A modern isotope record of changes in water and carbon budgets in a
groundwater-fed lake: Blue Lake, South Australia. Limnol. Oceanogr. 48, 2093-2105.
Hillaire-Marcell, C., Casanova, J., 1987. Isotopic hydrology and paleohydrology of the
Magadi (Kenya)-Natron (Tanzania) basin during Late Quaternary. Paleogeog., Paleoclim.,
Paleoecol. 58, 155-181.
Hoefs, J., 1980. Stable Isotope Geochemistry. 2
nd
Ed. Springer-Verlag, Berlim, 208 pp.
Hofmann, A., Tourani, A., Gaupp, R., 2000. Cyclicity of Triassic to lower Jurassic continental
red beds of the Argana Valley, Morocco: implications for paleoclimate and basin
evolution. Palaeogeogeog., Palaeoclim., Palaeoecol. 161, 229-266.
Holland, H.D., 1978. The Chemistry of the Atmosphere and Oceans. Wiley-Interscience, New
York, 351 pp.
Holmes, J.A., Street-Perrott, F.A., Allen, M.J., Fothergill, P.A., Harkness, D.D., Kroon, D.,
Perrott, R.A., 1997. Holocene palaeolimnology of Kajemarum Oasis, Northern Nigeria: an
143
isotopic study of ostracodes, bulk carbonate and organic carbon. J. Geol. Soc. London 154,
311-319.
Ingram, B.L., Ingle, J.C., Conrad, M.E., 1996. A 2000-yr record of Sacaramento-San Joaquin
River inflow to the San Francisco Bay estuary, California. Geology 24, 331-334.
In Sung, P., Kim, H.J., 2003. Palustrine calcretes of the Cretaceous Gyeongsang Supergroup,
Korea: Variation and paleoenvironmental implications. The Island Arc 12, 110-124.
Janaway, T.M., Parnell, J., 1989. Carbonate production within the Orcadian Basin, northern
Scotlabd: a petrographic and geochemical study. In: Talbot, M.R., Kelts, K (Eds.), The
Phanerozoic Record of Lacustrine Basins and their Environmental Signals. Palaeogeogr.
Palaeoclimatol. Palalaeoecol. 70, 89-105.
Johnson, T.C., Halfman, J.D., Showers, W.J., 1991. Paleoclimate of the past 4000 years at
Lake Turkana, Kenya, based on the isotopic composition of authigenic calcite.
Palaeogeog., Palaeoclim., Palaeoecol. 85, 189-198.
Jones, C.E., Jenkins, H.C., 2001. Seawater strontium isotopes, oceanic anoxic events, and
seafloor hydrothermal activity in the Jurassic and Cretaceous. Am. J. Sci. 301, 112-149.
Katz, A., Y. Kolodny, A. Nissenbaum, 1977. The geochemical evolution of Lake Lisan– Dead
Sea system: Geochim. Cosmochim. Acta 41, 1609-1626.
Kirkland, D.W., Evans, R., 1981. Source-rock potential of evaporitic environment. AAPG
Bull. 65, 181-190.
Kelts, K., Talbot, M., 1990. Lacustrine carbonates as geochemical archives of environmental
change and biotic/abiotic interactions. In: Tilzer, M.M., Serruya, C. (Eds.), Large lakes:
Ecological Structure and Function. Brock/Springer Series in Contemporary Bioscience,
Springer-Verlag, pp. 288-315.
Lima, M.R., 1982. Palinologia da Formação Codó, Maranhão. Boletim do Instituto de
Geociências-USP 13, 116-128.
Lima, M.R., Fulfaro, V.J., Bartorelli, A., 1980. Análise palinológica de sedimentos cretáceos
da região de Marabá, Estado do Pará. Boletim do Instituto de Geociências, USP 11, 55-
161.
Lister, G.S., Kelts, K., Chen, K.Z., Yu, J.-Q., Niessen, F., 1991. Lake Qingai, China: closed-
lake basin levels and the oxigen isotope record for ostracoda since the latest Pleistocene.
Palaeogeogr., Palaeoclimatol., Palaeoecol. 84, 141-162.
144
Martel, A.T., Gibling, M.R., 1991. Wave-dominated lacustrine facies and tectonically
controlled cyclicity in the Lower Carboniferous Horton Bluff Formation, Nova Scotia
Canada. In: Anadón, P., Cabrera, Ll., Kelts, K. (Eds.), Lacustrine Facies Analysis. Int.
