You are on page 1of 16

GEOHORIZONS

AUTHORS

Methodology for risking fault


seal capacity: Implications of
fault zone architecture
Roald B. Frseth, Eivind Johnsen, and
Susanne Sperrevik

ABSTRACT
We introduce a methodology for quantifying the risk associated
with a seal for fault-bounded prospects. Applying this methodology, the aspects of fault seal are confined within four main risk categories. The methodology allows comparable criteria to be applied
in the risking procedure to reduce uncertainty in fault seal assessments. As a foundation for the methodology, we combine onshore
and offshore data from large faults and demonstrate how architecture and the distribution of fault rocks may influence sealing
capacity. Despite the variable and complex structure of fault zones,
we have observed fault zone characteristics that appear in common
to the faults investigated, and we consider these factors to be crucial
in the risking of fault seal predictions. The fault zones in our database, typically bounded by external slip surfaces, represent two main
categories: (1) a layer of shale smear entrained into the fault zone
and derived from a thick shale source layer within the sequence
offset by the fault and (2) fault zones characterized by internal slip
surfaces, slivers of footwall and hanging-wallderived material rotated along the fault zone and commonly enclosed in a matrix of
shaly-silty fault gouge. This study highlights the disparity between
the complexity of actual faults and the abrasion-type shale gouge
ratio (SGR) algorithm currently used in the industry to estimate
sealing capacity of faults, which assumes that the seismically derived
throw is concentrated in a single fault plane. We discuss how this
may influence sequence juxtaposition across a fault, the associated
SGR values, and ultimately, the fault seal risking.

Copyright #2007. The American Association of Petroleum Geologists. All rights reserved.
Manuscript received May 18, 2006; provisional acceptance August 21, 2006; revised manuscript
received March 7, 2007; final acceptance March 8, 2007.
DOI:10.1306/03080706051

AAPG Bulletin, v. 91, no. 9 (September 2007), pp. 1231 1246

1231

Roald B. Frseth  Hydro ASA, Research


Centre, Box 7190, N-5020 Bergen, Norway;
roald.farseth@hydro.com
Roald B. Frseth received a Cand. Real in
structural geology from the University of Bergen
in 1971. He worked as a researcher at the University of Bergen (1971 1977) and has been
with Hydro ASA since 1977. He has worked extensively in the area of fault architecture in
onshore and offshore areas, fault seal analysis,
and the risking of fault seal capacity in prospect
analysis.
Eivind Johnsen  Eni Norge AS, Box 101
Forus, N-4064 Stavanger, Norway;
eivind.johnsen@eninorge.com
Eivind Johnsen has worked with Eni Norge AS
since 2003. He has a Cand. Scient. in structural
geology from the University of Bergen (1998).
From 1998 to 2003, he worked as an explorationist within operations and as a researcher
at Norsk Hydro Research Centre. His main
competence areas are seismic interpretation,
fault architecture, fault seal analysis, and
prospect analysis.
Susanne Sperrevik  Statoil ASA, Box 7200,
N-5020 Bergen, Norway; sperre@statoil.com
Susanne Sperrevik has worked with Statoil ASA
since 2005. She has a Cand. Scient. in structural geology from the University of Bergen
(1998) and a Dr. Scient. from the University of
Bergen (2002). From 1998 to 2005, she worked
as a researcher at Norsk Hydro Research Centre. Her main competence areas are fault seal
analysis, structural restoration, seismic interpretation, and prospect analysis.

ACKNOWLEDGEMENTS
The authors acknowledge permission given by
Hydro to use data collected in onshore outcrops and core data from wells drilled offshore
Norway, as well as the use of data from faultrelated traps that have been subject to postdrill
analysis. This article benefited from reviews
made by J. Steven Davis, Alfred Lacazette, and
David A. Pivnik. We also thank Ian Sharp of
Hydro ASA for his comments.

Figure 1. Schematic illustration


showing potential hydrocarbon traps
(fault blocks) resulting from normal
faults that offset a sand-shale sequence. To accumulate hydrocarbons
within a trap, there is generally the
requirement that faults represent
a seal against neighboring fault compartments. The methodology introduced in this study allows quantification of the risk associated with a seal
for self-separated and self-juxtaposed
reservoirs (modified from Frseth,
2006).

INTRODUCTION
Despite considerable research and publication on the
controls on fault seal capacity, there is still a significant
lack of precision regarding the implications of fault
zone architecture, the distribution of fault rock within
the fault zone, and its capacity to seal. Although this
is understandable because a host of variables affect
the sealing capacity, it remains a major weakness in our
ability to predict the sealing capacity of critical faults
and to quantify the risk associated with a seal for faultbounded traps. Limitations in seismic resolution also
introduce uncertainties with respect to factors that are
considered important in fault seal assessments.
The schematic illustration (Figure 1) shows a cross
section of potential hydrocarbon traps (fault blocks) resulting from normal faults that offset a sand-shale sequence. For fault blocks as traps and especially hangingwall traps, such as the fault block to the far left in
Figure 1, there is generally the requirement that hydrocarbons are sealed by faults against neighboring
fault compartments. In hydrocarbon exploration, it is
therefore vitally important to quantify the risk associated with a seal for fault-bounded prospects for predicting the maximum hydrocarbon column height that can
be supported by faults and, thereby, the size of the trap.
The risking of fault seal capacity is the probability that
the fault zone material is capable of sealing a specific
hydrocarbon column, and in the present account, we
consider the static sealing, i.e., where the seal remains an
effective barrier to cross-fault flow in a geological time
1232

Geohorizons

scale. In the industry, the shale gouge ratio (SGR) algorithm of Yielding et al. (1997) is currently applied in fault
seal assessments, and workers commonly refer to a minimum SGR value as a measure of the sealing capacity of
the fault. This study highlights the disparity between the
complexity of actual faults and the assumption of a single
fault plane that is the foundation in the SGR algorithm,
and we discuss how this may influence fault seal risking.
Faults associated with potential hydrocarbon traps in
an extensional setting represent two main groups based
on fault throw related to reservoir thickness (Figure 1):
The first group is self-separated reservoirs, in which
the reservoir is entirely separated from its continuation
across the fault. If the reservoir is juxtaposed against
shale across a fault, it results in a juxtaposition seal.
However, where reservoir A is self-separated, it may be
juxtaposed against reservoir B. If this is the case, a membrane seal along the fault itself is required to prevent the
leakage of hydrocarbons across the fault. A membrane
seal is a fault rock with high capillary entry pressure
(Watts, 1987) because of smear, cementation, cataclasis, or a combination of these. The shale that separates
reservoir A and reservoir B in the sequence (Figure 1)
represents a source layer with the potential to develop a
smear along faults that cut the sediments. If the shale
smear entrained into the fault zone forms a continuous
and impermeable membrane, the smear can separate
the two juxtaposed reservoir units. In the absence of
shale smear, the two juxtaposed reservoirs would be
expected to be hydraulically connected across the fault
plane. The second group is self-juxtaposed reservoirs, in

which the reservoir is partially juxtaposed against itself


across the fault. To prevent leakage, a membrane seal
along the sand-sand juxtaposition is required. In such
situations, an increased capillary entry pressure across
the fault may result from mechanical shearing with grain
reorganization, denser grain packing, and diagenetic reactions (e.g., Bjrlykke, 1999), as well as phyllosilicate
smearing (e.g., Fisher and Knipe, 2001).
Few published studies are based on the observation
of large-displacement (hundreds of meters) faults to describe their fault zone architecture and sealing properties. This category of faults, which are known to effectively seal large oil and gas columns (e.g., Frseth et al.,
1984; Nybakken, 1991; Zieglar, 1992; Jev et al., 1993;
Knott, 1993; Gibson, 1994; Kim et al., 2003; Frseth,
2006) and result in large cross-fault differences in the
aquifer pressure (e.g., Buhrig, 1989; Gaarenstroom et al.,
1993), represents seismically mappable features in the
subsurface. However, the internal structure and contents of fault zones occur on scales below the lateral
resolution of seismic data, and accordingly, large faults
are commonly risked as a single fault plane (Figure 1).
We present examples of large-displacement faults drilled
and cored offshore Norway, as well as examples of large
faults examined in outcrops. Despite the complex architecture of fault zones, we have observed fault zone characteristics that appear in common to the faults studied.
The recognition of such fault zone characteristics allows
most aspects of fault seal to be confined within a few
main categories, and in the proposed methodology,
each category is assigned a fault seal probability. The
aim is to present comparable criteria to be applied in
the risking procedure to ensure consistent risking and
to minimize the uncertainty in fault seal assessments
of fault-bounded prospects.

