Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Marine and Petroleum Geology 19 (2002) 1047–1071

www.elsevier.com/locate/marpetgeo

The role of depositional setting and diagenesis on the reservoir


quality of Devonian sandstones from the Solimões Basin,
Brazilian Amazonia
Rodrigo Dias Lima1, Luiz Fernando De Ros*
Instituto de Geociências, Universidade Federal do Rio Grande do Sul, Av. Bento Gonçalves, 9500, 91501-970 Porto Alegre, RS, Brazil
Received 31 July 2002; received in revised form 3 December 2002; accepted 7 December 2002

Abstract
Devonian sandstones of the Uerê Formation are important oil exploration targets in the Solimões Basin, western Brazilian Amazonia. The
basin fill comprises Ordovician to Permian sedimentary successions, Triassic diabase dykes and sills, and Cretaceous to Recent continental
deposits. This study deals with the Upper Devonian interval, which consists of sharp-based, progradational sandstones, attributed to a storm-
dominated shelf complex formed during an overall transgressive system tract, overlain by Frasnian-Famennian black shales. In spite of their
large lateral extent, the exploration of these sandstones is complicated by intense diagenesis, which strongly affected reservoir quality. The
main processes of porosity reduction are mechanical and chemical compaction and cementation by quartz overgrowths, carbonates (siderite
and dolomite) and fibrous illite. Porosity (up to 28%) was preserved by the inhibition of quartz overgrowth cementation and pressure
dissolution by grain-rimming, eogenetic, microcrystalline quartz and associated chalcedony, or smectite. Early diagenetic silica precipitation
is related to the dissolution of sponge spicules, which were concentrated in storm-reworked hybrid arenites and in interbedded spiculite
deposits. Locally, massive quartz cementation and recrystallisation occurred as a consequence of hot fluids circulation related to Triassic
magmatism.
q 2003 Elsevier Science Ltd. All rights reserved.
Keywords: Stratigraphy; Sandstone diagenesis; Reservoir quality

1. Introduction Although the vast Solimões Basin has been explored


throughout the past three decades, little is known about the
During 70s and 80s, diagenetic studies on clastic reservoir quality controls of the Devonian sandstones of the
reservoirs quality concentrated in the mechanisms of Uerê Formation. Most of the exploration efforts were
secondary porosity generation (Franks & Forester, 1984; concentrated in the Carboniferous sandstones of the Juruá
Giles & Marshall, 1986; Schmidt & McDonald, 1979; Formation, which contain the largest gas accumulations of
Surdam, Boese, & Crossey, 1984). On the other hand, last Brazil (close to 200 £ 109 m3 of gas in place, and
decade research dealt mostly with the diagenetic mechan- 11 £ 106 m3 of associated oil in place). The remote
isms of porosity preservation (Aase, Bjørkum, & Nadeau, geographic location of the Solimões Basin and the
1996; Bloch, Lander, & Bonell, 2002; Jahren & Ramm, extremely high costs of gas production and transportation
2000; Pittman, Larese, & Heald, 1992). The objective of this in the Amazonian forest has recently increased the interest
study is to unravel the controls on the observed preservation in the exploration for the Devonian sandstones, because
of porosity and variation of quality in the Devonian they contain mainly oil, occur throughout most of the basin,
sandstone reservoirs of the Solimões Basin (western and are closely associated to the major basin source rocks,
Brazilian Amazonia). the Jandiatuba shales (Mello, Koutsoukos, Mohriak, &
Bacoccoli, 1994). However, exploration of the Uerê
* Corresponding author. Fax: þ 55-51-3316-7047.
sandstones is complicated by the heterogeneous quality of
E-mail address: lfderos@inf.ufrgs.brs (L.F. De Ros).
1
National Petroleum Agency Grantee; present address: PETROBRAS these reservoirs, which range from highly porous (up to
Corporate University, Rio de Janeiro, RJ, Brazil. 28%) to extremely tight.
0264-8172/02/$ - see front matter q 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0264-8172(03)00002-3
1048 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Therefore, this study aims to understand the strati- Jurassic –Cretaceous transpressional tectonism, related to
graphic, depositional and petrologic controls on the quality the Andean orogenic events, which gave the area its present
of the Uerê reservoirs. The data, parameters and patterns configuration (Caputo & Silva, 1990; Fig. 1). Large
generated by this research will be applied in the construction compressional structures, in the form of reverse faults and
of predictive models to be used in the systematic exploration asymmetrical folds constitute the main hydrocarbon traps.
of the Devonian reservoirs. The basin was filled by six second-order depositional
sequences: Ordovician, Silurian-Devonian, Devonian-Car-
boniferous, Carboniferous-Permian, Cretaceous and Ter-
2. Geological setting tiary (Eiras et al., 1994; Fig. 2). The Devonian-Carboniferous
Sequence represents a transgressive-regressive cycle com-
The Solimões Basin is a large cratonic sag that covers posed of four successions, separated by regionally extensive
around 600 000 km2 of northwestern of Brazil, containing sequence boundaries, identified by miospore biostratigraphy
up to 4.5 km of marine to continental Palaeozoic deposits (Grahn, 1992; Loboziak, Melo, Quadros, Daemon, &
covered by Cretaceous to Tertiary continental deposits. Barrilari, 1994; Silva, 1987; Fig. 2). The most basal
Eohercynian extension caused normal faulting and uplift of succession is comprised of prograding Lower Devonian
the source areas of Brazilian Shield to the south, the deposits (Fig. 2) that occur only in the deeper Jandiatuba
Guyanas Shield to the north, and the Carauari Arch, which Sub-Basin and western portion of Juruá Sub-Basin. Over a
divided the Solimões Basin into the Jandiatuba and Juruá possibly transgressive surface (S2; Fig. 2), retro- to
sub-basins. The Carauari Arch exerted an important control aggradational, Eifelian-Early Givetian deposits occur at the
over the pre-Pennsylvanian sedimentation. Late Triassic to depocentre of Juruá Sub-Basin and eastern Jandiatuba Sub-
Early Jurassic diabase sills and dykes were intruded in the Basin. Early Givetian (?)-Frasnian sandstones and cherts
Devonian, Carboniferous and Permian sequences around (informally herein named Uerê sandstone) lay uncomform-
210 and 220 Ma ago, according to Ar/Ar dating (Szatmari, ably on Middle Devonian deposits and Precambrian
1996). The thermal pulse related to this basic magmatism metasediments and granitoids. The Uerê sandstones rep-
played an important role in hydrocarbon generation, resent a progradational to aggradational succession limited at
expulsion and migration from the Devonian source the base by a regional erosional surface (S3; Fig. 2).
rocks to Devonian and Carboniferous reservoirs (Caputo Radioactive Frasnian Jandiatuba shales regionally cover
& Silva, 1990). The extensional period was followed by the Uerê sandstones along the Devonian maximum flooding

Fig. 1. Location map of the Solimões Basin, with inset showing the São Mateus Field. Sampled wells shown as black dots. Wells with only log data in white.
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1049

Fig. 2. Simplified stratigraphic chart of the Solimões Basin (modified from Eiras et al., 1994) and detailed Devonian Sequence stratigraphy within the São
Mateus Field area (well NSM-1). Lithostratigraphic nomenclature informally adopted herein. LST, lowstand system tract; TST, transgressive system tract;
HST, highstand system tract. Two second-order sequence discontinuities (S1 and S5) bound the Devonian Sequence. S3 is a regionally correlatable surface,
corresponding to a transgressive erosion below the Uerê sandstones.

surface (MFS; Fig. 2), constituting the main source rock of regression known as the Hangenberg Event (Caplan &
the Basin (TOC av. 1.0%, up to 8.25%; Eiras, 1998; Grahn, Bustin, 1999; Streel et al., 2000). A regional erosive surface
1992; Mello et al., 1994; Silva, 1987). Towards the eastern (S5; Figs. 2 and 3(A)–(C)) separates the Carboniferous-
margin of the basin, the Jandiatuba shales are downlapped Devonian Sequence from thick (, 800 –1200 m) marginal
by a 50 m thick, progradational succession of Famennian marine evaporites, minor carbonate and siliciclastic deposits
sandstones rich in siliciceous sponge spicules interbedded of the Carboniferous-Permian Sequence (Eiras et al., 1994).
with shales and minor conglomerates of mud and chert
intraclasts reworked from the underlying Late Devonian
deposits, informally named Jandiatuba sandstones (Fig. 2). 3. Depositional facies of the Uerê sandstones
A marked erosive boundary (S4; Fig. 2) cuts the
Frasnian-Famennian shales and sandstones, and is covered The Uerê reservoir sandstones extend throughout the
by latest Famennian-Tournaisian glacial diamictites Devonian southern palaeomargin of the Solimões Basin
(Caplan & Bustin, 1999; Caputo & Crowell, 1985; Johnson, (Figs. 1 and 3(A) – (C)). The entire succession is 20 –60 m
Klapper, & Sandberg, 1985; Streel, Caputo, Loboziak, & thick, and is characterised by sharp-based, aggradationally to
Melo, 2000), corresponding to a worldwide glacio-eustatic progradationally stacked sandstone deposits. The presence
1050 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

of a basal erosional discontinuity (S3 surface) is suggested overlying Jandiatuba shales. Three facies associations
by (1) truncation of underlying Lower to Middle Devonian were identified in the Uerê sandstones, which are arranged
deposits and (2) lateral continuity of the individual facies in a coarsening-upward trend from the muddy facies
contained between the basal discontinuity surface and association 3 to sandy facies association 1. The following

Fig. 3. Well-log (gamma ray) cross-section of the Devonian Sequence through the São Mateus Field (A) and (B), and southern Solimões Basin (C). Correlation
surfaces are defined for the São Mateus Field area, with approximate correspondence to the regional stratigraphic boundaries in Fig. 2: SB1, to S1; TES, to S3;
SB2, to S4, and SB3, to S5. Datum is set at the MFS, which corresponds to the Jandiatuba shales. See Fig. 1 for location of wells and text for further
explanation.
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1051

Fig. 3 (continued )

interpretations are based on comparisons with facies and environment during storms with high-energy, long period
processes of modern shelves (Snedden & Nummedal, waves. Sponge spicule layers were derived from the
1991). disintegration of individual or small colonies of soft-bodied
Facies association 1 is composed of amalgamated, sharp- demosponges. Random spicule orientation and lack of
based sets of massive, parallel to low-angle cross-stratified, breakage suggest limited transport. Bioturbation on the top
fine to coarse sandstones (Fig. 4(A)). The individual sets are of storm sand layers and escape structures within laminated
0.05 –0.25 m thick, some with thin basal conglomeratic beds were produced by organisms that colonised the storm-
lags. Bioturbation is rare in this facies association, deposited sands during fair weather conditions (MacEa-
consisting mostly of isolated Skolithos and escape burrows. chern & Pemberton, 1992).
The setting of facies association 1 is interpreted as Facies association 3 is characterised by laminated black
amalgamated sands that were deposited in shallow (above shale interbedded with 0.1 –0.5 m thick, burrowed, massive-
fair-weather wave base) marine shelf environment during or parallel-laminated, fine to very-fine sandstones, and
intermittent storms. 0.05 – 0.15 m thick chert layers (Fig. 4(C)). The sandstone
Facies association 2 is composed of sharp-based, beds display sharp bases covered by imbricated mud
thinning-upwarding sandstones interbedded with chert and intraclasts (Fig. 4(B)). Facies associations 2 and 3 display
shale layers. Sandstones are fine- to medium-grained, and a low-diversity assemblage of trace fossils, dominated by
massive or parallel- to low-angle (swaley?) cross-stratified small (, 10 mm diameter) Planolites, Teichchinus, Sko-
(Fig. 4(E)). Coarse, sharp-based sandstone beds include lithos, and locally robust Thalassinoides and Cylindrichnus
imbricated mud intraclasts (Fig. 4(B)). The thickness of (Fig. 4(D)). Escape burrows are also commonly observed.
individual beds varies from 0.1 to 1.5 m. The sandstones are The high mud content and intense bioturbation of
locally bioturbated and soft-sediment-deformed (Fig. 4(D)). facies association 3 indicate deposition under deeper
Chert layers are 0.05 –0.2 m thick, and contain unbroken water, in which low-energy basinal conditions prevailed.
monoaxon and rare triaxon, randomly oriented sponge The deposition of the thin intercalated sandstone beds is
spicules that have been cemented and replaced by interpreted to have occurred below storm wave base, by
chalcedony and microquartz (Fig. 6(A)). Centimetre-thick episodic influx of shoreface sediments.
siderite-cemented intervals occur in sandstones that are The entire succession reflects the progradation of storm-
interbedded with shales. Facies association 2 is interpreted influenced shelf deposits, based on its coarsening-upward
to represent sands deposited in middle to lowermost shelf character, the presence of thin, non-amalgamated storm beds,
1052 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Fig. 4. Core photographs illustrating the main facies recognised in the Uerê sandstones. (A) Facies Association 1: tabular cross-stratified to massive sandstones.
Note mottled aspect due to carbonate cementation. (B) Facies Association 2: sharp-based, parallel- to low-angle inclined cross-stratified sandstones and sponge
spiculites (SS). Note mud intraclast lag. (C) Facies Association 3: sponge spiculites interbedded with shales. (D) Strongly bioturbated hybrid arenites from
Facies Association 2, with a Teichchinus (Te) trace within a Cruziana ichnofabric. (E) Stacking pattern of Facies Association 2, displaying an overall
coarsening-upward trend composed of sharp-based (arrows), thinning-upward, cm-thick cycles with low-angle cross-stratified, bioturbated to massive
sandstones and shales.