Ass. Sediment. Spec. Publ. 13, 223-244.
McKenzie, J.A., 1985. Carbon isotopes and productivity in the lacustrine and marine
enviroment. In: Stumm, W. (Ed.), Chemical Processes in Lakes. Wiley, New York, pp. 99-
118.
Mesner, J.C., Wooldrigde, L.C.P., 1964. Maranhão Paleozoic Basin and Cretaceous coastal
basins, north Brazil. AAPG Bull. 48,1475-1512.
Olsen, P.E., 1986. A 40-million-year lake record of Early Mesozoic orbital climatic forcing.
Science 234, 842-848.
Olsen, P.E., Kent, D.V., 1996. Milankovitch climate forcing in the tropics of Pangea during
the Late Triassic. Palaeogeog., Palaeoclim., Palaeoecol. 122, 1-26.
Paz, J.D.S., Rossetti, D.F., 2001. Reconstrução paleoambiental da Formação Codó (Aptiano),
borda leste da Bacia do Grajaú, MA. In: Rossetti, D.F., Góes, A.M., Truckenbrodt, W.
(Eds.), O Cretáceo na Bacia de São Luís-Grajaú. Museu Paraense Emilio Goeldi, Coleção
Friedrich Katzer, Belém, pp. 77-101.
Paz, J.D.S., Rossetti, D.F., submitted. Tectonically-driven lacustrine cycles: an example from
the Codó Formation (Late Aptian), northeastern Brazil. Geol. Mag.
Rezende, W.M., Pamplona, A.H.R.P., 1970. Estudo do desenvolvimento do Arco Ferrer-
Urbano Santos. PETROBRAS, Boletim Técnico 13, 5-14.
Rodrigues, R., 1995. A geoquímica Orgânica na Bacia do Parnaíba. Doctoral Thesis,
Universidade Federal do Rio Grande do Sul, Porto Alegre, 225 p.
Rosenmeier, M.F., Hodell, D.A., Brenner, M., Curtis, J.H., Guilderson, T.P., 2002. A 4000-
year lacustrine record of environmental change in the Southern Maya Lowlands, Petén,
Guatemala. Quaternary Res. 57, 183-190.
Rossetti, D.F., 2001. Arquitetura deposicional da Bacia de São Luís-Grajaú, meio norte do
Brasil. In: Rossetti, D.F., Góes, A.M., Truckenbrodt, W. (Eds.), O Cretáceo na Bacia de
São Luís Grajaú. Museu Paraense Emilio Goeldi, Coleção Friedrich Katzer, Belém, pp.
31-46.
145
Rossetti, D.F., Góes, A.M., 2000. Deciphering the sedimentological imprints of paleoseismic
events: an example from the Codó Formation, northern Brazil. Sediment. Geol. 135, 137-
156.
Rossetti, D.F., Truckenbrodt, W., 1997. Classificação estratigráfica para o Albiano-Terciário
Inferior (?) na Bacia de São Luís, MA. Boletim do Museu Paraense Emílio Goeldi (Série
Ciências da Terra) 9, 31-43.
Rossetti, D.F., Paz, J.D., Góes, A.M., in press. Facies analysis of the Codó Formation (Late
Aptian) in the Grajaú area, southern São Luís-Grajaú Basin. An. Acad. Bras. Cien.
Rossetti, D.F., Truckenbrodt, W., Santos-Júnior, A.E., 2001. Clima do Cretáceo no Meio-
Norte brasileiro. In: Rossetti, D.F., Góes, A.M., Truckenbrodt, W. (Eds.), O Cretáceo na
Bacia de São Luís Grajaú. Museu Paraense Emilio Goeldi, Coleção Friedrich Katzer,
Belém, pp. 67-76.
Rossetti, D.F., Paz, J.D., Góes, A.M., Macambira, M., 2000. A marine versus non-marine
origin for the Aptian-Albian evaporites of the São Luís and Grajaú basins, Maranhão state
(Brazil) based on sequential analysis. Rev. Bras. Geoc. 30, 642-645.
Russell, J.M., Johnson, T.C., Talbot, M.R. 2003. A 725-yr cycle in the climate of Central
Africa during the Late Holocene. Geology 31, 677-680.
Scotese, C.R., Gahagan, L.M., Larson, R.L., 1989. Plate tectonic reconstructions of the
Cretaceous and Cenozoic ocean basins. In: Scotese, C.R., Sager, W.W. (Eds.), Mesozoic
and Cenozoic Plate Reconstructions, Elsevier, Amsterdam, pp. 22-48.