METHODS APPLIED TO QUANTIFY FAULT


SEAL CAPACITY
To quantify the risk associated with fault-bounded prospects (Figure 1), two approaches are commonly applied.
One approach places emphasis on the prediction of the
presence and properties of membrane seals, and in general, measured properties of small faults are upscaled to
predict the fluid-flow properties of seismic-scale faults.
The character and properties of small-displacement
faults in siliciclastic sequences have been studied by several workers (Aydin, 1978; Pittman, 1981; Jamison and
Stearns, 1982; Gabrielsen and Koestler, 1987; Speksnijder,
1987; Knipe, 1992; Sassi et al., 1992; Lindsay et al., 1993;

Antonellini and Aydin, 1994; Berg and Avery, 1995;


Knipe et al., 1997; Fisher and Knipe, 1998, 2001; Hesthammer and Fossen, 1998; Gibson, 1998; Sperrevik
et al., 2000, 2002; Dewhurst et al., 2002; Lewis et al.,
2002; Kim et al., 2003). These studies include observations in outcrops, core measurements, and laboratory
studies to investigate how faults influence fluid-flow
properties (e.g., porosity, permeability, and capillary
properties).
The other approach takes little or no account of the
architecture of the fault zone, which is considered to
be unpredictable, but relies on a variety of algorithms to
predict fault seal capacity. Algorithms that have been
proposed focus on the likelihood of shale smearing within fault zones and generally express the relationship
between fault throw and the amount of clay and shale
in host rocks (Bouvier et al., 1989; Lindsay et al., 1993;
Skerlec, 1996; Fulljames et al., 1997; Yielding et al.,
1997; Fisher and Knipe, 2001). The algorithms most
commonly referred to are as follows: (1) SGR is a calculation of smear that is derived from the complete sedimentary succession that has slipped past a point on the
fault surface (Yielding et al., 1997), expressed as SGR
P
Vshale  Z=Throw, where, for a given interval,
V shale is the volumetric shale fraction and DZ the interval thickness; (2) the shale smear factor (SSF) of Lindsay
et al. (1993) estimates the possibility of a specific shale
layer in the sequence to be drawn along the fault zone
to form a thinned but continuous smear; SSF = Throw/
Thickness of the source layer.
The abrasion-type SGR algorithm of Yielding et al.
(1997) simply expresses that the more shaly the host
rock is, the greater the proportion of shale in the fault
zone, and therefore, the higher the capillary entry pressure. Shale gouge ratio values calculated using the software FAPS/TrapTester (Yielding et al., 1997) are empirically calibrated to pressure differentials across faults
in the same area (if available) to assess the fault seal
capacity for critical faults. An important aspect of fault
seal analysis applying this software is to investigate the
juxtaposition of lithologies across faults under consideration. Where a fault offsets a layered sedimentary succession composed of reservoir and nonreservoir lithologies, the construction of a juxtaposition diagram results
in a detailed juxtaposition plot that may contain several
potential seal and leak windows (Fristad et al., 1997;
Fulljames et al., 1997; Knipe, 1997; Yielding et al., 1997,
1999; Sverdrup et al., 2003). A crucial foundation in
the SGR algorithm is that the seismically derived throw
is concentrated to a single fault surface. However, there
is abundant evidence as demonstrated in this study as
Frseth et al.

1233

well as by other researchers (e.g., Gibson, 1994; Childs


et al., 1997; Foxford et al., 1998; Walsh et al., 1998;
Doughty, 2003) that fault zones contain several major
slip surfaces, and accordingly, a juxtaposition diagram
assuming a single fault will not represent the actual
across-fault situation. This is likely to have a major impact on calculated SGR values and, ultimately, the fault
seal risking.

FAULT ZONE ARCHITECTURE


We have access to an extensive database containing
information about large faults (seismic scale), both from
offshore and onshore areas. The database reflects a wide
variety of faults with respect to geographical area, type
of faulted rocks (siliciclastic rocks, carbonates), fault type
(normal and reverse), fault-plane geometry (planar and
listric), fault architecture, and burial depth. The type and
quality of information recorded from individual faults
will of course be dependent on whether the observed
faults occur in onshore or offshore areas. The onshore
faults where we have good outcrop exposures have been
investigated in detail with respect to architecture, lithology involved, and type of fault rock, but unfortunately,
these faults do not provide information about the total
sealing capacity of the large fault. For faults known
from oil and gas fields in offshore areas, the situation
is the opposite; we have relatively good control on the
sealing capacity (by recorded column heights and pressure differences across faults), but less control on the
fault architecture because we only have information from
seismic data (resolution issue) and relatively few wells
where large faults are drilled and cored. We therefore
combine onshore and offshore data to illustrate the architecture of large faults, and to discuss how architecture and the distribution of fault rocks may influence
the risk associated with a fault-bounded trap. The data
collected are the foundation for the risking methodology
suggested in this study. The examples we present are associated with large normal faults that offset interbedded
clastic stratigraphies as well as interbedded limestone
and shale. Based on observations from several areas, there
is no clear distinction with respect to fault zone architecture in these sedimentary successions. Our observations
indicate that wide zones with deformed rocks are typical
features of large normal faults, which developed when
sediments were poorly lithified. The zone with deformed
rocks is typically bounded by major slip surfaces.
Where two different sands are juxtaposed across a
fault (Figure 1), we have grouped the associated faults
1234

Geohorizons

into two categories. Category 1 consists of fault zones


with an intact layer of shale smear, which typically occurs where faults offset a sequence that contains a shale
source layer with a significant thickness (tens of meters).
Category 2 consists of fault zones containing several slip
surfaces with a complex mix of various lithologies and
fault rocks. Faults in this latter category do not have a
continuous smear from a thick source layer. All examples described in this article represent random cross sections of the fault under consideration. The architecture
of a fault changes along strike and dip and, based on our
observations, more frequently for category 2 types of
faults. However, in the risking of fault seal capacity, the
risk pertains to the location along the fault that has the
architecture most critical with respect to leakage across
the fault. The fault seal probability assigned to this specific part of the fault must reflect the height of the hydrocarbon column to be sealed by the fault.
Examples of Category 1 Fault Zones
Figure 2 shows cross sections of large faults with smear
derived from thick source layers. The first example is from
the Brage field in the northern North Sea (Figure 2a).
The Brage East fault that delineates the Brage field
to the east (Hage et al., 1987; Frseth, 1996; Aarland and
Skjerven, 1998) was drilled by production wells. Jurassic (BajocianVolgian) extension resulted in a normal
throw of 200250 m (660820 ft) across the fault at
reservoir level (Figure 2a). The interpretation of the
Brage East fault in the cross section is based on seismic
data and core data from wells. The fault juxtaposes oilbearing sandstones of the Lower Jurassic Statfjord Formation in the footwall (Brage field) against water-bearing
sandstones of the Middle Jurassic Brent Group in the
hanging wall (self-separated reservoir). A maximum pressure difference of 7 bars was recorded across the fault
zone. Oriented cores were taken, including the 20-m
(66-ft)-thick fault zone where it was penetrated by the
well. Two major slip surfaces, 0.2 and 0.8 m (0.6 and
2.6 ft) thick, respectively, have accommodated the
displacement that is about equally partitioned between
the two fault segments (Aarland and Skjerven, 1998).
Between the two slip surfaces, the shale-dominated
Lower Jurassic (Toarcian) Drake Formation is rotated
to become parallel to the footwall and hanging-wall
cutoff planes. The Drake Formation, with an initial thickness of about 100 m (330 ft), is the source for a smear
that is approximately 20 m (66 ft) thick at the well location. Hence, from the Brage East fault, cores are available to prove that smear has occurred, and for the actual

