You are on page 1of 28

Marine and Petroleum Geology 155 (2023) 106379

Contents lists available at ScienceDirect

Marine and Petroleum Geology


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

Role of depositional environment on clay coat distribution in deeply buried


turbidite sandstones: Insights from the Agat field, Norwegian North Sea
Fares Azzam a, *, Thomas Blaise a, Makfoula Dewla a, Patricia Patrier b, Daniel Beaufort b,
Ahmed Abd Elmola d, Benjamin Brigaud a, Eric Portier a, Jocelyn Barbarand a, Sylvain Clerc c
a
Université Paris-Saclay, CNRS, GEOPS, 91400, Orsay, France
b
Université de Poitiers, IC2MP - UMR 7285 - CNRS, 86073, Poitiers, France
c
Neptune Energy Norge AS, Vestre Svanholmen 6, 4313, Sandnes, Norway
d
The James Hutton Institute, Aberdeen, AB15 8QH, UK

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

Keywords: The preservation of reservoir quality in deeply buried sandstones largely depends on the presence of grain-
Reservoir quality coating clays that can inhibit quartz overgrowth and cementation. These grain-coating clays are the focus of
Grain-coating chlorite today’s hydrocarbon exploration as they enable new oil and gas discoveries in sedimentary basins. However, the
Turbidite
origin and distribution of these clays in turbidites are poorly documented in scientific literature. Our study
Sandstone
Depositional environment
addresses this knowledge gap by investigating the mechanisms and parameters governing the clay coat devel­
Quartz cement opment in turbidite sandstones, with a focus on the Agat Formation (Fm) in the Norwegian North Sea. Petro­
Diagenesis graphic observations, scanning electron microscopy, and X-ray diffraction analyses were used to identify the
Dissolution origin of clay minerals and their distribution, as well as to understand the diagenetic evolution of the sandstone
and its impact on reservoir quality. Based on the sedimentary facies description, the Agat Fm comprises four
principal members deposited in various depositional environments, ranging from distal lobe fringe to amal­
gamated proximal lobes and weakly confined channels. Chlorite and kaolinite are the main authigenic clay
minerals in these sandstones, associated with a variable amount of inherited detrital clays consisting in Fe-
bearing kaolinite and mixed-layer minerals composed of illite and hydroxyl-interlayered clay minerals formed
in acidic soil environments. Well-developed chlorite coats are associated with the highest reservoir quality (φ >
10%), while discontinuous chlorite coats are linked to extensive quartz cementation and poor reservoir quality
(φ < 5%). Samples with abundant pore-filling chlorite also exhibit poor reservoir quality (φ < 7%). Proximal lobe
deposits and weakly confined channels show well-developed chlorite coats, whereas discontinuous chlorite coats
are mainly found in proximal lobe deposits associated with high velocity escaping fluids during dewatering.
Pore-filling chlorites are typically present in distal lobe fringes and levee deposits. Chlorite in the Agat Fm is
mainly derived from the replacement of inherited detrital clays via an intermediate berthierine precursor
forming during early diagenesis. The dissolution of Fe-rich grains such as biotite and ferric detrital clays under
reducing conditions governed the kinetics of berthierine/chlorite growth. The late diagenesis is characterized by
extensive feldspar dissolution in a closed diagenetic system, leading to the precipitation of kaolinite and quartz
cement. The presence of chlorite coatings around detrital grains preserved the porosity from extensive quartz
cementation during deep burial, highlighting the role of chlorite in maintaining the quality of deeply buried
turbidite reservoirs.

1. Introduction properties of these sandstones are, however, highly variable owing to


the great diversity of morphology, internal structure, and composition
Turbidite sandstones are of great economic interest as they could (Goemaere et al., 2014). This variability is directly linked to the texture,
represent potential reservoirs for hydrocarbon accumulation and sites sedimentary structure, and mineralogy of sediments, as well as the
for the geological storage of CO2 (Pettingill, 1998). The petrophysical evolution they undergo after deposition. For instance, during burial

* Corresponding author.
E-mail addresses: Fares.199@live.com, Fares.azzam@universite-paris-saclay.fr (F. Azzam).

https://doi.org/10.1016/j.marpetgeo.2023.106379
Received 3 March 2023; Received in revised form 31 May 2023; Accepted 16 June 2023
Available online 25 June 2023
0264-8172/© 2023 Elsevier Ltd. All rights reserved.
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

diagenesis, the sediments are usually subjected to several trans­ and late diagenesis. For example, clay minerals can form coatings
formations, both physical (e.g., compaction) and chemical (e.g., around detrital grains and, therefore, may preserve the reservoir quality
cementation, dissolution, and recrystallization) (Burley and Worden, (Azzam et al., 2022; Bello et al., 2021; Saïag et al., 2016; Tang et al.,
2003). With the increase of burial depth, the sediments will undergo 2018; Virolle et al., 2020) by limiting the nucleation of secondary quartz
higher temperature and pressure conditions (>70 ◦ C, >2000 m). At this (Ajdukiewicz and Larese, 2012). A small amount of clay grain coats
point, the development of quartz overgrowth can drastically reduce the (2–5%) can inhibit quartz overgrowth at great depths, up to 4 km (Bloch
reservoir quality (Worden and Morad, 2000). Clay minerals in sandstone et al., 2002; Ehrenberg, 1993; Virolle et al., 2020). However, thick clay
can play a crucial role in controlling the reservoir quality during early coatings can sometimes reduce permeability (Azzam et al., 2022;

Fig. 1. (A) Simplified map showing the location of the study area indicated by a black square with the main structural elements of the North Sea (modified from the
NPD website). (B) Paleographical settings of the Agat Fm during the Albian-Aptian period showing sandstone deposition in slope settings sourced from the East, the
two red dashed square shapes show the location of the Duva field and Agat field (map modified from Gradstein et al. (2016)). (C) Lower Cretaceous stratigraphy of
the Northern North Sea (Gradstein et al., 2016).

2
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Hansen et al., 2021). They can also fill the primary porosity through deposited after the major rifting phase of the Middle to Late Jurassic,
infiltration during or shortly after deposition (Houseknecht and Ross, which created a series of under-filled rifts and sub-basins (Glennie,
1992; Porten et al., 2019) or through direct precipitation from pore 2009). The sand intervals were deposited in a series of channel-levee and
fluids during diagenesis (Azzam et al., 2022; Mansurbeg et al., 2020) intraslope lobe systems in deep water, continental slope environment
and consequently reduce the reservoir quality. Most of the clay minerals (Figs. 1B and 3) (Martinsen et al., 2005). Remote field test pressure
found in deeply buried sandstone (>3500 m) are authigenic, resulting measurements in the Agat succession have shown that the sands lack
from the transformation of detrital clay precursors or in-situ crystalli­ pressure communication between wells 35/3–2 and 35/3–4 but also
zation from pore fluids (Worden and Morad, 2009). The origin and indicate a lack of pressure between the upper reservoir zone, and the
distribution of authigenic clay minerals (particularly grain-coating deeper reservoir zone in well 35/3–4, indicating structural and strati­
clays) within turbidite reservoirs are poorly studied compared to sand­ graphic compartmentalization (Gulbrandsen, 1987). The Agat Fm is
stone from a shallow marine environment (Dowey et al., 2012; Man­ divided into ten members based on well data, gamma ray response, and
surbeg et al., 2009; Yezerski and Shumaker, 2018). Currently, there is a biostratigraphy (Agat 10, Agat 20, Agat 30, Agat 40, Agat 50, Agat 60,
lack of studies regarding the origin of authigenic grain-coating clays in Agat 70, Agat 80, Agat 100, and Agat 110). Only Agat 50, 70, 80, and
deep marine sands (Dowey et al., 2012). Moreover, understanding the 100 members were cored since they contain hydrocarbons and are
impact of clay coats on preserving or reducing reservoir quality is crucial studied in the present paper.
for optimizing resource extraction and management in deep marine
sands, which are important targets for petroleum exploration (Azzam 1.2. Tectono-stratigraphy
et al., 2022; Hansen et al., 2021; Porten et al., 2019). Therefore, the
main aim of this study is to define the parameters and processes that The Cretaceous succession in the Northern North Sea is dominated by
control the distribution of authigenic clays within turbidites and un­ the deposition of fine-grained lithologies (offshore marls and mud­
derstand their impact on reservoir quality. stones) from suspension settling (Fig. 1C) (Isaksen and Tonstad, 1989).
The present study’s focus is the turbidite sandstone of the Agat The sediments are transgressive, recording an overall deepening in the
Formation (Fm) of the Aptian to Albian age, located on the northeastern basin (Ziegler, 1975). However, several regressive events during
margin of the North Viking Graben in Block 35/3 in the Norwegian Aptian-Albian times and during the late Turonian that were related to
North Sea (Fig. 1A and B). The Agat Fm is considered one of the most the culmination of Austrian tectonic activity locally resulted in the
important hydrocarbon reservoirs in the lower Cretaceous succession in remobilization and deposition of coarse siliciclastic sediments in deep
the Norwegian North Sea and also holds promising potential for CO2 sub-basins (Crittenden et al., 1998; Oakman and Partington, 1998;
storage (Boe et al., 2002). Previous, petrographic studies on the Agat Fm Ziegler, 1975, 1992). During the Aptian-Albian times, gravity flows on
in the Duva field (block 36/7) show significant variation in reservoir the Måløy Slope created a series of slopes and basin floor sand systems
quality related to the distribution of authigenic clays, particularly in the known as the Agat Fm (Bugge et al., 2001; Jackson et al., 2008; Skibeli
form of chlorite grain coatings (Azzam et al., 2022; Hansen et al., 2021). et al., 1995). These slope systems were sourced from the east, from a
However, no studies have been done in the Agat Fm in the Agat field relatively small source area (5000–8000 km2), and the adjacent paleo
(block 35/3–5) that show more complex facies distribution (Lien et al., shelf was relatively narrow (20–30 km wide) (Bugge et al., 2001; Hansen
2006) compared to the rather massive homogeneous turbidite system in et al., 2021; Jackson et al., 2008; Martinsen et al., 2005). The
the Duva field (Azzam et al., 2022). The strong heterogeneity of the Agat morphology of the sub-basins (basin floor geometry, topography, and
formation in terms of depositional environment and reservoir quality paleo-bathymetry) was directly governed by the late-syn-rift configu­
offers an excellent opportunity to determine the origin, source, and ration. The Agat Fm was deposited at the base of a fault-related slope
distribution of authigenic clays within turbidites and study their impact that was cut by several incised canyons. The sedimentary materials were
on reservoir qualities. To achieve these goals, our study will cover the mainly derived from the remobilization and reworking of shallow ma­
following: rine sands (Martinsen et al., 2005; Sea et al., 2021).

(1) identifying the significant diagenetic reactions that the sandstone 2. Materials and methods
has undergone during burial
(2) determining the origin of authigenic clays and the mechanisms All samples are provided by Neptune energy A/S, together with
that govern their distribution within turbidite systems petrophysical data of well logs and core description reports (Made by
(3) assessing the role of diagenesis and authigenic clays on the VNG Norge AS, which was recently acquired by Neptune Energy). A total
reservoir quality of 37 core plugs were collected from four wells in block 35/3 (8 samples
(4) comparing and discussing the similarities and differences with from 35/3–2 well, 14 samples from 35/3–4 well, 9 samples from 35/3–5
other turbidite reservoirs well, and 6 samples from 35/3-7 S well). All samples were thin-sectioned
(5) building a diagenetic model linking diagenesis, depositional and polished. The thin sections were impregnated with blue epoxy.
facies, and reservoir quality Samples containing carbonate cement were stained with alizarin-
potassium ferricyanide. The samples were also stained with sodium
Addressing these objectives will improve the reservoir quality pre­ cobaltinitrite to reveal alkali feldspar.
diction of deeply buried turbidites and lower risks during hydrocarbon
exploration or CO2 storage. 2.1. Optical microscopy

1.1. Regional settings A Leica DM750P polarizing petrographic microscope was used to
observe thin sections under different magnifications. A composite
The Agat field (Block 35/3) is located on the Måløy Slope, 50 km microphotograph was created from a set of overlapping images for each
offshore Norway (Fig. 1A). It lies directly to the east of the Søgn Graben. thin section and then point-counted 250 times using Jmicrovision®
The Måløy Slope is dominated by N–S-striking normal faults formed software (Roduit, 2007) to determine grain types and porosity per­
during the Late Jurassic-to-Early Cretaceous rifting. Seven exploration centage. Additionally, the size of approximately 300 grains was
wells were drilled in Block 35/3: 35/3–1, 3–2, 3-3, 3–4, 3–5, 3–6, and 3- measured along their longer axis. Data was then imported to Gradistat ©
7 S. Hydrocarbons were encountered only in wells 35/3–2, 35/3–4, 35/ Simon Blott, a software calculating the grain size distribution, skewness,
3–5, and 35/3-7 S in deep-water sandstones of the Aptian-Albian age, and sorting (Blott and Pye, 2001). Sphericity and roundness were
called the Agat Fm. The Agat Fm is a post-rift hydrocarbon play visually defined following the chart of Krumbein (1963). The type of

3
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

grain-to-grain contact and the intensity of compaction were described presently at maximum burial depth. The thermal gradient in the study
for each sample using the methods of Taylor (1950). area is estimated at 3.4 ◦ C/100 m from the bottom hole temperature in
wells 35/3–2 and 35/3–4. The formation temperature is estimated to be
2.1.1. Scanning electron microscopy 118 ◦ C at 3600 m, according to VNG Norge.
Scanning Electron Microscopy (SEM) observations on thin sections
were carried out using a Phenom mini-SEM benchtop (10 keV in imaging 3. Sedimentology and sequence stratigraphy
mode and 15 keV for elementary analyses) equipped with an Energy-
Dispersive X-ray Spectroscopy analyzer (EDS) (GEOPS laboratory). 3.1. Depositional facies
The chemical composition of chlorites was determined using a JEOL
JSM IT500 scanning electron microscope equipped with secondary The sedimentological model of the Agat Fm has been discussed for
electron (SE) and backscatter electron (BSE) detectors and coupled with many years. First, Shanmugam et al. (1994) and Skibeli et al. (1995)
a Bruker Linxeye Energy Dispersive X-ray Spectrometer (EDS) (IC2MP proposed an interpretation based on limited core and seismic data
laboratory). The conditions of the analyses were as follows: 15 kV, 1 nA, suggesting deposition by a series of sandy slumps and mass flows in an
WD 11 mm, and 50 s of acquisition time. Elemental quantifications were upper slope environment near the shelf edge. However, based on new 3D
calibrated using the following natural reference materials: albite (Na, Al, seismic reflection data, the sandstone was described as turbidites,
Si), almandine (Mg, Fe), diopside (Ca), orthoclase (K), spessartine (Mn), deposited as a series of slope fans, channelized slope fans, and slope
and Ti metal (Ti) and corrected by the Phi-Rho-Z method. channel complexes (Bugge et al., 2001; Gradstein et al., 2016; Nystuen,
1999). Based on the previous core description made by VNG Norge (VNG
2.2. X-ray diffraction analysis internal report) and this study, six main facies have been identified in
the cored section of the Agat Fm (Table 1):
The bulk mineralogy of the sandstone was obtained using a PAN­
alytical X’Pert–Pro diffractometer with a Cu anticathode (Kα1 = 1.5405; 3.1.1. Massive/fluidized sandstones (facies MFst) (Fig. 2A, B, C)
45 kV; 40 mA) at Geosciences Paris-Saclay (GEOPS) laboratory. X-ray Description: These facies are mainly composed of fine to medium-
diffraction (XRD) analyses of randomly oriented powder were con­ grained, moderately to well-sorted sandstones. The sandstones are
ducted on thirty-five samples to characterize the bulk mineralogy based generally clean and massive, with common dewatering features such as
on dhkl reflections. The diffractograms were acquired between 2 and dish structures and fluidization pipes. Outsized clasts and coarse grains
65◦ 2θ at a step size of 0.02◦ 2θ and an acquisition time of 1 s. Semi- of claystone can be seen as ‘floating’ or aligned at specific intervals. The
quantitative estimates of the mineral phases from X-ray diffraction grain size generally decreases towards the top within individual meter-
data were carried out using Rietveld refinement with X’Pert HighScore scale beds. The sand bodies are sometimes overlain by claystones that
© PANalytical software. are cut by injected sands. In well 35/3–5, MFst facies are highly silty,
The clay fraction (<5 μm) was separated by sedimentation from 15 therefore they are classified as Massive fluidized silty sandstone (MFsst
samples using the protocol of Brown and Brindley (1980) and Moore and facies).
Reynolds (1997). First, the samples were crushed and then placed in an Interpretation: These facies record deposition from a high-
ultrasonic beaker. The <5 μm clay fraction was then extracted by col­ concentration turbidity current. Dewatering structures indicate high
lecting the first 5 cm after 39 min of settling. To identify the clay
mineralogy according to their d00l reflections, X-ray diffractograms of
oriented preparations were acquired in air-dried conditions (AD). The Table 1
presence of expandable clay minerals was checked using solvation with Classification and interpretation of the main sedimentary facies in the Agat Fm.
ethylene glycol (EG). Acquisitions were made on a Bruker D8 Advance Facies Subfacies Textures and Processes Depositional
diffractometer (40 kV and 40 mA) coupled with a copper source of sedimentary environment (
structures Fig. 3)
wavelength (Kα1) = 1.540598 Å with the following conditions: 2–35 2θ,
step size of 0.02◦ 2θ, acquisition time per step of 1 s. X-ray diffractograms MFst Mfs1 M-scale High- Stacked high Net
massive/ concentration to Gross channels
of oriented sample preparations were analyzed at 2–35 ◦ 2θ, step size of
fluidized sandy turbidites or lobes
0.02◦ 2θ, and acquisition time per step of 1s (IC2MP laboratory). The sandstones
protocol of Moore and Reynolds (1997) was used to estimate the Mfs2 Dm-scale
Reichweite parameter (ordering type) and the percentage of illite in I/S massive/
mixed layers. fluidized
sandstones
SRst Stratified and Mainly low-
2.3. Fourier transform infrared spectroscopy rippled concentration
sandstones sandy turbidites
The Fourier Transform Infrared (FTIR) spectra of kaolin minerals Pst Dm-scale pebbly Sandy debrites Channel-lag
sandstones and subordinate deposit
were acquired using a PerkinElmer FT-IR Microscope Spotlight 400. The
tractional flows
spectra were obtained in reflectance mode in the range 4000 to 550 Bst M-scale banded Moderate Low Net to Gross
cm− 1, with a resolution of 4 cm− 1, an interval of 1 cm-1, and 25 scans at silty sandstones concentration channel levees
20 ◦ C. Analysis was made using imaging mode with a spatial resolution muddy sand system or deposits
of 6 μm. FTIR was used to distinguish between kaolinite and dickite by turbidites of distal fringes of
Hst Hs1 Sand/mud Intermittent lobes
examining the absorption bands between 3500 and 3750 cm− 1 corre­
heteroliths with sandy turbidites
sponding to the hydroxyl (OH) stretching region (Prost et al., 1989). dm-scale and minor
sandstone beds debrites
2.4. Burial thermal history Hs2 Sand/mud Intermitent low
heteroliths with concentration
cm-scale turbidites
The burial history curve of the study area was reconstructed from sandstone beds
well 35/3–2 using Schlumberger PetroMod 1D ©. Data of heat flow WBm Non- to weakly Dilute low- Overbank muds/
history, erosion, and exhumation in the study are taken from Baig et al. bioturbated density turbidites Distal fan/Basinal
(2019) and Ben-Awuah et al. (2014). The net exhumation in the study mudrocks and suspension mudstone
deposits
area is negligible (Baig et al., 2019), indicating that the reservoir is