slightly reworked soft-bodied sponge remains and glaucony The general progradational pattern of the Uerê deposits
ooids and the widespread proximal Cruziana assemblage. was overlain by a MFS, represented by the transgressive
Jandiatuba shales (Silva, 1987). Two mechanisms could be
invoked to explain the deposition of these organic-rich
4. Stratigraphic framework shales: (1) encroachment of the epicontinental sea by an
oxygen-minimum zone during late Famennian flooding
The Devonian section was deposited in a very flat, (Savoy, 1992); (2) upwelling (Caplan & Bustin, 1996;
shallow and wide intracratonic depression, without a Savoy, 1992). High rates of primary productivity at the sea
distinct shelf/slope break. Therefore, the extremely low floor are evidenced by the abundance of sponges, a common
relief of the basin implies that a change of a few meters in constituent of siliceous deposits related to ancient upwelling
relative sea level would cause the flooding of large areas. cells (Caplan & Bustin, 1999; Parrish, 1982; Savoy, 1992).
There is a great lateral continuity (< 200 km) of the The possible stratigraphic settings for the progradation of
facies associations contained between the discontinuity storm-influenced shelf deposits underlain by erosional
surface S3 and the Jandiatuba shales (MFS; Fig. 3(A) –(C)). surfaces are those of forced-regressive (falling relative sea-
In their distal (NW) edge, there is a dominance of facies level stage) and lowstand complexes. Both are allocyclic,
associations 2 and 3. overlying genetically unrelated deposits and basally bounded
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1053

by sequence-stratigraphically important discontinuities Twenty-five thin-sections were carbon-coated and ana-


(Hunt & Tucker, 1992; Posamentier, 2002; Posamentier, lysed with a Cameca Camebax SX50 electron microprobe
Allen, James, & Tesson, 1992; Walker & Bergman, 1993; (EMP) equipped with four spectrometers and a back-
Walker & Wiseman, 1995). The evidence to reject a forced- scattered electron detector (BSE). The operating conditions
regressive scenario for the Uerê sandstones is that, in such were an acceleration voltage of 15 kV, beam diameter of
settings, the submarine component underlying the shallow 5 mm, beam current of 8 nA for the carbonates and, 12 nA
marine deposits extends only as far seaward as storm wave for the silicates. The standards used were anorthite (Ca, Al,
base reach, extending below that only as non-erosional Si), olivine (Mg, Fe, Mn), microcline (K), barite (Ba),
correlative conformities (MacEachern, Zaitlin, & Pember- strontianite (Sr) and jadeite (Na).
ton, 1999). Conversely, the base of Uerê sandstones is For the stable-isotope analyses of diagenetic carbonates,
markedly erosional throughout the basin. Additionally, it the bulk sample was powdered (, 200 mesh) and reacted
must be noted that in the distal areas (e.g. western Juruá Sub- with 100% phosphoric acid after 1 h at 25 8C for calcite and
Basin) Uerê sandstones basal boundaries (S3) remains at 50 8C for dolomite, and after 6 days at 50 8C for siderite.
erosional even where facies association 3 was deposited The evolved CO2 gas was analysed for carbon and oxygen
below fair-weather wave base. In the trangressively incised isotopes on a SIRA-12 mass spectrometer. The phosphoric
shoreface scenario the basal discontinuity is cut during acid fractionation factors used were 1.01025 for calcite at
relative sea-level stillstand prior to progradation, being after 25 8C (Friedman & O’Neil, 1977) and 1.009082 for siderite
modified by wave ravinement during transgression (Down- (Rosenbaum & Sheppard, 1986). The carbonate isotope data
ing & Walker, 1988). The resulting basal surface is erosive are presented in the normal d notation relative to PDB
and correlatable basinward, as observed in Uerê sandstones (Craig, 1957).
(MacEachern et al., 1999). Glaucony ooids that occur In this paper, the terms eo-, meso- and telodiagenesis are
dispersed in the Uerê sandstones (Fig. 6(B)) are also applied for the diagenetic stages sensu Schmidt and
consistent with a transgressive scenario. McDonald (1979).

5. Sampling and methods 6. Sandstone texture and composition

In this study we examined 50 m of cores taken from eight The Uerê sandstones range from very fine- to coarse-
oil wells. Thin sections prepared from 90 blue epoxy- grained, with predominance of fine to medium grain size. The
impregnated samples were examined with a petrographic sorting is usually moderate. There is a marked mixture of
microscope. The amounts of detrital and diagenetic com- well-rounded and angular grains in these sandstones,
ponents, and pore types, as well as the textural modal grain indicating a mixture of recycled and first-cycle sediments.
size and sorting parameters, were determined by counting 300 The sandstones are essentially subarkoses, and rarely
points in each of 65 selected representative thin sections. quartzarenites (Folk, 1968). The minor dissolution of
Sorting was estimated by comparison with the standard charts feldspars has not significantly modified the original average
of Beard and Weyl (1973). Packing proximity index was detrital composition from Q83F16.4L0.6 to present-day
determined by following Kahn (1956) procedures. Carbonate Q85.4F14L0.6 (Fig. 5). Original composition was reconstructed
cements were stained for identification with an acid solution
of alizarin red and potassium ferrocyanide (Dickson, 1965).
Microporosity was determined as the difference between
petrophysical porosity and porosity measured by petrographic
modal analysis. The percentage of types of detrital and
diagenetic constituents and of pores is expressed in relation to
bulk rock volume. Standard petrophysical nitrogen porosity
and air permeability values were acquired from 38 horizontal
plugs corresponding to part of the studied thin sections.
The morphology and the textural relationships
among minerals were examined in 11 gold-coated samples
with a JEOL JSM 5800 scanning electron microscope
(SEM) equipped with a Noran energy-dispersive spec-
trometer (EDS), using an accelerating voltage of 10 kV.
In order to identify the clay minerals present in the
sandstones X-ray diffraction analyses of the , 2 mm
fraction were performed in a Rigaku RU 200 diffractometer
in 16 oriented samples. The samples were air-dried, Fig. 5. Detrital composition of 65 representative Uerê sandstones (plotted in
ethylene glycol-saturated and heated at 490 8C for 4 h. the upper half of Folk (1968) diagram).
1054 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

from modal petrographic values of grain-replacive diagenetic Microquartz rims. Microquartz occurs as isopachous
constituents and grain-dissolution porosity. rims less than 20 mm thick, continuously covering the
Quartz grains are manly monocrystalline (av. 59%; max. detrital grains (Fig. 6(C) – (E)). In the sandstones were they
78%) with slightly undulose to abrupt extinction, whereas occur, such silica rims average 4% (up to 11%), but their
polycrystalline quartz occurs as trace to 4% (av. 0.6%). general average in the unit is 2% (Table 1). The rims show a
Detrital potassium feldspar (av. 5%; up to 10%) dominates layered structure of cryptocrystalline quartz (, 1 mm thick)
over plagioclase, which is totally albitised. Microcline and underlying a layer of randomly to parallel-oriented,
perthite are the most common types of potassium feldspar euhedral microquartz (1 –5 mm; Fig. 6(D) and (E)). Fibrous
(av. 3%, up to 7%, and av. 2%, up to 5%, respectively). The or ribbon-like illite and honeycombed mixed-layer illite-
main types of rock fragments are of fine micaceous smectite are intimately mixed with the microcrystalline
metamorphic, granitic-gneissic plutonic, strongly illitised quartz rims in nearly 35% of microquartz-cemented
felsic volcanic and mudrocks. sandstones (av. 1%; up to 5%; Fig. 6(E) and (F)).
Mud intraclasts occur in trace amounts (up to 3%) Intergranular pore-filling microquartz. Pore-filling
normally compacted to pseudomatrix (av. 0.7%; up to 4%) microquartz cement occurs as aggregates of euhedral,
and/or silicified (av. 1%; up to 11%). Trace amounts of randomly oriented microquartz crystals (av. 2%; up to
intraclasts of sponge spiculites occur in lag layers. 8%), similar in size and habit to grain-coating microquartz
Clay ooids occur dispersed in the sandstones (av. 0.4%; (Fig. 6(D)), being associated with small amounts of fibrous
up to 10%; Fig. 6(B)). The ooids show an approximately illite and ribbon-like illite-smectite.
glauconitic composition according to microprobe analyses, Isopachous chalcedony rims. Chalcedony rims occur in a
and typically contain nuclei of detrital heavy minerals, few sandstones containing silica/clay grain-coatings as 20–
quartz, feldspar, or expanded mica. 80 mm thick continuous rims, averaging 1% (up to 13%; av.
Bioturbation is widespread, and responsible for the mud 0.6%). Chalcedony surrounds mouldic pores after the
matrix with high organic matter content present in some of
dissolution of sponge spicules (20 – 80 mm long and 5–
the sandstones. Palinomacerals analysis reveals abundance
25 mm wide) (Fig. 6(A)). Locally, microquartz rim covers
of land plant fragments, as well as chitinozoans, scoleco-
isopachous chalcedony.
donts and acritarchs (Loboziak et al., 1994). This bioturba-
Intergranular pore-filling displacive chalcedony. This
tion matrix is commonly silicified, replaced by framboidal
cement occurs only in the hybrid arenites (Zuffa, 1980) with
pyrite or, rarely, by sphalerite. Heavy minerals are locally
sponge spicules or in pure sponge spiculites (av. 2%; up to
abundant (av. 1%; up to 6%), comprising mainly zircon,
34%), in which detrital grains comprise less than 10% of
tourmaline, sphene, and Fe– Ti oxides, the latter commonly
volume. Floating bioclasts (present-day mouldic pores) and
replaced by diagenetic TiO2.
siliciclastic grains and the high IGV values (< 40%)
The detrital composition of Uerê sandstones is consistent
with the intracratonic setting of the Solimões Basin, and with indicate the displacive character of this chalcedony cement.
the erosion of granitic and high-grade metamorphic Precam- Pore-filling chalcedony shows a spherulitic texture and
brian rocks continental blocks of Guyana and Brazilian shields contains abundant spherical micropores (10 – 20 mm), inter-
and recycling of pre-Middle Devonian sedimentary rocks. preted as mouldic pores after opal-CT lepispheres
(Fig. 7(A); Maliva & Siever, 1988). Spherulitic pore-filling
chalcedony covers (hence post-dates) isopachous chalced-
7. Diagenetic constituents ony rims that line detrital grains and mouldic pores after
sponge spicules.
The main diagenetic processes affecting the Uerê Quartz overgrowths. The volume of syntaxial quartz
sandstones are mechanical and chemical compaction, overgrowths (av. 4%; up to 10%) do not increase with depth,
authigenesis of various forms of silica, carbonates, and but rather vary widely from well to well. Quartz overgrowths
clay minerals, and dissolution and replacement of detrital are abundant both in shallow-buried, basin-margin sand-
grains by albite, carbonates, clay minerals, pyrite and TiO2. stones (av. 5%, up to 9% in RCA-1 at , 1940 m depth), and
in deeply buried, basin-centre sandstones (av. 5%, up to 10%
7.1. Silica in NSM-1 at , 3280 m depth; Fig. 7(B)). The occurrence of
poorly developed and discontinuous syntaxial quartz over-
The authigenesis of silica developed several habits in growths in early silica/clay-cemented sandstones (av. 2%; up
the Uerê sandstones, including (1) microquartz rims; to 8%; Fig. 7(C)) suggests that overgrowth distribution is
(2) intergranular pore-filling microquartz; (3) isopachous controlled by microquartz and authigenic clay. Overgrowths
chalcedony rims; (4) intergranular pore-filling displacive are mostly post-compactional, as evidenced by their absence
chalcedony; (5) syntaxial quartz overgrowths and (6) along intergranular contacts. In tightly compacted sand-
prismatic quartz outgrowths. The two latter habits are stones quartz overgrowths occlude most of the remaining
widespread, whereas the first four habits are restricted to intergranular porosity, and eventually engulf of fibrous illite
sandstones (originally) rich in sponge spicules. (Fig. 7(D)). The scarcity and tiny size of fluid inclusions in
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1055