Figure 2. Cross sections


of large faults with smear
derived from a thick
source layer (modified
from Frseth, 2006).
(a) A normal fault with
a throw of 200250 m
(660820 ft) at reservoir
level delineates the Brage
field in the northern North
Sea. The fault juxtaposes
oil-bearing sandstones of
the Lower Jurassic Statfjord
Formation in the footwall
(Brage field) against waterbearing sandstones of the
Middle Jurassic Brent Group
in the hanging wall. The
fault is a major seal caused
by shale smear along the
fault. (b) The Baba fault
in Sinai, Egypt, represents
the eastern margin of the
Suez rift. The fault offsets
alternating sandstone, interbedded shale, and
sandstone and limestone
of the prerift succession
(Moustafa, 1987). In the
cross section, displacement
is partitioned over several
major slip surfaces (inset).
The Esna Shale with an
original thickness of 50 m
(164 ft) is thinned to 12 m
(39 ft) at the exposed level
about midway between
footwall and hanging-wall
cutoffs for the western
fault segment that has a
throw of about 130 m
(426 ft) at this location.

displacement, the smear is interpreted to be continuous


to provide the seal and to account for the associated
across-fault pressure difference (Frseth, 2006).
The second example of category 1 faults is an onshore example from Sinai, Egypt (Figure 2b). Here, deformation features associated with large normal faults
that represent the eastern margin of the Suez rift are
well exposed. The faults offset alternating sandstone, interbedded shale, and sandstone and limestone (Robson,
1971; Moustafa, 1987; Sharp et al., 2000) of the prerift

succession (Figure 2b). The smear along large faults of


a Paleocene marine shale unit, the Esna Formation, is a
common phenomenon in this area. In the cross section,
the displacement is partitioned over several major slip
surfaces. The Esna Shale, with an original thickness of
50 m (164 ft) in this area, is thinned to 12 m (39 ft)
(Figure 2b, inset) at the exposed level about midway
between the footwall and hanging-wall cutoffs for the
western fault segment, with a throw of approximately
130 m (426 ft) at this location. Along strike, the shale
Frseth et al.

1235

Figure 3. Cross sections of


large faults characterized by wide
fault zones and complex fault
architecture. (a) A normal fault
within the Visund field in the
northern North Sea drilled and
partly cored where the throw
is 210 m (688 ft). The fault proved
to be a 60-m (196-ft)-wide deformation zone that contains two
relatively wide shear zones
with throws of 80 and 130 m
(262 and 426 ft), respectively.
The two shear zones represent
significant seals. (b) The Blue Ben
fault in Somerset at the southern
margin of the Bristol Channel,
United Kingdom, has a throw
of 220 m (721 ft) in a cliff exposure. The inset shows the likely
seismic expression of the Blue
Ben fault, assuming that the total
throw was concentrated to a
single fault. The exposed profile
reveals a much more complex
situation where two major slip
surfaces with throws of 85 and
125 m (278 and 410 ft), respectively, represent the external
boundaries of the 35-m (114-ft)wide fault zone. Several internal
slip surfaces juxtapose various
parts of the stratigraphy (modified from Frseth, 2006).

smear is gradually thinned for an increasing throw, and


at locations where the throw exceeds 230 m (754 ft),
no continuous shale is present along the Baba fault
(Frseth, 2006).
Examples of Category 2 Fault Zones
Figure 3 illustrates faults where smear from a thick
source layer is not evident, and the faults have a complex internal architecture. The lack of extensive shale
smear may result from the fact that the offset lithologies do not contain a thick source layer, or that the
fault throw is so large that it results in a discontinuous
smear at the location studied. Figure 3a is an interpretation of a normal fault within the Visund field in
1236

Geohorizons

the northern North Sea (Frseth et al., 1995) that was


penetrated by a well near the area of maximum throw
of 210 m (688 ft). The fault results from late Middle
Late Jurassic extension and juxtaposes a Middle Jurassic Brent Group sandstone reservoir of a hanging-wall
gas accumulation against the shale-dominated Dunlin
Group as well as Statfjord Formation sandstones in the
footwall (self-separated reservoir). The fault was mostly
cored and proved to be a 60-m (196-ft)-wide deformation zone with two relatively wide shear zones having
throw equal to 80 and 130 m (262 and 426 ft), respectively. In the well, the Brent Group occurs as a 25-m
(82-ft)-wide protolith (Caine et al., 1996) between the
two shear zones (Figure 3a). The shear zones represent
significant seals, and pressure differences of 3 27 bars

were recorded across the zones. The sealing capacity of


the shear zone to the east where sand-prone units are
juxtaposed across faults may result from fragmented
Brent Group lithologies and the enrichment of nonreservoir components along internal slip surfaces as seen
in the cores. There was no core recovery from the shear
zone to the west, where reservoir sands of the Brent
Group and the Statfjord Formation are juxtaposed. The
observed pressure differences associated with the shear
zones have been calibrated to SGR values that are in
the range of 30 50%.
The Blue Ben fault in Somerset at the southern margin of the Bristol Channel, United Kingdom, has a throw
of 220 m (721 ft) in a cliff exposure (Figure 3b), i.e., of
the same scale as the fault within the Visund field. Along
this part of the Bristol Channel, large normal faults as
well as subseismic faults that formed during Late Jurassic
extension (Peacock and Sanderson, 1991; Dart et al.,
1995) can be studied in exceptional exposures represented by cliffs and wave-cut platforms. The fault offsets
a Triassic Lower Jurassic stratigraphy represented by
marl, limestone, sandstone, siltstone, mudstone, and
shale. The Lower Jurassic section, dominated by limestone and shale, has a layer-cake stratigraphy (Dart et al.,
1995, and references therein), and accordingly, offset
across faults can be determined using the detailed stratigraphic template. Figure 3b (inset) shows the likely
seismic resolution of the Blue Ben fault and the apparent juxtaposition across the fault, assuming that the
total throw was concentrated to a single fault plane. The
outcrop profile reveals a much more complex situation
where two major slip surfaces with throw of 85 and
125 m (278 and 410 ft), respectively, represent the external boundaries of the 35-m (114-ft)-wide fault zone,
whereas internal slip surfaces juxtapose various parts
of the stratigraphy. Shear zones with significant rotation of bedding, lenses of the Blue Anchor Formation
enclosed in a silty matrix, and thin zones of smear that
extend for several meters are characteristic features
of the fault zone. The fault zone can be mapped along
strike for 350400 m (11481312 ft) on the foreshore
to demonstrate variations regarding width of the zone
and number of internal slip surfaces.

FAULT ZONE ARCHITECTURE AND PREDICTION


OF FAULT SEAL CAPACITY
Where sediments offset by seismic-scale faults contain
a shale source layer with a significant thickness (tens of
meters), there is a high probability of having a contin-

uous layer of sealing shale smear as long as the SSF is


below a critical threshold (Figure 2). Without the continuous smear from a thick source layer, but with the
likelihood of a fault zone containing several slip surfaces
with the mixing of various lithologies and fault rocks
(Figure 3), the architecture and content of the fault zone
become less predictable, and a decreased confidence is
attached to fault seal predictions.
Frseth (2006) applied data from the same database
as we used in this study to estimate the continuity of smear
associated with large faults of the category 1 type. Smear
along these faults is typically associated with fault segments that are slightly offset, where the overlap between
the segments creates an extensional dip relay (Childs
et al., 1996; Rykkelid and Fossen, 2002) (Figure 2) that
enables the source layer to be drawn out ductilely between the faults for an increasing slip. Work on faulting
in interbedded sand-shale sequences offshore Norway
as well as in other basins demonstrates that smears prevent leakage of hydrocarbons across large faults. These
are cases where two different sandstone units are juxtaposed across a fault (Figure 1), and the fault is proven to
trap a significant hydrocarbon accumulation (Figure 2a).
Based on core data, the incorporation of shale into the
fault zone is shown to represent the mechanism responsible for forming these seals (e.g., Smith, 1980; Bouvier
et al., 1989; Gibson, 1994; Frseth, 2006). Hence, data
available demonstrate the continuity of smears and also
that shale membranes entrained into fault zones form
effective seals.
Faults with complex fault zone architecture (category 2) and mixing of lithologies (Figure 3) lead to a
different approach. Because of the lack of measurements
of fluid-flow properties representing the effect of large
faults, workers have presented the analysis of the petrophysical properties of fault rocks derived from core-plug
measurements across small-displacement (centimeterscale) faults. This has allowed the construction of databases on fault rock characteristics in different host sediments under different geohistories. Processes introduced,
such as clay or phyllosilicate smearing, disaggregation,
cementation, and cataclasis that may alter the pore structure of faulted stratigraphy to give the associated faultrock types, emerge from this type of measurement (e.g.,
Knipe, 1997; Knipe et al., 1997; Fisher and Knipe, 1998,
2001; Gibson, 1998; Sperrevik et al., 2002). These workers emphasize the importance of knowing the fluidflow properties of small-scale faults, which they extrapolate and use as a tool to predict the fault rock properties
and sealing capacity of seismic-scale faults. Although
in theory, the capillary entry pressure of a fault rock is
Frseth et al.