4
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

rates of sedimentation and rapid sediment fall-out. Such deposits form succession up to 10 m thick in the core.
from rapid deceleration of flow, probably occurring at a break of slope. Interpretation: These stratified sandstones were deposited from low-
The presence of claystone drapes cut by injected sands capping some concentration tractional flows and are commonly found in the weakly
sand bodies suggests rapid abandonment followed by deposition of confined channels (Fig. 3). Development of stratification implies a
laterally extensive clays that acted as seals, allowing pore fluid pressure relatively low concentration of sand falling out from the decelerating
to build up in the sands before the seals ruptured and sand injection flow. Climbing-ripple cross-laminations are deposited by turbidity cur­
occurred. This facies appears mainly in the central part of the amal­ rents when suspended load fallout and bedload transport occur
gamated lobes and more locally in the weakly confined channels (Fig. 3). contemporaneously (Jobe et al., 2012). Thick sequences of climbing
ripples are likely generated by sustained flows with constant discharge,
3.1.2. Stratified and rippled sandstones (SRst) (Fig. 2D and E) also known as depletive steady flows (Kneller, 1995).
Description: These facies comprise planar- and cross-bedded to
rippled sandstones (ripples are <10 cm thick, showing a very low angle 3.1.3. Pebbly sandstones (Pst) (Fig. 2G)
of climb), which are generally moderately sorted to well sorted, clean to Description: The facies are generally found at the base of MFst facies.
silty, and fine to medium-grained. These stratified units locally become The deposits are clast to matrix-supported with clasts ranging in size
structureless as grain size contrast decreases. They occur in continuous from 2 mm to 30 mm. Clasts are preferentially aligned. They are

Fig. 2. Major sedimentary facies in the Agat Fm: (A) Clean structureless fine-grained sandstone; (B) fine-grained fluidized sandstone showing dish structures; (C)
fluidized sandstone showing pillars and pipes structure, the section is cemented by calcite; (D) Rippled sandstone showing cm-scale ripples where the stoss-side is
preserved; (E) Weakly stratified sandstone where stratification is defined by concentrations of Mica; (F) Banded sandstone showing ’Bands’ or laminae that are
defined by dark mica grains. The sandstones are 1–20 mm thick, generally fine-grained, and well-sorted. These facies are rich in mica and organic-rich material; (G)
Pebbly sandstone that is clast supported, the clast size ranges from 2 mm up to 30 mm and are generally preferentially aligned; (H) Sand-dominated heterolithic
interval showing sand injection features folded by compaction; (I) Clay dominated heterolithic interval composed of laminated claystone with wavy-bedded and
rippled very fine grained sand beds; (J) weakly to non-bioturbated siltstone and mudstone comprising folded fine-grained sandstone; (K) Core section composed of
mud-sized to silt-sized particles showing evidence of bioturbation (elliptical and dentritic features).

5
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 3. Depositional model of the Agat Formation showing the distribution of the main facies in the turbidite system. It is noteworthy that Agat 50, Agat 70, and Agat
80/100 represent separate turbidite systems and are not genetically related.

characterized by sharp bases. deformed stratification.


Interpretation: Probably deposited from energetic tractional currents Interpretation: The low bioturbation in these facies suggests rapid
related to the flows that incised the master channel. This facies appears accumulation of muddy turbidites, allowing little time for colonization
mainly at the base of the weakly confined channels (Fig. 3). by benthic microfauna or infaunal organisms. These facies are likely
deposited in anoxic conditions away from the main sandy fairway
3.1.4. Banded sandstones (Bst) (Fig. 2F) (overbank mud deposits all along the slope) (Fig. 3). The associated
Description: These facies comprise meter-scale beds of silty, lami­ muddy sandstones were probably deposited from concentrated, weakly
nated sandstone with common fluid escape features. Banding is defined cohesive flows that deformed as they were deposited.
by diffuse layers of siltier and cleaner sand. Bands are defined by dark Unlike the Agat field, the sedimentological model in the Duva field is
mica grains (0.1–3 mm) and intervening fine sand horizons are 1–20 mm dominated by massive unstratified coarse-grained deposits with very
thick. They are generally horizontal and pass transitionally up into thin fine-grained sediments interbeds (Azzam et al., 2022; Hansen et al.,
planar stratification of tractional appearance. 2021). The homogeneous sandstone succession in the Duva Field is
Interpretation: The relatively high silt content implies rapid depo­ related to the structural confinement of the Agat system within incised
sition from relatively muddy flows in the distal lobes (Fig. 3). The thick valleys, as confirmed by 3D seismic mapping conducted by Neptune
bedding and common fluid escape seen in these deposits imply rapid energy.
deposition from large-scale sediment gravity flows.

3.2. Sequence stratigraphy of the Agat section


3.1.5. Heterolith sandstones (Hst) (Fig. 2H and I)
Description: These facies are composed of 20–80% dm-scale to cm-
As mentioned before, the Agat section is divided into several mem­
scale sandstone beds intercalated with mudrocks. The sandstones are
bers (Agat 50 to 100). Interpretation of the depositional environment of
planar stratified to rippled. In some cases, the intercalated claystones are
each member was early made by VNG Norge using core description, well
injected by sands, whereas elsewhere, the claystones have been dis­
logs, and 3D seismic data (Fig. 4). The following section summarizes this
rupted and partly reworked by bioturbation.
interpretation:
Interpretation: These hetero-lithic deposits record episodic influxes
The Agat 50 interval is mainly composed of overbank mudstone,
of low-concentration sandy and muddy turbidites, with occasional hia­
interbedded with banded sandstone (Bst facies), debrites (Pst facies),
tuses allowing sparse colonization by microscopic epifauna and simple
and intruded heteroliths (Hst facies). This interval represents the distal
macroscopic infauna. They form proximally along the slope, in channel
part of an early deepwater lobe system prograding into the basin (Fig. 3).
levees or in distal fringe of lobes (Fig. 3).
The feeder channel system failed to progress far enough into the basin to
be recorded in the wells. The Agat 50 is followed by Agat 60, which was
3.1.6. Non-to weakly bioturbated mudrocks (WBm) (Fig. 2J and K)
dominated by the deposition of fine-grained sediments (siltstone and
Description: These facies are mainly composed of laminated clay­
mudstone) all over the slope (WBm facies). The sediment supply was
stone and siltstone with common rippled sand pinstripes. Minor slum­
limited during the Agat 60, probably due to a time of regional sea-level
ped, fluidized, highly silty sandstone and sandy siltstone beds are also
highstand. The sediment supply increased during the deposition of Agat
present, displaying prominent sandy pseudo-nodules and a weak,
70. This interval is composed of high Net:Gross sands dominated by

6
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 4. Stratigraphic correlation of deepwater sub-environments in the Agat 50 to Agat 100 interval, based on core sedimentology and gamma response (from VNG
Norge internal report).

amalgamated to weakly amalgamated structureless and dewatered Formation is mainly composed of quartz, K-feldspar, and Na-feldspar
sandstone (MFst facies) and weakly stratified sandstone overlain by a associated with variable amounts of clay minerals that consist in chlo­
muddy sandstone (SRst facies) and capped by siltstones and intruded rite, kaolinite, illite/mica, and mixtures of hydroxyl-interlayered min­
heteroliths (Hst facies). Coarser grains and pebbly sandstone occur to­ erals and mixed-layer minerals composed of illite and hydroxyl-
wards the base (Pst facies). The Agat 70 interval is interpreted as a lobe interlayered layers (mentioned as HIMs hereafter) and carbonates
complex, whereby each package represents an individual lobe (Fig. 3). (calcite, siderite, and traces of dolomite and ankerite). The mineralog­
The transition from Agat 70 to Agat 80 comprises siltstones-mudstone, ical composition also includes traces of pyrite, anatase, rutile, and
injected sandstones, deformed units, and debrites, which indicate a zircon. The semi-quantification of XRD using Rietveld refinement shows
relatively distal setting suggesting system backstepping. This transi­ some similarities between Agat 100 and Agat 80 units that differ from
tional unit is overlain by the Agat 80, which is dominantly composed of Agat 70 and Agat 50 (Fig. 5). The Agat 50 and Agat 70 contain more
stratified and massive (Mfst and SRst facies) sand units with fine-grained quartz and Na-feldspar, fewer K-feldspars, and mica. On the other hand,
thin beds separating the main sand units. The Agat 80 is interpreted as a Agat 80 and Agat 100 contain less quartz and are highly rich in K-
weakly-confined channelized succession (Fig. 3). The preservation of the feldspar with only a small amount of Na-feldspar. Calcite is heteroge­
fine-grained interbeds is probably related to channel migration (avul­ neously distributed across the Agat Fm but is highly concentrated in
sion) and vertical aggradation. The transitional unit is laterally exten­ Agat 100 (av. 15%) and Agat 80 (av. 17%), while being less abundant in
sive and was recorded in all wells suggesting that it may be associated Agat 70 (av. 9.5%) and scarce in Agat 50 (av. 0.6%). Similarly, siderite is
with relative sea level rise. Therefore, Agat 70 and 80 are not genetically more common in Agat 100 (Av. 1.3%) and Agat 80 (Av. 1.4%), while
related but rather represent two separated deep-water systems. The being scarce in Agat 70 (av. 0.2%). However, the highest concentration
channelized Agat 80 system is abruptly overlain by Agat 100, which of siderite is recorded in Agat 50 (av. 2.2%). Kaolinite and chlorite are
shows strong heterogeneity of lithology across the studied wells. In the present in all samples in varying amounts ranging from 2% up to 21%,
South (well 35/3–5), Agat 100 is composed of medium-grained sand­ with an average of 6%.
stone, whereas in the North (well 35/3–4 and 35/3–2), the grain size is
rather fine-grained dominated by siltstone, showing intercalation with
fine-grained sandstone. Agat 100 is interpreted as channel complex in 4.2. Petrography: detrital mineralogy
well 35/3–5, whereas in well 35/3–4, it is interpreted as a levee deposit
(Fig. 3). Based on petrographic observations, most of the samples are sub-
mature, showing a moderate angularity, and are moderately well sor­
4. Results ted. The grain size ranges from lower fine sand to upper fine sand
(166–268 μm) except for the Agat 100 in well 35/3–5, where sandstone
4.1. Bulk mineralogical composition is mainly composed of lower medium sand (300–400 μm) (Fig. 4B).
Compaction is weak to moderate but is locally strong in samples con­
Based on XRD analysis (Fig. 5), the bulk mineralogy of the Agat taining more significant amounts of ductile grains (mica, glauconite, and
pseudomatrix). This is the case of Agat 50 and some samples in Agat 70,

7
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

polycrystalline (av. 18%, 9%, 10%, and 18% in Agat 50, 70, 80, 100,
respectively). Feldspars mainly comprise K-feldspars in Agat 80 and 100
(Av. 7.5% of K-feldspars and 2.8% of Na-plagioclase) but are dominated
by Na-feldspars in Agat 50 and 70 (Av. 4.6% of K-feldspars and 5.2% of
Na-plagioclase). Most of K- and Na-feldspar grains have undergone
partial to total dissolution. However, some K-feldspar and plagioclase
grains with evidence of sericitization show low dissolution features.
Lithics are largely metamorphic rock fragments, but locally igneous and
sedimentary rock fragments were observed. The metamorphic rock
fragments consist of polycrystalline quartz showing schistose patterns.
The rigid framework also includes heavy minerals (zircon, garnet,
opaques, rutile, tourmaline) and bioclasts. Heavy minerals, particularly
zircon and rutile, are widespread in Agat 70 (up to 1.5%) while being
rare in the other Agat units. Bioclasts comprise brachiopod fragments,
some tests of foraminifera, and echinoid spines.

