Davies 2001

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Sequential dip-slip fault movement during rifting: a new model for the

evolution of the Jurassic trilete North Sea rift system


R. J. Davies1, J. D. Turner2 and J. R. Underhill2
1
ExxonMobil Exploration Co., 233 Benmar, Houston, Texas 77060, USA
(e-mail: richard_davies@email.exxonmobil.com)
2
Department of Geology and Geophysics, Grant Institute, The University of Edinburgh, West Mains Road,
Edinburgh EH9 3JW, UK (e-mail: jturner@glg.ed.ac.uk/jru@glg.ed.ac.uk)

ABSTRACT: Controversy has long surrounded the kinematics of faulting in the


Middle–Late Jurassic North Sea trilete rift system. Integration of structural styles and
subsidence analysis derived from well-constrained seismic interpretation enables a
new, unified model to be proposed in which strike-slip was negligible, dip-slip
extension predominated throughout the rifting episode and normal faults were active
sequentially not synchronously. Extension was initiated on N–S and NNE–SSW
trending faults during the Bathonian and Callovian, NE–SW and E–W structures
during the Oxfordian and NW–SE faults during the Kimmeridgian and Volgian. The
results allow us to speculate that fault activity was driven by variations in the
prevailing far-field stress regime that were superimposed upon a trilete junction that
formed as consequence of Middle Jurassic thermal doming. Significantly, rotation
of the stress field during rifting is similar in other rifts, such as the Afro-Arabian
system.

KEYWORDS: stress, sequence, kinematics, dip-slip fault, rotation (geology)

INTRODUCTION contemporaneously may in fact be flawed and that sequential


It is generally accepted that active extension in the North Sea fault activity occurred instead. The systematic variation in the
(Fig. 1) took place during the Permo-Triassic and during the timing of initiation of fault movement during rifting appears to
Middle and Late Jurassic (Roberts et al. 1995; Faerseth 1996) have depended upon the orientation of the faults themselves.
and that the formation of the North Sea trilete rift system This is an important part of a new model we propose, in which
during the latter event was preceded by the development of a Middle and Late Jurassic rifting started with the development of
Middle Jurassic thermal dome (Ziegler 1990; Underhill & a thermal dome and the formation of a trilete junction centred
Partington 1993). The rift itself is arguably one of the most on the dome, followed by progressive sequential dip-slip
studied ones in the world, with over 2000 well penetrations and movement on N–S/NNE–SSW, NE–SW, E–W and NW–SE
a comprehensive coverage of 2D and 3D seismic data. Despite intra-rift extensional faults. We finally speculate as to what
this, no agreement has been reached with regard to the process may have driven the sequential faulting.
kinematics of Middle and Late Jurassic rifting.
Middle to Late Jurassic rifting was essentially accommodated
by movement on four fault sets (Fig. 1). An explicit or implicit PREVIOUS STUDIES
assumption of all of the recent models that purport to explain Interpretation of subsurface data suggests that Jurassic rifting
the kinematics of the system is that all fault sets in the area was initiated during the Bathonian in the Viking Graben, with
around the triple junction were active at the same time. Making the remainder of the Jurassic being dominated by active faulting
this assumption has forced many of these models, whether in all three rift arms. During this extensional episode a
proposing a constant or changing extension direction, to first-order transgression took place over Northwest Europe
include oblique-slip activity for one or more of the active fault (Jacquin et al. 1998). This led to the retreat of extensive Middle
sets. This is fundamental, as this interpretation underpins what Jurassic delta plain and shoreface systems due, in part, to the
we believe is an important misconception that still exists within subsidence generated by the deflation of the North Sea dome
the published literature regarding Middle and Late Jurassic rift (Underhill & Partington 1993). Marine environments were
development, namely that some or all of the faults were the established throughout the North Sea and this period was
result of either oblique-slip or strike-slip fault movement. dominated by the deposition of shallow marine shoreface sands
In this paper we use 2D and 3D seismic and well data from and associated sediments in a shelf setting (Fig. 2; Rattey &
the UK sector of the North Sea triple junction area to propose Hayward 1993). During the Kimmeridgian and Volgian the
a new unified model for the post-dome rifting phase. We shoreface systems were replaced by turbidites and debris flows,
do this by first examining the validity of the evidence for after rifting had resulted in the development of significant
oblique-slip and/or strike-slip extension rather than simply bathymetric relief (Rattey & Hayward 1993). Extension ceased
dip-slip extension. We then put forward stratigraphic data that in the Ryazanian and was followed by post-rift subsidence
suggest that the assumption that all fault sets were active during the Cretaceous and Cenozoic.

Petroleum Geoscience, Vol. 7 2001, pp. 371–388 1354-0793/01/$15.00  2001 EAGE/Geological Society of London
372 R. J. Davies et al.

Fig. 1. A structure map of the


North Sea trilete rift system (after
Bartholomew et al. 1993) showing the
intersection of four fault populations
which strike in N–S/NNE–SSW,
NE–SW, E–W and NW–SE directions.

Fig. 2. Stratigraphic column, including


(a) the Middle and Late Jurassic and
(b) the pre-Middle Jurassic.
IAU, Intra-Aalenian or Mid
Cimmerian unconformity; ICU,
Intra-Callovian unconformity;
IOU, intra-Oxfordian unconformity
(after Davies et al. 1999).

We focus on the development of the immediate triple fault orientations can be identified within the triple junction
junction area (Fig. 1) – the Moray Firth rift arm, the Central area, the axes of the three rift arms themselves strike N–S/
Graben (as far south as Quadrants 29 and 30) and the Viking NNE–SSW, E–W and NW–SE. All faults – except the NE–SW
Graben (as far north as Quadrant 9). Although four intra-rift faults of the Inner Moray Firth – are broadly parallel to one of
Jurassic rifting of the North Sea 373

Fig. 3. Seismic line over the Puffin high within the Central Graben showing a series of high-angle deep faults (marked X) cross-cutting the
Rotliegendes above which there are a series of lower angle faults cross-cutting the Mesozoic section (marked Y). Similar fault patterns elsewhere
within the Central Graben have been interpreted as strike-slip faults with high-angle faults at depth linked through a salt section to lower angle
faults within the carapace (Sears et al. 1993). This gives the appearance of a flower structure. In this particular example we interpret the
predominant sense of motion to be dip-slip with de-coupling between the deeper and shallower faults and potentially gravitational movement
of the Mesozoic carapace downdip from east to west, to the immediate left of the well penetration. Structural geometries are also controlled by
salt withdrawal and or dissolution (Z) above deeper Rotliegendes faults (Hodgson et al. 1992).