1237

independent of thickness (Hubert, 1953; Watts, 1987),


it is not a straightforward process to predict the properties of large-scale faults from small-scale faults, for
example, as the distribution of fault rocks within a wide
fault zone of category 2 becomes unpredictable.
None of the fault seal algorithms commonly applied
are, in themselves, a measure of sealing capacity of the
fault. Instead, these algorithms provide estimates of the
relative likelihood of shale smear being developed at
the fault surface (Yielding et al., 1997). To use calculated SSF and SGR values as estimates of seal capacity,
they have been calibrated in data sets where sealing
behavior is documented. Frseth (2006) applied data
from large faults to establish the threshold between
continuous and discontinuous smears and argued that
with SSF equal to 4 or smaller, a continuous smear along
large faults is expected. Gibson (1994) and Aydin and
Eyal (2002) presented data from large faults and also
concluded that an SSF of approximately 4 represents a
threshold between continuous and discontinuous smears.
Outcrop data (e.g., Lindsay et al., 1993; Frseth, 2006)
and experiments (Sperrevik et al., 2000) show that continuous smears may form along small faults (<10-m
[<33-ft] throw) for much higher SSF values. Accordingly, SSFs derived from small faults cannot be applied
to predict the continuity of smear along large faults
(Frseth, 2006).
In situations where smear from a single source layer
is unlikely, the abrasion-type SGR algorithm of Yielding
et al. (1997) is commonly applied. Calibration for several data sets from different basins have indicated that
an SGR of about 20% (volume of shale in the slipped
interval) is a typical threshold between minimal acrossfault pressure difference and significant seal (Fristad
et al., 1997; Yielding et al., 1997; Foxford et al., 1998;
Yielding, 2002; Bretan et al., 2003). To assess the sealing
capacity of faults that delineate a fault-bounded trap
(Figure 1), calculated SGR values should ideally be calibrated to pressure differentials across faults in the same
area. In general, such information is not available, and
published calibration diagrams based on data from other
areas or basins are commonly applied. A major weakness
in applying published calibration diagrams (Yielding,
2002; Bretan et al., 2003) is the fact that it is not known
how the V shale (volumetric shale fraction) for smear attributes in the sedimentary sequence have been calculated. Differences in SGR values may originate from the
differences in methods applied, and caution must be
taken when results from different areas and companies
are compared. In general, only the fraction of clay-shale
and phyllosilicates in the sequence are considered in
1238

Geohorizons

SGR calculations. However, several lithologies apart


from clay-shale, such as coal, silt, and carbonate, may
contribute to the low-permeable fault gouge (Frseth,
2006), and accordingly, such lithologies should be included in the smear summation. Several authors have
discussed the problems of using a global calibration diagram and concluded that the relationship between the
clay content and petrophysical properties of faults is not
obvious over the total range and could be more complex
than previously acknowledged (e.g., Fisher and Knipe,
2001; Sperrevik et al., 2002; Bretan et al., 2003). From
published data, it is evident that one global calibration
of SGR versus sealing capacity is not applicable in all
areas. In the database we have used in this study, we have
examples of faults from fields where the SGR versus
sealing capacity are compatible with published calibration plots, but there are also examples of the opposite.
Faults within the Njord field (Figure 4) offshore midNorway are examples that contradict the global calibration of SGR versus fault seal capacity. Low SGR values (<10%) are associated with parts of the fault surface
where the seismic interpretation shows sand-sand juxtaposition, i.e., potential leak points. However, pressure
differentials across the faults demonstrate that they represent seals. The faults, with throw of some 20 40 m
(66131 ft), represent wide zones (throw/thickness ratio of 3.7 and 2.8 respectively, at well locations), and
from core data, the fault displacement is interpreted to
be partitioned onto several slip surfaces. The increased
proportion of nonreservoir components incorporated
because of several slip surfaces, combined with a relative
deep burial (>2800 m; >9186 ft) and cementation, contributes to the sealing capacity of these fault zones. An
onshore analog to these types of faults is found at Hartley
Steps, Northumberland, United Kingdom, where the
fault is well exposed in a coastal section (Figure 5). The
fault offsets an Upper Carboniferous sequence with
sand, silt, shale, and coal layers (Jones, 1967). The cumulative offset of 17 m (55 ft) is partitioned onto several faults with throw from 6 m (19 ft) down to the
centimeter scale, over a 10-m (33-ft)-wide zone. The
fault with the largest throw has a maximum width
of 1 m (3.3 ft) in the outcrop and contains a sheared
melange of various wall rocks (Figure 5, inset). The
southeastern part of this zone is a 0.150.25-m (0.49
0.82-ft)-thick gouge bounded by slip surfaces, and the
gouge is derived primarily from shale and coal layers
in the hanging wall numbered 7, 8, and 9. The faultbounded zone northwest of this gouge is maximum at
0.8 m (2.62 ft) in thickness and composed of rotated
and steeply inclined lenses of shale with ironstone bands

Figure 4. Faults within the Njord field offshore mid-Norway represent seals, although they juxtapose sand units and have low SGR
values. Cores from the wells show wide fault zones where slip surfaces represent the external boundaries of the fault zone. Internal
slip surfaces separate lenses of reservoir sandstones. Thin shale and coal source layers have been smeared for some distance along some
of the faults. (a) A fault with a throw of about 20 m (66 ft) juxtaposes reservoir sandstone units of the Lower Jurassic Tilje Formation.
(b) A fault with a throw of about 40 m (131 ft) juxtaposes Lower Jurassic (Tilje Formation) and Upper Triassic (Are Formation) reservoir
sandstones.

(5) and sandstone (6) that are enclosed in a matrix derived


from shale and coal layers. This observation of entrainment of various lithologies by mechanical wear have the
characteristics of the abrasion-type SGR algorithm of
Yielding et al. (1997). However, this example demonstrates that fault seal analysis applying the SGR algorithm, inferring a single slip surface, will not represent
the actual across-fault situation, and calculated SGR
values will not represent the fault seal capacity. Like
the Njord field faults, the actual sealing capacity will be
enhanced because of a large number of slip surfaces
in combination with cementation and the smear from
thin shale and coal layers along intrafault zone faults.
Workers commonly refer to SGR values as a measure of the sealing capacity of faults, and that increasing SGR should equate to increasing seal capacity. However, as discussed above, application of the SGR method

requires an understanding of the likely errors. We do


agree that high shale content in the fault zone is favorable for the sealing capacity, but we also emphasize the
importance of other factors that have to be considered to
quantify the risk associated with a seal for fault-bounded
prospects. The next section describes a methodology
that aims to include the factors most important in the
risking of fault seal capacity.

METHODOLOGY TO RISK FAULT SEAL CAPACITY


With knowledge of the fault displacement and sediment
distribution within a prospect, it is possible to predict the
juxtaposition across fault zones, the likely association of
fault rocks (Figure 1), and to risk the fault seal capacity.
The risking of the fault seal has to be conducted for the
Frseth et al.