4.2.2. Ductile framework


The ductile framework is mainly composed of mica (0.5–19.5%),
glauconitic grains (0.5–9%), detrital clay matrix (traces up to 21%), and
traces of black to brownish elongated macerals. Mica grains comprise
biotite (Av. 3.4% in Agat 70, 80, 100 and av. 7.4% in Agat 50),
muscovite (Av. 1.2% in Agat 70, 80, 100 and av. 2.3% in Agat 50), and
glauconite (Av. 1% in all Agat members). Both biotite and muscovite
show variable degrees of deformation due to compaction (Fig. 6E and F).
Extensive replacement of biotite by chlorite, siderite, and pyrite is
commonly observed (Fig. 9C and D). Glauconitic grains are comparable
in size to the associated quartz grains. They appear as discrete olive-
green colored grains showing a globular shape (Fig. 6E). Detrital clay
matrix (up to 21%) includes ductile lithic fragments that comprise
degraded igneous fragments, and sedimentary rock fragments consisting
of variably deformed mudstone/siltstone clasts.
Detrital clay matrix also includes pore-filling clays that are
commonly observed in Agat 100 (brownish matrix) (Fig. 6C). They also
form coatings around detrital grains that are partially chloritized
(Fig. 8C and D). SEM-EDS analyses of these clays (Table 4) indicate a
high content of Al2O3 (av. 22.8%) as well as a high content of iron
expressed as Fe2O3 (av. 9%). When plotted in the MR3+- 2 R3+- 3 R2+
ternary diagram (Velde, 2000), it can be inferred that they belong to the
compositional field of dioctahedral phyllosilicates, which encompasses
minerals like illite, dioctahedral smectites, and kaolinite/donbassite
(Fig. 10).
Samples containing high amounts of these clays show a very broad
diffraction peak with maxima near 11.2 Å (~7.8 ◦ 2θ) that shift to 13.1 Å
(~6.9 ◦ 2θ) after EG saturation. Such XRD behavior is typical of mixed-
layer minerals composed of illite and expandable clay phases (Moore
and Reynolds, 1997). However, the chemistry of these clays is not
compatible with a regular illite/smectite (I/S). The relatively high Fe2O3
Fig. 5. Semi-quantification of the bulk mineralogical composition from X-ray content in these clays is more consistent with Fe-rich hydrox­
diffraction using Rietveld refinement. The samples are selected from different
y-interlayered clay minerals (HIMs) interlayering with illite (Virolle
wells. HIMs: hydroxyl-interlayered clay minerals.
et al., 2022). HIMs typically form under oxidizing conditions in
moderately acidic soil environments (Meunier, 2007). Barnhisel and
where the presence of mica and pseudomatrix significantly accentuates Bertsch (1989) defined HIMs as a mixture between smectite, vermicu­
the effects of compaction (Fig. 6E and F). lite, and pedogenic or aluminous chlorite. The formation of these clays is
believed to occur during the weathering process in acidic soils by the
4.2.1. Rigid framework interlayering of hydroxy-metal complexes (Al3+, Mg2+, Fe2+/3+) into
The point-counting results of the different Agat members are sum­ expansible 2:1 phyllosilicates such as smectite, and vermiculite (Dietel
marized in Tables 2 and 3. Point counting results show that the sand et al., 2019; Georgiadis et al., 2020).
grains are mainly composed of quartz (35%–68%) associated with K- SEM observations of these HIMs-rich-matrix show that they contain
and Na-feldspar (3.5%–16.5%) and minor lithic rock fragments (0.5%– traces of detrital kaolinite forming small individual crystallites (Fig. 8A
11.6%). Using the quartz-feldspar-lithic (QFL) ternary diagram classi­ and B). These detrital kaolinites contain relatively high amounts of
fication of Folk (1980), most of the samples fall in the subarkose and Fe2O3 (av. 4%) as opposed to authigenic kaolinites, which are only
arkose domains (Fig. 6A). Some samples in the Agat 100 fall in the composed of Al and Si. Traces of detrital kaolinite were also observed in
feldspathic litharenite domain showing a significant increase in lithic some samples forming long vermiform booklets replaced by chlorite
fragments. (Fig. 8G and H). Fe-rich kaolinite is generally associated with HIMs in
Quartz grains are mainly monocrystalline (av. 34%, 45%, 41%, and soil profiles (Virolle et al., 2022), which confirm the pedogenetic origin
32% in Agat 50, 70, 80, and 100, respectively) with subordinate of these clays.

8
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 6. (A) QFL (Quartz– Feldspar – Lithics) ternary diagram of the different sandstone samples in the Agat Fm; (B) Cumulative grain size frequency plot for the main
Agat members; (C) Mineralogical assembly from Agat 100 (3231 m) showing sandstone that is coarse grained and poorly sorted, sand grains are composed of
polycrystalline and monocrystalline quartz and K-feldspar, the primary intergranular pores are filled with chlorite and I/S-kaolinite-rich matrix; (D) Mineralogical
assembly from Agat 80 (3466 m) showing fine grained sandstone with high porosity (mainly primary), that is locally filled with kaolinite developing near secondary
pores, all the grains are covered with thin continuous chlorite coating; (E) Mineralogical assembly from Agat 70 (3500 m) showing fine grained sandstone that are
strongly compacted due to the high amount of biotite and muscovite and the extensive development of quartz overgrowth; (F) mineralogical assembly from Agat 50
(3692 m) showing fine to medium grained sandstone that are locally strongly compacted, the porosity is filled with chlorite; Legends: Quartz polycrystalline and
monocrystalline (Qp, Qm), K-feldspar (K-felds), Primary porosity (PP), Secondary pores (SP), Dissolved felspar (Dis-felds), Biotite (B), Muscovite (M), Glauconite (G),
Pore-filling chlorite (PFC), chlorite coat (CC), Quartz overgrowth (QO), Kaolinite (K), Pore-filling detrital clays (PFDC), Titanium oxide (TOx).

Only one of the studied samples (i.e., Agat 100–3406 m) shows the superstructure reflection at very low diffraction angles (29.2◦ ) (Moore
presence of discrete smectite (peak at 14.7 Å in air-dried conditions and and Reynolds, 1997).
16.9 Å after EG saturation) (Fig. 5). The lack of corrensite (C/S mixed
layers) in this sample is confirmed by the absence of corrensite

9
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

4.3. Petrography: authigenic mineralogy F). Kaolinite is characterized by d00l reflections at 7.13 Å and 3.57 Å. The
d00l reflections of kaolinite overlap with those of dickite. Dickite is
Authigenic minerals consist primarily of non-expandable clay min­ identified using randomly oriented XRD preparations, marked by the
erals (mainly kaolinite and chlorite), carbonate cement (calcite, siderite, following reflections: 2.32 Å, 2.21 Å, 1.97 Å (Fig. 12A) (Beaufort et al.,
and traces of dolomite and ankerite), and a small amount (<0.5%) of 1998; Moore and Reynolds, 1997). The presence of dickite (in addition
pyrite, microcrystalline quartz, and anatase. Quartz overgrowth is pre­ to kaolinite) in sample 3707 m was also confirmed using FTIR spec­
sent mainly in Agat 50 and 70 in variable amounts (traces up to 5.1%). troscopy, marked by a change in the IR spectra in the
hydroxyl-stretching band region characterized by broadening and
4.3.1. Authigenic clay minerals shifting of peaks from 3701 cm to 1 to 3712 cm-1 and from 3650 cm to 1
to 3655 cm-1 and weakening at 3668 cm-1 (Fig. 12B and C) (Beaufort
4.3.1.1. Kaolinite/dickite. Kaolinite is present in most of the samples et al., 1998; Lanson et al., 2002).
(traces up to 5%) as pore-occluding cement. It occurs as a loosely packed
vermiform booklet, mainly found within primary intergranular porosity 4.3.2. Chlorite
or near degraded/dissolved feldspar grains (Fig. 8E). In addition to Chlorite has a relatively complex distribution across the Agat Fm and
kaolinite, dickite was detected with kaolinite in the deepest samples is observed in all the samples in variable amounts (traces up to 12%). It
(Agat 50–3707 m) but appears less abundant below 3500 m. Authigenic can be found in different forms: (1) as a grain-coating, forming rosette-
kaolinite booklets range in length from 10 μm up to 60 μm (Fig. 8E and like crystals around detrital grains (Fig. 9A and B); (2) as a pore filling,

Fig. 7. X-ray diffraction traces of oriented clay fraction specimens (<5 μm) for some representative samples of the different Agat members (AD: Air-dry, EG: Ethylene
glycol-solvated). MLM: mixed-layer clay minerals composed of illite and hydroxyl-interlayered layers.

10
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

occluding the primary intergranular pores or, in rare cases filling comprising Massive/fluidized sandstones (MFst facies), Stratified and
degraded plagioclase grains (Fig. 9A and B); (3) as a replacement of rippled sandstones (SRst facies), and to a lesser extent, Pebbly sand­
detrital kaolinite/HIMs (Fig. 9G and H), and mica, particularly biotite stones and Heterolith sandstones (Pst and Hst). Chlorite Microfacies 3 is
(Fig. 9C and D). Chlorite is identified by the following d00l reflections: found in both Agat 70 and 80 in Massive/fluidized sandstones (MFst
14.13 Å, 7.05 Å, 4.70 Å, and 3.54 Å. In all samples, the intensity of the facies). Chlorite Microfacies 4 is found in all Agat members but mainly in
even reflections (002 and 004) is higher compared to the odd reflections Agat 100, composed of mud-rich channel complex deposits and in the
(001 and 003), indicating iron-rich chlorite (Fig. 7) (Brown and Brind­ distal lobe deposits of Agat 50 (Hst facies and Bst facies).
ley, 1980). Chlorite polytypes were characterized by randomly oriented
XRD preparation. The typical peaks of polytype Ib (β = 90◦ ) have been 4.3.5. Carbonate cements
identified in most of the studied samples marked by the reflections at Calcite.
2.67 Å, 2.50 Å, and 2.15 Å (Fig. 12A) (Bailey, 1988; Moore and Rey­ Calcite (traces up to 35%) is the most important pore-occluding
nolds, 1997). However, a small component of chlorite IIb was detected cement in the Agat Fm. Based on the core description and XRD data,
in one sample (3447 m), marked by the following reflections at 2.01 Å, approximately 10% of the total volume of the Agat Fm is cemented by
2.59 Å, and 2.66 Å. calcite. Calcite is very common in Agat 100 (av. 15%) and Agat 80 (av.
20%), while being less developed in Agat 70 (av. 9%) and Agat 50
4.3.3. Chlorite elemental composition (traces). Two generations of calcite cement were identified in the stud­
The chemistry of chlorites was measured in the different Agat ied samples: (1) calcite 1 (pre-compactional calcite) is a mixture of
members using SEM-EDS analysis (Fig. 10A, Table 4). Chlorite coats are ferroan and non-ferroan calcite. It occurs only in a few samples. It shows
iron-rich, with an average FeO value of 35%. However, a strong varia­ a sparry texture with a crystal size ranging from 20 μm up to 180 μm
tion of chlorite chemistry was noted between the different Agat mem­ (Fig. 13A). Calcite 1 can occupy up to 35% of the intergranular volume
bers and between pore-filling chlorites and chlorite coats. In Agat 50 and and mainly occurs between loosely compacted grains showing floating
Agat 100, the chemical composition of chlorite coats appears quite texture. It also shows a high degree of fracturing and pressure-solution,
similar, with an average SiO2, Al2O3, and XFe (Fe/(Fe + Mg + Mn)) of where some siderite crystals are developed (Fig. 13A). Calcite 2 (post-
33.5%, 25.3%, and 0.75, respectively. In Agat 70 and 80 chlorites, the compactional calcite) is mainly ferroan in composition and occurs as a
average of SiO2, Al2O3, and XFe is 31.2, 24.7, and 0.71, respectively. blocky cement phase, filling both primary and in some cases, secondary
Overall, the Mg content is higher in Agat 70 and Agat 80 while being pores developed after plagioclase dissolution (Fig. 13C). However, it
lower in Agat 50 and Agat 100. Additionally, pore-filling chlorites does not appear to fill secondary porosity created by K-feldspar disso­
contain less Fe than chlorite coats, but both show relatively similar Mg lution (Fig. 13B). Ferroan calcite is also observed enclosing grain-
content. The chemistry of chlorites was plotted on the MR3+-2 R3+-3 R2+ coating chlorite (Fig. 13D). This generation of calcite occupies up to
ternary diagram (Fig. 10B). The latter suggests that many of the 20% of the intergranular volume of the samples and mainly develops
analyzed chlorites show mixing with other mineral phases. Abnormally between compacted grains.
low octahedral occupancy coupled with significant interlayer charge Siderite.
and excess of silicon in tetrahedral sheets is typical of a mixture of The average siderite content in the studied samples is 2.5%. How­
chlorite with detrital clays such as HIMs and Fe-bearing kaolinite ever, this proportion increases in Agat 50, reaching up to 13.4%. Also,
(Fig. 10B). siderites are highly concentrated in certain levels within Agat 100,
reaching up to 8%, considerably reducing the porosity. Siderite occurs in
4.3.4. Chlorite distribution a variable amount in Agat 80 and Agat 70 but is generally less than 3%.
Chlorite shows a complex distribution across the studied samples. It appears as single or locally clustered rhombohedral-shaped crystals
The size of the coating ranges from 2 to 10 μm. Grain-coat coverage closely associated with biotite (Fig. 9C and D) and chlorite (Fig. 9G and
ranges from 10% up to 100%. Based on the size of the coating, the grain- H). It mainly replaces biotite but can also be found as coatings around
coat coverage, and the ratio of chlorite coat to pore-filling chlorite, the detrital grains (Fig. 8A and B). SEM analysis shows that siderite is rich in
samples are classified into four microfacies: magnesium with an average MgO of ~10%. The core of the siderite is
often dissolved, forming intracrystalline porosity.
1) Microfacies 1: Partially coated grains (grain-coat coverage ranging Dolomite and ankerite.
from 10% up to 70%) with commonly developed quartz overgrowth, These two minerals were only detected using XRD analysis. Dolomite
generally lacking pore-filling chlorite (Fig. 11A). is detected in one sample (Agat 100–3238 m). Ankerite is found in some
2) Microfacies 2: completely coated grains (grain-coat coverage >95%) samples in Agat 80 and 100.
lacking clay bridging between detrital grains with a relatively ho­
mogenous size of chlorite coats and lacking pore-filling chlorite and 4.3.6. Other minor diagenetic minerals
quartz overgrowth (Fig. 11B). In addition to the authigenic clays and carbonates, SEM observations
3) Microfacies 3: completely coated grains (grain coat coverage >90%) reveal diagenetic pyrite, microcrystalline quartz, quartz overgrowth,
showing clay bridging between detrital grains, the size of the coats is albite, and titanium dioxide (anatase) present in small amounts
heterogenous, quartz overgrowth is absent, and pore-filling chlorite (<0.5%).
is low to moderate (Fig. 11C). Pyrite is mainly found as framboidal aggregates in some sedimentary
4) Microfacies 4: sandstone dominated by pore-filling chlorites with a lithic fragments and organic fragments. It can also be found at the sur­
relatively small amount of chlorite coats heterogeneously distributed face of detrital grains in association with chlorite (Fig. 9A and B). SEM
around detrital grains, quartz overgrowth is variable (traces up to observation also reveals the presence of microcrystalline quartz forming
5.5%) (Fig. 11D). coating around detrital grains. The crystals are randomly oriented,
measuring roughly between 2 and 10 μm in length (Fig. 13E). These
Some samples can show a mix of these microfacies. Fig. 11E and F microcrystalline quartz rims are mixed with chlorite and relicts of HIMs
shows the distribution of chlorite microfacies in the different Agat (Fig. 13E and F). Quartz overgrowth, on the other hand, appears as large
members and in the main sedimentary facies. Overall, Chlorite Microf­ euhedral crystals engulfing early microcrystalline quartz and chlorite
acies 1, 2, and 3 are mainly observed in Agat 70 and 80. Chlorite (Fig. 13F). Quartz overgrowth is heterogeneously distributed in the
microfacies 1 is dominant in Agat 70, composed of proximal lobe de­ studied samples but is largely developed in the Agat 70 and 50 (traces up
posits corresponding mainly to Massive/fluidized sandstones (facies to 5%). Finally, anatase is observed in all the studied samples in a small
MFst). Chlorite Microfacies 2 is dominant in Agat 80 and in Agat 70, amount. It comprises either single crystals or small clusters of crystals,

11
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 8. (A–B) SEM image of detrital matrix in Agat


100 (3262 m). The detrital matrix is composed of
kaolinite and hydroxy-interlayered clay minerals
(HIMs). Note that the outer layer of the matrix is
replaced by siderite and chlorite; (C) SEM Image of
quartz grains coated by detrital clays (HIMs), which is
partially replaced by chlorite in the outer layer in
contact with the porosity. See Table 4 for the chemical
composition of detrital clay coats; (D) SEM image of
detrital clay coats (I/S) forming the inner part of
chlorite coat; (E) SEM image of authigenic kaolinite
filling the primary intergranular pores near dissolved
plagioclase grain; (F) High magnification SEM image
of authigenic kaolinite showing vermicular habits; (G)
SEM image of vermiform clays being replaced by
chlorite. These morphologies are observed locally in
some samples forming pseudo-matrix; (H) SEM image
of kaolinite-like mineral being totally replaced by
chlorite.

12
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 9. (A) SEM image of chlorite coating the surface of


quartz grains and filling the intergranular porosity. The
coating ranges in size from 2 μm up to 4 μm; (B) chlorite
filling and coating partially dissolved plagioclase grain.
Note the presence of framboidal pyrite associated with
chlorite; (C–D) SEM image of deformed biotite grains
being replaced by siderite and chlorite; (E) SEM image of
clay coats replaced by chlorite forming aggregates
around a detrital grain. (F) zoom on the aggregates
showing the presence of silt-sized biotite and mica grains
dispersed in an early detrital clay matrix that has been
replaced by chlorite; (G–H) SEM image of chlorite and
siderite replacing a detrital clay matrix composed of
detrital kaolinite (partially chloritized) and fibrous clays
(HIMs).