the three rift zones. Fault growth and linkage occurred during
extension and rifting within the North Sea (e.g. Dawers &
Underhill 2000; McLeod et al. in press). However, we aim to
examine the timing of faulting on the most prominent intra-rift
faults that dominate the structural architecture of the three rift
arms, disregarding minor faults in relay ramps and breached
relay ramps which may have been controlled by local stress
conditions.
Over the last decade the primary areas of focus for those that
have proposed that strike-slip and oblique-slip faulting were
important aspects of Jurassic rifting have been the Central
Graben and parts of the Moray Firth. The majority of the
supportive evidence presented has been 2D and 3D seismic
data acquired over the Central Graben (Bartholomew et al.
1993; Sears et al. 1993; Eggink et al. 1996). Both the Triassic and
Jurassic rift phases have affected sections containing Permian
Salt (Penge et al. 1993). Key lines of evidence for postulating
oblique-slip and strike-slip faulting were:
(1) the apparent anastomosing and, in places, rhombohedral
planform geometry of Jurassic faults, above what were
considered to be long-lived, repeatedly reactivated base-
ment shear zones;
(2) in cross-sectional view, the vertical linkage (through salt
cover in some cases) of the zones of basement shear Fig. 4. (a) Geoseismic section from the Central Graben (after
with splaying low-angle, shallower faults, dissecting the Helgesen 1999), that crosses (b) a series of NE–SW striking faults at
Mesozoic cover, with these being interpreted as flower base Cretaceous level and (c) N–S and E–W striking faults at
Rotliegendes level. The common interpretation has been to link the
structures; high-angle, sub-vertical deeper faults (marked X) to shallower faults
(3) the generation of apparent uplift and subsidence between that are of lower angle (45) – marked Y – giving apparent flower
the basement shear zones. geometries. This structural configuration can be accounted for
without making this linkage and Helgesen (1999) suggests that there
Additional supporting evidence was the apparent similarity is no direct hard linkage between the upper and lower faults.
between the planform geometry of the faults as imaged near the Helgesen’s interpretation is consistent with dip-slip extension within
base Cretaceous level and those created in sandbox modelling this part of the Central Graben.
374 R. J. Davies et al.

Fig. 5. 2D regional seismic line from the Inner Moray Firth that traverses a number of NE–SW striking faults (marked in red). Note the major
normal fault on the western side of the line, separating the Sutherland Terrace from the Beatrice half-graben. The Beatrice block is dissected by
a number of small offset planar normal faults (marked X). Note there is little or no clear evidence for rotation on the fault block during the
Callovian and Bathonian (between Top Triassic to Top A Sand seismic picks). Also note some evidence for syn-extensional thickening of the
J2.1–J2.3 interval, which gives an indication that extension was occurring during the Late Oxfordian to Early Volgian (marked Y). All the planar
normal faults are of Jurassic age. The flower structure centred on the Great Glen fault (marked Z) is much younger and has no effect on Jurassic
sediment thickness variations (Underhill 1991a, 1991b).

of oblique-slip faults (Eggink et al. 1996). The development of Coward 1995; Eggink et al. 1996; Davies et al. 1999; Erratt et al.
flower structures was considered to be the result of an obliquity 1999).
between interpreted NW–SE orientated basement lineaments
and what was considered to be an overall E–W extension
direction. In plan view the manifestation of the low-angle faults FAULT KINEMATICS ASSESSED FROM SEISMIC
at near base Cretaceous level was that of a series of linked fault DATA
traces that had a rhombohedral geometry.
These observations led to the interpretation that the Central The role of Permian salt
Graben was the result of oblique-slip faulting. A strike-slip In the Central Graben, Permian salt has had a profound effect
origin for the Inner Moray Firth originated from a regional on the structural and stratigraphic evolution of the Permo-
synthesis that made the assumption that rifting occurred at the Triassic and Jurassic successions. Where salt is present, two
same time in each rift arm and that the faults within the Central areas offer contrasting structural configurations: intra-rift
Graben and the Viking Graben were dip-slip normal faults highs/rift margins with thin salt sections and intra-rift basinal
(Roberts et al. 1990). For their kinematic model to work, lows with thick salt sections.
Roberts et al. (1990) had to advocate significant strike-slip Within the areas of thin salt, grounding of pods of Triassic
movement on one basin bounding fault of the Inner Moray sediments occurred probably due to loading, thin-skinned
Firth – the Great Glen Fault. Many local, field-scale studies extension and reactive and active diapirism; the salt commonly
have, in fact, determined that fault arrays were predominantly deformed into a complex series of salt walls (Wakefield et al.
dip-slip (see Abbots 1991). There have been fewer proponents 1993). Jurassic sediments were subsequently deposited above
of dip-slip fault activity and extension orthogonal to the rift the salt walls, in accommodation space generated by salt
zone (e.g. Underhill 1991a; Erratt 1993; Davies et al. 1999; evacuation and/or dissolution. This dynamic interplay between
Erratt et al. 1999). With the exception of the Inner Moray the underlying basement (top Rotliegendes) structure, salt
Firth, Roberts et al. (1990) were also strong advocates of deposition and subsequent loading and salt movement has been
predominantly dip-slip extension in the North Sea region. encapsulated within salt tectonic models, such as the ‘podology’
As a result of the uncertainty regarding fault kinematics, two model (Hodgson et al. 1992) involving low-angle faults that cut
schools of thought have developed with regard to the control- the Mesozoic section as a result of differential loading of the
ling extension direction. One advocates a constant extension Permian salt by Triassic sediments. Jackson & Vendeville
direction during Jurassic times (Speksnijder 1987; Badley et al. (1992) have also shown that salt walls and diapirs can develop
1988; Roberts et al. 1990; Ziegler 1990; Bartholomew et al. 1993; in similar settings due to syn-sedimentary thin-skinned exten-
Brun & Tron 1993; Faerseth 1996) and the other advocates a sion. Complex faults and folds may also have developed as a
changing extension direction (Doré & Gage 1987; Thomas & result of forced folding of the Mesozoic carapace above
Jurassic rifting of the North Sea 375