Fig. 6. (A) Optical photomicrograph of a sponge spiculite with mouldic pores after sponge spicules, rimmed and partially filled by chalcedony; plane light.
(B) Backscattered electrons (BSE) image of a glaucony ooid. (C) Optical photomicrograph of a sandstone with thin microcrystalline quartz rims. (D) Secondary
scanning electrons microscopy (SEM) image of a sandstone cemented by microcrystalline quartz rims. (E) SEM image detail showing rims of oriented
microquartz crystals. (F) SEM image of a grain surface rimmed by microcrystalline quartz and clay and minor fibrous illite (IS); note cryptocrystalline quartz
(Qc), as well as tiny quartz crystals grown on top of the clay coating (arrow); Qo: quartz outgrowth.

the quartz overgrowths rendered the acquisition of homo- microquartz, commonly engulfing fibrous illite as scat-
genisation temperature impossible. tered crystals, extending into mouldic, intra- and inter-
Quartz outgrowths. Prismatic macroquartz outgrowths granular pores mostly in sandstones cemented by early
occur on quartz grains discontinuously coated by silica/clay (av. 1%; up to 6%; Fig. 7(C)).
1056 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Table 1
Statistical summary of the petrographic and petrophysical parameters of the Uerê sandstones. Parameters that are distinctive among petrofacies are italicised

Petrofacies; number of samples Petrofacies A; n ¼ 29 Petrofacies B; Petrofacies C; Whole unit; n ¼ 59


n ¼ 19 n ¼ 11

Total Average Max Average Max Average Max Average Max

Detrital quartz 50.8 63.0 69.4 79.3 64.1 75.0 59.4 79.3
Quartz monocrystalline 50.3 62.7 68.6 78.3 63.6 74.3 58.8 78.3
Quartz polycrystalline 0.5 1.7 0.8 4.3 0.5 1.3 0.6 4.3
Quartz in plutonic r.f. 0.0 0.7 0.0 0.0 0.0 0.0 0.0 0.7
Detrital feldspar 5.2 11.0 5.4 10.0 5.8 9.0 5.4 11.0
Detrital K-feldspar 5.2 10.3 5.3 10.0 5.7 9.0 5.3 10.3
Orthoclase 0.2 2.0 0.1 0.7 0.1 1.0 0.1 2.0
Microcline 3.1 5.0 3.0 6.0 3.5 6.7 3.2 6.7
Perthite 1.9 5.3 2.2 4.0 2.0 3.3 2.0 5.3
K-feldspar in plutonic r.f. 0.1 2.0 0.1 0.7 0.1 0.7 0.1 2.0
Detrital plagioclase 0.0 0.7 0.0 0.3 0.0 0.3 0.0 0.7
Plagioclase monocrystalline 0.0 0.3 0.0 0.3 0.0 0.3 0.0 0.3
Plagioclase in plutonic r.f. 0.0 0.7 0.0 0.0 0.0 0.0 0.0 0.7
Plutonic r.f. 0.2 2.7 0.1 0.7 0.1 0.7 0.1 2.7
Total fine-crystalline lithics 0.3 1.0 0.4 1.3 0.1 1.0 0.3 1.3
Volcanic r.f. 0.0 0.3 0.0 0.3 0.1 0.7 0.0 0.7
Chert r.f. 0.1 0.7 0.1 0.7 0.0 0.0 0.1 0.7
Mudrock fragment 0.2 1.0 0.2 1.0 0.0 0.3 0.2 1.0
Metamorphic rock fragment 0.0 0.3 0.1 0.7 0.0 0.0 0.0 0.7
Mica 0.5 3.7 0.1 0.7 0.1 1.0 0.3 3.7
Heavy minerals 1.1 3.0 0.8 4.3 1.1 6.0 1.0 6.0
Mud intraclast 0.2 2.7 0.0 0.0 0.0 0.3 0.1 2.7
Silica intraclast (spiculite) 0.1 2.0 0.0 0.3 0.0 0.0 0.0 2.0
Clay ooid 0.1 2.0 0.8 10.0 0.5 4.0 0.4 10.0
Silica bioclast 0.0 0.7 0.0 0.0 0.0 0.0 0.0 0.7
Mud pseudomatrix þ bioturbation matrix 0.8 4.0 0.6 4.3 0.4 3.0 0.7 4.3
Diagenetics total 26.1 49.1 16.4 21.7 20.3 28.0 21.8 49.1
Quartz overgrowth 2.3 7.7 3.6 6.7 7.6 10.3 3.7 10.3
Quartz outgrowth 1.3 6.3 0.3 2.7 0.1 0.7 0.7 6.3
Quartz filling moldic pore 0.2 3.3 0.0 0.0 0.0 0.0 0.1 3.3
Chalcedony displacive cement 1.8 33.8 0.0 0.0 0.0 0.0 0.9 33.8
Chalcedony rims 1.2 13.3 0.0 0.3 0.0 0.0 0.6 13.3
Chalcedony spherullite 0.1 1.3 0.1 1.0 0.0 0.0 0.1 1.3
Microquartz rims 4.1 10.7 0.0 0.0 0.0 0.0 2.0 10.7
Microquartz pore-filling 2.0 8.3 0.0 0.0 0.0 0.0 0.9 8.3
Clay coatings 0.9 4.7 0.0 0.0 0.0 0.0 0.4 4.7
Silicified secondary matrix 1.9 11.0 0.1 1.0 0.2 2.0 1.0 11.0
Anhydrite coarse intergranular 0.5 2.7 0.3 1.0 0.1 0.3 0.3 2.7
Anhydrite replacing feldspar grain 0.0 0.3 0.0 0.0 0.1 0.7 0.0 0.7
Anhydrite replac. quartz grain 0.0 0.3 0.0 0.0 0.0 0.0 0.0 0.3
Dolomite replacive intergranular 1.4 6.3 2.3 5.3 1.1 4.7 1.6 6.3
Dolomite replacing feldspar grain 0.2 1.0 0.4 2.0 0.3 1.3 0.3 2.0
Dolomite replacing quartz grain 0.3 1.3 0.3 1.3 0.1 0.3 0.3 1.3
Dolomite replacing secondary matrix 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Siderite intergranular 0.1 2.3 0.0 0.0 0.0 0.0 0.0 2.3
Siderite replacing grain 0.0 0.7 0.0 0.0 0.0 0.0 0.0 0.7
Illite intergranular fibrous 0.7 2.3 1.9 5.7 0.8 3.3 1.1 5.7
Illite coatings 0.4 2.3 1.1 4.3 0.3 2.3 0.6 4.3
Illite in feldspar grain 0.1 1.3 0.2 0.7 0.1 0.7 0.1 1.3
Illite after kaolinite 0.3 4.0 0.0 0.0 0.0 0.3 0.2 4.0
Illite in mica 0.0 0.3 0.0 0.0 0.0 0.0 0.0 0.3
TiO2 intergranular 0.5 1.7 0.6 2.7 1.1 3.7 0.7 3.7
TiO2 replacing grain 0.1 1.0 0.1 1.0 0.3 0.7 0.2 1.0
Pyrite framboidal intergranular 0.8 2.3 0.1 1.0 0.8 3.0 0.5 3.0
Pyrite coarse intergranular 0.1 0.7 0.1 0.7 0.1 0.7 0.1 0.7
Pyrite replacing grain 0.3 2.7 0.1 1.0 0.3 2.0 0.3 2.7
Pyrite framboidal in feldspar 0.2 0.7 0.1 0.7 0.2 1.0 0.1 1.0
Albite replacing plagioclase 2.4 5.0 1.8 3.3 2.7 5.0 2.2 5.0
Albite replacing K-feldspar 1.8 4.0 1.7 4.0 3.1 6.0 2.0 6.0
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1057

Table 1 (continued)
Petrofacies; number of samples Petrofacies A; n ¼ 29 Petrofacies B; Petrofacies C; Whole unit; n ¼ 59
n ¼ 19 n ¼ 11

Total Average Max Average Max Average Max Average Max

Albite overgrowth 0.1 1.0 0.4 1.3 0.7 1.7 0.3 1.7
Bitumen 0.0 0.3 0.4 1.3 0.3 1.0 0.2 1.3
Macroporosity 14.9 25.0 6.4 11.7 7.6 13.3 10.7 25.0
Intergranular 9.8 16.3 4.0 8.3 5.3 9.7 7.0 16.3
Intragranular in feldspar 2.1 4.3 1.7 3.0 1.4 3.3 1.9 4.3
Intragranular in quartz grain 0.8 3.0 0.2 1.0 0.2 1.3 0.5 3.0
Intragranular in mica 0.0 0.7 0.0 0.0 0.0 0.3 0.0 0.7
Intragranular in heavy mineral 0.1 0.7 0.1 0.7 0.0 0.3 0.1 0.7
Dissolution of pseudomatrix 0.5 5.0 0.0 0.7 0.2 2.0 0.3 5.0
Dissolution of cement 0.1 2.6 0.0 0.3 0.1 1.3 0.1 2.6
Mouldic 1.1 6.3 0.0 0.0 0.0 0.0 0.5 6.3
Fracture 0.1 1.0 0.2 1.3 0.0 0.3 0.1 1.3
Oversized 0.3 3.3 0.3 1.3 0.3 1.0 0.3 3.3
Petrophysical porosity 15.3 28.0 9.2 13.8 12.3 17.8 12.8 28.0
Petrophysical permeability 67.6 1153.6 4.7 29.1 7.2 20.9 35.6 1153.6
Microporosity 15.3 25.0 6.4 11.7 7.2 13.3 10.8 25.0
Secondary porosity 5.1 10.1 2.4 5.0 2.3 8.0 3.7 10.1
Intergranular volume 31.6 45.1 16.0 21.0 19.2 26.0 24.1 45.1
Grain volume 68.4 78.7 84.0 90.7 80.8 84.3 75.9 90.7
Cement total 18.5 41.9 11.6 17.7 13.0 19.3 15.2 41.9
Carbonate total 9.5 77.0 2.9 6.0 1.5 6.0 5.9 77.0
Silica total 11.1 24.0 4.1 7.3 7.6 11.0 8.1 24.0
Grain replacement total 7.7 19.0 8.6 14.3 8.8 14.0 8.2 19.0
Modal grain size (mm) 0.2 0.4 0.3 0.5 0.2 0.4 0.2 0.5
Sorting 0.9 2.0 0.8 2.0 0.7 1.0 0.8 2.0
Packing index (Pp) (G/G) 15.1 26.0 40.4 80.0 37.4 54.0 27.6 80.0
Original porosity (Beard & Weyl, 1973) 33.3 40.2 33.9 39.7 34.4 38.8 33.7 40.2
Compactacional porosity loss (%) COPL 18.5 26.0 28.3 32.4 27.2 32.6 23.4 32.6
Cementational porosity loss (%) CEPL 8.8 25.7 4.0 6.2 4.8 6.8 6.5 25.7