Smoot, J.P., Olsen, P.E., 1994. Climatic cycles as sedimentary controls of rift-basin lacustrine
deposits in the Early Mesozoic Newark Basin based on continuous core. In: Lomando,
A.J., Schreiber, B.C., Harris, P.M. (Eds.) Lacustrine Reservoirs and Depositional Systems.
SEPM Core Workshop, 19, 239-295.
Steenbrink, J., Van Vugt, N, Kloosterboer-Van Hoeve, M.L., Hilgen, F.J., 2000. Refinement
of the Messinian APTS from sedimentary cycle patterns in the lacustrine Lava section
(Servia Basin, NW Greece). Earth Planet. Sci. Lett. 181, 161-173.
Stiller, M., Hutchinson, G.E., 1980. The waters of Meron: a study of Lake Huleh, 1. Stable
isotopic composition of carbonates of 54 m core: paleoclimatic and paleotrophic
implications. Arch. Hydrobiol. 89, 275-302.
146
Stiller, M., Rounick, J.S., Shasha, S., 1985. Extreme carbon-isotope enrichments in
evaporating brines. Nature 316, 434-435.
Stuiver, M., 1970. Oxygen and carbon isotope ratios of fresh-water carbonates as climatic
indicator. J. Geophys. Res. 75, 5247-5257.
Tan, F.C., Hudson, J.D.,1974. Isotopic studies on the palaeoecology and diagenesis of the
Great Estuarine Series (Jurassic) of Scotland. Scottish J. Geol. 10, 91-128.
Talbot, M.R., 1990. A review of the paleohydrological interpreatation of carbon and oxygen
isotopic ratios in primary lacustrine carbonates. Chem. Geol. 80, 261-279.
Talbot, M.R., Kelts, K., 1990. Paleolimnological signatures from carbon and oxygen isotopic
ratios in carbonates from organic carbon-rich lacustrine sediments. In: Katz, B.J. (Ed.),
Lacustrine Basin Exploration: Case Studies and Modern Analogs. AAPG Memoir 50, 99-
112.
Tucker, M.E., Wright, V.P., 1990. Carbonate Sedimentology. Blackwell Science, Oxford, 482
pp.
Turner, J.V., Fritz, P., Karrow, P.F., Warner, B.G., 1983. Isotopic and geochemical
composition of marl lake waters and implications for radiocarbon dating of marl lake
sediments. Can. J. Earth Sci. 20, 599-615.
Urey, H.C. 1947. The thermodynamic properties of isotopic substance. Geochim. Cosmochim.
Acta 48, 562-581.
Valero-Garcés, B.L., Kelts, K., Ito, E., 1995. Oxigen and carbon isotopes trends and
sedimentological evolution of a meromictic and saline lacustrine system: the Holocene
Medicine Lake basin, North American Great Plains, USA. Palaeogeogr., Palaeoclimatol.,
Palaeoecol. 177, 253-278.
Warren, J. K., 1999. Evaporites: Their Evolution and Economics: Blackwell Scientific,
Oxford, 438 pp.