basement horsts (Erratt 1993). Irrespective of the exact struc- Permian salt and the overlying Triassic and Jurassic strata
tural model, it is clear that a range of complex structural (Fig. 4a). Mapping by others (Helgesen 1999) also shows there
configurations developed due to the influence of salt. Some of to be two separate, rather than linked fault sets. Helgesen
these configurations have been interpreted to be the result of mapped a lower set of high-angle faults that cut the basement
oblique-slip and strike-slip faulting. (top Rotliegendes) and a shallower series of planar faults that
cut the Mesozoic section. In Block 22/30 the faults that cut the
Palaeozoic strata have E–W and N–S orientations (Fig. 4b, c).
Dip-slip versus oblique-slip and strike-slip faulting In plan view, these deeper faults are oblique to the overlying
Although we do not provide an exhaustive structural interpret- Mesozoic faults. The shallower faults have an overall NW–SE
ation of the Central Graben, the two examples presented below orientation, but are arcuate in many places. They appear to
are within the central parts of the rift zone and are represen- control the geometry of the Upper Jurassic and Cretaceous
tative of the structural architectures seen elsewhere. Our successions that infill the minibasins that formed above the
interpretation considers the faults to be normal dip-slip faults rotated fault blocks (Fig. 4a). The Permian salt is of variable
that, in some cases, have listric geometries. Such examples have thickness and forms pillow and ‘detached’ pillow structures
been considered to be oblique-slip and strike-slip faults, inter- (Fig. 4a). The base of the Triassic is tilted and has grounded on
pretations that have underpinned the development of an the footwall highs of the underlying Rotliegendes fault blocks.
oblique-rifting model for the Central Graben (Bartholomew Helgesen’s (1999) interpretation of how this geometry devel-
et al. 1993; Sears et al. 1993). This, in turn, has been integrated oped is that Late Jurassic extension led to N–S and E–W
into a kinematic model for the triple junction as a whole striking faults at Rotliegendes level. Salt diapirism and evacua-
(Bartholomew et al. 1993). This model would have the three tion, along with movement on the shallower faults that cut the
rifts arms cross-cut by numerous strike-slip and oblique-slip Mesozoic section, led to the grounding and rotation of the
faults, with the North Sea being some kind of obliquely-rifted Triassic and Jurassic strata on the edges of the basement blocks
triple junction. and these acted as pivot points (Fig. 4a). In common with our
structural interpretation of the Puffin high, Helgesen (1999)
does not interpret the deep through-going sub-vertical faults as
Puffin intra-rift high connecting with shallower flower structures. Helgesen’s struc-
The Puffin high is an intra-rift high, located in the axis of the tural configuration is more consistent with dip-slip extension
Central Graben (Fig. 1). It is one structure used as evidence by than pervasive wrench faulting. We therefore suggest that it is
the proponents of oblique-slip and strike-slip faulting (Eggink possible to interpret both the Puffin and Helgesen examples
et al. 1996). On the Puffin high, a relatively thin Permian salt within the context of dip-slip rather than oblique-slip and
section separates the Mesozoic section from the Rotliegend strike-slip faulting. In both cases the fault sets are separated by
basement. Our mapping of 3D data over the Puffin high a salt unit with no through-going ‘hard’ linkage between the
suggests that the salt separates a deep, high-angle fault set that deep, high-angle faults and the shallower low-angle faults.
dissects the top Rotliegendes reflector from a shallower series Much of the structural complexity within the Puffin high can be
of low-angle faults (25–45) in the Mesozoic section (Fig. 3). attributed to the effect of salt wall formation and subse-
The low-angle faults are not aligned in this cross-sectional view quent salt dissolution or collapse, above faulted and tilted
with the underlying high-angle faults that cut the basement. We Rotliegendes blocks.
interpret these shallow faults to have formed due to gravita-
tionally driven movement of the Mesozoic carapace on the salt,
above an inclined Rotliegendes basement. We do not link the Inner Moray Firth
high-angle, deep faults with the low-angle, shallower faults. The Inner Moray Firth is a further area where contradictory
These and similar configurations imaged on the western margin accounts of the dominant fault kinematics exist (cf. Roberts
of the Central Graben (Stewart et al. 1999) are consistent with et al. 1990; Underhill 1991a; Frost & Rose 1996) and the debate
predominantly dip-slip movements both of the faults offsetting remains active (Argent et al. 2000; Cains 2000). One important
the basement and the faults in the Mesozoic strata (Fig. 3). difference with the Central Graben is the absence of a thick
Although there are no alignments of deep and shallow faults Cretaceous and Cenozoic post-rift sedimentary cover. This
and no mappable connecting fault planes between the two, improves seismic resolution, yielding the best imaging of
recent interpretations commonly show a single high-angle fault Jurassic faults and successions in the North Sea and clear
strand at depth, splaying to a series of lower angle faults in the examples of Late Jurassic syn-sedimentary growth faults
shallow section (Bartholomew et al. 1993; Sears et al. 1993; (Underhill 1991a, b).
Eggink et al. 1996). Such interpretations have resulted in The 2D seismic line we present shows a dissected half-
structural configurations similar to a classic flower structure, graben wedge (Fig. 5). This seismic line does, in contrast to the
that would be consistent with strike-slip and oblique-slip fault others we present (Figs 6, 8 and 9), show the characteristic
kinematics. The high-angle faults that cut the basement, some features of a classic flower structure that would support an
of which are the bounding faults to the Puffin high, were interpretation of strike-slip fault movement (Underhill &
considered to be sinistral oblique-slip faults by Eggink et al. Brodie 1993). The high-angle faults within the flower all link at
(1996). Bartholomew et al. (1993) and Sears et al. (1993) also depth to form a single fault trace that we interpret to be
maintained that the bounding faults of the Puffin high were essentially vertical. The component faults to the flower struc-
zones of basement shear. ture cut though the entire Mesozoic section. Although the
structure is a characteristic of strike-slip faulting, we interpret
Axis of the Central Graben the planar normal faults to have formed during the Jurassic and
Where the Permian salt is thicker, within intra-basinal lows for the flower structure to have had no role in Jurassic rifting,
example, a different structural configuration is seen. In the axis forming during Tertiary compression (Thomson & Underhill
of the Central Graben (Blocks 22/30 and 29/4 and 29/5) 1993; Underhill & Brodie 1993). Only where Jurassic faults
acquisition of 3D seismic data is now revealing the crucial have been inverted and reactivated during later Tertiary
relationship between the basement (top Rotliegendes), the compression (Thomson & Underhill 1993; Argent et al. 2000)
376 R. J. Davies et al.

this case the Great Glen Fault, to have had a major strike-slip
component of motion during the Jurassic. This interpretation
was challenged by Underhill (1991a) who advocated a minimal
component of strike-slip movement during the Jurassic and this
was also later the conclusion of Roberts & Holdsworth (1999).
Our interpretation of the Central and Moray Firth grabens is
that intra-rift faults are predominantly dip-slip in origin and not
through-going sub-vertical strike-slip or oblique-slip faults.
In the Moray Firth and the Viking Graben the role of
basement lineaments in controlling the exact location of
Mesozoic faults remains unclear. If a basement lineament
control were dominant in the Moray Firth or Viking Graben we
would not expect to see the systematic and progressive fault
growth and linkage histories that have been recognized recently
in the Northern North Sea and elsewhere (e.g. Dawers &
Underhill 2000; Davies et al. 2000; McLeod et al. in press).
Recent unpublished work has also demonstrated that Triassic
faults in the Northern North Sea dip in the opposite sense to
Jurassic faults and were not reactivated during Jurassic rifting.
In addition, in the Central Graben, the Helgesen (1999)
interpretation would indicate that the exact orientation of the
Jurassic faults is not directly controlled by any underlying
basement lineaments.