7.2. Clay minerals rims (Fig. 7(D)). Such eogenetic smectite is abundant around
mouldic pores after sponge spicules and feldspar grains.
Authigenic clay minerals (trace to 9%) include mainly Illite-smectite and illite also formed by the transformation of
illite and mixed-layer illite-smectite that occur as pore- detrital smectite from bioturbation matrix or pseudomatrix
lining rims and pore-filling cement, and as replacement formed by the compaction of mud intraclasts.
of mud matrix and detrital feldspar. Illite was formed Chlorite occurs as interlocked pseudohexagonal crystals
both by direct neoformation in the pores and by in the shallow-buried quartz-cemented sandstones
replacement of kaolinite and smectite precursors. Total (, 1940 m depth) in well RCA-1, and as thin and
Illitic clay contents average 2%, of which 0.6% (up to discontinuous rims (< 5 mm) covered by quartz over-
4%) occurs as discontinuous rims on detrital grains growths (Fig. 8(A)). Bulk XRD data of these sandstones
(Fig. 7(E)), 0.2% (up to 1%) occurs as replacement of revealed that clay minerals average 5% (trace to 8%), in
detrital feldspars, and 1% (up to 6%) occurs as which chlorite comprises 85% and I/S 15%.
intergranular neoformed cement. Unsilicified pseudoma-
trix was replaced by illite. Illite replacement of kaolinite 7.3. Carbonates
(av. 0.2%; up to 4%) is pseudomorphic, preserving the
vermicular and booklet habits of the kaolinite aggregates. Ferroan dolomite and ankerite dominate over siderite and
Such illitised kaolinite occurs adjacent to secondary calcite (Fig. 9). Fe-dolomite/ankerite is widespread as
pores after feldspar dissolution (Fig. 7(F)). Illite replaces scattered spots, which give a mottled appearance to the
the feldspar grains along cleavage boundaries and sandstones (Fig. 4(A)). Locally, it occurs as partial to
twinning planes. pervasive strata-bound (0.01 – 0.1 m thick) cement in the
As determined by XRD analysis, mixed-layer illite- proximity of organic matter-rich shale layers. Fe-dolomite/
smectite clays are well-ordered, with 75– 85% illite, and ankerite shows a blocky to poikilotopic habit (0.1 – 10 mm)
were formed by the transformation of pre-compactional habit, occurring as intergranular, locally replacive and
smectite coatings intimately intergrown with the microquartz displacive cement in sandstones with high IGV values
1058 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Fig. 7. (A) SEM image of microquartz-rimmed spherical pores interpreted as opal-CT lepisphere moulds. (B) Optical photomicrograph of a sandstone
massively cemented by quartz overgrowths. (C) SEM image of a sandstone with grains partially rimmed by microcrystalline quartz, with discontinuous
overgrowths and outgrowths. (D) SEM image of a sandstone cemented by quartz overgrowths engulfing illite-smectite clays. (E) BSE image showing
well-developed fibrous illite rims. (F) BSE image showing illitised kaolinite engulfing Fe-dolomite (white) in a microquartz-rimmed sandstone.

(32 –37%; Fig. 8(B)). Oversized patches of dolomite are the Ankerite occurs locally in tightly compacted to
product of complete replacement of spicules, identified in recrystallised sandstones and metasandstones from
BSE images as darker, slightly magnesium-enriched ‘ghosts’ SMT-2 and SMT-3 wells area. Ankerite is anhedral to
(Fig. 8(C)). Commonly, Fe-dolomite/ankerite poikilotopic subhedral in habit with curved crystal faces and undulose
patches shows oxidised borders. Fe-dolomite/ankerite is extinction (saddle ankerite) occurring as veins cross-
absent within mouldic pores after feldspar dissolution. cutting the metasandstones. Vein-filling saddle ankerite
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1059

Fig. 8. (A) SEM image showing chlorite rims partially engulfed by quartz overgrowths. (B) BSE image showing poikilotopic Fe-dolomite cement replacing
microquartz rims. (C) BSE image of a sandstone cemented by Fe-dolomite with ‘ghosts’ of replaced sponge spicules. (D) Optical photomicrograph showing a
metasandstone with mica and ankerite cementation along fractures. (E) BSE image showing Fe-dolomite (medium gray) engulfing siderite (white). (F) Optical
photomicrograph of a strongly pressure-dissolved sandstone with thin illite coatings (petrofacies B).

(1.0 – 2.5 mm thick) is often associated with stylolites along framboidal pyrite and engulfed by Fe-dolomite/ankerite
mica/clay-rich laminae, replacing considerable portion of (Fig. 8(E)). Locally, siderite forms spherulitic aggregates
these rocks as irregular patches (Fig. 8(D)). Saddle ankerite is (20 – 60 mm), and replaces quartz, feldspar and mud
non-stoichiometric [Ca1.15(Fe0.42, Mg0.37, Mn0.12)(CO3)2]. intraclasts. Many Uerê siderites are extremely Mg-rich
Siderite cement occurs in loose-packed sandstones (MgCO3 av. 18 mol%; up to 32 mol%; sideroplesites,
(packing proximity index , 20%; IGV , 32%) as scat- Deer, Howie, & Zussman, 1962). Siderite crystals are
tered rhombs (5.0 –60 mm), commonly associated with irregularly zoned in terms of Fe and Mg contents.
1060 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Authigenic titanium oxides occur mostly as intergranular


aggregates of scattered euhedral crystals of rutile
(2 –10 mm), replacing grains and filling secondary pores
after detrital Ti – Fe heavy minerals.
Anhydrite occurs as minor (av. 0.4%; up to 3%), coarsely
crystalline (10 –25 mm), post-compactional intergranular
cement that partially replaces feldspar and quartz grains and
overgrowths.
Bitumen occurs either infilling pores or as dark brown
coatings on the surface of quartz overgrowths and smectite-
illite rims (av. 0.2%; up to 1%).

8. Compaction and porosity

Conspicuous evidence of mechanical compaction


Fig. 9. Microprobe composition of representative carbonate cements in includes the fracturing of quartz and feldspar grains
Uerê sandstones. (commonly healed by authigenic quartz and albite,
respectively) and by the deformation of ductile grains
7.4. Albite (mica, mud intraclasts, glauconitic ooids). Chemical
compaction occurred by pressure dissolution both along
Diagenetic albite with a typical pure end-member intergranular contacts and stylolites (Fig. 8(F)). Compaction
composition is common (av. 2%; up to 6%), occurring as is limited in sandstones cemented by abundant eogenetic
replacement of, and as overgrowths on feldspar grains. Both silica or carbonate cementation. Locally, in sandstones
detrital K-feldspar and plagioclase were albitised, with cemented by microquartz rims, pressure dissolution is
replacement occurring after partial grain dissolution, as limited to intergranular sutured and concave-convex con-
evidenced by the variable amounts of dissolution voids
tacts with rare, low-amplitude (, 0.1 mm), pyrite-lined
within these grains. Albitised K-feldspar and plagioclase
stylolites along laminations enriched in detrital heavy
were distinguished according to the petrographic and
minerals and micas. Sandstones devoid of early microquartz
chemical criteria described by Morad, Bergan, Knarud and
cement display pervasive chemical compaction, represented
Nystuen (1990). Albitised K-feldspars are untwinned, with
by abundant stylolites (0.25 –2 mm amplitude), in places
irregular extinction and turbid appearance due to
amalgamated in dissolution seams a few mm thick.
the presence of numerous, small fluid inclusions and
Thin section porosity (av. 11%; up to 25%) varies widely
microporosity. The albitised feldspars are composed of a
with cement type. In sandstones cemented by eogenetic
large number of small (10 –50 mm), lath-like albite crystals.
The albitised plagioclase grains show irregular polysyn- microquartz, macroporosity averages 15%, whereas in
thetic twins, due to their replacement by numerous sandstones devoid of such cement and with scarce (, 5%)
elongated albite crystals (0.02 – 0.1 mm) arranged parallel or abundant (5 – 10%) quartz overgrowth, macroporosity
to the twinning and cleavage planes of the host grains. averages 6 and 8%, respectively (Table 1). About two-thirds
In addition to the replacive habit, albite also occurs as of the macroporosity is primary intergranular, with a
overgrowths on plagioclase grains (av. 0.3%; up to 1.7%). In maximum value of 16% (av. 7%). Secondary porosity
some cases, the overgrowths surround intragranular and comprises dominantly mouldic pores generated by the
mouldic pores that resulted from the post-overgrowth dissolution of detrital feldspar (av. 2%; up to 4%), and
dissolution of detrital plagioclase cores. Albite also occur as sponge spicules (av. 0.6%; up to 6%). The latter type is
small (2–5 mm) discrete crystals associated with illitic clays. volumetrically important in hybrid arenites in the form of
elongate (20 – 80 mm long and 5 – 25 mm wide) pores
7.5. Other diagenetic constituents rimmed by microquartz, chalcedony or smectitic clay.
Other secondary pore types include those generated by
Pyrite is widespread and occurs as intergranular the dissolution of pseudomatrix (av. 0.3%; up to 5%), silica
framboids (, 8 mm; av. 0.7%; up to 3%) and, rarely, as cement (chalcedony, intergranular microquartz aggregates;
coarse euhedral crystals (8 –20 mm; av. 0.3%; up to 0.7%). av. 0.1%; up to 3%), and oversized pores (av. 0.3%; up to
Grain-replacive pyrite is also common, occurring along 3%) formed by the dissolution of detrital quartz and feldspar
feldspar fractures and cleavages (av. 0.3%; up to 1%), and grains. Microporosity is widespread in the sandstones,
within mud intraclasts, heavy minerals, and glauconitic averaging 3% (largest values, up to 11%, in microquartz/
ooids and concentrated along stylolites (av. 0.3%; up to 3%). chalcedony-cemented sandstones).
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1061