13
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 10. Chemical composition of chlorite coats and pore-filling chlorites plotted on (A) a scatter plot of Fe vs. Mg and on (B) a MR3+- 2R3+ -3 R2+ ternary diagram
(Velde, 1985).

which are present within primary intergranular pores. In SEM, anatase porosity loss, OP is the original porosity before compaction estimated
crystals are well-developed and display a characteristic bipyramidal based on grain size and sorting using the chart of (Beard and Weyl,
habitus (Fig. 13H). 1973), CEM (%) is the volume of cement, and IGV (%) is the inter­
granular volume (total volume minus the volume of framework grain).
Fig. 14E shows that porosity was equally lost by compaction and by
4.4. Petrophysical properties
cementation. Porosity is strongly lost by carbonate cementation and
moderately lost by the presence of pore-filling chlorites and quartz
Macroporosity (from point counting) is dominated by primary
cement. In weakly cemented sandstones, 50–80% of the original
intergranular porosity (Av. ~ 3.5%) and subordinate secondary porosity
porosity was lost by compaction with an average of 62%, while 20–40%
(Av. ~1.6%). The highest porosity and permeability are recorded in
of the porosity was plugged by cementation. The cementation in these
Agat 70 and 80, followed by Agat 100, while the poorest reservoir
sandstones was mainly related to the precipitation of kaolinite. The
quality is generally found in Agat 50 (Fig. 14A). A strong relationship is
growth of chlorite coat has a minor impact on the intergranular volume.
found between chlorite microfacies and the porosity and permeability
(Fig. 14B). Samples dominated by microfacies 2 and 3 show high
5. Discussion
intergranular porosity (Av. 10%) and well-connected pores (10–500
mD). The intergranular porosity in these samples generally lacks pore-
5.1. Paragenetic sequence
filling chlorites and quartz overgrowth, and the compaction is rela­
tively weak to moderate. Samples dominated by microfacies 1 and 4
A paragenetic sequence is constructed based on textural relation­
show a porosity of less than 7% and permeability ranging from 0.5 mD
ships between the different authigenic phases (Fig. 15). The diagenetic
up to 10 mD. The intergranular porosity in microfacies 4 is poorly
evolution is represented along with the burial history of the Agat Fm
connected due to great amounts of ductile grains, mainly pore-filling
(well 35/3–2) reconstructed using Petromod© software. Diagenetic
chlorites and micas, causing greater compaction. In microfacies 1, the
history is subdivided into two main stages: (1) An early stage that takes
intergranular porosity is reduced by the extensive growth of quartz
place at a relatively shallow depth under the influence of depositional
cement. Samples containing large amounts of calcite cement show the
pore waters and bacterial activity, (2) A late stage, which occurs after
poorest reservoir quality. The porosity in these samples is mainly related
the main phase of compaction at a higher temperature.
to secondary pores (Av. 0.5%) developed from the dissolution of un­
Early diagenesis is characterized by the mechanical infiltration of
stable grains (largely Na-feldspar). In terms of sedimentary facies,
detrital clays, grain rearrangement and compaction, precipitation of
massive fluidized sandstones (Mfst) and stratified sandstone (SRst) show
calcite 1, precipitation of microcrystalline quartz, Co-precipitation of
the highest reservoir quality (Fig. 14C). The reservoir quality is gener­
siderite, pyrite, and chlorite. Early diagenesis also involved some
ally poor in Heterlolitic sandstone (Hst) and banded sandstone (Bst) and
dissolution of feldspars (mainly plagioclase) and probably early calcite
variable in pebbly sandstone (Pst). Calcite cementation was detected in
cement. The Agat Fm contains variable amounts of detrital clays forming
all types of sedimentary facies. A positive correlation is found between
the inner part of the chlorite coat, or as a pore-filling matrix. These clays
secondary porosity and kaolinite (R2 = 0.76) (Fig. 14D). The proportion
are composed of mixed layer illite and HIMs associated with detrital
of kaolinite increases with the increase of secondary porosity related to
kaolinite. Mechanical infiltration of these clays likely occurs during or
the dissolution of feldspar grains.
shortly after sediment deposition through dewatering processes (see
The relative amount of porosity loss by compaction and cementation
section 5.2.1). This was followed by local precipitation of calcite 1,
was estimated by using the following equations of Lundegard (1992):
which forms between loosely packed grains occupying an intergranular
(100 − OP) ∗ IGV volume between 30 and 35%, suggesting precipitation before significant
COPL = OP (1)
(100 − IGV) loss of porosity by compaction could occur. Early calcite cement was
reported in the Duva field and in other hydrocarbon fields in the North
CEPL = (OP − COPL) ∗
CEM
(2) sea (Azzam et al., 2022; Saigal and Bjørlykke, 1987; Worden et al.,
IGV 2020). These studies have suggested precipitation from the dissolution
of soluble carbonate phases such as aragonite and high Mg calcite. Minor
where COPL is the compaction porosity loss, CEPL is the cementation

14
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 11. Chlorite distribution in the Agat Fm: (A) Microfacies 1: samples dominated by thin discontinuous chlorite coats and well-developed quartz overgrowth; (B)
Microfacies 2: samples dominated by thin to moderate continuous chlorite coats, quartz overgrowth is scarce; (C) Microfacies 3: samples dominated by continuous
chlorites coats showing bridges between detrital grains with common pore-filling chlorite; (D) Microfacies 4: samples dominated by pore-filling chlorites; (E–F)
Histogram showing the proportion of chlorite microfacies in the different Agat members and in the different sedimentary facies.

microcrystalline quartz is detected on the surface of some quartz grains, than 50 ◦ C (Block Vagle et al., 2003). In the Agat Fm, microcrystalline
forming a thin coating (1–5 μm). These microcrystals are absent at grain quartz closely associated with chlorite and relicts of HIMs, acted as a
contacts suggesting precipitation after the main phase of compaction. barrier against further coarsening of quartz overgrowth (Fig. 13F). Early
Microcrystalline quartz in sandstones is generally attributed to the diagenetic minerals also include framboidal pyrite and siderite, which
dissolution of biogenic silica (e.g., sponge spicules) (Hendry and Trewin, are closely associated with chlorite (Fig. 9B, C, G). The coexistence of
1995; Jahren and Ramm, 2009; Lima and De Ros, 2002; Worden et al., siderite, framboidal pyrite, and chlorite indicates formation in a
2012). They generally form at shallow depths, at a temperature of less reducing environment (Azzam et al., 2022). Framboidal pyrite is

15
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 12. (A) X-ray diffraction traces of randomly oriented clay fractions (<5 μm) for some representative samples from the different Agat members; (B) Compilation
of FTIR spectra of kaolin group minerals in the hydroxyl-stretching band region for some representative samples; (C) SEM image of blocky dickite crystals filling
primary porosity (sample 35/3–2 at 3707 m).

generally formed by bacterial reduction of sulfates (Folk, 2005). How­ ranging from ~85 ◦ C to ~105 ◦ C (Azzam et al., 2022). Bioclast and
ever, the amount of pyrite is low compared to siderite, indicating a low Ca-plagioclase are scarce in the Agat formation. Therefore, the main
concentration of sulfur and, thus, weak development of the source of calcite 2 is probably related to the dissolution of more soluble
sulfate-reducing zone in the Agat Fm. Under such conditions, reduced early calcite cements like calcite 1. Calcite 2 is followed by strong
iron will tend to form siderite and Fe-clays like berthierine (Azzam et al., feldspar dissolution, mainly affecting K-feldspar grains and the residual
2022; Camprubí and Canet, 2009; Longstaffe, 2003; Taylor and Mac­ plagioclase grains. This is confirmed by the fact that secondary pores
quaker, 2011). Early detrital clays likely transformed into berthierine at created by the dissolution of K-feldspars are rarely filled by calcite 2
shallow depths under reducing conditions, as suggested by the presence (Fig. 13B). This dissolution phase is associated with the precipitation of
of chlorite polytype Ib in the studied samples (see section 5.2.1) (Azzam kaolinite and quartz overgrowth which fills primary intergranular
et al., 2022; Beaufort et al., 2015). The final stage of early diagenesis was porosity (Figs. 8E and 13G). Fluid inclusions data in quartz overgrowth
characterized by the progressive chloritization of berthierine at tem­ indicate trapping temperatures ranging between 80 ◦ C and 121 ◦ C in
peratures close to 50 ◦ C (Beaufort et al., 2015). wells 35/3–2 and 35/3–4 (Petrea and Molenaar, 2014/VNG Norge in­
The middle to late diagenesis is marked by extensive cementation by ternal report). Traces of hydrocarbon were also reported in fluid in­
calcite 2 and by the dissolution of feldspar grains (largely K-feldspars) clusions in quartz overgrowth, suggesting that hydrocarbon
associated with pore-filling kaolinite/dickite, late quartz overgrowth emplacement started prior to or during quartz cementation. Based on
and anatase. Calcite 2 encloses all the early diagenetic events (chlorite, core descriptions, hydrocarbons are only detected in calcite-free sand­
siderite, and rarely quartz overgrowth), suggesting precipitation during stone intervals. This suggests that hydrocarbon emplacement occurred
late diagenesis. Fluid inclusion data obtained on similar calcites from the after calcite 2 cementation. Finally, bipyramidal anatase crystals are
nearby Duva field show that calcite 2 precipitated at a temperature seen enclosing earlier grain-coating radial chlorite (Fig. 13H), indicating

16
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 13. (A) photomicrograph of a sample (well 35/3–5, 3245 m) cemented by apparently calcite 1, note that siderite develops along calcite crystal borders
indicating that it postdates calcite 1 cement; (B) photomicrograph of a sample cemented by blocky calcite 2, which has a homogeneous texture. Note that calcite 2
does not fill secondary pores created by K-feldspar dissolution (C) photomicrograph in cross-polarized light showing calcite 2 replacing early dissolved plagioclase
grains; (D) SEM image showing calcite 2 postdating chlorite coats; (E) SEM image showing grain-coating microcrystalline quartz which has seeded upon a detrital
quartz grain; (F) SEM image of quartz overgrowth engulfing early microcrystalline quartz and chlorite grain coating; (G) SEM image of porosity filled by kaolinite/

17
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

dickite. Note the presence of quartz overgrowth, which seems to be stopped against kaolinite. (H) SEM image of anatase crystal showing pyramidal shape engulfing
chlorite crystals.

precipitation during late diagenesis. Anatase usually results from the Fe-grains like biotite, glauconite, goethite, Fe-ooids and other
degradation and re-precipitation of TiO2 from Ti-bearing minerals such ferromagnesian grains (Azzam et al., 2022; Beaufort et al., 2015;
as ilmenite, rutile, or biotite. Despite being widely regarded as immo­ Virolle et al., 2020, 2022; Worden et al., 2020). Berthierine can
bile, many studies have proved that Ti can be mobilized and dissolved in then be transformed into Fe-chlorite with increasing temperature
fluids rich in organic acids (Parnell, 2004; Pe-Piper et al., 2011; Schulz through solid-state transformation (Beaufort et al., 2015).
et al., 2016). Hydrocarbon emplacement in the Agat Fm during late (3) The third process is the precipitation of chlorites directly from
diagenesis has probably led to the dissolution of Ti-bearing minerals, pore fluids. This process was proposed in sandstone associated
allowing the crystallization of anatase. with evaporites and carbonates. Chlorites can be formed directly
from magnesian brines derived from the alteration of evaporitic/
carbonate sequences (Rossel, 1982; Seeman, 1982; Goodchild
5.2. Origin of authigenic clay minerals
and Whitaker, 1986; Janks et al., 1992; Purvis, 1992; Gaupp
et al., 1993; Searl, 1994). This process was also suggested by
5.2.1. Chlorite
Anjos et al. (2009) and Umar et al. (2011a) in volcaniclastic
sandstones, where chlorites can directly originate from the
5.2.1.1. Review of chlorite formation in deep-water sandstones. The alteration of volcaniclastic rock fragments. However, supersatu­
occurrence of chlorite in turbidite reservoirs was only reported in a few ration of the solutions with respect to chlorite is not likely in
examples worldwide (Table 5). From these examples, multiple sources low-temperature diagenetic systems (T < 70 ◦ C) characterized by
and formation mechanisms of chlorite have been proposed, and the very slow heating and compaction rates (Beaufort et al., 2015).
effectiveness of chlorite in preserving reservoir quality was highly var­
iable. Indeed, the development of chlorite in turbidite deposits can be
5.2.1.2. Origin and source of chlorite in the Agat Formation. Previous
influenced by many parameters, such as turbidite composition and
studies on chlorite authigenesis in the Agat Fm in the Duva field (block
texture, depositional environment, and sediment provenance. These
36/7) demonstrated that chlorite results from the solid-state trans­
parameters will not only control the presence of chlorite but will also
formation of a berthierine precursor (Azzam et al., 2022). The authors
control their chemical composition and how they are distributed in deep
showed that berthierine originated from the alteration of Fe-grains
marine siliciclastic reservoirs. Based on Table 5, three main processes
(Fe-ooids, Fe-illite, Glauconites) and fine fraction materials rich in Fe
have been proposed for chlorite authigenesis in deep marine settings:
originally formed at the shelf and transported into deep marine settings
by turbidity currents. Chlorite in the Duva field is extremely rich in Fe
(1) The first process implies the mechanical infiltration of chlorite
(XFe >0.88), forming in conjunction with phosphatic minerals like
clay precursors, probably by dewatering processes (Bello et al.,
apatite. However, the mineralogy of the Agat Fm in the Agat field is very
2021; Houseknecht and Ross, 1992; Porten et al., 2019). The
different from that of the Duva field. Fe-grains are dominated by biotite,
nature of the clay precursor depends on the geological settings of
muscovite, and rare glauconite grains. Contrary to the Duva field, the
the study area, particularly on the composition of the bedrock
absence of Fe-ooids, Fe-illite and the scarcity of glauconite in the Agat
and the climate conditions (Akinlotan et al., 2022).
field indicates sedimentation under different environmental conditions.
Tri-octahedral smectite (saponite) is a very common clay pre­
Polytype Ib (β = 90◦ ) is dominant in both the Agat field and the Duva
cursor for chlorite (mainly Mg-rich chlorite) in sediment depos­
field, supporting a transformation from a berthierine precursor (Azzam
ited in semi-arid to evaporitic conditions (Anjos et al., 2009; Link
et al., 2022). However, chlorite in the Agat field is rich in magnesium
and Welton, 1982a; Spain, 1992). Saponite can transform into
(XFe ~ 0.71–0.75), suggesting that the chemistry of the fluids and the
chlorite via a chlorite/smectite (corrensite) mixed layer with
nature of clay precursors were different from that of the Duva field.
increasing depth and temperature. On the other hand, dioctahe­
Based on XRD analysis, the studied samples do not contain any traces of
dral smectites are likely to evolve into I/S mixed layers and ul­
corrensite, therefore excluding a transformation from tri-octahedral
timately illite if K+ is available during diagenesis (Abd Elmola
smectite (saponite). If corrensite was completely transformed into
et al., 2020; Bello et al., 2021; Chang et al., 1986).
chlorite at the present burial temperature (105 ◦ C–129 ◦ C), then poly­
(2) The second process is the replacement of an inherited 7 Å, Fe-rich
type IIb should be dominant (Beaufort et al., 2015). However, IIb pol­
green serpentine-like clay mineral (e.g., berthierine) by Fe-rich
ytype is a minor component in the studied samples, indicating that only
chlorite (14 Å) through mixed-layer serpentine-like/chlorite.
trace amounts of Fe-chlorite were derived from corrensite or probably
Green clay minerals generally form in shallow marine settings
had a detrital origin. In order to grow at low temperatures (<50 ◦ C),
(estuarine, deltaic, or inner shelf settings) under warm/tropical
chlorite must overcome considerable activation energy of nucleation,
climate conditions (Ehrenberg, 1993; Ku and Walter, 2003; Odin,
and the pore fluids must be extremely supersaturated with respect to
1988). In sediments, they occur in different habits: ooids,
chlorite (Worden et al., 2020). However, in the presence of metastable
pre-compactional coatings, and as matrix/pore-filling (Odin,
berthierine, chlorite can easily form with increasing temperature
1988). These clay minerals are regarded as metastable and tend
(>50 ◦ C) through solid-state transformation (Beaufort et al., 2015 and
to evolve into Fe-chlorite (chamosite) upon burial (Beaufort
references therein).
et al., 2015; Hillier, 1994; Odin, 1988). Chloritization of ber­
The formation of berthierine in the Agat Fm begins with the disso­
thierine has been proposed as the origin of chlorite in many North
lution of early detrital clay precursors. Petrographic observation shows
Sea reservoirs (Azzam et al., 2022; Ehrenberg, 1993). Berthierine
that chlorite replaces kaolinite in many of the studied samples (Fig. 8G
in turbidites can be either detrital, forming at the shelf in shallow
and H, and 7G, H). In addition, many samples contain fibrous mixed
marine sands and subsequently remobilized into deep settings by
layer clays composed of illite and HIMs, forming coatings around
turbidity currents, or developing in-situ by replacing detrital
detrital grains. Chlorite appears to develop on these coated grains
Fe-grains/clays in the presence of dissolved ferrous iron under
(Fig. 8A–B, C, D, and 9H). The composition of these mixed layer clays is
reducing conditions. Bhattacharyya (1983) and Virolle et al.
close to detrital clay coats that are found in the Lower Cretaceous
(2022) demonstrated that kaolinite, smectite, and Fe-rich inter­
shallow marine sandstones of the Wealden Group in the Paris Basin
layered clay minerals could transform into berthierine if ferrous
(Virolle et al., 2022). The Wealden sandstones are rich in pedogenetic
iron is available. Iron can be supplied from the alteration of

18
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 14. Cross-plot of permeability versus point counted porosity (Macroporosity) coded with (A) Agat members, (B) chlorite microfacies, and (C) sedimentary facies;
(D) Cross-plot of secondary porosity versus kaolinite abundance; (E) Plot of intergranular volume (IGV) versus intergranular mineral cement (CEM) coded with Agat
members (diagram after Houseknecht, 1987).