147
6. CONCLUSÕES GERAIS
1. A Formação Codó consiste em depósitos dominantemente continentais nas áreas
estudadas, onde se distinguem dois contextos deposicionais: a) lagos estáveis, bem
estratificados e salinos, com períodos de fechamento e conseqüente anoxia do sistema
prevalecendo na margem leste da Bacia de São Luís-Grajaú, com evaporitos restritos à porção
central nas partes mais profundas da bacia lacustre; e b) complexos de sabkha/saline pan na
porção sul da Bacia de São Luís-Grajaú, onde as condições climáticas e/ou tectônicas
favoreceram o desenvolvimento de lagos efêmeros, com a formação de extensas planícies de
lama contendo abundante precipitação de evaporitos centrada nas partes marginais do sistema;
2. A Formação Codó na área de estudo depositou-se de forma cíclica, formando ciclos
de arrasamento ascendente, interpretados como registro de períodos de expansão e contração
do sistema lacustre-sabkha-saline pan, gerados por flutuações climáticas e tectônicas;
3. Ciclos climáticos foram revelados por diversos pares milimétricos de fácies
deposicionais, arranjados em estilo semelhante a varves;
4. Ciclos tectônicos foram revelados por: a) alta variabilidade de fácies; b) limitada
extensão lateral dos ciclos; c) mudanças aleatórias na espessura e freqüência dos ciclos; e d)
correspondência com zonas de deformação sísmica sindeposicional;
5. O Sr mostrou-se como um parâmetro paleoambiental sensível, registrando valores
muito distintos daqueles esperados para águas marinhas neoaptianas;
148
6. Os valores de Sr também tiveram boa correlação com os ciclos deposicionais de
menor freqüência (i.e., ciclos de ordem superior), sendo sua variação atribuída a períodos de
contração e expansão da bacia lacustre;
7. O S mostrou-se menos sensível como indicador paleoambiental. Embora os valores
sejam condizentes com salmoura de origem continental na área de Codó, isto não se repetiu na
área de Grajaú, onde os valores diferem em uma faixa mais ampla do que aquela esperada em
ambientes com influência continental. É possível que os valores de S distintos entre as duas
áreas estudadas estejam refletindo variações nas condições de oxigenação do sistema
deposicional, com valores maiores na área de Codó, onde condições anóxicas prevaleceram,
incentivando a proliferação de bactérias sulfato-redutoras que influenciaram fortemente o
fracionamento do S e deixaram o sulfato residual da água do lago enriquecido
preferencialmente em
34
S.
8. Este trabalho demonstrou que flutuações nas taxas de subsidência, causadas por
atividade sísmica sin-sedimentar pode responder por variações nos valores dos isótopos de C e
O muito semelhantes àquelas observadas em sucessões resultantes formadas por flutuações de
ordem climática;
9. Diferentes estilos e/ou intensidades de pulsos sísmicos, alternados com deposição
sedimentar, causam mudanças no nível do lago e promovem períodos de inundação e
dessecação do sistema deposicional durante episódios de estiramento e compressão,
respectivamente. A deposição de carbonatos com valores de δ
13
C e δ
18
O preferencialmente
mais leves, associada com a instalação preferencial de lagos perenes na fase de inundação, e
de valores de δ
13
C e δ
18
O mais pesados, associados com a instalação de lagos efêmeros
durante dessecação, foram bem marcantes no ciclos de ordens intermediária e superior.
10. A análise de fácies detalhada e o mapeamento de padrões de empilhamento estratal
são a chave para a interpretação de sucessões progradantes como no caso da Formação Codó.
A combinação deste tipo de informações com dados isotópicos de C, O, Sr e S propicia
interpretações mais refinadas de sistemas deposicionais continentais onde ocorrem depósitos
de evaporitos e calcários. Além disto, este tipo de integração possibilita a disponibilização de
informações visando-se reconstituir condições paleohidrológicas do sistema deposicional e
dos fatores controladores de sua evolução.
Livros Grátis
( http://www.livrosgratis.com.br )
Milhares de Livros para Download:
Baixar livros de Administração
Baixar livros de Agronomia
Baixar livros de Arquitetura
Baixar livros de Artes
Baixar livros de Astronomia
Baixar livros de Biologia Geral
Baixar livros de Ciência da Computação
Baixar livros de Ciência da Informação
Baixar livros de Ciência Política
Baixar livros de Ciências da Saúde
Baixar livros de Comunicação
Baixar livros do Conselho Nacional de Educação - CNE
Baixar livros de Defesa civil
Baixar livros de Direito
Baixar livros de Direitos humanos
Baixar livros de Economia
Baixar livros de Economia Doméstica
Baixar livros de Educação
Baixar livros de Educação - Trânsito
Baixar livros de Educação Física
Baixar livros de Engenharia Aeroespacial
Baixar livros de Farmácia
Baixar livros de Filosofia
Baixar livros de Física
Baixar livros de Geociências
Baixar livros de Geografia
Baixar livros de História
Baixar livros de Línguas
Baixar livros de Literatura
Baixar livros de Literatura de Cordel
Baixar livros de Literatura Infantil
Baixar livros de Matemática
Baixar livros de Medicina
Baixar livros de Medicina Veterinária
Baixar livros de Meio Ambiente
Baixar livros de Meteorologia
Baixar Monografias e TCC
Baixar livros Multidisciplinar
Baixar livros de Música
Baixar livros de Psicologia
Baixar livros de Química
Baixar livros de Saúde Coletiva
Baixar livros de Serviço Social
Baixar livros de Sociologia
Baixar livros de Teologia
Baixar livros de Trabalho
Baixar livros de Turismo