9. Discussion

9.1. Reservoir petrofacies

Three petrofacies were recognised in the sandstones,


based on their packing, porosity, and cementation (Table 1):
(1) petrofacies A: porous (. 15%) sandstones with micro-
quartz grain-rim; (2) petrofacies B: tight (, 10%), strongly
compacted to conspicuously quartz-cemented sandstones;
(3) petrofacies C: moderately porous (10 –15%), partially
quartz-cemented sandstones. These petrofacies are thus
defined based on petrographic and petrophysical character-
istics of the sandstones, which are in turn strongly controlled
by diagenesis, without stratigraphic or environmental
connotations. The three petrofacies are easily recognised
on a plot of intergranular volume (IGV) versus volume of Fig. 11. Plot of petrophysical helium porosity versus horizontal air
permeability for the three reservoir petrofacies. The highest porosity and
silica cement (including microquartz grain-coating and permeability values correspond to sandstones cemented by microcrystalline
syntaxial quartz overgrowths; Fig. 10). Petrofacies A is thus silica from petrofacies A and the lowest values, to tightly compacted
characterised by IGV . 20% and various amounts of petrofacies B sandstones. Quartz cemented petrofacies B sandstones show
microcrystalline silica cement. Petrofacies B an C are intermediate porosity and permeability values.
characterised by IGV , 20%, and are distinguished based
on an amount of quartz overgrowth cement of , 5% for 9.2. Porosity reduction processes
petrofacies B and between 5 and 10% for petrofacies C.
The petrophysical characteristics of each petrofacies are A plot of cement volume versus intergranular volume
also consistent with the above-mentioned definition
(Houseknecht, 1987; Fig. 13(A)) reveals that the reduction
(Fig. 11). Strongly compacted and quartz-cemented
of porosity occurred through an interplay of compaction and
petrofacies B and C show low porosity values (, 15%)
cementation. In petrofacies A sandstones with abundant,
compared to porous petrofacies A sandstones (15 – 28%),
pre-compactional eogenetic silica cement, porosity
which also display better permeabilities.
reduction was dominantly due to cementation. Conversely,
The diagenetic evolution pathways of the three defined
porosity reduction was dominantly due to chemical
petrofacies are schematically represented in Fig. 12, which
compaction in petrofacies B (Fig. 8(F)), as well as in
depicts the major diagenetic conditions encountered by the
petrofacies C sandstones with post-compactional quartz
Uerê sandstones, discussed below, and the resulting
overgrowths (Fig. 7(B)).
reservoir quality products.
However, by plotting calculated indices that take into
consideration the reduction in bulk rock volume due to
chemical compaction (Lundegard, 1992), a more realistic
evaluation of the relative roles of compaction and
cementation in porosity reduction is obtained (Fig. 13(B)).
A plot of corrected compactional versus ‘cementational’
porosity reduction indices reveals that compaction was
overall more important than cementation in reducing
porosity in the whole unit, including in petrofacies A
sandstones, with the exception of a few samples
(Fig. 13(B)). The ‘original depositional porosity’ values
were obtained from Beard and Weyl (1973) parameters.
The assumed average original porosity of petrofacies A is
33.5% (i.e. moderately to well-sorted sandstones with
0.2 mm average modal grain size). These sandstones show
average intergranular volume of 27.3% and average
cement volume of 18%. According to the calculated
compactional and cementational porosity loss indices,
these silica-cemented sandstones have lost on average
Fig. 10. Plot of the silica cement volume (chalcedony, microcrystalline and
19.4% porosity (around 57% of the original porosity) due
macrocrystalline quartz) versus intergranular volume, showing a clear to compaction and 8.5% porosity (around 25% of the
distinction between the three reservoir petrofacies. original porosity) due to cementation.
1062 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Fig. 12. Diagenetic evolution pathways of Uerê sandstones for the defined reservoir petrofacies.

9.3. Marine eodiagenesis quartz (Williams & Crerar, 1985; Williams, Parks, &
Crerar, 1985). The process starts with biogenic silica
Marine eodiagenesis was strongly influenced by the dissolution (point 1 in Fig. 14(A)) and supersaturation of
presence of siliceous sponge spicules. Siliceous sponge pore fluids with respect to opal-CT and quartz. Relative
spicules were originally abundant in the Late Devonian rates of opal-A and opal-CT nucleation govern the extent
sediments of the Solimões Basin, but have been to which silica activity is buffered near point 2 in
extensively dissolved, recrystallised or replaced by Fig. 14(A), and hence the net increase in surface area.
microcrystalline quartz during eodiagenesis and shallow From the point when further changes in specific surface
mesodiagenesis. Mouldic pores after dissolved spicules area have negligible effect on opal-A solubility, opal-CT
are common within microquartz- and chalcedony-cemen- of the surface area # that of point 3 in Fig. 14(A) can
ted sandstones of petrofacies A, attesting to a major form. With progressively less opal-A remaining, silica
redistribution of silica from biogenic sources to eogenetic activity starts to decline and opal-CT nucleation becomes
cements (Fig. 14(A)). Thermodynamic and kinetic subordinate to growth. Opal-CT crystals with lower
models indicate that progressive silica transformations surface area form at the expense of smaller and less
in marine deposits occur from the phase of highest ordered ones. Such competitive crystal growth process is
entropy (amorphous opal) to the phase of lowest entropy known as Ostwald ripening.
(quartz) in a dissolution-reprecipitation pathway. The A specific surface area of 150 m2/g was assumed for
generalised diagenetic sequence is: opal-A (biogenic the sponge spicules in the Uerê sandstones, what is
silica) ! disordered opal-CT ! ordered opal-CT ! significantly lower than the specific surface area of
cryptocrystalline quartz or chalcedony ! microcrystalline microporous radiolaria and diatoms tests (, 250 m2/g)
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1063

Fig. 14. Diagrammatic interpretation of the genetic relationships between


diagenetic silica polymorphs with the evolution of pore-fluid silica activity
and specific surface area (Williams et al., 1985): (A) commonly observed
silica polymorphs evolution, and (B) evolution of silica polymorphs in clay-
bearing sandstones.

respect to quartz nucleation and growth buffer the silica


activity until it falls back towards the quartz solubility
line with decreasing pore-water silica saturation. As the
system evolves down this line, cryptocrystalline quartz
is gradually replaced by epitaxial microcrystalline
quartz (point 5 in Fig. 14(A)). The dominance of
Fig. 13. (A) Plot of cement volume versus intergranular volume (House-
knecht, 1987) and (B) plot of compactional porosity loss (COPL) versus cryptocrystalline and microcrystalline quartz suggests
cementational porosity loss (CEPL) (Lundegard, 1992) for 65 representa- that the dissolution rate of precursor phases (i.e.
tive Uerê sandstones. See text for comments. biogenic silica, opal-A and/or opal-CT) was fast enough
to sustain silica saturation at the microquartz saturation
assumed by Williams et al. (1985), and higher than that level, resulting in numerous crystals instead of larger
of opal-CT lepispheres. Evidence of intermediate opal- ones (Jahren & Ramm, 2000). The common occurrence
CT precipitation in the Uerê sandstones includes the of 1 – 5 mm crystals could be explained as the relatively
presence of spheroidal pores within microquartz and low microquartz solubility have kept the driving force
chalcedony cements that are probably moulds of for crystal growth by Ostwald ripening at minimal
dissolved opal-CT lepispheres (Fig. 7(A); Astin, 1987; levels. This eogenetic pathway showed by sponge
Hendry & Trewin, 1995). The occurrence of micro- spicules-rich sandstones is similar to the diagenetic
quartz or chalcedony as rims around such opal-CT transformation of deep ocean siliceous oozes (Aase
moulds suggests that locally petrofacies A sandstones et al., 1996; Knauth, 1994).
followed the pathway 1 ! 2 ! 3 during opal-A/opal-CT However, this idealised sequence of diagenetic trans-
transition (Fig. 14(A)). Precipitation of cryptocrystalline formations does not correspond precisely to what is observed
quartz began once the thermodynamic drive for Ostwald in petrofacies A sandstones, where different silica poly-
ripening of the opal-CT became insignificant (point 4 in morphs commonly coexisted. Opal-A and chalcedony have
Fig. 14(A)). Relative rates of opal-CT dissolution with coexisted, allowing the preservation of sponge spicules
1064 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

moulds, as well as opal-CT and microquartz, allowing the large amounts of bioturbation matrix inhibited silica
preservation of moulds of lepispheres. Such coexistence may cementation.
have been allowed by the activity of Mg2þ in solution, as Marine eodiagenesis included the precipitation of
well as by the presence of Al, Mg, Fe and Mn hydroxides and coarsely crystalline Fe-dolomite/ankerite and siderite.
alkali cations adsorbed onto grain surfaces. These factors are High content of Mg showing in most siderites are
considered to significantly affect silica reactivity and the rate consistent with precipitation from marine pore-water
of opal-A ! opal-CT reaction (Dove & Crerar, 1990; (Mozley, 1989). Locally, spherulitic siderite aggregates
Kastner, Keene, & Gieskes, 1977 and references therein). suggest that precipitation was triggered by bacterial
The original amount and shape of the sponge spicules is activity (Huggett, Dennis, & Gale, 2000). Fe-dolomite is
preserved within areas cemented by eogenetic dolomite or a common eodiagenetic carbonate in organic-rich marine
Mg-rich siderite (Fig. 8(C)). The consumption of Mg2þ and sediments if sulfate is rapidly depleted prior to major
increase in alkalinity during carbonate precipitation would compaction (Curtis & Coleman, 1986; Morad, 1998).
thus facilitate opal-CT and microquartz nucleation.
Such differences in the evolution pathway of eogenetic 9.4. Meteoric eodiagenesis
silica precipitation in petrofacies A sandstones may be
related to the common occurrence of eogenetic clay The observed feldspar dissolution and precipitation of
coatings. The presence of clay is known to speed up the authigenic kaolinite (Fig. 7(F)) was probably caused by
replacement of opal-CT by quartz (Chang & Yortsos, 1994; meteoric flushing related to the late Pennsylvanian-Permian
Siever & Woodford, 1973; Williams et al., 1985), because uplift (Fig. 2). Although for the good lateral connectivity of
clay adsorbs silica, and thus inhibits opal-CT nucleation. the sandstones, the extent of kaolinite precipitation varies
The observed clay coatings are composed of illite-smectite, widely, apparently due to the patterns of meteoric flow from
probably evolved from eogenetic smectite. Eogenetic basin margin recharge areas. Kaolinite rarely replaces
dissolution of biogenic silica promotes the precipitation directly the feldspar grains, but instead fills intergranular
of K-rich smectite in estuarine and marine environments and feldspar dissolution pores. This suggests a certain
(Michalopoulos & Aller, 1995; Michalopoulos, Aller, & degree of aluminium mobilisation with bulk porosity
Reeder, 2000; Wollast & de Broeu, 1971). Clay mineral enhancement in the sandstones. Widespread oxidation of
precipitation and concomitant dissolution of biogenic silica the earlier Fe-dolomite/ankerite poikilotopic cement, even
exert control on dissolved Si concentration, and thus on the in tight compacted sandstones of petrofacies B also suggests
diagenetic evolution of silica polymorphs in marine an environment of oxidizing fluids previous to effective
sediments. Additionally, such clays constitute a sink of Si burial, compaction and quartz cementation in the
and K (and perhaps Mg and Fe) in the pore-water fluids, sandstones.
and so may further influence early silica diagenesis
(Michalopoulos et al., 2000). 9.5. Mesodiagenesis
The presence of eogenetic clay coatings in sandstones of
petrofacies A may have induced the evolution pathway The precipitation of quartz overgrowths and outgrowths
1 ! 4 in Fig. 14(B). Detrital clays that adsorbed silica at occurred after effective burial and compaction. This is
higher pore-water silica concentrations at points 1 –4 in indicated by (1) the intergrowth of quartz with fibrous illite
Fig. 14(B) were at points 4 and 5 out of equilibrium with the (Fig. 7(D)); (2) the relatively small intergranular volume
new pore-water concentrations and began to desorb silica and quartz overgrowths volume of (petrofacies B: quartz
(Chaika & Williams, 2000). This process buffered pore overgrowth av. 3.6% and IGV av. 16.7%; petrofacies C:
water at low silica concentrations (, 20 mg/kg), which quartz overgrowths av. 7.6% and IGV av. 19.7%); (3) the
favoured microquartz precipitation at shallow burial. There- partial replacement of post-compactional, poikilotopic
fore, the precipitation of eogenetic clays may have favoured anhydrite by quartz.
the precipitation of quartz relative to opal-CT both by From the silica sources commonly considered for
decreasing concentration of dissolved silica, thus quartz cementation (McBride, 1989), five could be
competing by the silica with opal-CT, and by inhibiting invoked as potential sources for the precipitation of
opal-A ! opal-CT transformation and early silica nuclea- quartz overgrowths in Uerê sandstones: (1) the pressure
tion along coated grain surfaces. dissolution of detrital quartz along intergranular contacts
Variations in the amounts of early silica cements in and stylolites; (2) the dissolution of opaline skeletons and
petrofacies A depend thus largely on the relative amount of spicules of sponges; (3) the dissolution or alteration of
detrital clay, biogenic silica, and above all, authigenic clay detrital silicates, mainly of feldspars; (4) the illitisation of
coatings. In the presence of abundant detrital clays from detrital and authigenic smectite or kaolinite; (5) the
bioturbation or mud intraclasts, or mostly eogenetic convection of hot fluids related to magmatism.
smectite coatings, opal-A to opal-CT transformation was Pressure dissolution was an important source of silica
inhibited and microquartz formed directly from biogenic for quartz overgrowth cementation in petrofacies B
silica dissolution. However, thicker smectite coatings or sandstones, as indicated by the common occurrence of
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1065