TIMING OF FAULT MOVEMENT ASSESSED FROM


SEISMIC DATA
2D and even recent 3D seismic data are rarely of enough
resolution to allow for division of the Jurassic into seismic
sequences so detailed as to indicate the timing of fault move-
ment. This is commonly the result of the deterioration in
seismic resolution due to the depth of burial and/or the
presence of a chalk cover. The Jurassic section can also be thin
on the footwalls of extensional faults and may be dissected by
the through-going fault planes. This makes the mapping of
stratal units across fault blocks difficult. It is also easy to
overestimate the duration of active extension using seismic data
alone (Prosser 1993). To overcome these pitfalls we integrate
the interpretation of the seismic data with stratigraphic infor-
mation from wells. We have selected seismic lines orthogonal
to the strike of faults from optimal areas within the three rift
Fig. 6. (a) Seismic line from the Nevis South fault block, with (b) arms in order to make a qualitative assessment of the timing of
structure map at near base Cretaceous level. Two large offset planar fault block rotation.
normal faults of N–S and NNE–SSW trend bound the block, which
shows limited internal fault dissection. The fault on the western side
shows signficant offset (1 second TWT). The intra-Aallenian uncon-
formity is mapped across the block. Younger intra-Callovian and N–S/NNE–SSW faults
intra-Oxfordian unconformities are related to fault block rotation. The Viking Graben is dominated by faults that have a N–S and
Both correspond to truncated footwall successions and significantly NNE–SSW orientation (Fig. 1). In the central sector of the rift
younger divergent hanging wall wedges (marked X and Y). Fault
block rotation took place during the Bathonian, Callovian and arm, in the Bruce–Beryl embayment, a series of broadly N–S
Oxfordian. and NNE–SSW striking faults bound the Beryl and South
Nevis hydrocarbon accumulations. The South Nevis Field
occurs in a N–S orientated fault block (Fig. 6) that has a
would it be acceptable to interpret some strike-slip movement, Jurassic extensional history. Crestal wells have proven a rela-
which of course would post-date the Jurassic rifting, upon tively complete Jurassic succession in the footwall. The intra-
which we focus here. Mapping of the depocentres also Aalenian unconformity (the result of the older Middle Jurassic
demonstrates that their form is that of an extensional thermal doming) is mapped across the fault block. In cross-
half-graben, an interpretation which is consistent with the sectional and plan view the main bounding fault has a geometry
absence of local structural and sedimentological variation along that is consistent with dip-slip fault movement and shows no
the strike of the planar faults. This, along with basic mapping of convincing evidence for oblique-slip or strike-slip motion. Two
these structures, shows that they are the result of dip-slip not footwall unconformities of Callovian and Oxfordian ages are
oblique-slip or strike-slip faulting. The flower structure itself is mapped at the crest of the fault and correspond down-dip to
related to the Great Glen Fault, which was interpreted to have slightly younger, thickened hanging wall wedges adjacent to the
been a major basin bounding strike-slip fault by Roberts et al. Shetland Platform. We interpret both to be the result of
(1990). In their model the implicit assumption that Jurassic extensional movement on the fault itself and associated foot-
faults were active contemporaneously required some faults, in wall uplift. The evidence for two, thickened, hanging wall
Jurassic rifting of the North Sea 377

Fig. 7. Geoseismic section across the


western part of the South Viking
Graben (after Cockings et al. 1992). A
large offset planar normal fault shows
that fault rotation initiated during the
Bathonian (Bathonian interval shows
some truncation) in the footwall. There
is clear thickening of the Callovian and
possibly Lower and Middle Oxfordian
into the basin-bounding fault.
Truncation of the Callovian to the east
indicates that there was movement on
both bounding faults and intra-basinal
faults. These data would indicate that
the Early Callovian and Middle
Callovian were times of active fault
block rotation, which probably
continued into the Early, Middle and
Late Oxfordian.

wedges of Bathonian and Callovian/Oxfordian age suggests


that these were times of fault block rotation of the South Nevis
Field fault block.
Some 120 km to the south of the Bruce–Beryl embayment,
within the South Viking Graben, the Fladen Ground Spur
forms another N–S striking high that separates the Moray Firth
and Viking Graben rift arms (Fig. 1). A series of linked N–S
striking faults form the eastern boundary of the high and this
fault linkage has resulted in the formation of a single 40–50 km
long fault strand that has a significant throw (c. 3 km). Similarly
orientated extensional faults to the east of what is effectively the

Fig. 8. (a) Seismic line over NE–SW striking faults from (b) the
Ettrick Field, within the Outer Moray Firth. Note that the Oxfordian Fig. 9. (a) Seismic line from a 3D dataset shot over the Elgin and
section is divergent into the planar normal fault in the centre of the Shearwater Fields. (b) Faults have an overall NW–SE strike, but are
line (marked X). Kimmeridgian and Volgian sections drape the fault commonly also arcuate in planform. Fault block rotation occurred
blocks. Poor reflector continuity possibly associated with lateral during the Kimmeridgian and Volgian (marked X). Some thicken-
facies changes adjacent to the fault in the SE make the mapping of ing of the Bathonian section, into the footwall is also apparent
some of the reflectors difficult. (marked Y).
378 R. J. Davies et al.