19
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 15. Burial history and paragenetic sequence of the major diagenetic events in the Agat Formation.

clays like Fe-rich Kaolinite and hydroxy-interlayered clay minerals (e.g., the Lower Cretaceous (Lidmar-Bergstrom et al., 1997). Microbial ac­
hydroxy-interlayered smectite, hydroxy-interlayered vermiculite). Ac­ tivity during early diagenesis played a critical role in maintaining
cording to Virolle et al. (2022), these clays are the main precursors for reducing conditions, allowing the release of Fe2+ from the dissolution of
berthierine growth during early diagenesis. Detrital clays can transform Fe-bearing grains (biotite and Ferric detrital clays such as HIMs), leading
into berthierine through dissolution-recrystallization processes under to the nucleation and growth of berthierine (Azzam et al., 2022; Mur­
reducing conditions (Bourdelle et al., 2021; Mosser et al., 2010; Rivard akami et al., 2004; Sugimori et al., 2009; Virolle et al., 2022). However,
et al., 2013). The Wealden sandstones in the Paris Basin offer a good low concentrations of sulfate and aqueous HCO3- are necessary to
analog to the Agat Fm at a shallow depth. Detrital clays in the Agat Fm develop berthierine at the expense of pyrite and siderite (Azzam et al.,
likely have a pedogenetic origin (kaolinitic and hydroxy interlayered 2022). With increasing temperature (>50 ◦ C), berthierine can transform
smectite/vermiculite clays) related to the weathering of soil profiles into chlorite through solid-state transformation. This transformation is
developed under a humid climate on the Norwegian mainland during nearly complete at the present-day reservoir temperature

20
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

(105 ◦ C–129 ◦ C). promoted microfacies 1 in some parts of the Agat 70, probably because
the initial clay content is low and the force of escaping fluids was high
5.2.1.3. Mechanism of chlorite clay precursors incorporation in turbidites. enough to remove all clay particles from the pore network, leaving some
The incorporation of chlorite clay precursors in sandstone, in the form of residual clay coats around the sand grains. Chlorite microfacies 2 and 3
coating or pore-filling, can be related to several mechanisms depending are mainly observed in weakly confined channel complex deposits (Agat
on the depositional environment. Detrital clays can be incorporated into 80) but also in Agat 70. These deposits are dominated by a complex of
fluvial and alluvial sandstones via infiltration by muddy waters (Moraes massive, fluidized, and stratified sands with common mudstone in­
and De Ros, 1990; Tang et al., 2018). In shallow marine and deltaic terbeds. The force of escaping fluids was moderate in these deposits,
settings, detrital clays can be formed by the rapid mineralization of only transporting clays that are in the center of the pore network while
Fe-rich clays near major river mouths (Azzam et al., 2022; Ehrenberg, leaving those along grain boundaries. The texture of Agat 80 is slightly
1993; Griffiths et al., 2021; Hansen et al., 2021; Worden et al., 2020). In finer than Agat 70, probably ensuring a more even distribution of
estuarine settings, the biological activity of diatoms can produce bio­ detrital clays around the sand grains, promoting, therefore, microfacies
films that act as a glue between detrital clays and sand grains (Duteil 2. Microfacies 4 is observed in all deposits but mainly in the distal lobe
et al., 2020; Virolle et al., 2020, 2021; Wooldridge et al., 2017; Worden deposits (Agat 50) and in the mud-rich channelized system (Agat 100).
et al., 2020). The distribution of chlorite in the Agat Fm is very het­ The distal lobe deposits of the Agat 50 are mainly composed of banded
erogeneous. Some samples are dominated by pore-filling chlorites (Bst) and heterolithic (Hst) facies. These deposits are generally less
(Microfacies 4), whereas others are dominated by chlorite in the form of permeable (poorly drained) than lobe and channel deposits, therefore,
coating (Microfacies 2 and 3). Chlorite coating coverage is good on escaping fluids were less efficient in redistributing the clay matrix. The
average (>80%). However, some samples show poor coverage (<40%) channel complexes of the Agat 100 are particularly rich in silt and mud
by chlorite and are cemented by quartz (Microfacies 1). Such complex and are poorly drained, promoting, therefore, microfacies 4.
distribution of chlorite can be either related to heritage, in which Our results show that chlorite distribution is directly linked to the
detrital clays are incorporated into the sediment prior to their deposition depositional environment (lobe, channel, or levee). Channel deposits
in deep marine settings or infiltrated into the sandstone during or after and lobe deposits are both prominent settings for the development of
deposition. continuous chlorite coats in the Agat Fm. However, discontinuous
In the Duva field, the Agat sandstones are massive, structureless, chlorite coats are more commonly encountered in amalgamated lobe
coarse, and poorly sorted, transported in a confined turbidite system deposits. Finally, distal lobe deposits, levee deposits, and mud-rich
(Azzam et al., 2022; Hansen et al., 2021). Based on Azzam et al. (2022), channel complexes are prominent settings for pore-filling chlorites.
chlorite coat precursors in this sandstone are inherited from shallow
marine environments. However, in the Agat field, the sandstones are 5.2.2. Kaolinite/dickite
fine grained and well sorted, showing complex sedimentary facies dis­ In the Agat Fm, a fair correlation is found between secondary
tribution dominated by massive fluidized and cross-stratified sandstone. porosity created by feldspar dissolution and kaolinite. This indicates that
The very fine texture and the good sorting in the Agat field compared to feldspar dissolution was the main source of Al for kaolinite precipitation.
the Duva field suggests a longer transportation distance from the sedi­ However, it should be noted that feldspar grains are generally preserved
ment sources to the site of deposition (Bouma, 2000). In such a context, in the samples cemented by calcite 2. Kaolinite is scarce in these samples
the abrasion of detrital clay coating will be strong, and the preservation (<0.5%), suggesting that feldspar dissolution in permeable zones was
rate of detrital clay coats will be low (Verhagen et al., 2020). In addition, minor during early diagenesis and became extensive during late
the scarcity of Fe-grains such as glauconite, Fe-ooid, and peloids sug­ diagenesis.
gests that low Fe-mineralization was taking place on the shelf (Griffiths Many experimental studies have demonstrated that acidic pore wa­
et al., 2021). Therefore, chlorite coat distribution in the Agat field is less ters are prominent for the dissolution of feldspar (Helgeson et al., 1984;
likely related to heritage. Matteo et al., 2018; Zhang et al., 2009). Some studies suggest that
Porten et al. (2019) demonstrated that mechanical infiltration of meteoric water can circulate deeply into the sediments and promote
detrital clays by dewatering processes has the potential to form dissolution reactions through the production of carbonic acids (Bjor­
continuous clay coats in turbidite deposits. The sedimentary structures lykke, 1983; Mansurbeg et al., 2020; Yuan et al., 2017). Meteoric-water
of the Agat Fm in the Agat field are very similar to those described by incursion usually happens during early diagenesis or telogenesis, when
Porten et al. (2019) from the deep-marine sandstones of the Springar sediments are shallowly buried (Mansurbeg et al., 2012, 2020). How­
Formation (Vøring Basin, Norwegian Sea). Dewatering occurs when ever, the Agat Fm has deposited at least 60–80 km from the paleoshore
fluids trapped in sediments are over-pressurized due to compaction/­ line (Copestake et al., 2003), making meteoric incursion during early
consolidation. As a result, the fluids will escape out of the sediment, diagenesis unlikely. Other authors have proposed that organic acids
causing significant changes in the grain fabric. Escaping fluids can produced by the maturation of sedimentary organic matter can increase
transport or remove clay particles and redistribute them as continuous the acidity of porewater and cause mineral dissolution (Ehrenberg,
clay coats in well-drained zones or fill the pores systems in a poorly 1991; Schmidt and Mcdonald, 1979; Surdam et al., 1989; Yuan et al.,
drained zone. As suggested by Porten et al. (2019), the formation of well 2015, 2019). Organic acid-rich fluids are usually produced when the
continuous clay coat (case of chlorite microfacies 2 and 3) requires kerogen enters the oil window and reaches peak expulsion at a tem­
continuous percolation of pore water throughout the sediment. How­ perature close to 100 ◦ C (Andresen et al., 1994). This process is more
ever, the force of escaping fluids must be moderate to preserve clay coats probable in our case since it is synchronous with the timing of feldspar
along grain borders. The formation of discontinuous clay coats dissolution in the Agat Fm.
(microfacies 1) is favored in high permeability sands, where the force of The reaction of K-feldspar dissolution and kaolinite precipitation can
escaping fluids is strong enough to remove clays from the pore network be written as follow:
but also those along grain borders. On the other hand, sediments
K–AlSi3O8 (K-feldspar) + H+ + 0.5H2O = 0.5Al2Si2O5 (OH)4 (kaolinite)+
dominated by pore-filling clays (microfacies 4) are favored in poorly
2SiO2 (aq) + K+ (1)
drained, fine-grained sandstone and clay-rich deposits.
In the Agat Fm, chlorite microfacies 1 is mainly observed in proximal One may note that this reaction is associated with the release of
lobe deposits (Agat 70) characterized by massive-fluidized sandstones aqueous SiO2, which can form authigenic quartz if the fluids become
(MFst facies). Amalgamated lobes are generally thick, massive, and supersaturated with respect to quartz. The precipitation temperature of
fluidized, with small intercalations with mudstone. Dewatering has quartz overgrowth obtained from fluid inclusions (85 ◦ C–120 ◦ C)

21
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

(Petrea and Molenaar, 2014) is within the window of K-feldspar disso­ Approximately 8% of the total volume of Agat 70 is cemented by calcite,
lution reactions, suggesting that kaolinite precipitation was associated significantly reducing the reservoir quality.
with quartz cementation. It is important to note that kaolinite rarely In the weakly confined channels (Agat 80), the reservoir quality is a
appears to fill secondary pores created by feldspar dissolution. Kaolinite function of chlorite distribution and calcite cementation. The Agat 80
precipitation generally occurs near dissolved grains or within isolated has more heterogeneity in terms of sedimentary facies and structure
intergranular pores. This indicates that the dissolution by-products (e.g., showing low to moderate amounts of pore-filling chlorite. Quartz
Na+, K+, Al3+, SiO2(aq)) are transferred into solution and then cement is scarce due to the presence of chlorite coats that completely
distributed across the intergranular pore system to be finally precipi­ cover most of the sand grains. Calcite cement is pervasive in the Agat 80,
tated as secondary minerals. In a closed diagenetic system, the precipi­ with about 20% of the total volume being cemented by calcite. This
tation of authigenic clays (e.g., kaolinite) will help maintain the solution strongly reduces the permeability and connectivity of the sand bodies.
undersaturated with respect to feldspar by continuously removing the The dissolution of feldspar was associated with the precipitation of
solutes from the aqueous solution, hence allowing more feldspar to be kaolinite, with limited impact on the reservoir quality.
dissolved (Xiao et al., 2018). In the Duva field, the extensive dissolution The Agat 100 has a similar reservoir quality to the Agat 50 despite
of feldspar is rarely associated with the precipitation secondary min­ the difference in the deposition environment. The reservoir quality was
erals, suggesting that dissolution happens in a relatively open diagenetic greatly impacted by the presence of a high amount of pore-filling
system with high fluid flow, probably related to faulting (Azzam et al., chlorite and by calcite cementation. The increase of pore-filling chlo­
2022). The extensive precipitation of kaolinite associated with quartz rite in the Agat 100 is mainly related to the high volumes of detrital clays
cement in the Agat field indicates a more closed diagenetic system with originally present in these sediments. Quartz overgrowth is scarce due to
low fluid flows. the presence of pore-filling chlorite, which prevents extensive quartz
Dickite is detected along kaolinite in the deepest part of well 35/3–2 cementation while increasing the volume of ineffective porosity. Like
at a depth close to 3700 m (T~129 ◦ C) but appears as a minor compo­ Agat 80, the dissolution of feldspar occurs in a relatively closed diage­
nent below 3500 m. Dickite is the stable polytype of kaolinite, it pro­ netic system associated with kaolinite precipitation. Calcite cementation
gressively replaces kaolinite with increasing depth and temperature. is pervasive in the Agat 100, with about 15% of the total volume being
Kaolinite is nearly completely conversed into dickite at a depth greater cemented by calcite.
than 4000 m (Beaufort et al., 1998; Lanson et al., 2002). The presence of Distal lobe deposits are poor exploration targets for hydrocarbon
dickite with kaolinite in sample 3700 m could be explained by its exploration or CO2 storage. The sandstone is relatively thin (meter to a
exposition to a higher temperature and burial depth compared to the few meters in thickness), and the reservoir quality is low due to the high
other studied samples. ductile content filling the intergranular pores. Lobe deposits are very
good exploration targets since they are massive, thick, and have overall
5.3. Reservoir quality evolution pathways of the Agat Formation better connectivity. However, amalgamated lobes have more risks of
developing quartz overgrowth due to the discontinuous chlorite coating.
A conceptual diagenetic model of the main Agat members is pro­ Weakly confined channels are good exploration targets since they show
posed in Fig. 16. The reservoir quality of the Agat Fm in block 35/3–5 overall good reservoir quality. The presence of chlorite coats has pro­
was controlled by several eogenetic and mesogenetic alterations mainly tected the reservoir against quartz overgrowth during burial. However,
related to the initial mineralogy and depositional environment. All the calcite cementation is important in these sandstones, which significantly
Agat members have experienced similar diagenetic reactions. However, reduces the connectivity of the sand bodies. Finally, the channel system
the intensity of these reactions varies significantly as a function of the of Agat 100 is a mild exploration target due to the strong intercalation
depositional environment. with mudstone layers and the high amount of pore-filling clays (mainly
In the distal lobe deposits (Agat 50), the reservoir quality was mainly chlorite and kaolinite), and the extensive cementation by calcite.
influenced by mechanical compaction and by the presence of a large
amount of detrital clays in the pore network. Late diagenesis has a 6. Conclusion
limited effect on reservoir quality. The dissolution of feldspar grains has
created secondary pores, which increased the overall porosity. However, The Agat Fm records the progressive filling in continental slope
this dissolution was associated with kaolinite precipitation in the pri­ environment of sub-basins associated with rift systems during the Lower
mary intergranular pores, which reduced permeability. Quartz over­ Cretaceous. The Agat section comprises four principal members that
growth developed locally where chlorite coats are absent, slightly reflect different depositional environments and sandstone types: Agat 50
reducing permeability. Siderites are very common in Agat 50, but they (distal lobe environment), characterized by heterolithic and banded
mainly replace biotite grains, therefore, they have limited impact on the sandstones, Agat 70 (proximal lobe deposits) and Agat 80 (weakly
reservoir quality. confined channels), formed by high-density turbidity currents and
In proximal lobe deposits (Agat 70), the reservoir quality is mainly a dewatering processes, and Agat 100 (amalgamated mud-rich channels)
function of chlorite coat distribution. Chlorite coats are thin in the Agat characterized by a complex of fluidized, stratified, medium to coarse
70 (traces to a few micrometers), showing heterogeneous distribution sandstones. Sandstone diagenesis is mainly influenced by the authi­
around detrital grains (grain coverage of 40% up to 100%). The Agat 70 genesis of chlorite and kaolinite, in addition to calcite cementation and
is generally massive and clay-poor, allowing dewatering fluids to flow feldspar dissolution. Chlorite affects the reservoir quality both positively
faster through pore networks. The velocity of escaping fluids was high in (as grain coats) and negatively (as pore fillings). Samples with well-
some parts of the Agat 70, leading to discontinuous chlorite coats. When developed chlorite coats display the highest reservoir quality, while
chlorite grain coverage is low, extensive quartz cementation forms samples with discontinuous chlorite coats or high amounts of pore-
during late diagenesis, which significantly reduces the primary porosity filling chlorite show poor reservoir quality. The distribution of chlorite
(5–7%). On the other hand, samples with good chlorite grain coverage in the Agat Fm is directly related to the depositional environment and
show the highest porosity (12–15%). The presence of a thin continuous the intensity of the dewatering processes. Well-developed chlorite coats
chlorite coating in these samples has retained more intergranular mac­ are found in amalgamated lobe deposits and weakly confined channels.
ropores than in those with partial chlorite coatings. However, the Discontinuous chlorite coats are mainly found in proximal lobe deposits
presence of mica in the Agat 70 has reduced to some degree the porosity associated with high-velocity escaping fluids during dewatering. On the
due to compaction upon burial. The dissolution of feldspar has a limited other hand, distal lobe deposits, levee deposits, and mud-rich channels
effect on the reservoir quality. The created secondary pores are associ­ are dominated by pore-filling chlorite related to the poorly drained
ated with kaolinite precipitation, which reduces permeability. character of these deposits. The origin of chlorite in the Agat Fm is

22
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Fig. 16. Conceptual model of the diagenetic evolution pathways of the main Agat members in calcite-free intervals.