sutured intergranular contacts and discrete stylolites. fluid circulation, to which the role of intrabasinal
Such features are however, rare to absent in pervasively magmatism may have been essential. There is a similarity
quartz-cemented petrofacies C sandstones, indicating that between the K/Ar and Ar/Ar ages of authigenic illite in
internal pressure dissolution was not the only source for the Carboniferous sandstones (180 – 220 Ma; Mizusaki
mesogenetic quartz cementation. et al., 1990) and those of the Triassic diabase intrusions
The dissolution of sponge spicules was the main source (Szatmari, 1996). This suggests that the widespread
for the microcrystalline eogenetic silica cements, but was an precipitation of illite and co-precipitated quartz over-
unlikely source for mesogenetic quartz overgrowths, since growths, both in the Carboniferous and Devonian
spicule remnants were preserved from eogenetic dissolution sandstones, may have occurred in connection with the
only in areas pervasively cemented by eogenetic carbonates circulation of fluids related to the rapid subsidence and
or silica. Likewise, the dissolution of detrital feldspars was high heat flow promoted by the magmatism, Such
not an important silica source for mesogenetic quartz, enhanced thermal regime would induce active fluid
considering the limited volume of feldspars affected, and convection through faults and permeable beds, such as
the fact that the main episode of feldspar dissolution took what is interpreted for quartz cementation in other basins
place during meteoric eodiagenesis, and not during (De Ros, Morad, Broman, Césero, & Gomez-Gras, 2000;
mesodiagenesis. Girard, Deynoux, & Nahon, 1989; Gluyas, Grant, &
The illitisation of eogenetic smectite and kaolinite may Robinson, 1993; Summer & Verosub, 1992).
have partially supplied silica to mesogenetic quartz The effects of magmatism on the alteration of the
precipitation. Assuming K-feldspar as the most likely Devonian sandstones may have been locally intensified. A
source of potassium, the transformation of smectite in illite seismic anomaly within the São Mateus field has long been
can be described as follows: described as a ‘basement high’. Cores taken at , 3200 m in
the SMT-2 and SMT-3 wells have recovered metasand-
KAlSi3 O8 þ 2K0:3 Al1:9 Si4 O10 ðOHÞ2 stones with granoblastic texture, extensively recrystallised
K-feldspar smectite
quartz phengitic mica and breccias cemented by replacive
! 2K0:8 Al1:9 ðAl0:5 Si3:5 ÞO10 ðOHÞ2 þ 4SiO2ðaqÞ ð1Þ saddle ankerite (Fig. 8(D)). However, the similarity of the
illite quartz gamma-ray logs of these rocks with those of the Uerê
sandstones (Fig. 3(A)), and petrographic analyses suggest
According to the quantified amounts of illitic clays, the
that these intervals correspond to hydrothermally altered
transformation of an average of 1.8 vol% of smectite would
Uerê sandstones. Such focused circulation of hydrothermal
produce , 0.6 vol% of quartz.
fluids, presumably connected with the Triassic magmatism,
Furthermore, illitisation of kaolinite may also supply
promoted extensive quartz recrystallisation, as well as mica
silica for mesogenetic quartz precipitation:
and ankerite formation, similarly to what is observed in
Al2 Si2 O5 ðOHÞ4 þ KAlSi3 O8 ! Kal3 Si3 O10 ðOHÞ2 several present and ancient geothermal systems influenced
kaolinite K-feldspar illite by advective (hydrothermal) fluid flow related to magmatic
activity (McDowell & Paces, 1985; Pitman, Henry, &
þ 2SiO2ðaqÞ þ H2 O ð2Þ Seyler, 1998; Schiffman, Bird, & Elders, 1985; Searl, 1994
quartz
and references therein).
However, this reaction is of limited importance since The albitisation of feldspars and precipitation of albite
illitised kaolinite occurs significantly only in eogenetic overgrowths, anhydrite and minor ankerite are interpreted to
silica-cemented petrofacies A. According to Eq. (2), the be deep mesogenetic phases that post-date the Triassic
alteration of 0.2 vol% kaolinite would produce , 0.1 vol% magmatism. The origin and chemical composition of fluids
quartz. Thus, the intergrowth of illite and quartz cements from which anhydrite and albite precipitated cannot be
may presumably reflect: (1) release of silica due to the unravelled from the available data. However, the mesoge-
reaction of kaolinite and smectite to form illite, and (2) co- netic anhydrite in the Uerê sandstones was probably derived
precipitation from high-temperature ($ 100 –140 8C), Si – from the dissolution of anhydrite in the Carboniferous and
K – Al-charged fluids (Ehrenberg & Nadeau, 1989). Permian evaporitic sequences (, 850 – 1100 m thick; Eiras
Therefore, internal silica sources seem to be insufficient et al., 1994), which are in contact with the Devonian through
to explain the observed volumes of mesogenetic quartz large reversed faults. During Jurassic to Early Cretaceous
cementation. Quartz precipitation must have thus involved Solimões Basin, inversion and extensive uplift promoted
the subsurface circulation of Si-charged fluids. The meteoric flushing of the overlying evaporitic sequences, as
relatively small mud/sand ratio of the Devonian sequence is characteristic of many other occurrences of late-stage
(Figs. 2 and 3) suggests that the illitisation of smectites in sulphate cements in deeply buried sandstones (Dworkin &
the Devonian shales (Boles & Franks, 1979) was of Land, 1994; Gluyas, Jolley, & Primmer, 1997). Evaporite-
limited importance as source for mesogenetic quartz sourced brines may also account for the observed feldspar
cements. The precipitation of the observed quartz volumes albitisation and precipitation of albite overgrowths.
would therefore, involve enhanced, probably convective, Additionally to external Na supply, albitisation of feldspar
1066 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

(mostly of K-feldspar) occurs when the aNaþ/aHþ ratio 9.6. Paragenetic sequence
increase relative to aKþ/aHþ ratio in pore fluids, due to
transformation of smectite to illite and to authigenesis of The sequence of diagenetic processes in the sandstones is
illite (Morad, 1986). Both vitrinite reflectance (av. , 1% presented in Fig. 15. Due to the complex paragenetic
Ro) and illite transformation reaction suggest that Uerê relationships among the authigenic constituents and to
sandstones were subject to maximum temperatures about the limitation of geothermometric and geochronologic data,
120 8C, which are within the temperature range for albite only a schematic representation of the diagenetic evolution
authigenesis (Morad et al., 1990). could be achieved. The simplified paragenetic diagram

Fig. 15. Diagram of the paragenetic sequence and burial history of the Uerê sandstones in the São Mateus oil field. See text for explanation.
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1067

depicts the general evolution of Uerê sandstones, and should 9.7. Porosity preservation
be examined together with the diagenetic evolution path-
ways of each petrofacies illustrated in Fig. 12. The Two major mechanisms of porosity reduction have
paragenetic evolution is framed within a burial history of been recognised within the Uerê sandstone: (1) mechan-
São Mateus Field reservoirs, representative for Uerê ical and chemical compaction (2) mesogenetic quartz
Formation along basin centre. cementation. The rate of porosity reduction due to quartz
Four major stages of diagenetic evolution were recog- cementation is a function of temperature and availability
nised, including (Fig. 15): (1) marine eodiagenesis, of clean quartz grains surfaces for the growth of quartz
dominated by carbonate cementation, dissolution of sponge overgrowth cement (Walderhaug, 2000). Two mechan-
spicule and silica pore-lining cementation, the latter two isms may have prevented a more extensive compaction
processes being restricted to petrofacies A; (2) meteoric and quartz cementation in the studied reservoirs: (1)
eodiagenesis, which induced heterogeneous feldspar dissol- early hydrocarbon emplacement (2) eogenetic micro-
ution and kaolinite precipitation; (3) shallow mesodiagen- quartz grain-rimming.
esis (1900 – 2600 m depth), marked by heterogeneous It is generally accepted that the displacement of aqueous
chemical compaction and quartz cementation; and (4) pore fluids by hydrocarbons can inhibit the progress of
deep mesodiagenesis (. 2600 m depth), marked by illite diagenetic reactions. The overall lack of mesogenetic quartz
authigenesis and further quartz precipitation, probably cementation of petrofacies A sandstones could be in part
connected to Triassic magmatism. Thermal maturation of attributed to early emplacement oil. However, the Uerê
Devonian shales and hydrocarbon emplacement on reser- sadstones are immediately overlain by the Jandiatuba oil-
voirs took place at this time (Fig. 15; Mello et al., 1994). generator shales, and their burial history at basin depocentre
Extensive eogenetic silica authigenesis, represented by shows that hydrocarbon had migrated to reservoirs during
microquartz and chalcedony cementation, silicification of deep (, 3400 m) burial (Fig. 15). Additionally, the similar
pseudomatrix, plus bioturbation matrix, and of smectite volumes of quartz cementation and porosity above and
below oil/water contacts suggest that hydrocarbon
coatings, was promoted significant dissolution of sponge
migration has not halted diagenesis. Therefore, other factors
spicules, which took place soon after deposition. The
than early hydrocarbon emplacement probably control the
precipitation of pyrite, magnesian siderite and Fe-dolomi-
porosity preservation and distribution in the Uerê reservoirs.
te/ankerite occurred subsequent to silica cementation.
There is a direct correspondence between high values of
Sponge spicules were replaced by eogenetic carbonate
porosity and permeability in the reservoirs, and the presence
cement, leaving ‘ghosts’ discernible only in BSE images
of microcrystalline quartz rims (Fig. 16). The action of
(Fig. 8(C)), which indicate the large original amount of
microquartz rims in inhibiting the precipitation
sponge spicules within petrofacies A sandstones, which
of quartz overgrowths is known in a series of occurrences
would characterize them as hybrid arenites. The absence of
(Aase et al., 1996; Bloch et al., 2002; Osborne & Swarbrick,
eogenetic carbonate cement within mouldic pores after 1999; Ramm, Forsberg, & Jahren, 1997). Additionally,
dissolved feldspar grains and the oxidised borders of the the presence microquartz rims helps to stabilize the
poikilotopic carbonate cement spots is consistent with the intergranular contacts, thus increasing the resistance of the
eogenetic precipitation of the carbonates prior to meteoric rock to pressure dissolution and preserving porosity.
water influx. Therefore, it is concluded that the major control on
Meteoric eodiagenesis was responsible for feldspar porosity evolution in the Uerê sandstones was the eogenetic
dissolution and minor kaolinite precipitation. Rapid burial silica cementation and its inhibitory role on quartz
due to deposition of the thick Carboniferous-Permian cementation and pressure dissolution.
sequence promoted heterogeneous mechanical and chemi-
cal compaction, which affected mostly petrofacies B and C 9.8. Exploration significance
sandstones devoid of eogenetic cementation. During
subsequent shallow mesodiagenesis widespread, syn- to The main mechanism of porosity preservation in the
pos-compactional quartz overgrowths were precipitated Uerê sandstones is the inhibition of quartz overgrowth and
within petrofacies B and C, while restricted quartz out- pressure dissolution by microquartz rims. Thus, the most
growths were formed in petrofacies A. important factor to constrain reservoir quality in an oil
The occurrence of quartz outgrowths engulfing fibrous exploration approach is the distribution of this eogenetic
illite, which in turn cover earlier overgrowths in petrofacies silica cementation.
B and C, suggests a recurrence of quartz precipitation during Eogenetic silica cements (grain-rimming and pore-filling
mesodiagenesis. Feldspar albitisation and minor albite microquartz and chalcedony) and associated smectitic clay
overgrowths, as well as precipitation of anhydrite engulfing coatings occur specifically in sandstones interpreted as
late quartz outgrowths and illite coatings are probably storm deposits. These deposits were originally hybrid
connected to deep mesogenetic fluids derived from the arenites (Zuffa, 1980) rich in sponge spicules which, upon
overlying Carboniferous evaporites. dissolution of the spicules both within the arenites and in
1068 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

microquartz and/or chalcedony, and typically did not


evolved into good reservoirs (Fig. 12).