graben-bounding fault have smaller throws and a Jurassic Callovian, with activity potentially continuing through to the
section is preserved on the footwalls of these faults. In this Late Oxfordian. The NE–SW striking fault block from
published example the Upper Oxfordian to Lower Ryazanian the Ettrick Field in the Moray Firth was active during the
drapes the extensional fault and the older Middle Oxfordian to Oxfordian and Kimmeridgian and the NW–SE striking
Lower Callovian is characterized by a divergent hanging wall fault bounding the Shearwater Field shows only minor
wedge (Fig. 7; Cockings et al. 1992). This and the Nevis South pre-Kimmeridgian activity and appears to have been active
Field fault, are both examples of N–S striking faults that occur mainly during the Kimmeridgian and Volgian. These obser-
along the western margin of the Viking Graben. They both vations on fault timing are consistent with the changes in the
show a consistent timing of fault activity: with initiation during style of sedimentation that took place during the Late Jurassic
the Bathonian, probably on a series of unlinked faults, and that we discuss later.
continued movement during the Callovian and probably into
the Late Oxfordian.
Cross-cutting fault relationships
In some parts of the triple junction only one fault population
NE–SW faults can be mapped, for example NE–SW striking faults predomi-
In contrast to the South Viking Graben, the Inner Moray Firth nate within the Inner Moray Firth (Fig. 1). In other areas two or
and much of the Outer Moray Firth are dominated by faults even three fault populations occur together, for example in the
that strike NE–SW. Where the Jurassic succession is particu- North Viking Graben and the Witch Ground Graben. In such
larly thick, for instance within the central part of the rift arm, areas, seismic data quality permitting, the relative timing of fault
the Jurassic interval can be subdivided and the Jurassic exten- activity can be determined by identifying cross-cutting relation-
sional histories of the faults can be assessed. The Ettrick Field ships. For example the Viking Graben is characterized by
in Blocks 20/2 and 20/3 is located to the east of the Inner N–S/NNE–SSW and NE–SW striking faults (Faerseth et al.
Moray Firth basin in the South Halibut Trough (Fig. 1). Here, 1997; Fig. 1). In the North Viking Graben it has been
no Callovian or older Jurassic rocks are known to be preserved established that N–S/NNE–SSW faults are cross-cut by
(Andrews et al. 1990) and the Jurassic section is dissected by a NE–SW orientated faults, indicating that N–S/NNE–SSW
series of NE–SW intra-rift tilted fault blocks that dominate the faults pre-date NE–SW faults (Faerseth et al. 1997). In their
structural architecture of this part of the rift. In the Outer interpretation some of the N–S faults were interpreted to be the
Moray Firth, in contrast to the examples of N–S striking faults result of the Permo-Triassic phase of extension and the
from the Viking Graben (Figs 6 and 7), it is the Oxfordian and NE–SW faults are considered to be the result of a Jurassic
Lower Kimmeridgian section that thins onto the footwall extension. In the Witch Ground Graben (Fig. 1), WSW–ENE
and thickens onto the hanging wall of NE–SW striking faults striking faults (for example the Theta Graben) intersect
(Fig. 8). In this case the Volgian succession drapes the footwall NW–SE striking faults. Although we have found no definitive
of the extensional fault, which appears to tip out within examples of cross-cutting relationships on seismic data, Boldy
Kimmeridgian-aged sediments. This indicates that, at least on & Brealey (1990) used the thickness variations of Middle and
this fault, but most probably within this part of the rift zone, Upper Jurassic strata to show that WSW–ENE faults pre-date
Oxfordian and Kimmeridgian were times of active fault block NW–SE faults. These observations may be further evidence
rotation and footwall uplift (Davies 1995). Our example that faulting did not occur on all faults at the same time, with
from the Inner Moray Firth (Fig. 5) also shows NE–SW the actual timing of movement being different and related to
orientated faults that were active during the Late Oxfordian, the fault orientation.
Kimmeridgian and Early Volgian.
STRATIGRAPHIC ANALYSIS OF WELLS LOCATED
NW–SE faults IN FAULT HANGING WALLS
In the Central Graben, despite the structural complexity due to Seismic data provide a qualitative ‘first-order’ approximation
the presence of the Permian salt, discrete and coherent fault for the timing of faulting. Information on the sedimentary
blocks have been imaged on 2D and 3D seismic data (Figs 3, 4 response to extension is preserved in the stratigraphic record,
and 9; see also Jeremiah & Nicholson 1999, fig. 10). One such however, and can provide more detail of the timing of active
rotational fault block now defines the Shearwater Field in Block extension than the seismic data alone. This can be done by
22/30b, which is bounded to the north and northeast by a fault examining the sedimentary successions in the hanging walls
that has an arcuate planform geometry at near base Cretaceous adjacent to extensional faults and looking for systematic
level. On the southwestern flank of the structure, Volgian and changes in, for instance, the thickness of the syn-extensional
Kimmeridgian sediments thicken into the hanging wall of the fill. Extension could occur with concomitant changes in water
adjacent extensional fault, which is the northeastern boundary depth but with little syn-extensional sedimentary infill and so
of the Elgin Field. In this example, little or no thickening is water depth estimations have been made to account for this.
seen within the pre-Kimmeridgian section, although the Middle Our stratigraphic analysis is based upon well penetrations and
Jurassic may thicken slightly towards the footwall, possibly these have been used in subsidence analysis and for the
indicating that subsidence then may have been controlled by construction of isopach maps of individual stratigraphic units.
other faults or more likely by an earlier phase of post-rift, Other supportive data can be gleaned from documented
thermally driven subsidence. The timing of active extension can constraints that have now been placed upon the timing of the
be constrained on the basis of divergent seismic reflector development of rift-flank unconformities (Davies et al. 1999).
relationships as occurring during the Kimmeridgian and
Volgian.
The examples from the Moray Firth, Viking Graben and Stratigraphic database
Central Graben show different timings of the initiation of active High-resolution subsidence analysis was undertaken for 98
extension. The N–S striking faults on the western margin of the wells in the South Viking Graben, the triple junction and the
Viking Graben were active first, during the Bathonian and Moray Firth. These wells were selected on the basis that (a) they
Jurassic rifting of the North Sea 379

Fig. 10. Middle Jurassic to earliest


Cretaceous subsidence history – Well
12/21-3 (see Fig. 1 for location). Total
(sediment loaded) subsidence is plotted
without a water depth correction
(dashed line) and with a minimum and
maximum water depth correction
(shaded area between circles). Open
circles are total subsidence with a
minimum water depth correction and
solid circles have a maximum water
depth correction applied. The further
these circles plot from the dashed line,
the deeper the water depth. Subsidence
is plotted as depth (in metres) to
compaction basement (in this case
taken as base Permian) against time (in
millions of years). Sedimentation rates
are plotted as a histogram along the
base of the plot (measured in metres
per million years). Vertical lines split
the plot up into the Middle Jurassic,
Late Jurassic and earliest Cretaceous.
Changes in subsidence and
sedimentation rates reflect the timing of
active extension and faulting (see main
text for details).