23
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

mainly related to the replacement of early detrital clays, mainly Fe- interests or personal relationships that could have appeared to influence
bearing kaolinite and various types of Al–Fe-rich hydroxy interlayered the work reported in this paper.
clay minerals (HIMs) and biotite through an intermediate berthierine
precursor. The chloritization reaction was promoted through the Data availability
establishment of reducing conditions during early diagenesis and
through the dissolution of biotite and Fe3+- bearing clays. The late Data will be made available on request.
diagenesis was associated with the extensive dissolution of feldspar
grains. This reaction occurs in a relatively closed diagenetic system Acknowledgments
allowing the precipitation of kaolinite/dickite and quartz cement. The
origin of the dissolution was related to the circulation of acidic fluids This study was funded by Neptune Energy. The authors thank the
during late diagenesis, probably derived from the maturation of organic- Norwegian Petroleum Directorate (NPD) for providing access to core
rich shales. The Agat Fm offers an excellent example of how a deposi­ material from the Agat field. The authors would also like to thank Serge
tional environment can control the distribution of chlorite and reservoir Miska and Olivier Dufaure for assistance with XRD analysis and Julius
quality in turbidite reservoirs. The similarity of the Agat Fm with other Nouet for assistance with SEM. Paris-Saclay University benefits from the
turbidite reservoirs in the North Sea suggests that it can be considered as Schlumberger Software Donation Program. We are grateful to Schlum­
an analog for predicting chlorite coat emplacement in deep marine berger for the PetroMod software license. We are also grateful to the
settings. anonymous reviewers for their constructive comments and suggestions
that improved this paper. We would also like to thank the editor and the
Declaration of competing interest associate editor Evren Unsal.

The authors declare that they have no known competing financial

Appendix A
Table 2
Point counting results of detrital mineralogy, authigenic mineralogy, and porosity for the main Agat members from wells 35/3-7 S and 35/3-5.

Well 35/3-7 S 35/3-5

Agat member Agat 80 (n = 10) Agat 70 (n = 4) Agat 100 (n = 6)

Depth(mDD) 3620 m–3659 m 3662 m–3669 m 3231 m–3283 m

Facies association MFst MFst, SRst MFst, MFsst, SRst

Min Max Average Min Max Average Min Max Average

Detrital mineralogy
Quartz monocrystalline 30.50 49.00 45.79 49.50 54.50 52.50 21.36 42.50 32.40
Quartz polycrystalline 5.00 11.00 8.42 7.00 12.00 9.25 13.28 27.67 18.21
K-feldspar 8.50 19.00 13.17 4.5 11.5 8.38 3.14 7.43 5.16
Plagioclase 1.00 6.00 3.50 3.5 10 7.25 1.28 3.03 2.11
Lithics 0.50 2.00 1.13 0.50 1.50 0.88 0.83 11.62 7.32
Heavy minerals 0.00 1.50 0.58 0.00 1.50 0.50 0.00 0.30 0.10
Muscovite 0.00 5.00 2.04 0.50 1.50 0.88 0.10 1.67 0.72
Biotite 0.50 8.00 3.58 0.50 3.50 1.38 1.24 8.33 3.67
Glauconite 1.00 9.00 3.29 0.00 2.50 1.38 0.51 3.33 1.66
Organic fragments 0.00 8.00 1.83 0.00 0.00 0.00 0.00 1.67 0.74
Detrital clay pseudo-matrix 0.50 28.00 3.42 0.50 4.00 2.38 0.10 21.36 4.53
Authigenic mineralogy
Secondary quartz 0.00 2.00 0.25 0.00 1.00 0.33 0.00 3.73 1.53
Calcite cement 0.00 16.00 3.96 0.00 3.00 0.88 0.00 17.67 6.33
Dolomite 0.00 0.00 0.00 0.00 1.00 0.33 0.00 0.00 0.00
Siderite 0.00 2.00 0.45 0.00 0.00 0.00 0.81 8.84 3.63
Pyrite 0.00 0.50 0.13 0.00 0.00 0.00 0.10 0.10 0.10
Titanium oxides 0.00 2.00 0.70 1.00 2.50 1.75 0.20 0.70 0.50
Kaolinite 0.00 3.50 2.05 2.00 5.00 3.50 0.00 3.73 1.53
Chlorite 0.00 12.00 4.68 2.00 3.50 2.50 1.26 14.47 6.91
Chlorite coat 0.00 10.00 3.43 2.00 3.50 2.50 0.42 2.04 1.57
Pore filling chlorite 0.00 1.99 1.25 0 0 0.00 0.84 12.43 5.34
Macroporosity 0.50 8.50 2.70 0.00 12.50 6.13 0.49 11.66 6.05
Primary porosity 0.00 7.00 2.15 0.00 12.00 5.25 0.00 8.33 3.15
Secondary porosity 0.50 2.50 1.25 0.50 3.00 1.63 0.49 5.47 2.90
Average grain size 166.00 196.00 180.00 210.00 215.00 213.00 208.90 449.30 345.68
Sandstone classification Subarkose Subarkose Feldspathic litharenite, Sublitharenite, Subarkose

24
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Table 3
Point counting results of detrital mineralogy, authigenic mineralogy, and porosity for the main Agat members from wells 35/3–2 and 35/3–4.

WELL 35/3-2 35/3-4

Agat member Agat 50 (n = 3) Agat 70 (n = 3) Agat 70 (n = 2) Agat 80 (n = 3) Agat 100 (n =


1)

Depth(mDD) 3692 m–3707 m 3600 m–3625 m 3500 m–3524 m 3447 m–3480 m 3410 m

Facies association Bst, Hst MFst MFst, Hst MFst Hst

Min Max Average Min Max Average Min Max Average Min Max Average

Detrital mineralogy
Quartz monocrystalline 23.02 55.12 34.34 32.61 37.25 34.93 46.73 49.52 48.13 29.39 51.75 38.09 31.86
Quartz polycrystalline 11.90 31.84 18.78 14.22 17.83 16.03 2.00 5.03 3.52 10.00 14.88 11.88 8.85
K-Feldspar 0.50 2.94 1.35 4.05 4.83 4.44 3.81 4.72 4.27 2.70 9.74 6.74 5.08
Plagioclase 1.50 5.00 3.67 3.50 4.21 3.86 4.29 5.33 4.81 1.12 4.05 2.80 3.05
Lithics 2.36 11.94 5.95 3.43 5.22 4.33 0.50 0.50 0.50 1.75 4.96 2.86 3.98
Heavy minerals – – – 0.50 1.50 0.80 1.74 1.92 1.84 0.80 1.20 0.90 0.50
Muscovite 1.97 3.97 2.31 0.43 1.47 0.95 1.01 1.43 1.22 0.10 2.96 1.67 0.88
Biotite 1.49 19.05 7.37 2.94 4.35 3.65 4.52 5.71 5.12 2.07 3.50 3.04 8.41
Glauconite 0.00 1.98 0.83 0.53 1.42 0.97 0.5 2.86 1.68 0.41 3.15 1.84 0.88
Organic fragments 1.00 4.72 2.7 0.96 2.45 1.71 2.02 3.29 2.65 0.10 0.75 0.23 0.44
Detrital clay pseudo- 0.10 1.2 0.7 0.10 0.50 0.30 0.50 1.01 0.75 – – – 0.5
matrix
Authigenic mineralogy
Secondary quartz 2.56 5.47 4.40 0.00 7.10 1.35 1.01 4.76 2.89 0.00 2.10 0.53 0.00
Calcite cement 0.40 0.50 0.45 0.60 25.00 15.03 0.00 0.48 0.24 5.59 36.73 22.92 34.00
Dolomite 0.00 0.00 0.00 0.00 0.00 0.00 – – – – – – 0.00
Siderite 2.49 13.49 8.21 0.00 1.48 0.51 1.01 2.80 1.91 0.00 1.40 0.44 0.50
Pyrite 0.05 0.20 0.10 0.05 0.15 0.10 0.05 0.10 0.07 0.00 0.15 0.05 0.50
Titanium oxides 0.58 0.73 0.65 0.73 0.96 0.81 0.87 0.96 0.92 0.58 0.68 0.62 1.12
Kaolinite 1.94 4.23 3.25 0.00 2.17 1.09 2.51 4.29 3.40 0.00 2.45 1.08 0.00
Chlorite 11.29 12.49 11.76 0.98 3.47 2.23 7.62 9.55 8.59 1.65 7.40 3.91 1.77
Chlorite coat 1.49 3.70 2.44 0.98 1.30 1.14 3.33 3.52 3.43 1.65 4.44 2.56 1.77
Pore filling chlorite 9.80 11.00 10.27 0.00 2.17 1.09 4.29 6.03 5.16 2.45 2.96 2.71 0.00
Macroporosity 3.00 3.18 3.13 0.99 6.72 3.12 5.7 13.57 9.635 0.00 4.90 1.63 0.00
Primary porosity 1.94 3.57 2.62 0.00 4.35 2.03 3.8 10.28 7.04 0.00 4.20 1.14 0.00
Secondary porosity 0.40 2.36 1.09 0.37 2.37 1.09 1.9 3.29 2.595 0.00 1.11 0.45 0.00
Average grain size 205.00 268.00 227.00 225.00 261.00 244.00 204.40 220.50 212.45 175.90 201.20 191.90 177.50
Sandstone classification Subarkose Subarkose Subarkose Subarkose, lithic arkose Subarkose

Table 4
Average chemical compositions of chlorite coats in the different Agat members compared to chlorite from the Duva field and detrital clays. Total iron has been
arbitrarily considered as ferric in detrital clays and as ferrous in authigenic chlorite. The structural formula is calculated based on 14 oxygens for chlorite, 11 oxygens
for detrital clay coat, and 7 oxygens for kaolinite. Oct: octahedral occupancy, Int: interlayer charge, XFe= Fe/(Fe + Mg).

Agat 50 (N = 13) Agat 70 (N = 15) Agat 80 (N = 14) Agat 100 (N = 13) Duva field (N = 16) Detrital clay coat (N = 7) Detrital kaolinite (N = 5)

SiO2 33.5 31.2 29.7 32.0 32.2 47.1 46.1


Al2O3 25.3 24.7 23.5 23.9 22.3 21.6 34.7
FeO 33.4 35.3 38.0 36.2 39.9 0.0 0.0
Fe2O3 0.0 0.0 0.0 0.0 0.0 8.8 3.9
MgO 6.0 7.7 7.3 5.9 4.3 2.7 0.1
TiO2 0.1 0.0 0.0 0.2 0.0 0.5 0.0
MnO 0.0 0.0 0.0 0.0 0.0 0.0 0.1
Na2O 0.0 0.0 0.0 0.0 0.0 0.0 0.0
K2O 1.3 0.6 0.8 1.1 0.2 5.7 0.1
CaO 0.0 0.0 0.0 0.0 0.4 0.0 0.0
Si 3,3 3,2 2,9 3,2 2,9 3.4 3.2
Al total 2,9 2,6 2,7 2,8 2,5 1.9 2.8
Mg 0,7 1,0 1,1 0,8 0,6 0.3 0.0
Fe2+ 2,2 2,5 2,8 2,5 3,8 0.0 0.0
Fe3+ 0.0 0.0 0.0 0.0 0.0 0.5 0.2
Ti 0,0 0,0 0,0 0,0 0,0 0.0 0.0
Mn 0,0 0,0 0,0 0,0 0,0 0.0 0.0
Na 0,0 0,1 0,1 0,0 0,0 0.0 0.0
K 0,2 0,1 0,1 0,1 0,0 0.5 0.0
Ca 0,0 0,0 0,0 0,0 0,0 0.0 0.0
Oct 5,1 5,3 5,5 5,3 5,8 2.0 2.2
XFe 0,8 0,7 0,7 0,8 0,9 – –
Int 0.1 0.2 0.2 0.2 0.0 0.4 0.0

25
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Table 5
Examples of chlorite-bearing turbidite sandstone.

Location Formation Age Depth Type of Effectiveness of chlorite on Origin of chlorite References
chlorite reservoir quality
preservation

Arkoma basin, Atoka Carboniferous >3 km Fe- Positive Infiltrated clay precursor Houseknecht and Ross
USA Formation chlorite (?) (1992)
Delaware basin, Cherry canyon Permian 1.6–2.1 Fe- Negative Smectite precursor Spain (1992)
USA Formation km chlorite
Santa Ynez basin, Matilija Eocene Outcrop Fe–Mg- Negative Smectite precursor (Link and Welton,
USA Sandstone chlorite 1982b)
Norwegian Agat Formation Lower Cretaceous >2.5 km Fe- Positive/Negative Berthierine precursor (Azzam et al., 2022;
Continental Shelf, chlorite Hansen et al., 2021)
Norway
Norwegian Tofte Early Jurassic >3,9 km Fe- Positive Berthierine precursor Ehrenberg (1993)
Continental Shelf, Formation chlorite
Norway
Norwegian Ty Formation Palaeocene >2.3 km Fe- Positive Infiltrated clay precursor Hurst and Buller
Continental Shelf, chlorite (?) (1984)
Norway
West of Shetland, Valia and Palaeocene ~3 km Fe–Mg- Positive ? Sullivan et al. (1999)
UK Sullom chlorite
Formation
Haymana basin, Haymana Upper Cretaceous >1.7 km Fe- Negative ? Gecer et al. (2019)
Turkey Formation chlorite
Kirthar Fold Belt, Pab Formation Maastrichtian >2.7 km Fe- Positive Alteration product of Umar et al. (2011b)
Pakistan (Upper Cretaceous) chlorite volcanic fragments
Santos Basin, Itajaí-Açu Upper Cretaceous >4 km Fe–Mg- Positive Alteration product of Anjos et al. (2009)
Brazil Formation chlorite volcanic fragments/
smectite precursor
Nelson Oil Field, UK Sele Formation Paleocene 2.7–2.8 Fe–Mg- Positive Smectite precursor Bello et al. (2022)
Central North Sea km chlorite