1239

Figure 5. A fault zone at Hartley Steps, Northumberland, United Kingdom, offsets an Upper Carboniferous sequence with sand, silt,
shale, and coal layers. The cumulative offset of 17 m (55 ft) is partitioned onto several slip surfaces with throw equal to 6 m (19 ft)
down to the centimeter scale over a 10-m (33-ft)-wide zone. The fault with the largest throw has a maximum width of 1 m (3.3 ft) in
the outcrop and contains a sheared melange of various wall rocks (inset). The southeastern part of this zone is a 0.150.25-m (0.49
0.82-ft)-thick gouge bounded by slip surfaces, and the gouge is derived primarily from the shale and coal layers in the hanging wall
numbered 7, 8, and 9. The fault-bounded zone northwest of this gouge is a maximum of 0.8 m (2.62 ft) in thickness and is composed
of rotated and steeply inclined lenses of shale with ironstone bands (5) and sandstone (6) that are enclosed in a matrix derived from
shale and coal layers.

prognosed height of the hydrocarbon column. The methodology we introduce encompasses most aspects of fault
seal assessments where normal faults offset a sand-shale
sequence, with major fault blocks as potential hydrocarbon traps (Figure 1). The fault seal diagram (Figure 6)
enables a rapid visual assessment of prospect risk and a
consistent fault risk to be incorporated into risking procedures. In the diagram, potential fault seal types are
confined within four main categories with respect to fault
seal probability (P FS), and a range in P FS represents each
category. Fault seals are associated with self-juxtaposed
and self-separated reservoirs, respectively (Figures 1, 6).
The two end members are represented by an upper P FS
(0.7 1.0) for a juxtaposition seal and a lower P FS (0.0
0.3) where clean reservoir sandstone is partially selfjuxtaposed across a fault. For these two categories, the
risk pertains to a single slip surface. The range 0.3 < P FS <
0.7 in the diagram represents self-separated reservoirs
with the apparent communication of sandstones across
a fault zone. A fault seal for these situations may result
from either the smear of a thick source layer (Figure 2)
1240

Geohorizons

or from the mixing of fault rocks within a complex fault


zone (Figure 3). For each of the four fault seal categories, parameters positive or detrimental to the quality
of the particular type of fault seal are listed (Figure 6).
The evaluation of these specific parameters as well as
the general fault-seal modifiers should be used to narrow the range in uncertainty regarding fault seal probability. The required data are routinely available in
prospect analysis.
The P FS values we use to establish the four main
fault seal categories are based on an in-house database.
The database contains postdrill analysis routinely done
to evaluate critical parameters (e.g., fault seal) by comparing the risk (P FS) assigned to the parameter before
drilling with the result from the drilling. Hence, the
database generated contains the systematic collection
of data to be used in the risking of fault seal capacity
for new prospects. The risking ranges in the fault seal
diagram (Figure 6) have been empirically calibrated,
applying faults where we have relatively good control
on the fault architecture as well as the sealing capacity

Figure 6. Aspects of fault seal


assessment for large faults in
mixed clastic sequences are
confined within four main categories with respect to overall
fault seal probability (P FS). The
two end members are represented by an upper P FS (0.7
1.0) for a juxtaposition seal and
a lower P FS (0.0 0.3) where
clean reservoir sandstone is
partly self-juxtaposed across a
fault. In these two categories,
the risk pertains to a single slip
surface. The range 0.3 < P FS <
0.7 in the diagram is associated
with self-separated reservoirs
where deformation processes
may create fault-rock membrane sealing. For each fault
seal condition given as a range
in P FS, parameters positive or
detrimental to the quality of the
particular type of fault seal are
listed. The evaluation of these
specific parameters as well as
several general fault seal
modifiers (lower right) should
be used to narrow the range in
uncertainty.

by recorded column heights and pressure differences


across faults. Postdrill analyses are available for more
than 100 fault-bounded prospects that include footwall and hanging-wall traps (Figure 1) and traps containing hydrocarbons as well as failures. Data are mainly
from the Norwegian shelf, and traps are related to Jurassic extension and normal faults that offset Jurassic
Triassic siliciclastic sediments (Figures 2a, 3a, 4). The
traps that contain hydrocarbons represent a wide range
of column heights and underfilled traps, as well as traps
filled to spill. The risking ranges we use as a basis for

predrill fault seal analysis have proved to be very applicable offshore Norway. However, workers involved in
the risking of fault seal and having sufficient empirical
data from other rift basins might find it necessary to
adjust risking ranges because of basin specifics.
Fault Seal Modifiers
Factors such as maximum burial, thickness of the fault
zone, area of critical juxtaposition, in-situ principal
stress, and fault reactivation are important with respect
Frseth et al.

1241

to fault seal capacity and should be considered in the


risking process. Maximum burial is very influential,
and increasing efficiency of fault seals with depth has
been reported from several basins (e.g., Hindle, 1989;
Knott, 1993; Gibson, 1994; Hesthammer et al., 2002;
Sperrevik et al., 2002; Yielding, 2002; Bretan et al.,
2003). The reduction of pore-throat radii and increased
capillary entry pressure for fault membranes with increasing depth may result from increased confining
pressure and mechanical and chemical compaction.
With a burial of greater than or equal to 3000 m (9800 ft)
and more, temperatures exceed 90jC, and quartz cementation may occur to increase the sealing capacity
of faults (e.g., Fisher and Knipe, 2001). Because other
factors than smear may contribute to the net fault seal
capacity when sediments are deeply buried, a fault at a
deep level should be assigned a higher fault seal probability than a fault at shallow level, although they provide the same SGR.
An increased capillary entry pressure of a fault zone
and the extent to which it retards fluid flow is interpreted by several researchers to be positively correlated
with the thickness of the zone (e.g., Scholz, 1987; Evans,
1990; Knott, 1994; Childs et al., 1997; Sperrevik et al.,
2002). The thickness of the fault zone may become important, especially in cases without the smear from a
thick source layer, where several major slip surfaces and
a mixture of fault rocks within the fault zone may enhance the fault seal capacity (Figures 3 5).
The area of self-juxtaposed reservoir sandstone or the
area of juxtaposition of the reservoir against other permeable lithologies over the fault surface (self-separated
reservoir) above the hydrocarbon-water contact is a critical factor in the prospect evaluation (e.g., Bailey et al.,
2006). With several areas with sand-sand juxtaposition,
some situations may represent candidates for fatal leakage. Hence, a large area of sand-sand juxtaposition with
several potential leak windows should be regarded as
higher risk in fault seal assessments than faults where
the critical juxtaposition is restricted to a small area, provided that the minimum SGR values and burial depths
are comparable for the two situations.
Fault reactivation or inversion has a complex influence on sealing behavior, with both area enhancement
and breakdown occurring (e.g., Jones and Hillis, 2003).
The orientation and magnitude of the in-situ stress, as
well as pore pressure and fault geometry, should be used
to assess the likelihood of reactivation of critical faults
(Bailey et al., 2006). Critically stressed faults, i.e., faults
that are close to reactivation within the current stress
system, are more likely to be conductive to hydrocarbons.
1242

Geohorizons

Noncritically stressed faults are more likely to remain


sealed (e.g., Bolas and Hermanrud, 2002; Wiprut and
Zoback, 2002; Jones and Hillis, 2003).

0.7 < P FS < 1.0 Juxtaposition Seal


Faults are likely to seal if they juxtapose reservoir sandstone against an impermeable lithology across a single
slip surface (Hubert, 1953; Smith, 1980; Watts, 1987;
Nybakken, 1991; Knott, 1993; Gibson, 1994), and a
juxtaposition seal (Figure 1) is generally assigned a high
fault seal probability (Figure 6). However, the assumed
impermeable lithology across the fault may contain thin
and localized permeable layers to allow leakage across
the fault if the fault plane in itself is not a barrier to fluid
flow. Because the assumption of a single fault plane to
give a juxtaposition seal is based on seismic mapping, it
involves an uncertainty caused by limitations in seismic
resolution. Accordingly, the probability of a fault seal
should rarely be assigned a P FS value of 1.0. With indications of a more complex fault architecture (e.g., from
wells penetrating large faults in nearby areas) and the
possibility of localized sand-sand communication across
the fault, the probability of a fault seal decreases significantly to be assigned a range of 0.30.6 (Figure 6).