penetrated a Jurassic section, (b) they were not drilled as the accommodation space available for sediment accumulation
deviated wells and did not encounter any seismically imaged during the Jurassic would have been the result of the compac-
faults and so that measured stratigraphic thicknesses are real tion of underlying, older successions. A failure to correct for
and (c) biostratigraphic resolution was deemed to be good. the effects of this compaction would lead to overestimates of
These data have allowed a quantitative, statistical analysis of the the amount of Jurassic subsidence. All wells were therefore
subsidence patterns in the hanging wall wells. In addition decompacted to the base of the Permian to remove most of
non-decompacted isopach maps are based upon between 182 the effects of any unpenetrated sediment and to provide a
and 392 well penetrations (depending on which stratigraphic common starting point for the subsidence history calculations.
unit was being considered). These were constructed in order to Unpenetrated Mid–Lower Jurassic, Triassic and Permian sedi-
understand regional changes in the geometry of stratigraphic ment thicknesses were obtained from seismic data, by corre-
units that make up the Middle and Upper Jurassic fill of the lation from nearby wells and from published work (e.g.
triple junction area. For both the subsidence analysis and for Andrews et al. 1990; Frostick et al. 1992). The effects of
the construction of regional isopach maps, each well was Cenozoic erosion in the Inner Moray Firth mean that Jurassic
divided into datable units. Biostratigraphic data from individual sediments are over-compacted at the present day compared to
wells, allied to newly established sequence stratigraphic corre- what would be predicted using standard porosity–depth func-
lations (Stephen et al. 1993; Davies et al. 1996; Stephen & tions. A failure to correct for this would lead to an underesti-
Davies 1998; Davies et al. 1999), were used to subdivide the mate of Jurassic subsidence rates. The Cenozoic denudation
Jurassic–earliest Cretaceous succession in each well. This pro- estimates of Hillis et al. (1994), were used to correct for these
vides a more detailed, consistent and accurate stratigraphic effects. Further details of the subsidence analysis methodology
subdivision than would be possible if one was to rely upon the we use are given by Turner (1997).
uncorroborated chronostratigraphic subdivisions posted on Changes in background subsidence rates driven by tectonic
composite logs of variable vintage. processes, such as active extension, are reflected in the sedi-
mentary record in two ways: by variations in the thickness of
sediment that accumulated and by changes in depositional
Subsidence analysis and its application environments. During periods of accelerated subsidence rates,
Subsidence histories were calculated using standard back- as expected for hanging wall sites during periods of active
stripping (subsidence analysis) methods (after Steckler & Watts extension, there may be an increase in sediment accumulation
1978). The same stratigraphic criteria were used to subdivide all rates, a deepening of water depths or both (Fig. 10). The
wells, so that the subsidence results from different wells and balance between these effects will depend upon the rate of
areas can be directly compared and any spatial or temporal sediment supply. If sediment supply rates do not keep pace
variations in subsidence could be identified. Stratigraphic ages with background subsidence rates then water depths will
were converted to numerical ages using the geological time- increase. Conversely, if sediment supply rates exceed subsid-
scale of Harland et al. (1990). Present-day sediment thicknesses ence rates, water depths will decrease. It is therefore important
were decompacted using the exponential porosity–depth deter- to look at variations in both sediment accumulation rates and
minations of Sclater & Christie (1980) for each of the main water depth.
lithologies identified.
Corrections were also made, where appropriate, for the Errors associated with subsidence analysis
effects on calculated Jurassic subsidence of the compaction of The single largest potential error in most subsidence studies
deeper unpenetrated strata and of Cenozoic denudation. Part of comes from uncertainties in estimating the water depths of
380 R. J. Davies et al.

deposition and variations in water depth through time. For the


areas studied here, the limited seismic resolution due in part to
structural complexity, precludes the determination of water
depth through the identification of clinoform geometries.
Therefore we use fossils, especially benthic microfossils, fossil
assemblages, sedimentary facies, sedimentary structures, dis-
tinctive organic and inorganic geochemical signatures, rock
colours and mineral assemblages as environmental indicators.
A comparison with modern environments helps to quantify
possible water depth ranges but it is not possible to make
precise water depth determinations. Whilst likely water depth
ranges are fairly narrow and reliable for shallow water depths,
uncertainties of several hundred metres or more are introduced
for open marine, slope settings. In this study we have concen-
trated on identifying the timing of initiation of active extension,
which is recorded as an increase in subsidence rates and is
reflected in both sediment thickness and water depth changes.
As water depths increase it becomes progressively harder to
discriminate between changes in basement subsidence and
changes in sediment supply. Two key changes in water depth
are used in this study to identify increases in subsidence rates,
and hence the initiation of active faulting: the transition from
non-marine or marginal marine conditions to shallow marine
environments and the change from stable, shallow marine shelf
environments to deep, open marine conditions. Both can be
identified with confidence in all wells where the relevant
sections are preserved.

Subsidence analysis: qualitative observations


The analysis of Middle–Late Jurassic subsidence from represen-
tative wells in different parts of the study area records similar
overall patterns of subsidence (Fig. 11). The differences in their
detailed subsidence histories, however, can only be explained by Fig. 11. Subsidence curves Middle Jurassic to earliest Cretaceous
variations in the timing of extension and normal faulting. All subsidence histories for a selection of hanging wall wells (see Fig. 1)
for this period includes the main Middle–Late Jurassic syn-rift stage.
wells analysed record a period of non-deposition or erosion All wells preserve a record of Middle Jurassic uplift, erosion and
during the early part of the Middle Jurassic (the Mid-Cimmerian doming (the Mid Cimmerian Unconformity, Underhill & Partington,
unconformity). Non-marine to marginal marine Middle Jurassic 1993, 1994). The time at which subsidence resumed at each site
sediments overlie the Mid-Cimmerian unconformity in most varies, partly controlled by their distance from the centre of the
wells. All sites throughout the study area also record a steady dome and partly by the timing of initiation of active extension in
deepening of water depths, from non-marine conditions during each location. The transition from moderate rift initiation subsidence
to rapid peak rifting subsidence also occurred at different times in
the Bathonian (161 Ma) to slope conditions in all hanging wall different locations controlled by the orientation of nearby normal
locations by Kimmeridgian times (153 Ma). This increase in faults. See main text for further discussion.
water depth was accompanied by very high sedimentation rates
in most wells. In Well 12/21-3 from the Inner Moray Firth, for
example, the change from shelf to slope conditions during the
early part of the Late Jurassic was accompanied by sedimenta- the timing of changes in subsidence. There are differences in
tion rates of 150–400 m Ma1 (Fig. 10). This combination of the age of sediments overlying the Mid-Cimmerian (intra-
deepening water depths and high sedimentation rates is evi- Aalenian) unconformity. Bajocian to Bathonian sediments are
dence for a period of rapid basement subsidence, consistent present above the unconformity along most of the length of the
with the effects of extension. Viking Graben (e.g. in Wells 9/13-34 and 16/17-6) and in the
The Late Oxfordian to Early Volgian (155.9–150.1 Ma) was a west of the Moray Firth (in Well 12/21-3). The time gap at
period of rapid basement subsidence in all wells, although there the unconformity is much longer in the east of the Moray Firth
were major variations in the timing of initiation of the rapid closer to the triple junction: Early–Middle Oxfordian and Late
subsidence. In some wells (e.g. 12/21-3 and 15/23-9 – Fig. 11) Oxfordian sediments lie above the unconformity in Wells
a period of slow to moderate subsidence preceded the main 20/2-4 and 15/23-9, respectively. There were also major
phase of rapid basement subsidence. The timing of this shift differences in the initiation of rapid basement subsidence. Rapid
from moderate to rapid subsidence varied. Towards the end of subsidence began first, during Bathonian times, in the Viking
the Volgian and during the Ryazanian, deep marine conditions Graben (Wells 9/13-34 and 16/17-6). This contrasts with the
existed throughout the study area (and persisted during the Moray Firth (Wells 12/21-3, 20/2-4), where the Bathonian to
Cretaceous and into the Cenozoic in most wells). Uncertainties Callovian were periods of moderate basement subsidence
in constraining the precise water depths mean that it is before a major acceleration in subsidence during the Oxfordian.
impossible to decide whether subsidence rates began to slow at Lastly, rapid subsidence did not begin until the later part of the
the same time in all wells. Kimmeridgian and Early Volgian in the Witch Ground Graben
Superimposed on this overall pattern of Middle–Late Jurassic and the centre of the triple junction (Wells 15/23-9 and
subsidence and sedimentation are other significant variations in 22/2-2). Overall this represents a difference in the timing of the
Jurassic rifting of the North Sea 381