References temperature geological systems – a review. Clay Miner. 50, 497–523. https://doi.
org/10.1180/claymin.2015.050.4.06.
Bello, A.M., Jones, S., Gluyas, J., Acikalin, S., Cartigny, M., 2021. Role played by clay
Abd Elmola, A., Asaad, A., Patrier, P., Beaufort, D., Ballini, M., Descostes, M., 2020. Clay
content in controlling reservoir quality of submarine fan system, Forties Sandstone
mineral signatures of fault-related fluid flows in a sandstone reservoir: a case study
Member, Central Graben, North Sea. Mar. Petrol. Geol. 128 https://doi.org/
from the Teloua Formation, Tim Mersoї Basin, Niger. J. Afr. Earth Sci. 168, 103840
10.1016/j.marpetgeo.2021.105058.
https://doi.org/10.1016/j.jafrearsci.2020.103840.
Bello, A.M., Jones, S.J., Gluyas, J., Al-Ramadan, K., 2022. Impact of grain-coating clays
Ajdukiewicz, J.M., Larese, R.E., 2012. How clay grain coats inhibit quartz cement and
on porosity preservation in paleocene turbidite channel sandstones: nelson oil field,
preserve porosity in deeply buried sandstones: observations and experiments. Am.
UK central North Sea. Minerals 12. https://doi.org/10.3390/min12050555.
Assoc. Petrol. Geol. Bull. 96, 2091–2119. https://doi.org/10.1306/02211211075.
Ben-Awuah, J., Andriamiaja, S., Mijinyawa, A., Ali, A., Siddiqui, N.A., Adda, G.W., 2014.
Akinlotan, O.O., Moghalu, O.A., Hatter, S.J., Okunuwadje, S., Anquilano, L.,
Effect of some input parameters on 3D basin and petroleum systems modelling: a
Onwukwe, U., Haghani, S., Anyiam, O.A., Jolly, B.A., 2022. Clay mineral formation
case study of the Norwegian section of the northern North Sea. Res. J. Appl. Sci. Eng.
and transformation in non-marine environments and implications for Early
Technol. 7, 3746–3762. https://doi.org/10.19026/rjaset.7.730.
Cretaceous palaeoclimatic evolution: the Weald Basin, Southeast England.
Bhattacharyya, D.P., 1983. Origin of berthierine in ironstones. Clay Clay Miner. 31,
J. Palaeogeogr. 11, 387–409. https://doi.org/10.1016/j.jop.2022.04.002.
173–182. https://doi.org/10.1346/ccmn.1983.0310302.
Andresen, B., Throndsen, T., Barth, T., Bolstad, J., 1994. Thermal generation of carbon
Bjorlykke, K., 1983. Diagenetic reactions in sandstones. Sediment Diagenes 169–213.
dioxide and organic acids from different source rocks. Org. Geochem. 21,
https://doi.org/10.1007/978-94-009-7259-9_3.
1229–1242. https://doi.org/10.1016/0146-6380(94)90166-X.
Bloch, S., Lander, R.H., Bonnell, L., 2002. Anomalously high porosity and permeability in
Anjos, S.M.C., De Ros, L.F., Silva, C.M.A., 2009. Chlorite authigenesis and porosity
deeply buried sandstone reservoirs: origin and predictability. Am. Assoc. Petrol.
preservation in the upper cretaceous marine sandstones of the santos basin, offshore
Geol. Bull. 86, 301–328. https://doi.org/10.1306/61eedabc-173e-11d7-
eastern Brazil. Clay Miner. Cem. Sandstones 289–316. https://doi.org/10.1002/
8645000102c1865d.
9781444304336.ch13.
Block Vagle, G., Hurst, A., Dypvik, H., 2003. Origin of quartz cements in some sandstones
Azzam, F., Blaise, T., Patrier, P., Abd-Elmola, A., Beaufort, D., Portier, E., Brigaud, B.,
from the jurassic of the inner moray firth (UK). In: Sandstone Diagenesis,
Barbarand, J., Clerc, S., 2022. Diagenesis and reservoir quality evolution of the
pp. 375–389. https://doi.org/10.1002/9781444304459.ch22.
Lower Cretaceous turbidite sandstones of the Agat Formation (Norwegian North
Blott, S.J., Pye, K., 2001. GRADISTAT: a grain size distribution and statistics package for
Sea): impact of clay grain coating and carbonate cement. Mar. Petrol. Geol. 142,
the analysis of unconsolidated sediments. Earth Surf. Process. Landforms 26,
105768 https://doi.org/10.1016/j.marpetgeo.2022.105768.
1237–1248. https://doi.org/10.1002/ESP.261.
Baig, I., Faleide, J.I., Mondol, N.H., Jahren, J., 2019. Burial and exhumation history
Boe, R., Magnus, C., Terje, P., Rindstad, O., Bjorn, I., 2002. CO2 Point Sources and
controls on shale compaction and thermal maturity along the Norwegian North Sea
Subsurface Storage Capacities for CO2 in Aquifers in Norway.
basin margin areas. mar. pet. geol. 104, 61–85. https://doi.org/10.1016/j.
Bouma, A.H., 2000. Coarse-grained and fine-grained turbidite systems as end member
marpetgeo.2019.03.010.
models: applicability and dangers. Mar. Petrol. Geol. 17, 137–143. https://doi.org/
Bailey, S.W., 1988. X-ray diffraction identification of the polytypes of mica, serpentine,
10.1016/S0264-8172(99)00020-3.
and chlorite. Clay Clay Miner. 363 (36), 193–213. https://doi.org/10.1346/
Bourdelle, F., Dubois, M., Lloret, E., Durand, C., Addad, A., Bounoua, S., Ventalon, S.,
ccmn.1988.0360301.
Recourt, P., 2021. Kaolinite-to-Chlorite conversion from Si,Al-rich fluid-origin veins/
Barnhisel, R.I., Bertsch, P.M., 1989. Chlorites and hydroxy-interlayered vermiculite and
Fe-rich carboniferous shale interaction. Minerals 11. https://doi.org/10.3390/
smectite. In: Minerals in Soil Environments. SSSA Book Series, pp. 729–788. https://
min11080804.
doi.org/10.2136/sssabookser1.2ed.c15.
Brown, G., Brindley, G.W., 1980. X-Ray diffraction procedures for clay mineral
Beard, D.C., Weyl, P.K., 1973. Influence of texture on porosity and permeability of
identification. Cryst. Struct. Clay Miner. Their X-Ray Identif. 305–360. https://doi.
unconsolidated sand. am. assoc. pet. geol. bull. https://doi.org/10.1306/819a4272-
org/10.1180/MONO-5.5.
16c5-11d7-8645000102c1865d.
Bugge, T., Tveiten, B., Bäckström, S., 2001. The depositional history of the cretaceous in
Beaufort, D., Cassagnabere, A., Petit, S., Lanson, B., Berger, G., Lacharpagne, J.C.,
the northeastern North Sea. Nor. Pet. Soc. Spec. Publ. 10, 279–291. https://doi.org/
Johansen, H., 1998. Kaolinite-to-dickite reaction in sandstone reservoirs. Clay
10.1016/S0928-8937(01)80018-7.
Miner. 33, 297–316.
Burley, S., Worden, R., 2003. Sandstone Diagenesis: Recent and Ancient. International
Beaufort, D., Rigault, C., Billon, S., Billault, V., Inoue, A., Inoue, S., Patrier, P., 2015.
Association Of Sedimentologists Reprints. https://doi.org/10.1002/prot.20838.
Chlorite and chloritization processes through mixed-layer mineral series in low-

26
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Camprubí, A., Canet, C., 2009. Berthierine and chamosite hydrothermal: genetic guides analysis of the Kyrre Fm (Upper Cretaceous), Måløy Slope, offshore Norway. Mar.
in the Peña Colorada magnetite-bearing ore deposit, Mexico. Earth Planets Space 61, Petrol. Geol. 25, 663–680. https://doi.org/10.1016/j.marpetgeo.2007.12.007.
291–295. https://doi.org/10.1186/BF03352910. Jahren, J., Ramm, M., 2009. The porosity-preserving effects of microcrystalline quartz
Chang, H.K., Mackenzie, F.T., Schoonmaker, J., 1986. Comparisons between the coatings in arenitic sandstones: examples from the Norwegian continental shelf. In:
diagenesis of dioctahedral and trioctahedral smectite, Brazilian offshore basins. Clay Spec. Publ. Int. Ass. Sedimentol, pp. 271–280. https://doi.org/10.1002/
Clay Miner. https://doi.org/10.1346/CCMN.1986.0340408. 9781444304237.ch18.
Copestake, P., Sims, A.P., Crittenden, S., Hamar, G.P., Ineson, J.R., Rose, P.T., Jobe, Z.R., Lowe, D.R., Morris, W.R., 2012. Climbing-ripple successions in turbidite
Tringham, M.E., 2003. Lower cretaceous. In: Evans, D., Graham, C., Armour, A., systems: depositional environments, sedimentation rates and accumulation times.
Bathurst, P. (Eds.), The Millennium Atlas: Petroleum Geology of the Central and Sedimentology 59, 867–898. https://doi.org/10.1111/j.1365-3091.2011.01283.x.
Northern North Sea. Geological Society of London, pp. 191–211. Kneller, B., 1995. Beyond the turbidite paradigm: physical models for deposition of
Crittenden, S., Cole, J.M., Kirk, M.J., 1998. The distribution of Aptian sandstones in the turbidites and their implications for reservoir prediction. Geol. Soc. London, Spec.
central and northern North Sea (UK sector) - a lowstand systems tract play. Part 2: Publ. 94, 31–49. https://doi.org/10.1144/GSL.SP.1995.094.01.04.
distribution and exploration strategy. J. Petrol. Geol. 21, 187–211. https://doi.org/ Krumbein, W.C., S, L.L., 1963. Stratigraphy and Sedimentation, second ed. W.H.
10.1111/j.1747-5457.1998.tb00653.x. Freeman, San Francisco.
Dietel, J., Gröger-Trampe, J., Bertmer, M., Kaufhold, S., Ufer, K., Dohrmann, R., 2019. Ku, T.C.W., Walter, L.M., 2003. Syndepositional formation of Fe-rich clays in tropical
Crystal structure model development for soil clay minerals – I. Hydroxy-interlayered shelf sediments, San Blas Archipelago, Panama. Chem. Geol. 197, 197–213. https://
smectite (HIS) synthesized from bentonite. A multi-analytical study. Geoderma 347, doi.org/10.1016/S0009-2541(02)00363-7.
135–149. https://doi.org/10.1016/j.geoderma.2019.03.021. Lanson, B., Beaufort, D., Berger, G., Bauer, A., Cassagnabère, A., Meunier, A., 2002.
Dowey, P.J., Hodgson, D.M., Worden, R.H., 2012. Pre-requisites, processes, and Authigenic kaolin and illitic minerals during burial diagenesis of sandstones: a
prediction of chlorite grain coatings in petroleum reservoirs: a review of subsurface review. Clay Miner. 37, 1–22. https://doi.org/10.1180/0009855023710014.
examples. Mar. Petrol. Geol. 32, 63–75. https://doi.org/10.1016/j. Lidmar-Bergstrom, K., Olsson, S., Olvmo, M., 1997. Palaeosurfaces and associated
marpetgeo.2011.11.007. saprolites in southern Sweden. Palaeosurfaces recognition. Reconstr.
Duteil, T., Bourillot, R., Grégoire, B., Virolle, M., Brigaud, B., Nouet, J., Braissant, O., Palaeoenvironmental Interpret 95–124.
Portier, E., Féniès, H., Patrier, P., Gontier, E., Svahn, I., Visscher, P.T., 2020. Lien, T., Midtbø, R.E., Martinsen, O.J., 2006. Depositional facies and reservoir quality of
Experimental Formation of Clay-Coated Sand Grains Using Diatom Biofilm deep-marine sandstones in the Norwegian Sea. Nor. Geol. Tidsskr. 86, 71–92.
Exopolymers XX, pp. 1–6. https://doi.org/10.1130/g47418.1/5074182/g47418. Lima, R.D., De Ros, L.F., 2002. The role of depositional setting and diagenesis on the
pdf. reservoir quality of Devonian sandstones from the Solimões Basin, Brazilian
Ehrenberg, S.N., 1993. Preservation of anomalously high porosity in deeply buried Amazonia. Mar. Petrol. Geol. 19, 1047–1071. https://doi.org/10.1016/S0264-8172
sandstones by grain-coating chlorite: examples from the Norwegian continental (03)00002-3.
shelf. Am. Assoc. Petrol. Geol. Bull. 77, 1260–1286. https://doi.org/10.1306/ Link, M.H., Welton, J.E., 1982a. Sedimentology and reservoir potential of Matilija
f4c8e062-1712-11d7-8645000102c1865d. Sandstone: an Eocene sand-rich deep-sea fan and shallow-marine complex,
Ehrenberg, S.N., 1991. Kaolinized, potassium-leached zones at the contacts of the Garn California. Am. Assoc. Petrol. Geol. Bull. https://doi.org/10.1306/03b5a97b-16d1-
Formation, Haltenbanken, mid-Norwegian continental shelf. Mar. Petrol. Geol. 8, 11d7-8645000102c1865d.
250–269. https://doi.org/10.1016/0264-8172(91)90080-K. Link, M.H., Welton, J.E., 1982b. Sedimentology and reservoir potential of Matilija
Folk, R.L., 2005. Nannobacteria and the formation of framboidal pyrite: textural Sandstone: an Eocene sand-rich deep-sea fan and shallow-marine complex,
evidence. J. Earth Syst. Sci. https://doi.org/10.1007/BF02702955. California. Am. Assoc. Petrol. Geol. Bull. 66, 1514–1534. https://doi.org/10.1306/
Gecer, A., Buyukutku, A., Caetano, P.S., Rocha, F.T., Kıbrıs, M.E., Albayrak, M., 2019. 03b5a97b-16d1-11d7-8645000102c1865d.
Reservoir potential of the haymana formation submarine-fan sandstones in the Longstaffe, F.J., 2003. Berthierine. In: Middleton, G.V., Church, M.J., Coniglio, M.,
haymana basin of Turkey. J. Pet. Explor. Prod. Technol. https://doi.org/10.1007/ Hardie, L.A., Longstaffe, Frederick J. (Eds.), Encyclopedia of Sediments and
s13202-019-0666-1. Sedimentary Rocks. Springer Netherlands, Dordrecht, pp. 64–66. https://doi.org/
Georgiadis, A., Dietel, J., Dohrmann, R., Rennert, T., 2020. Review Article. What are the 10.1007/978-1-4020-3609-5_26.
nature and formation conditions of hydroxy-interlayered minerals (hims) in soil? Lundegard, P.D., 1992. Sandstone porosity loss; a big picture view of the importance of
J. Plant Nutr. Soil Sci. 183, 12–26. https://doi.org/10.1002/jpln.201900283. compaction. J. Sediment. Res. 62, 250–260. https://doi.org/10.1306/D42678D4-
Glennie, K.W., 2009. Petroleum geology of the north sea: basic concepts and recent 2B26-11D7-8648000102C1865D.
advances: fourth edition. Petroleum Geology of the North Sea: Basic Concepts and Mansurbeg, H., Caja, M.A., Marfil, R., Morad, S., Remacha, E., Garcia, D., Martín-
Recent Advances: Fourth Edition. https://doi.org/10.1002/9781444313413. Crespo, T., El-Ghali, M.A.K., Nystuen, J.P., 2009. Diagenetic evolution and porosity
Goemaere, É., Katsch, A., Eschghi, I., Dreesen, R., 2014. Geological record and destruction of turbiditic hybrid arenites and siliciclastic sandstones of foreland
depositional setting of Palaeozoic oolitic ironstones in Western Europa. Anthropol. basins: evidence from the Eocene Hecho Group, Pyrenees, Spain. J. Sediment. Res.
Praehist. 125, 23–43. 79, 711–735. https://doi.org/10.2110/jsr.2009.060.
Gradstein, F.M., Waters, C.N., Charnock, M., Munsterman, D., Hollerbach, M., Mansurbeg, H., De Ros, L.F., Morad, S., Ketzer, J.M., El-Ghali, M.A.K., Caja, M.A.,
Brunstad, H., Hammer, Ø., Vergara, L., 2016. Stratigraphic guide to the cromer Othman, R., 2012. Meteoric-water diagenesis in late Cretaceous canyon-fill turbidite
knoll, shetland and chalk groups, north sea and Norwegian sea. Newslett. Stratigr. reservoirs from the Espírito Santo Basin, eastern Brazil. Mar. Petrol. Geol. https://
49, 73–280. https://doi.org/10.1127/nos/2016/0071. doi.org/10.1016/j.marpetgeo.2012.03.009.
Griffiths, J., Worden, R.H., Utley, J.E.P., Brostrøm, C., Martinius, A.W., Lawan, A.Y., Al- Mansurbeg, H., Morad, S., Al Suwaidi, M., Qurtas, S., Tveiten, O.G., Shahrokhizadeh, S.,
Hajri, A.I., 2021. Origin and distribution of grain-coating and pore-filling chlorite in Harchegani, F.K., 2020. Meteoric-water incursion into marine turbditic sandstones:
deltaic sandstones for reservoir quality assessment. Mar. Petrol. Geol. 134, 105326 evidence from the andrew formation (paleocene), UK central graben, North sea. Mar.
https://doi.org/10.1016/j.marpetgeo.2021.105326. Petrol. Geol. 118, 104428 https://doi.org/10.1016/j.marpetgeo.2020.104428.
Gulbrandsen, A., 1987. Agat. In: Geology Ofthe Norwegian Oil and Gas Fields. Graham Martinsen, O.J., Lien, T., Jackson, C., 2005. Cretaceous and Palaeogene turbidite systems
and Trotman, London, pp. 363–370. in the North Sea and Norwegian Sea basins: source, staging area and basin
Hansen, H.N., Løvstad, K., Lageat, G., Clerc, S., Jahren, J., 2021. Chlorite coating physiography controls on reservoir development. Pet. Geol. Conf. Proc. 6,
patterns and reservoir quality in deep marine depositional systems – example from 1147–1164. https://doi.org/10.1144/0061147.
the Cretaceous Agat Formation, Northern North Sea, Norway. Basin Res. 33, Matteo, E.N., Huet, B., Jové-Colón, C.F., Scherer, G.W., 2018. Experimental and
2725–2744. https://doi.org/10.1111/bre.12581. modeling study of calcium carbonate precipitation and its effects on the degradation
Helgeson, H.C., Murphy, W.M., Aagaard, P., 1984. Thermodynamic and kinetic of oil well cement during carbonated brine exposure. Cement Concr. Res. https://
constraints on reaction rates among minerals and aqueous solutions. II. Rate doi.org/10.1016/j.cemconres.2018.03.016.
constants, effective surface area, and the hydrolysis of feldspar. Geochem. Meunier, A., 2007. Soil Hydroxy-interlayered minerals: a re-interpretation of their
Cosmochim. Acta. https://doi.org/10.1016/0016-7037(84)90294-1. crystallochemical properties. Clay Clay Miner. 55, 380–388. https://doi.org/
Hendry, J.P., Trewin, N.H., 1995. Authigenic quartz microfabrics in Cretaceous 10.1346/ccmn.2007.0550406.
turbidites; evidence for silica transformation processes in sandstones. J. Sediment. Moore, D.M., Reynolds, R.C.j., 1997. X-Ray Diffraction and the Identification and
Res. 65, 380–392. https://doi.org/10.1306/d42680cc-2b26-11d7- Analysis of Clay Minerals, second ed. Oxford University Press.
8648000102c1865d. Moraes, M.A.S., De Ros, L.F., 1990. Infiltrated clays in fluvial Jurassic sandstones of
Hillier, S., 1994. Pore-lining chlorites in siliciclastic reservoir sandstones: electron Reconcavo Basin, northeastern Brazil. J. Sediment. Petrol. 60, 809–819. https://doi.
microprobe, SEM and XRD data, and implications for their origin. Clay Miner. 29, org/10.1306/212f928c-2b24-11d7-8648000102c1865d.
665–679. https://doi.org/10.1180/claymin.1994.029.4.20. Mosser, R., Cathelineau, M., Guillaume, D., Charpentier, D., Rousset, D., Barres, O.,
Houseknecht, D.W., Ross, L.M., 1992. Clay minerals in atokan deep-water sandstone Michau, N., 2010. Effects of temperature, ph, and iron/clay and liquid/clay ratios on
facies, arkoma basin: origins and influence on diagenesis and reservoir quality. Orig. experimental conversion of dioctahedral smectite to berthierine, chlorite,
Diagenesis, Petrophysics Clay Miner. Sandstones 227–240. https://doi.org/10.2110/ vermiculite, or saponite. Clay Clay Miner. 58 https://doi.org/10.1346/
pec.92.47.0227. CCMN.2010.0580212.
Hurst, A., Buller, A.T., 1984. Dish structures in some Paleocene deep-sea sandstones ( Murakami, T., Ito, J.-I., Utsunomiya, S., Kasama, T., Kozai, N., Ohnuki, T., 2004. Anoxic
Norwegian sector, North Sea): origin of the dish-forming clays and their effect on dissolution processes of biotite: implications for Fe behavior during Archean
reservoir quality. J. Sediment. Petrol. 54, 1206–1211. https://doi.org/10.1306/ weathering. Earth Planet Sci. Lett. 224, 117–129. https://doi.org/10.1016/j.
212F859E-2B24-11D7-8648000102C1865D. epsl.2004.04.040.
Isaksen, D., Tonstad, K., 1989. A revised cretaceous and tertiary lithostratigraphic Nystuen, J.P., 1999. Submarine sediment gravity flow deposits and associated facies:
nomenclature for the Norwegian North Sea. NPD Bull 5. core examples from the Agat Member. In: Extended Abstracts Bergen Conference.
Jackson, C.A.L., Barber, G.P., Martinsen, O.J., 2008. Submarine slope morphology as a Norwegian Petroleum Society, pp. 211–215.
control on the development of sand-rich turbidite depositional systems: 3D seismic