0.5 < P FS < 0.7 Clay-Shale Smear


Fault zones where the fault rock is a smear from a thick
source layer (Figures 2, 6) consistently have high capillary
entry pressures and provide good seals (Smith, 1980;
Bouvier et al., 1989; Gibson, 1994; Frseth, 2006). The
main uncertainty associated with a seal is the spatial
continuity of the smear. With an SSF equal to or smaller
than 4, there is a high probability of forming an intact
layer of sealing shale smear along the fault (Frseth,
2006). With a positive evaluation of parameters considered important to provide a continuous shale membrane
(Figure 6), we have experienced that the likelihood of
fault seal will approach an upper P FS value of 0.7. With
indications that the uppermost fault segment is offset
toward the footwall (seismic resolution issue; Koldoye
et al., 2003), it results in a contractional dip relay (Peacock and Zhang, 1993; Childs et al., 1996; Rykkelid and
Fossen, 2002). Although a thick source layer was involved in the faulting, this situation will have a reduced
smear potential because the shale may be faulted or
squeezed out of the subhorizontal part of the fault zone
(e.g., Koldoye et al., 2003; Frseth, 2006). Accordingly,

this situation has a reduced smear potential, and the


fault seal probability is reduced to the range 0.3 0.6
(Figure 6).
0.3 < P FS < 0.6 Possibility of Leak Windows Caused by
Sand-Sand Juxtaposition
For this risking range, seismic interpretation has excluded
the possibility of a juxtaposition seal or a self-juxtaposed
reservoir (Figure 6). In addition, the spatially continuous
smear from a thick source layer is considered unlikely.
In the fault seal diagram, three fault seal scenarios represent this range in P FS. The risk for prospectivity is
caused primarily by the possibility of leak windows because of sand-sand communication across fault zones resulting from sand lenses entrained within the fault zone
(Figure 4). Because the three fault seal scenarios representing this risking range have several common characteristics, the risking procedure will be largely the same,
and important features are as follows:
A continuous clay-shale membrane from a single
source layer is unlikely because of either the lack of a
thick source layer, large SSF (>5) to result in a discontinuous smear, or the presence of a contractional dip
relay.
 Smear may result from relatively thin multiple shale
units or other lithologies with smear potential such
as coal, micaceous sand, silt, and carbonates (Frseth,
2006). With such lithologies present within the faulted
stratigraphy, smears supplied from source beds both
in the footwall and hanging wall may merge into
a composite smear. With smear from thin multiple
units, the possibility of leaky points increases as compared to smear from a single and thick source interval,
although the two situations have similar SGR values.
 Although sand lenses are entrained into a fault zone,
the lenses may be enclosed in a matrix of fine-grained
material (Figures 4, 5) in which case, cross-leakage
becomes less likely.
 If the faulting process is likely to generate fault rocks
by cementation and diagenesis or cataclasis, this will
enhance the fault seal capacity.


To reduce uncertainty in the range 0.3 < P FS < 0.6,


parameters positive or detrimental to fault seal capacity
are listed and should be evaluated (Figure 6). The maximum burial depth and the SGR are considered the most
influential factors. Shale gouge ratio values of 0.25%
or more may indicate the probability of a seal, whereas
SGR values less than 0.20%, in general, point to a mini-

mal sealing capacity. For a specific SGR value, the fault


seal capacity increases if the temperature exceeds 90jC,
i.e., a burial of 3000 m (9842 ft) or more with the possibility of quartz cementation.
0.0 < P FS < 0.3 Self-Juxtaposed Reservoir
A range of 0 0.3 is assigned for the fault seal probability where clean reservoir sandstone is partially selfjuxtaposed across a single slip surface (Figures 1, 6)
and the rocks are buried at shallow depths (<2500 m;
<8200 ft). As a result of mechanical shearing with grain
reorganization and denser grain packing, as well as diagenetic reactions, a fault plane will be less permeable
than the adjacent host rock (e.g., Bjrlykke, 1999). Fault
seal processes, likely to act under such conditions, are
termed disaggregation zones in clean sandstones
( < 15% clay) and phyllosilicate smearing, where the
amount of phyllosilicates are greater than 15% (Knipe,
1997; Knipe et al., 1997; Fisher and Knipe, 1998, 2001).
Offshore Norway, faults at shallow depth that selfjuxtapose relatively clean reservoir sandstones are, in
several places, known to contain leak windows that allow cross-fault communication (e.g., Knott, 1993; Fristad
et al., 1997). Gibson (1994) presented a similar conclusion from the Columbus Basin, offshore Trinidad, where
faults that juxtapose a sandstone unit against itself do
not significantly limit fluid communication during migration and entrapment.

CONCLUSIONS
In the industry, a recurring problem is to quantify the
risk associated with a seal for fault-bounded traps. The
risking of fault seal capacity is the probability that the
fault zone material is capable of sealing a specific hydrocarbon column. Several factors are known to influence
the fault seal capacity, and the architecture of fault zones
can be highly variable, both along individual faults and
between faults. However, in the risking of fault seal
capacity, the risk pertains to the location along the fault
that has the architecture most critical with respect to
leakage across the fault. The fault seal probability assigned to this specific part of the fault should reflect
the height of the hydrocarbon column to be sealed by
the fault as well as the prognosed type of hydrocarbon.
In the methodology we introduce, all aspects of fault
seal related to extensional settings are confined within
four main risk categories with respect to fault seal probability (P FS), and based on an empirical database, each
Frseth et al.

1243

category is represented by a range in P FS. The methodology allows comparable criteria to be applied in the
risking procedure and reduce uncertainty in fault seal
assessments. As a foundation for the methodology, we
provide data from large faults drilled and cored offshore
Norway and from outcrops in study areas onshore. We
combine onshore and offshore data to illustrate architecture of large faults and to demonstrate how architecture
and the distribution of fault rocks may influence sealing
capacity.
Our fault data represent different sequences, rheological properties, and deformation conditions. However, despite the variable and complex structure of fault
zones, we have observed fault zone characteristics that
appear in common to the faults investigated, and we
consider these factors to be important in the risking of
fault seal predictions. The fault zones in our database,
typically bounded by external slip surfaces, represent two
main categories: (1) a layer of shale smear entrained into
the fault zone and derived from a thick shale source layer
within the sequence offset by the fault; and (2) fault
zones characterized by internal slip surfaces, slivers
of footwall- and hanging-wallderived material rotated
along the fault zone and commonly enclosed in a matrix of shaly-silty fault gouge.
In the industry, the abrasion-type SGR algorithm
is currently applied, and workers commonly refer to a
minimum SGR value as a measure of the sealing capacity of the fault. This study highlights the disparity
between the complexity of actual faults and the SGR
algorithm, which assumes that the seismically derived
throw is concentrated to a single fault plane. We demonstrate how this may influence sequence juxtaposition across a fault, the associated SGR values, and ultimately, the fault seal risking.

REFERENCES CITED
Aarland, R. K., and J. Skjerven, 1998, Fault and fracture characteristics of a major fault zone in the northern North Sea: Analysis
of 3D seismic and oriented cores in the Brage field (Block 31/4),
in M. P. Coward, T. S. Daltaban, and H. Johnson, eds., Structural
geology in reservoir characterization: Geological Society (London) Special Publication 127, p. 209 229.
Antonellini, M., and A. Aydin, 1994, Effect of faulting on fluid flow
in porous sandstones: Petrophysical properties: AAPG Bulletin,
v. 78, p. 355 377.
Aydin, A., 1978, Small faults formed as deformation bands in
sandstones: Pure and Applied Geophysics, v. 116, p. 913 930.
Aydin, A., and Y. Eyal, 2002, Anatomy of a normal fault with shale
smear: Implications for fault seal: AAPG Bulletin, v. 86, p. 1367
1381.
Bailey, W. R., J. Underschulz, D. N. Dewhurst, G. Kovack, S.