(a)

(b)
Fig. 12. (a) Quantative analysis of Middle–Late Jurassic syn-rift subsidence – cumulative frequency plots. Changes in the relative timing of peak
fault activity for faults of different orientations can be quantified by looking at variations in the timing of the most rapid syn-rift subsidence in
hanging wall wells. The main syn-rift interval (Middle–Late Jurassic, Callovian–Early Volgian) was subdivided into five time periods (Callovian,
Early/Middle Oxfordian, Late Oxfordian, Kimmeridgian and Early Volgian) for 66 wells drilled in hanging walls within (b) the study area –
including the South Viking Graben, triple junction, Outer and Inner Moray Firth. The further to the right that a time period plots the greater
its relative importance (see the discussion in the main text).The Late Oxfordian time period plots to the right of the other time periods in the
top plots of (a) indicating that, for N–S and NNE–SSW orientated faults, a greater proportion of syn-rift subsidence took place during the Late
Oxfordian than during the other time periods. This picture changes for the other fault sets. Whilst the Late Oxfordian is still dominant for
NE–SW faults, the Late Oxfordian and Early Volgian are equally important for E–W faults, with the Early Volgian dominant for NW–SE faults.
By comparing the relative position of the different time periods in these plots it is apparent that the age of peak syn-rift subsidence youngs in
the following order: N–S/NNE–SSW to NE–SW to E–W to NW–SE. This corresponds to the results obtained from other datasets in this
paper, suggesting that the different fault sets had different movement histories, with fault initiation and peak displacement rates taking place at
different times for each fault set. These results are consistent with a clockwise rotation in far-field stresses during the syn-rift.

initiation of rapid subsidence of at least 10 million years, from propose that these overall similarities and local variations in
the Bathonian in the South Viking Graben to the Early Volgian the timing of Middle–Late Jurassic subsidence reflect the
in the Witch Ground Graben and triple junction. These interaction between the collapse and gradual subsidence of
observations, made from what we consider to be representative the North Sea dome (Underhill & Partington 1993, 1994) and
wells, are similar to those made using seismic data alone. We the initiation of extension and large scale normal faulting.
382 R. J. Davies et al.

The age of the strata sub-cropping the Mid Cimmerian c33% of wells, c20% in 43% of wells, c30% or less in 52%
Unconformity, and hence the time gap that it represents, of wells and c40% in 70% of wells. This is a higher overall
increases towards the centre of the North Sea dome and this proportion of syn-rift subsidence than that recorded for the
pattern is, to a large extent, mirrored by the age of marine other four time, periods indicating the relative importance of
sediments onlapping the unconformity (Underhill & Partington Late Oxfordian subsidence.
1993, 1994). Each of the Jurassic rift arms preserves a record of What emerges are clear and systematic variations in the
gradual Middle to Late Jurassic marine onlap towards the triple relative importance of different time intervals for the different
junction, reflecting the overall, regional subsidence of the fault orientations. Whilst for both N–S/NNE–SSW and
dome. NE–SW faults the Late Oxfordian was the period of most rapid
The flooding of the dome did not, however, take place at the subsidence, the Early Volgian was the main period of rapid
same rate down each of the rift arms. Instead the Viking subsidence for NW–SE faults. There is a gradual decrease in
Graben flooded much more quickly than the Moray Firth. the relative importance of the earlier parts of the rift phase
Deep marine conditions were well established down most of (Callovian and Oxfordian) and increase in the importance of
the length of the South Viking Graben, to within less than the later periods (Kimmeridgian and Early Volgian) moving
50 km of what we propose was the dome centre, by Callovian from N–S/NNE–SSW to NE–SW to E–W and finally NW–SE
times (e.g. Well 16/17-6 in Fig. 11). In contrast, marine faults. This suggests a systematic variation in the movement
conditions in the Moray Firth during the Callovian were history of these different fault sets.
restricted to the western part (Well 12/21-3 in Fig. 11), 200 km The subsidence results used in this analysis have not been
away from the dome centre. The area of marine conditions corrected for the effects of variations in palaeo-water depths.
expanded gradually from west to east along the remainder of Such corrections would require a range of possible water
the Moray Firth rift arm during the Oxfordian and Early depths to be assigned, rather than a single value. Given the
Kimmeridgian (Underhill & Partington 1993, 1994 – Fig. 11). uncertainty in determining precise water depths within this
This observation, and the record of rapid Bathonian subsidence range we elected to analyse the tightly constrained record of
in the Viking Graben (Fig. 11), suggests that Bathonian subsidence preserved by sedimentary thickness variations alone.
subsidence in the South Viking Graben was controlled by Basin-wide patterns of water depth variations reinforce the
extension and normal faulting as well as regional dome subsid- evidence presented here for variations in the timing of rapid
ence. Fault-controlled Bathonian–Callovian sedimentation is subsidence and, hence, fault movement, depending on fault
clearly apparent on seismic data from the South Viking Graben orientations.
(Fig. 7, Cockings et al. 1992). In contrast, Bathonian–Callovian
subsidence in the Inner Moray Firth shows no evidence for
faulting and the moderate subsidence recorded is consistent Isopach maps
with the gradual decay of the North Sea dome. High quality Isopach maps of (a) Bajocian to Early Callovian, (b) Early
seismic data available from the Inner Moray Firth (Fig. 5) Callovian to Middle Oxfordian, (c) Middle Oxfordian to Late
also clearly show that the Bathonian–Callovian sequence was Oxfordian and (d) Kimmeridgian to Early Volgian strata were
unaffected by syn-sedimentary faulting. used to try to assess whether the systematic changes in the
These observations raise the question – why should orientation of the active fault sets are manifested regionally by
extension have begun earlier in the South Viking Graben than variations in the thickness of the stratigraphic units. The close
the Moray Firth? The most obvious difference between the two similarity between these units and those used in the statistical
areas is the orientation of the main bounding faults, mostly analysis (Fig. 13) allows for direct comparison.
N–S in the Viking Graben and NE–SW in the Inner Moray During the Middle Jurassic a N–S orientated depocentre
Firth. The possible link between fault orientation and the developed in the South Viking and Central grabens (Erratt
timing of rapid subsidence, using seismic data (Figs 6, 7, 8, and et al. 1999; Fig. 13a). The Central Graben was probably the
9), was further tested through a quantitative analysis of our southernmost extension of a major N–S orientated depocentre
subsidence results. which extended from the Viking Graben into the Central North
Sea with, to the east and west of this central axis, Middle
Jurassic sediments probably only being preserved locally within
Quantitative analysis of subsidence hanging wall depocentres. During the Oxfordian a NE–SW
Quantitative analysis of Middle–Late Jurassic subsidence orientated depocentre/rift zone developed in the Moray Firth,
patterns was performed using 66 hanging wall wells from the extending from the South Halibut Trough into the Witch
Moray Firth, South Viking Graben and triple junction areas Ground Graben and Fisher Bank Basin and as far south as the
(Fig. 12). The main Middle–Late Jurassic syn-rift interval was Gannet/Guillemot embayment of the Central Graben – in the
subdivided into five time periods for this analysis (Callovian, region of Block 21/25 (Fig. 13b, c). Some of the isopach
Early–Middle Oxfordian, Late Oxfordian, Kimmeridgian, Early patterns may be the result of post-depositional uplift and
Volgian). The percentage of the total syn-rift subsidence that erosion. Thus, we interpret the NW–SE orientated depositional
took place during each period was calculated for each well. thick in the Central Graben to reflect the effect of truncation
These results are presented as cumulative frequency plots during periodic rift flank uplift in areas such as the Forties
for each of the main fault trends identified regionally (N–S/ Montrose High. During the Kimmeridgian and Early Volgian, a
NNE–SSW, NE–SW, E–W and NW–SE), based on the third depocentre of NW–SE orientation appears to overprint
orientation of nearby major faults. Some wells are close to the first two basin geometries (Fig. 13d). This basin was broadly
faults with different orientations and so are included in more equivalent to the present-day Central Graben. The northern
than one plot. For each time period the cumulative frequency end of the Central Graben stretched as far north as Quadrant
plot displays the percentage of wells recording different 15 during the Kimmeridgian and Volgian. Some of these
proportions of the total syn-rift subsidence. The further to the changes in isopach geometry are subtle, because depocentres
right a time period plots the greater its relative importance. For generated, for example during the Oxfordian, may have been in
example, for wells adjacent to N–S/NNE–SSW faults the Late existence during the Kimmeridgian, when the third series of
Oxfordian accounts for c10% or less of syn-rift subsidence in NW–SE orientated topographic lows also developed.
Jurassic rifting of the North Sea 383