10. Conclusions

The study of Upper Devonian Uerê sandstones in eastern


Solimões Basin yielded important clues to the depositional
and diagenetic controls on the preservation of porosity and
quality evolution of these shallow-marine reservoirs:
(1) The stratigraphy and facies associations suggest
deposition within an overall progradational regime,
followed by a major transgression. The storm reworking
of nearshore siliciceous sponge biostromes allowed the
deposition of spiculites and spicule-rich (hybrid) arenites
bellow the fair-weather wave base depths.
(2) The main diagenetic processes affecting the Uerê
sandstones are the authigenesis of various forms of silica,
particularly microquartz rim cements, which occur orig-
inally in spicule-rich, hybrid arenites, and mechanical and
chemical compaction.
(3) Three reservoir petrofacies were defined, based on the
packing, porosity, and types of cementation: (i) petrofacies
A: porous sandstones (. 15%) with microquartz rims; (ii)
petrofacies B: tight (, 10% porosity), strongly compacted,
moderately quartz-cemented (, 5%) sandstones; (iii) pet-
rofacies C: moderately porous (10 – 15%), conspicuously
Fig. 16. Plot of thin section macroporosity versus horizontal air
permeability, with symbols representing the amounts of microcrystalline
quartz-cemented (. 5%) sandstones.
silica rims in sandstones from petrofacies A. Some samples devoid of (4) Four major diagenetic evolution stages were
microquartz present relatively high macroporosity values due to the presence recognised: (1) marine eodiagenesis, dominated by the
of clay coatings, which preserved porosity from quartz overgrowths, dissolution of sponge spicules and precipitation of silica
however, decreasing permeability. (restricted to petrofacies A), and siderite/dolomite;
interbedded spiculites (now spiculitic cherts), became (2) meteoric eodiagenesis, responsible for heterogeneous
porous, microquartz-cemented sandstones. Therefore, the feldspar dissolution and kaolinite precipitation; (3) shallow
areas with high frequency of storm reworking are prone to mesodiagenesis (1900 –2600 m depth), marked by hetero-
have more levels of spicule-rich deposits, and thus of geneous chemical compaction and quartz overgrowth
derived porous microquartz cemented sandstones. The cementation and (4) deep mesodiagenesis (. 2600 m
growth of the siliciceous sponges biostromes is interpreted depth), marked by illite authigenesis and further quartz
precipitation, and probably connected to the convection of
to have occurred in nearshore settings, where reworking by
hot fluids related with Triassic magmatism. Thermal
storms allowed the deposition of hybrid sands between the
maturation of Devonian shales and hydrocarbon emplace-
fair-weather and storm wave base (i.e. facies association 2).
ment in reservoirs took place at this time.
There is a direct relationship between the volume of
(5) Mouldic pores after spicules, which are widespread
microquartz rims, of preserved porosity and of per-
within petrofacies A microquartz- and chalcedony-cemen-
meability values (Fig. 16). It must be considered, however,
ted sandstones, attest to a major redistribution of silica from
that there is an optimum thickness for the microquartz rims biogenic sources to eogenetic cements. A thermodynamic-
effectively preserve the porosity without seriously com- kinetic model explains the progressive diagenetic
promising the permeability (Aase et al., 1996; Bloch et al., silica transformation from the phase of highest entropy
2002; Ramm et al., 1997). Thick rims obstruct the pore (amorphous opal) to the phase of lowest entropy (quartz) in
throats, and extremely thin rims are not efficient in a dissolution-reprecipitation pathway. The scarcity of opal-
preventing overgrowths and pressure dissolution. In the CT remnants is probably related to the occurrence of
Uerê sandstones, the optimum rim thickness is between eogenetic smectitic clay coatings, which accelerated
5 and 10 mm, and the optimum volume is 4 – 6%, which is the opal-CT/quartz transformation and/or competitively
apparently related to sands with less than 10 vol% of adsorbed silica, thus inhibiting opal-CT nucleation.
original spicules contents. Hybrid sands extremely rich in (6) The distribution of mesogenetic quartz overgrowth
spicules were cemented by thick rims and/or pore-filling cementation is widespread but extremely heterogeneous
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1069

from thin section to layer scale. This is attributed to Evolução das Bacias Sedimentares (pp. 169 –193). Rio de Janeiro:
inhibition of overgrowth by the presence of eogenetic PETROBRAS.
Chaika, C., & Williams, L. A. (2000). Density and mineralogy variations as
microcrystalline quartz rims, and to the convection of hot a function of porosity in Miocene Monterey Formation oil and gas
fluids connected to Triassic magmatism, which co-precipi- reservoirs in California. AAPG Bulletin, 85, 149 –167.
tated late quartz and illite. Chang, J., & Yortsos, Y. C. (1994). Lamination during silica diagenesis-
(7) The major mechanism of porosity preservation is the effects of clay content and Ostwald ripening. American Journal of
inhibition of quartz overgrowth cementation and pressure Science, 294, 137–172.
Craig, H. (1957). Isotopic standards for carbon and oxygen correction
dissolution by eogenetic microquartz rims. The distribution of factors for mass spectrometric analysis of carbon dioxide. Geochimica
the eogenetic silica cements (grain-rimming and pore-filling Cosmochimica Acta, 12, 133 –149.
microquartz, chalcedony) and associated clay coatings is Curtis, C. D., & Coleman, M. L. (1986). Controls on the precipitation of
directly related to storm layers rich in sponge spicules. early diagenetic calcite, dolomite and siderite concretions in complex
depositional sequences. In D. L. Gautier (Ed.), Roles of organic matter
in sediment diagenesis (pp. 23–33). SEPM Special Publication 38,
Society of Economic Paleontologists and Mineralogists.
Acknowledgements Deer, W. A., Howie, R. A., & Zussman, J. (1962) (Vol. 5). Non-silicates.
Rock-forming minerals, London: Longman, p. 371.
De Ros, L. F., Morad, S., Broman, C., Césero, P., & Gomez-Gras, D.
We gratefully thank PETROBRAS, in special Humberto (2000). Influence of uplift and magmatism on distribution of quartz and
Pampolha Lima and R. Nonato M. Cunha for access to illite cementation: Evidence from Siluro-Devonian sandstones of the
samples, data, information, resources, and for the license to Paraná Basin, Brazil. In R. Worden, & S. Morad (Eds.), Quartz
publish this work. Special acknowledgements are paid to the cementation in sandstones (26) (pp. 231 –252). Special Publications of
support of the Brazilian National Petroleum Agency—ANP International Association of Sedimentologists, Oxford: Blackwell.
Dickson, J. A. D. (1965). A modified staining technique for carbonates in
(grant and research funds to R.D. Lima) and the National thin section. Nature, 205, 587.
Research Council—CNPq (grant to L.F. De Ros). We Dove, P. M., & Crerar, D. A. (1990). Kinetics of quartz dissolution in
acknowledge the use of the support and analytical facilities electrolyte solutions using a hydrothermal mixed flow reactor.
of the Institute of Geosciences of Rio Grande do Sul Federal Geochimica Cosmochimica Acta, 54, 1609–1625.
Downing, K. P., & Walker, R. G. (1988). Viking formation, Joffre Field,
University. We thank Dr S. Morad, from Uppsala University,
Alberta: Shoreface origin of long, narrow sand body encased in marine
for his revision on a previous version of the manuscript. mudstones. AAPG Bulletin, 72, 1212–1228.
Dworkin, S. I., & Land, L. S. (1994). Petrographic and geochemical
constraints on the formation and diagenesis of anhydrite cements,
Smackover sandstones, Gulf of Mexico. Journal of Sedimentary
References
Research, A64, 339–348.
Ehrenberg, S. N., & Nadeau, P. H. (1989). Formation of diagenetic illite in
Aase, N. E., Bjørkum, P. A., & Nadeau, P. H. (1996). The effect of grain- sandstones of the Garn Formation, Haltenbanken area, mid-Norwegian
coating microquartz on preservation of reservoir porosity. AAPG continental shelf. Clay and Minerals, 24, 233–253.
Bulletin, 80, 1654–1673. Eiras, J. F (1998). Geology and petroleum system of the Solimões Basin,
Astin, T. R. (1987). Petrology (including fluorescence microscopy) of Brazil . 1998 AAPG International Conference and Exibition (pp. 446).
cherts from the Portlandian of Wiltshire, UK—evidence of an episode Rio de Janeiro.
of meteoric water circulation. In J. D. Marshall (Ed.), Diagenesis of Eiras, J. F., Becker, C. R., Souza, E. M., Gonzaga, F. G., Silva, J. G. F.,
sedimentary sequences. Geological Society Special Publication 36, Daniel, L. M. F., Matsuda, N. S., & Feijó, F. J. (1994). Bacia do
Oxford: Blackwell, pp. 73–84. Solimões. Boletim de Geociências da PETROBRAS, 8, 17–45.
Beard, D. C., & Weyl, P. K. (1973). Influence of texture on porosity and Folk, R. L. (1968). Petrology of sedimentary rocks. Austin, TX: Hemphill,
permeability of unconsolidated sand. AAPG Bulletin, 57, 349 –369. 107 p.
Bloch, S., Lander, R. H., & Bonell, L. (2002). Anomalously high porosity Franks, S. G., & Forester, R. W. (1984). Relationships among secondary
and permeability in deeply buried sandstones reservoirs: Origin and porosity, pore fluid chemistry and carbon dioxide, Texas Gulf Coast. In
predictability. AAPG Bulletin, 86, 301–328. D. A. McDonald, & R. C. Surdam (Eds.), Clastic diagenesis. AAPG
Boles, J. R., & Franks, S. G. (1979). Clay diagenesis in Wilcox sandstones Memoir 37, American Association of Petroleum Geologists.
of southwest Texas: Implications of smectite diagenesis on sandstone Friedman, I., & O’Neil, J. R. (1977). Compilation of stable isotopic
cementation. Journal of Sedimentary Petroleum, 49, 55 –70. fractionation factors of geochemical interest. In M. Fleischer (Ed.),
Caplan, M. L., & Bustin, R. M. (1996). Factors governing organic matter Data of geochemistry (pp. 12USGS Professional Paper 440-KK, United
preservation potential in marine petroleum source rocks from States Geological Survey.
palaeocontinental margins: Evidence from Upper Devonian to Lower Giles, M. R., & Marshall, J. D. (1986). Constraints on the development of
Carboniferous Exshaw Formation. Canadian Petroleum Geology secondary porosity in the subsurface: Re-evaluation of processes.
Bulletin, 44, 474 –494. Marine and Petroleum Geology, 3, 243 –255.
Caplan, M. L., & Bustin, R. M. (1999). Devonian-Carboniferous Girard, J.-P., Deynoux, M., & Nahon, D. (1989). Diagenesis of the Upper
Hangenberg mass extinction event, widespread organic-rich mudrock Proterozoic siliciclastic sediments of the Taoudeni Basin (West Africa)
and anoxia: Causes and consequences. Palaeogeology, Palaeoclimatol- and relation to diabase emplacement. Journal of Sedimentary
ogy, Palaeoecology, 148, 187 –207. Petroleum, 59, 233 –248.
Caputo, M. V., & Crowell, J. C. (1985). Migration of glacial centers across Gluyas, J. G., Grant, S. M., & Robinson, A. G. (1993). Geochemical
Gondwana during the Paleozoic Era. Geological Society of American evidence for a temporal control on sandstone cementation. In A.
Bulletin, 96, 1020–1036. Horbury, & A. Robinson (Eds.), Diagenesis and basin development
Caputo, M. V., & Silva, O. B. (1990). Sedimentação e tectônica da Bacia do (pp. 23–33). AAPG Studies in Geology, The American Association of
Solimões. In G. P. Raja Gabaglia, & E. J. Milani (Eds.), Origem e Petroleum Geologists.
1070 R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071