1244

Geohorizons

Mildren, and M. Raven, 2006, Multi-disciplinary approach to


fault and top seal appraisal; Pyrenees-Macedon oil and gas fields,
Exmouth Sub-basin, Australian Northwest Shelf: Marine and
Petroleum Geology, v. 23, p. 241 259.
Berg, R. R., and A. H. Avery, 1995, Sealing properties of Tertiary
growth faults, Texas Gulf Coast: AAPG Bulletin, v. 79, no. 3,
p. 375 393.
Bjrlykke, K., 1999, An overview of factors controlling rates of
compaction and fluid flow in sedimentary basins, in B. Jamtveit
and P. Meakin, eds., Growth, dissolution and pattern formation
in geosystems: Dordrecth, Kluwer, p. 381 404.
Bolas, H. M. N., and C. Hermanrud, 2002, Rock stresses in sedimentary basins Implications for trap integrity, in A. G. Koestler
and R. Hunsdale, eds., Hydrocarbon seal quantification: Norwegian Petroleum Society (NPF ) Special Publication 11, p. 17
35.
Bouvier, J. D., C. H. Kaars-Sijpesteijn, D. F. Kluesner, C. C.
Onyejekwe, and R. C. Van der Pal, 1989, Three-dimensional
seismic interpretation and fault sealing investigations, Nun River
field, Nigeria: AAPG Bulletin, v. 73, p. 1397 1414.
Bretan, P., G. Yielding, and H. Jones, 2003, Using calibrated shale
gouge ratio to estimate hydrocarbon column heights: AAPG
Bulletin, v. 87, p. 397 413.
Buhrig, C., 1989, Geopressured Jurassic reservoirs in the Viking
Graben: Modeling and geological significance: Marine and Petroleum Geology, v. 6, p. 31 48.
Caine, J. S., J. P. Evans, and C. B. Foster, 1996, Fault zone architecture
and permeability structure: Geology, v. 24, p. 1025 1028.
Childs, C., A. Nichol, J. J. Walsh, and J. Watterson, 1996, Growth
of vertically segmented normal faults: Journal of Structural
Geology, v. 18, p. 1389 1397.
Childs, C., J. J. Walsh, and J. Watterson, 1997, Complexity in fault
zone structure and implications for fault seal prediction, in P.
Mller-Pedersen and A. G. Koestler, eds., Hydrocarbon seals:
Importance for exploration and production: Norwegian Petroleum Society (NPF) Special Publication 7, p. 61 72.
Dart, C. J., K. McClay, and P. N. Hollings, 1995, 3D analysis of
inverted extensional fault systems, southern Bristol Channel
basin, UK, in J. G. Buchanan and P. G. Buchanan, eds., Basin
inversion: Geological Society (London) Special Publication 88,
p. 393 413.
Dewhurst, D. N., R. M. Jones, R. R. Hillis, and S. D. Mildren, 2002,
Microstructural and geomechanical characterisation of fault
rocks from the Carnarvon and Otway basins: Australian
Petroleum Production and Exploration Association Journal,
v. 42, p. 167 186.
Doughty, P. T., 2003, Clay smear seals and fault sealing potential of
an exhumed growth fault, Rio Grande rift, New Mexico: AAPG
Bulletin, v. 87, p. 427 444.
Evans, J. P., 1990, Thickness-displacement relationships for fault
zones: Journal of Structural Geology, v. 12, p. 1061 1065.
Frseth, R. B., 1996, Interaction of Permo-Triassic and Jurassic extensional fault blocks during the development of the northern
North Sea: Journal of the Geological Society (London), v. 153,
p. 931 944.
Frseth, R. B., 2006, Shale smear along large faults: Continuity of
smear and the fault seal capacity: Journal of the Geological
Society (London), v. 163, p. 741 751.
Frseth, R. B., K. A. Oppeben, and A. Sbe, 1984, Trapping
styles and associated hydrocarbon potential in the Norwegian
North Sea, in M. T. Halbouty, ed., Future petroleum provinces
of the world: AAPG Memoir 40, p. 585 597.
Frseth, R. B., T. S. Sjblom, R. J. Steel, T. Liljedahl, B. E. Sauar,
and T. Tjelland, 1995, Tectonic controls on Bathonian Volgian
syn-rift successions on the Visund fault block, northern North
Sea, in R. J. Steel, V. Felt, E. Johannesen, and C. Mathieu, eds.,

Sequence stratigraphy of the north-west European margin:


Norwegian Petroleum Society (NPF) Special Publication 5,
p. 467 490.
Fisher, Q. J., and R. J. Knipe, 1998, Fault sealing processes in siliciclastic sediments, in G. Jones, Q. Fisher, and R. J. Knipe, eds.,
Faulting, fault sealing and fluid flow in hydrocarbon reservoirs:
Geological Society (London) Special Publication 148, p. 117
134.
Fisher, Q. J., and R. J. Knipe, 2001, The permeability of faults
within siliciclastic petroleum reservoirs of the North Sea and
Norwegian continental shelf: Marine and Petroleum Geology,
v. 18, p. 1063 1081.
Foxford, K. A., J. J. Walsh, J. Watterson, I. R. Garden, S. C. Guscott,
and S. D. Burley, 1998, Structure and content of the Moab fault
zone, U.S.A., and its implications for fault prediction, in G.
Jones, Q. J. Fisher, and R. J. Knipe, eds., Faulting, fault sealing
and fluid flow in hydrocarbon reservoirs: Geological Society
(London) Special Publication 148, p. 87 103.
Fristad, T., A. Groth, G. Yielding, and B. Freeman, 1997, Quantitative fault seal prediction: A case study from Oseberg Syd,
in P. Mller-Pedersen and A. G. Koestler, eds., Hydrocarbon
seals: Importance for exploration and production: Norwegian
Petroleum Society (NPF) Special Publication 7, p. 107 124.
Fulljames, J. R., L. J. J. Zijerveld, R. C. M. W. Franssen, G. M.
Ingram, and P. D. Richard, 1997, Fault seal processes, in P.
Mller-Pedersen and A. G. Koestler, eds., Hydrocarbon seals:
Importance for exploration and production: Norwegian Petroleum Society (NPF) Special Publication 7, p. 51 79.
Gaarenstroom, L., R. A. J. Tromp, M. C. de Jong, and A. M.
Brandenburg, 1993, Overpressure in the central North Sea:
Implications for trap integrity and drilling safety, in J. P. Parker,
ed., Petroleum geology of northwest Europe: Proceedings of the
4th Conference: Geological Society (London), p. 1305 1313.
Gabrielsen, R. H., and A. G. Koestler, 1987, Description and structural implications of fractures in Late Jurassic sandstones of the
Troll field, northern North Sea: Norsk Geologisk Tidsskrift,
v. 67, p. 371 381.
Gibson, R. G., 1994, Fault-zone seals in siliciclastic strata of the
Columbus Basin, offshore Trinidad: AAPG Bulletin, v. 78,
p. 1372 1385.
Gibson, R. G., 1998, Physical character and fluid-flow properties of
sandstone-derived fault zones, in M. P. Coward, T. S. Daltaban,
and H. Johnson, eds., Structural geology in reservoir characterization: Geological Society (London) Special Publication 127,
p. 83 97.
Hage, A., K. Bomstad, and J. E. Strand, 1987, Brage field, in A. M.
Spencer, ed., Geology of the Norwegian oil and gas fields:
Norwegian Petroleum Society, 493 p.
Hesthammer, J., and H. Fossen, 1998, The use of dipmeter data to
constrain the structural geology of the Gullfaks field, northern
North Sea: Marine and Petroleum Geology, v. 15, p. 549 573.
Hesthammer, J., P. A. Bjrkum, and L. Watts, 2002, The effect of
temperature on sealing capacity of faults in sandstone reservoirs: Examples from the Gullfaks and Gullfaks Sr fields,
North Sea: AAPG Bulletin, v. 86, p. 1733 1751.
Hindle, A. D., 1989, Downthrown traps of the NW Witch Ground
Graben, UK North Sea: Journal of Petroleum Geology, v. 12,
p. 405 418.
Hubert, M. K., 1953, Entrapment of petroleum under hydrodynamic conditions: AAPG Bulletin, v. 37, p. 1954 2026.
Jamison, W. R., and D. W. Stearns, 1982, Tectonic deformation of
Wingate Sandstone, Colorado, National Monument: AAPG
Bulletin, v. 66, p. 2584 2608.
Jev, B. I., C. H. Kaars-Sijpesteijn, M. P. A. M. Peters, N. L. Watts,
and J. T. Wikie, 1993, Akaso field, Nigeria: Use of integrated
3-D seismic, fault slicing, clay smearing, and RFT pressure data