27
F. Azzam et al. Marine and Petroleum Geology 155 (2023) 106379

Oakman, C.D., Partington, M.A., 1998. Petroleum geology of the North Sea, in: Taylor, K.G., Macquaker, J.H.S., 2011. Iron minerals in marine sediments record
cretaceous. In: GLENNIE, K.W. (Ed.), Petroleum Geology of the North Sea, Basic chemical environments. Elements 7, 113–118. https://doi.org/10.2113/
Concepts and Recent Advances, fourth ed. Blackwell Scientific, Oxford, pp. 294–349. gselements.7.2.113.
https://doi.org/10.1016/s0037-0738(00)00100-7. Umar, M., Friis, H., Khan, A.S., Kassi, A.M., Kasi, A.K., 2011a. The effects of diagenesis on
Odin, G.S., 1988. Green Marine Clays: Oolitic Ironstone Facies, Verdine Facies, Glaucony the reservoir characters in sandstones of the late cretaceous pab formation, kirthar
Facies and Celadonite-Bearing Facies - a Comparative Study, Green Marine Clays: fold belt, southern Pakistan. J. Asian Earth Sci. 40, 622–635. https://doi.org/
Oolitic Ironstone Facies, Verdine Facies, Glaucony Facies and Celadonite-Bearing 10.1016/j.jseaes.2010.10.014.
Facies - a Comparative Study. https://doi.org/10.1016/0012-8252(90)90077-9. Umar, M., Friis, H., Khan, A.S., Kassi, A.M., Kasi, A.K., 2011b. The effects of diagenesis
Parnell, J., 2004. Titanium mobilization by hydrocarbon fluids related to sill intrusion in on the reservoir characters in sandstones of the late cretaceous pab formation,
a sedimentary sequence. Scotland. Ore Geol. Rev. 24, 155–167. https://doi.org/ kirthar fold belt, southern Pakistan. J. Asian Earth Sci. https://doi.org/10.1016/j.
10.1016/j.oregeorev.2003.08.010. jseaes.2010.10.014.
Pe-Piper, G., Karim, A., Piper, D.J.W., 2011. Authigenesis of titania minerals and the Velde, B., 2000. Clay Minerals: a Physico-Chemical Explanation of Their Occurrence.
mobility of Ti: new evidence from pro-deltaic sandstones, cretaceous scotian basin, Elsevier.
Canada. J. Sediment. Res. 81, 762–773. https://doi.org/10.2110/jsr.2011.63. Verhagen, I.T.E., Crisóstomo-Figueroa, A., Utley, J.E.P., Worden, R.H., 2020. Abrasion of
Petrea, C., Molenaar, N., 2014. Analyses of Fluid Inclusions in Quartz Overgrowths from detrital grain-coating clays during sediment transport: implications for diagenetic
Wells 35/3-2 and 35, vols. 3–4. Project No. G1011. clay coats. Sediment. Geol. 403, 105653 https://doi.org/10.1016/j.
Pettingill, H.S., 1998. Worldwide turbidite exploration and production: a globally sedgeo.2020.105653.
immature play with opportunities in stratigraphic traps. Day. https://doi.org/ Virolle, M., Brigaud, B., Beaufort, D., Patrier, P., Abdelrahman, E., Thomas, H.,
10.2118/49245-MS. Portier, E., Samson, Y., Bourillot, R., Féniès, H., 2022. Authigenic berthierine and
Porten, K.W., Warchoł, M.J., Kane, I.A., 2019. Formation of detrital clay grain coats by incipient chloritization in shallowly buried sandstone reservoirs: key role of the
dewatering of deep-water sands and significance for reservoir quality. J. Sediment. source-to-sink context. GSA Bull. https://doi.org/10.1130/b35865.1.
Res. 89, 1231–1249. https://doi.org/10.2110/jsr.2019.65. Virolle, M., Brigaud, B., Féniès, H., Bourillot, R., Portier, E., Patrier, P., Derriennic, H.,
Prost, R., Dameme, A., Huard, E., Driard, J., Leydecker, J.P., 1989. Infrared study of Beaufort, D., 2021. Preservation and distribution of detrital clay coats in a modern
structural OH in kaolinite, dickite, nacrite, and poorly crystalline kaolinite at 5 to estuarine heterolithic point bar in the gironde estuary (Bordeaux, France).
600 K. Clay Clay Miner. 37, 464–468. J. Sediment. Res. 91, 812–832. https://doi.org/10.2110/JSR.2020.146.
Rivard, C., Pelletier, M., Michau, N., Razafitianamaharavo, A., Bihannic, I., Virolle, M., Féniès, H., Brigaud, B., Bourillot, R., Portier, E., Patrier, P., Beaufort, D.,
Abdelmoula, M., Ghanbaja, J., Villiéras, F., 2013. Berthierine-like mineral formation Jalon-Rojas, I., Derriennic, H., Miska, S., 2020. Facies associations, detrital clay
and stability during the interaction of kaolinite with metallic iron at 90 ◦ C under grain coats and mineralogical characterization of the Gironde estuary tidal bars: a
anoxic and oxic conditions. Am. Mineral. 98, 163–180. https://doi.org/10.2138/ modern analogue for deeply buried estuarine sandstone reservoirs. Mar. Petrol. Geol.
am.2013.4073. 114, 104225 https://doi.org/10.1016/j.marpetgeo.2020.104225.
Roduit, N., 2007. Un Logiciel D’analyse D’images Pétrographiques Polyvalent. Wooldridge, L.J., Worden, R.H., Griffiths, J., Thompson, A., Chung, P., 2017. Biofilm
Saïag, J., Brigaud, B., Portier, É., Desaubliaux, G., Bucherie, A., Miska, S., Pagel, M., origin of clay-coated sand grains. Geology 45, 875–878. https://doi.org/10.1130/
2016. Sedimentological control on the diagenesis and reservoir quality of tidal G39161.1.
sandstones of the upper cape hay formation (permian, bonaparte basin, Australia). Worden, R., French, M., Mariani, E., 2012. Amorphous nanofilms result in growth of
Mar. Petrol. Geol. 77, 597–624. https://doi.org/10.1016/j.marpetgeo.2016.07.002. misoriented microcrystalline quartz cement maintaining porosity in deeply buried
Saigal, G.C., Bjørlykke, K., 1987. Carbonate cements in clastic reservoir rocks from sandstones. Geology 40, 179–182. https://doi.org/10.1130/g32661.1.
offshore Norway-relationships between isotopic composition, textural development Worden, R.H., Griffiths, J., Wooldridge, L.J., Utley, J.E., Lawan, A.Y., Muhammed, D.D.,
and burial depth. Geol. Soc. Spec. Publ. 36, 313–324. https://doi.org/10.1144/GSL. Simon, N., Armitage, P.J., 2020. Chlorite in sandstones. Earth Sci. Rev. 204, 103105
SP.1987.036.01.22. https://doi.org/10.1016/j.earscirev.2020.103105.
Schmidt, V., Mcdonald, D.A., 1979. Texture and recognition of secondary porosity in Worden, R.H., Morad, S., 2009. Clay minerals in sandstones: controls on formation,
sandstones. Asp. Diagenes. 209–225. https://doi.org/10.2110/PEC.79.26.0209. distribution and evolution. In: Clay Mineral Cements in Sandstones. https://doi.org/
Schulz, H.-M., Wirth, R., Schreiber, A., 2016. Nano-crystal formation of Tio2 polymorphs 10.1002/9781444304336.ch1.
brookite and anatase due to organic—inorganic rock–fluid interactions. J. Sediment. Worden, R.H., Morad, S., 2000. Quartz cementation in oil field sandstones: a review of
Res. 86, 59–72. https://doi.org/10.2110/jsr.2016.1. the key controversies. In: Quartz Cementation in Sandstones. John Wiley & Sons,
Sea, N.N., Hansen, H.N., Jahren, J., 2021. Chlorite Coating Patterns and Reservoir Ltd, pp. 1–20. https://doi.org/10.1002/9781444304237.ch1.
Quality in Deep Marine Depositional Systems – Example from the Cretaceous Agat Worden, Richard H., Morrall, G.T., Kelly, S., Mc Ardle, P., Barshep, D.V., 2020b.
1–20. https://doi.org/10.1111/bre.12581. A renewed look at calcite cement in marine-deltaic sandstones: the Brent Reservoir,
Shanmugam, G., Lehtonen, L.R., Straume, T., Syvertsen, S.E., Hodgkinson, R.J., Heather Field, northern North Sea, UK. Appl. Anal. Tech. To Pet. Syst. https://doi.
Skibeli, M., 1994. Slump and debris-flow dominated upper slope facies in the org/10.1144/SP484-2018-43.
Cretaceous of the Norwegian and northern North Seas (61-67◦ N): implications for Xiao, M., Yuan, X., Cheng, D., Wu, S., Cao, Z., Tang, Y., Xie, Z., 2018. Feldspar
sand distribution. Am. Assoc. Petrol. Geol. Bull. https://doi.org/10.1306/a25fe3e7- dissolution and its influence on reservoirs: a case study of the lower triassic
171b-11d7-8645000102c1865d. baikouquan formation in the northwest margin of the junggar basin, China.
Skibeli, M., Barnes, K., Straume, T., Syvertsen, S.E., Shanmugam, G., 1995. A sequence Geofluids 2018, 1–19. https://doi.org/10.1155/2018/6536419.
stratigraphic study of lower cretaceous deposits in the northernmost North Sea. Nor. Yezerski, D.J., Shumaker, N., 2018. PS improving prediction of porosity preservation in
Pet. Soc. Spec. Publ. 5, 389–400. https://doi.org/10.1016/S0928-8937(06)80077-9. thermally-stressed deep marine sandstones: a synthesis of grain-coating chlorite
Spain, D.R., 1992. Petrophysical evaluation of a slope fan/basin-floor fan complex: observations. In: AAPG 2017 Annual Convention and Exhibition.
cherry canyon formation, ward county, Texas1. Am. Assoc. Petrol. Geol. Bull. 76, Yuan, G., Cao, Y., Jia, Z., Gluyas, J., Yang, T., Wang, Y., Xi, K., 2015. Selective
805–827. https://doi.org/10.1306/BDFF88E4-1718-11D7-8645000102C1865D. dissolution of feldspars in the presence of carbonates: the way to generate secondary
Sugimori, H., Yokoyama, T., Murakami, T., 2009. Kinetics of biotite dissolution and Fe pores in buried sandstones by organic CO2. Mar. Petrol. Geol. 60, 105–119. https://
behavior under low O2 conditions and their implications for Precambrian doi.org/10.1016/j.marpetgeo.2014.11.001.
weathering. Geochem. Cosmochim. Acta 73, 3767–3781. https://doi.org/10.1016/j. Yuan, G., Cao, Y., Schulz, H.M., Hao, F., Gluyas, J., Liu, K., Yang, T., Wang, Y., Xi, K.,
gca.2009.03.034. Li, F., 2019. A review of feldspar alteration and its geological significance in
Sullivan, M., Coombes, T., Imbert, P., Ahamdach-Demars, C., 1999. Reservoir quality and sedimentary basins: from shallow aquifers to deep hydrocarbon reservoirs. Earth Sci.
petrophysical evaluation of Paleocene sandstones in the West of Shetland area. Pet. Rev. https://doi.org/10.1016/j.earscirev.2019.02.004.
Geol. Conf. Proc. 5, 627–633. https://doi.org/10.1144/0050627. Yuan, G., Cao, Y., Zhang, Y., Gluyas, J., 2017. Diagenesis and reservoir quality of
Surdam, R.C., Crossey, L.J., Hagen, E.S., Heasler, H.P., 1989. Organic-inorganic sandstones with ancient “deep” incursion of meteoric freshwater——an example in
interactions and sandstone diagenesis. Am. Assoc. Petrol. Geol. Bull. https://doi.org/ the Nanpu Sag, Bohai Bay Basin, East China. Mar. Petrol. Geol. https://doi.org/
10.1306/703c9ad7-1707-11d7-8645000102c1865d. 10.1016/j.marpetgeo.2017.02.027.
Tang, L., Gluyas, J., Jones, S., 2018. Porosity preservation due to grain coating illite/ Zhang, Y., Zeng, J., Yu, B., 2009. Experimental study on interaction between simulated
smectite: evidence from buchan formation (upper devonian) of the ardmore field, UK sandstone and acidic fluid. Petrol. Sci. 6, 8–16. https://doi.org/10.1007/s12182-
North Sea. Proc. Geol. Assoc. 129, 202–214. https://doi.org/10.1016/j. 009-0002-3.
pgeola.2018.03.001. Ziegler, P.A., 1992. North Sea rift system. Tectonophysics 208, 55–75. https://doi.org/
Taylor, J., 1950. Pore-space reduction in sandstones. Am. Assoc. Petrol. Geol. Bull. 34, 10.1016/0040-1951(92)90336-5.
701–716. Ziegler, P.A., 1975. Geologic evolution of North Sea and its tectonic framework. Am.
Assoc. Petrol. Geol. Bull. 59, 1073–1097. https://doi.org/10.1306/83d91f2e-16c7-
11d7-8645000102c1865d.

28

You might also like