Gluyas, J., Jolley, L., & Primmer, T. J. (1997). Element mobility during Michalopoulos, P., & Aller, R. (1995). Rapid clay mineral formation in
diagenesis; sulfate cementation of Rotliegent sandstones, southern Amazon Delta sediments—reverse weathering and oceanic elemental
North Sea. Marine and Petroleum Geology, 14, 1001–1011. cycles. Science, 270, 614–617.
Grahn, Y. (1992). Revision of Silurian and Devonian strata of Brazil. Michalopoulos, P., Aller, R. C., & Reeder, R. J. (2000). Conversion of
Palinology, 16, 35 –61. diatoms to clay during early diagenesis in tropical, continental shelf
Hendry, J. P., & Trewin, N. H. (1995). Authigenic quartz microfabrics in muds. Geology, 28, 1095– 1098.
Cretaceous turbidites: Evidence for silica transformation processes in Mizusaki, A. M. P., Anjos, S. M. C., Wanderley, J. W., Filho Silva, O. B.,
sandstones. Journal of Sedimentary Research, A65, 380 –392. Costa, M. G. F., Lima, M. P., & Kawashita, K. (1990). Datação K/Ar de
Houseknecht, D. W. (1987). Assessing the relative importance of ilitas diagenéticas. Boletim de Geociências da PETROBRAS, 4, 237–252.
compaction processes and cementation to reduction of porosity in Morad, S. (1986). Albitization of K-feldspar grains in Proterozoic arkoses
sandstones. AAPG Bulletin, 71, 633–642. and greywackes from southern Sweden. Neues Jahrbuch for Miner-
Huggett, J., Dennis, P., & Gale, A. (2000). Geochemistry of early siderite alogie Monatshefte, 4, 145– 156.
cements from the Eocene succession of Whiteclif Bay, Hampshire Morad, S. (1998). Carbonate cementation in sandstones: Distribution
Basin, UK. Journal of Sedimentary Research, 70, 1107–1117. patterns and geochemical evolution. In S. Morad (Ed.), Carbonate
Hunt, D., & Tucker, M. E. (1992). Stranded parasequences and the forced cementation in sandstones (26) (pp. 1–26). Special Publications of
regressive wedge systems tract: Deposition during base-level fall. International Association of Sedimentologists, Oxford: Blackwell.
Sedimentary Geology, 81, 1– 9. Morad, S., Bergan, M., Knarud, R., & Nystuen, J. P. (1990). Albitization of
Jahren, J., & Ramm, M. (2000). The porosity-preserving effects of detrital plagioclase in Triassic reservoir sandstones from the Snorre Field,
microcrystalline quartz coatings in arenitic sandstones: Examples from Norwegian North Sea. Journal of Sedimentary Petroleum, 60, 411 –425.
the Norwegian continental shelf. In R. H. Worden, & S. Morad (Eds.), Mozley, P. S. (1989). Relation between depositional environment and the
Quartz cementation in sandstones (29) (pp. 271 – 280). Special elemental composition of early diagenetic siderite. Geology, 17, 704–706.
Publications of International Association of Sedimentologists, Oxford: Osborne, M. J., & Swarbrick, R. E. (1999). Diagenesis in North Sea HPHT
Blackwell. clastic reservoirs—consequences for porosity and overpressure pre-
Johnson, J. G., Klapper, G., & Sandberg, C. A. (1985). Devonian eustatic dicition. Marine and Petroleum Geology, 16, 337–353.
fluctuations in Euramerica. Geological Society of American Bulletin, Parrish, J. T. (1982). Upwelling and petroleum source beds, with reference
96, 567 –587. to the Paleozoic. AAPG Bulletin, 66, 750– 774.
Kahn, J. S. (1956). The analysis and distribution of the properties of Pitman, J. K., Henry, M., & Seyler, B. (1998). Reservoir quality and
packing in sand-size sediments. 1. On the measurement of packing in diagenetic evolution of Upper Mississippian rocks in the Illinois Basin:
sandstones. Journal of Geology, 64, 385–395. influence of a regional hydrothermal fluid flow event during late
Kastner, M., Keene, J. B., & Gieskes, J. M. (1977). Diagenesis of siliceous diagenesis. Geological Survey Professional Paper, 1597, Washington:
oozes. I. Chemical controls on the rate of opal-A to opal-CT United States Government Printing Office, p. 24.
transformation—an experimental study. Geochimica Cosmochimica Pittman, E. D., Larese, R. E., & Heald, M. T. (1992). Clay coats:
Acta, 41, 1041–1059. Occurrence and relevance to preservation of porosity in sandstones. In
Knauth, L. P. (1994). Petrogenesis of chert. In P. J. Heaney, C. T. Prewitt, & D. W. Houseknecht, & E. D. Pittman (Eds.), Origin, diagenesis, and
G. V. Gibbs (Eds.), Silica: physical behavior, geochemistry and petrophysics of clay minerals in sandstones (pp. 241–264). SEPM
materials applications (pp. 233–258). Reviews in Mineralogy, Miner- Special Publication 47, Society of Economic Paleontologists and
alogica Society of America. Mineralogists.
Loboziak, S., Melo, J. H. G., Quadros, L. P., Daemon, R. F., & Barrilari, Posamentier, H. W. (2002). Ancient shelf ridges—a potentially significant
I. M. R (1994). Devonian-Dinantian Miospore Biostratigraphy of the component of the transgressive systems tract: Case study from offshore
Solimões and Parnaı́ba Basins (with considerations on the Devonian of northwest Java. AAPG Bulletin, 86, 76–106.
the Parná Basin). Internal report. PETROBRAS/CENPES/DIVEX/ Posamentier, H. W., Allen, G. P., James, D. P., & Tesson, M. (1992). Forced
SEBIPE. regressions in a sequence stratigraphic framework: Concepts, examples,
Lundegard, P. D. (1992). Sandstone porosity loss—a big picture view of the and exploration significance. AAPG Bulletin, 76, 1687–1709.
importance of compaction. Journal of Sedimentary Petroleum, 62, Ramm, M., Forsberg, A. W., & Jahren, J. S. (1997). Porosity-depth trends in
250– 260. deeply buried Upper Jurassic reservoirs in the Norwegian Central
MacEachern, J. A., & Pemberton, S. G. (1992). Ichnological aspects of Graben: An example of porosity preservation beneath the normal
Cretaceous shoreface succession and shoreface variability in the economic basement by grain-coating microquartz. In J. A. Kupecz, J. G.
Western Interior Seaway of North America. In S. G. Pemberton (Ed.), Gluyas, & S. Bloch (Eds.), Reservoir quality prediction in sandstones
Applications of ichnology to petroleum exploration, a core workshop and carbonates (pp. 177 –200). AAPG Memoir 69, The American
(pp. 57–84). SEPM, Core Workshop, Society of Economic Paleontol- Association of Petroleum Geologists.
ogists and Mineralogists. Rosenbaum, J. M., & Sheppard, S. M. F. (1986). An isotopic study of
MacEachern, J. A., Zaitlin, B. A., & Pemberton, S. G. (1999). A sharp- siderites, dolomites and ankerites at high temperatures. Geochimica
based sandstone of the Viking Formation. Journal of Sedimentary Cosmochimica Acta, 50, 1147–1150.
Research, 69, 876 –892. Savoy, L. (1992). Environmental record of Devonian-Mississippian
Maliva, R. G., & Siever, R. (1988). Pre-Cenozoic nodular cherts: Evidence carbonate and low-oxygen facies transitions, southernmost Canadian
for Opal-CT precursor and direct quartz replacement. American Journal rocky mountains and northernmost Montana. Geological Society of
of Science, 288, 798 –809. American Bulletin, 104, 1412–1432.
McBride, E. F. (1989). Quartz cement in sandstones: A review. Earth Schiffman, P., Bird, D. K., & Elders, W. A. (1985). Hydrothermal
Science Reviews, 26, 69–112. mineralogy of calcareous sandstones from the Colorado River delta in
McDowell, S. D., & Paces, J. B. (1985). Carbonate alteration minerals in the Cerro Prieto geothermal system, Baja California, Mexico. Minerals
the Salton Sea geothermal system, California, USA. Minerals Magazine, 49(Part 3), 435 –449.
Magazine, 49(352), 469–479. Schmidt, V., & McDonald, D. A. (1979). The role of secondary porosity in
Mello, M. R., Koutsoukos, E. A. M., Mohriak, W. U., & Bacoccoli, G. the course of sandstone diagenesis. In P. A. Scholle, & P. R. Schluger
(1994). Selected petroleum systems in Brazil. In L. B. Magoon, & (Eds.), Aspects of diagenesis (pp. 175 –207). SEPM Special Publi-
W. G. Dow (Eds.), The petroleum system—from source to trap (pp. cation 29, Society of Economic Paleontologists and Mineralogists.
499– 512). AAPG Memoir 60, The American Association of Petroleum Searl, A. (1994). Diagenetic destruction of reservoir potential in shallow
Geologists. marine sandstones of the Broadford Beds (Lower Jurassic), north-west
R.D. Lima, L.F. De Ros / Marine and Petroleum Geology 19 (2002) 1047–1071 1071

Scotland: Depositional versus burial and thermal history controls on Szatmari, P (1996). Datação 40Ar/39Ar do vulcanismo Mesozóico nas
porosity destruction. Marine and Petroleum Geology, 11, 131–147. bacias do Solimões e Amazonas. Internal Report. PETROBRAS/
Siever, R., & Woodford, N. (1973). Sorption of silica by clay minerals. CENPES/DIVEX /SEMBA.
Geochimica Cosmochimica Acta, 37, 1851–1880. Walderhaug, O. (2000). Modeling quartz cementation and porosity in
Silva, O. B (1987). Análise da Bacia do Solimões (revisão litoestrati- Middle Jurassic Brent Group sandstones of the Kvitebjorn Field,
gráfica, magmatismo e geoquı́mica). MSc Thesis, Universidade Federal Northern North Sea. AAPG Bulletin, 84, 1325–1339.
de Ouro Preto, 158 pp. Walker, R. G., & Bergman, K. M. (1993). Shannon Sandstone in Wyoming:
Snedden, J. W., & Nummedal, D. (1991). Origin and geometry of storm- A shelf-ridge complex reinterpreted as lowstand shoreface deposits.
deposited beds in modern sediments of the Texas continental shelf. Journal of Sedimentary Petroleum, 63, 839 –851.
In D. J. P. Swift, G. F. Oertel, R. W. Tillman, & J. A. Thorne (Eds.), Walker, R. G., & Wiseman, T. (1995). Lowstand shorefaces, transgressive
Shelf sand and sandstone bodies: geometry, facies, and sequence incised shorefaces, and forced regressions: Examples from the Viking
stratigraphy (14) (pp. 283–308). Special Publications of International Formation, Joarcam area, Alberta. Journal of Sedimentary Research, 1,
Association of Sedimentologists, Oxford: Blackwell. 132– 142.
Streel, M., Caputo, M. V., Loboziak, S., & Melo, J. H. G. (2000). Late Frasnian- Williams, A. W., & Crerar, D. A. (1985). Silica diagenesis. II. General
Famennian climates based on palynomorphs analysis and the question of mechanisms. Journal of Sedimentary Petroleum, 55, 312–321.
the Late Devonian glaciation. Earth Science Reviews, 52, 121–172. Williams, A. W., Parks, G. A., & Crerar, D. A. (1985). Silica
Summer, N. S., & Verosub, K. L. (1992). Diagenesis and organic maturation diagenesis. I. Solubility controls. Journal of Sedimentary Petroleum,
of sedimentary rocks under volcanic strata, Oregon. AAPG Bulletin, 76, 55, 301–311.
1190–1199. Wollast, R., & de Broeu, F. (1971). Study of the behavior of dissolved silica
Surdam, R. C., Boese, S. W., & Crossey, L. J. (1984). The chemistry of in the estuary of the Scheldt. Geochimica Cosmochimica Acta, 35,
secondary porosity. In R. C. Surdam, & D. A. McDonald (Eds.), Clastic 613– 620.
diagenesis (pp. 127 –149). AAPG Memoir 37, American Association of Zuffa, G. G. (1980). Hybrid arenites: Their composition and classification.
Petroleum Geologists. Journal of Sedimentary Petroleum, 50, 21 –29.

You might also like