on fault trapping and dynamic leakage: AAPG Bulletin, v. 77,


p. 1389 1404.
Jones, J. M., 1967, The geology of the coast section from Tynemouth
to Seaton Sluice: Transactions of the National Historical Society
of Northumbria, v. 16, p. 153 192.
Jones, R. M., and R. R. Hillis, 2003, An integrated, quantitative
approach to assessing fault-seal risk: AAPG Bulletin, v. 87,
p. 507 524.
Kim, J.-W., R. R. Berg, J. S. Watkins, and T. T. Tieh, 2003,
Trapping capacity of faults in the Eocene Yegua Formation,
East Sour Lake field, southeast Texas: AAPG Bulletin, v. 87,
p. 415 425.
Knipe, R. J., 1992, Faulting processes and fault seal, in R. M. Larsen,
H. Brekke, B. T. Larsen, and E. Talleraas, eds., Structural and
tectonic modeling and its application to petroleum geology: Norwegian Petroleum Society (NPF) Special Publication 1, p. 325
342.
Knipe, R. J., 1997, Juxtaposition and seal diagrams to help analyze
fault seals in hydrocarbon reservoirs: AAPG Bulletin, v. 81,
p. 187 195.
Knipe, R. J., Q. J. Fisher, G. Jones, M. R. Clennell, A. B. Farmer, A.
Harrison, B. Kidd, E. McAllister, J. R. Porter, and E. A. White,
1997, Fault seal analysis: Successful methodologies, application and future directions, in A. G. Koestler and R. Hunsdale,
eds., Hydrocarbon seal quantification: Norwegian Petroleum
Society (NPF) Special Publication 11, p. 15 40.
Knott, S. D., 1993, Fault seal analysis in the North Sea: AAPG
Bulletin, v. 77, p. 778 792.
Knott, S. D., 1994, Fault zone thickness versus displacement in the
Permo-Triassic sandstones of NW England: Journal of the Geological Society (London), v. 151, p. 17 25.
Koldoye, B. A., A. Aydin, and E. May, 2003, A new process-based
methodology for analysis of shale smear along normal faults in
the Niger Delta: AAPG Bulletin, v. 87, p. 445 463.
Lewis, G., R. J. Knipe, and A. Li, 2002, Fault seal analysis in unconsolidated sediments: A field study from Kentucky, U.S.A.,
in A. G. Koestler and R. Hunsdale, eds., Hydrocarbon seal quantification: Norwegian Petroleum Society (NPF) Special Publication 11, p. 243 253.
Lindsay, N. G., F. C. Murphy, J. J. Walsh, and J. Watterson, 1993,
Outcrop studies of shale smears on fault surfaces: International Association of Sedimentologists Special Publication 15,
p. 113 123.
Moustafa, A. R., 1987, Drape folding in the Baba-Sidri area, eastern
side of the Suez rift, Egypt: Journal of Geology, v. 31, p. 15
27.
Nybakken, S., 1991, Sealing fault traps An exploration concept in
a mature petroleum province: Tampen Spur, northern North
Sea: First Break, v. 9, p. 209 222.
Peacock, D. C., and D. J. Sanderson, 1991, Displacements, segment
linkage and relay ramps in normal fault systems: Journal of
Structural Geology, v. 13, p. 721 733.
Peacock, D. C. P., and X. Zhang, 1993, Field examples and numerical modeling of oversteps and bends along normal faults in
cross-section: Tectonophysics, v. 234, p. 147 167.
Pittman, E. D., 1981, Effects of fault-related granulation on porosity
and permeability of quartz sandstones, Simpson Group (Ordovician), Oklahoma: AAPG Bulletin, v. 65, p. 2381 2387.
Robson, D. A., 1971, The structure of the Gulf of Suez (Clysmic)
rift, with special reference to the eastern side: Journal of the
Geological Society (London), v. 8, p. 359 369.
Rykkelid, E., and H. Fossen, 2002, Layer rotation around vertical
fault overlap zones: Observations from seismic data, field examples, and physical experiments: Marine and Petroleum Geology,
v. 19, p. 181 192.
Sassi, W., S. E. Livera, and B. P. R. Caline, 1992, Reservoir

Frseth et al.

1245

compartmentalization by faults in Cormorant Block IV, U. K.,


northern North Sea, in R. M. Larsen, H. Brekke, B. T. Larsen,
and E. Talleraas, eds., Structural and tectonic modeling and its
application to petroleum geology: Norwegian Petroleum Society (NPF) Special Publication 1, p. 355 364.
Scholz, C. H., 1987, Wear and gouge formation in brittle faulting:
Geology, v. 15, p. 493 495.
Sharp, I. R., R. L. Gawthorpe, J. R. Underhill, and S. Gupa, 2000,
Fault-propagation folding in extensional settings: Examples of
structural style and synrift sedimentary response from the Suez
rift, Sinai, Egypt: Geological Society of America Bulletin,
v. 112, p. 1877 1899.
Skerlec, G. M., 1996, Risking fault seal in the Gulf Coast (abs.):
AAPG Annual Meeting Program, v. 5, p. A131.
Smith, D. A., 1980, Sealing and non-sealing faults in Louisiana Gulf
Coast salt basin: AAPG Bulletin, v. 64, p. 145 172.
Speksnijder, A., 1987, The structural configuration of Cormorant
Block IV in context of the northern Viking Graben structural
framework: Geolgie Mijenbouw, v. 65, p. 357 379.
Sperrevik, S., R. B. Frseth, and R. H. Gabrielsen, 2000, Experiments on clay smear formation along faults: Petroleum Geoscience, v. 6, p. 113 123.
Sperrevik, S., P. A. Gillespie, Q. J. Fisher, T. Halvorsen, and R. J.
Knipe, 2002, Empirical estimation of fault rock properties, in
A. G. Koestler and R. Hunsdale, eds., Hydrocarbon seal quantification: Norwegian Petroleum Society (NPF) Special Publication 11, p. 109 125.
Sverdrup, E., J. Helgesen, and J. Vold, 2003, Sealing properties of

1246

Geohorizons

faults and their influence on water-alternating-gas injection


efficiency in the Snorre field, northern North Sea: AAPG Bulletin, v. 87, p. 1437 1458.
Walsh, J. J., J. Watterson, A. Heath, and C. Childs, 1998, Representation and scaling of faults in fluid flow models: Petroleum
Geosciences, v. 4, p. 241 251.
Watts, N. L., 1987, Theoretical aspects of cap-rock and fault seals
for single- and two-phase hydrocarbon columns: Marine and
Petroleum Geology, v. 4, p. 274 307.
Wiprut, D., and M. D. Zoback, 2002, Fault reactivation, leakage
potential, and hydrocarbon column heights in the northern
North Sea, in A. P. Koestler and R. Hunsdale, eds., Hydrocarbon seal quantification: Norwegian Petroleum Society (NPF)
Special Publication 11, p. 2003 2019.
Yielding, G., 2002, Shale gouge ratio calibration by geohistory, in
A. G. Koestler and R. Hunsdale, eds., Hydrocarbon seal quantification: Norwegian Petroleum Society (NPF) Special Publication 11, p. 1 15.
Yielding, G., B. Freeman, and D. P. Needham, 1997, Quantitative
fault seal prediction: AAPG Bulletin, v. 81, p. 897 917.
Yielding, G., J. A. verland, and G. Byberg, 1999, Characterization
of fault zones for reservoir modelling: An example from the
Gullfaks field, northern North Sea: AAPG Bulletin, v. 6, p. 925
951.
Zieglar, D. L., 1992, Hydrocarbon columns, buoyancy pressures,
and seal efficiency: Comparisons of oil and gas accumulations
in California and the Rocky Mountain area: AAPG Bulletin,
v. 76, p. 501 508.

You might also like