Infill of the basins was interrupted by the development of Bathonian–Callovian and Oxfordian times. This is especially
regional unconformities that have been interpreted to have noticeable for wells drilled in the hanging walls to NW–SE
been the result of footwall and rift flank uplift during extension orientated faults (e.g. 15/23-9 in the Witch Ground Graben
(Davies et al. 1999). When looked at in more detail two regional and 22/2-2 in the triple junction, Fig. 11). In these areas the
unconformities are identified within the syn-rift succession. main acceleration in subsidence rates linked to the onset of
These are intra-Callovian and intra-Middle Oxfordian in age major faulting took place during the Early Volgian. In the
(Davies et al. 1999). The occurrence of at least two Middle to Witch Ground Graben the Early Volgian was also marked by
Late Jurassic syn-rift unconformities suggests that multiphase increased water depths, with the flooding of Piper Formation
extension did occur (Davies et al. 1999). shallow marine shelf sands by open marine Kimmeridgian Clay
Formation shales and turbidites of the Galley, Ettrick and
Claymore members. A NW–SE orientated thick overprints the
INTEGRATION OF SEISMIC AND NE–SW trend (Fig. 13d, marked X).
STRATIGRAPHIC OBSERVATIONS
Beryl Embayment and South Viking Graben
TECTONO-STRATIGRAPHIC EVOLUTION OF
In both of these areas the Bathonian and Callovian were times THE NORTH SEA RIFT
of rapid sedimentation and deepening water depths, indicative
of a major acceleration in basement subsidence rates consistent Sequential intra-rift dip-slip faulting
with the effects of active faulting. This agrees with the seismic The Inner Moray Firth and the Central Graben have been key
data (Figs 6 and 7) showing that faulting probably started areas for the development of complex kinematic models for
during the Bathonian (Cockings et al. 1992; Wood et al. 1999). rifting that involved essentially contemporaneous fault move-
Subsidence rates in the Viking Graben continued to increase ment on the main fault sets and oblique-slip and strike-slip
during Callovian times (Fig. 11) as water depths and sedimen- faulting (Bartholomew et al. 1993; Sears et al. 1993; Eggink et al.
tation rates increased as a result of larger-scale fault movement; 1996). Our examination of some of the key data suggests that
Callovian strata in the South Viking Graben can clearly be seen there is little or no convincing evidence for the classic features
to thicken into N–S trending faults (Figs 6 and 7). The of exposed and analogue modelled strike-slip and oblique-slip
Callovian was also marked by the development of complex fault systems. We suggest that the evidence is at least ambigu-
depositional environments that reflected the local topography ous or capable of an alternative interpretation, this being that
generated in the South Viking Graben by active faulting, on a local and regional scale sequential dip-slip faulting took
including non-marine sediments (Sleipner Formation), shallow place. The most direct evidence for this comes from seismic
marine sands and open marine shales (Hugin and Heather data over the fault blocks themselves. The most significant
formations) and deep marine turbidites (Ling Member; reasons for the lack of consensus on fault and rift kinematics
Cockings et al. 1992). The isopach of the Bathonian and Lower are first, the lack of seismic resolution in the Central Graben
Callovian (Fig. 13a) shows that the preserved section is pre- and the complexities of salt tectonics and, second, a misinter-
dominantly restricted to a N–S axis within the South Viking pretation of the timing of movement on the Great Glen Fault.
Graben and the Central Graben. Bathonian to Callovian The statistical analysis of subsidence within the hanging walls
subsidence patterns elsewhere in the study area (e.g. Inner of the four fault populations within the triple junction area is,
Moray Firth) show no evidence for either rapid differential however, perhaps the most compelling evidence for sequential
subsidence or syn-sedimentary faulting (Figs 5 and 11). faulting. On the basis of the statistical analysis of subsidence
and isopach patterns it seems to be reasonable that the initial
interpretation of sequential faulting, made on the basis of the
Moray Firth seismic data alone, can be applied throughout the triple
In the Inner Moray Firth, after moderate subsidence rates junction region. Although uncertainty remains with regard to
and a gradual increase of water depths during Bathonian to the timing of the cessation of fault activity in some cases, we
Callovian times, there was a dramatic increase in subsidence speculate that the switching of activity from one fault set to
rates during the Oxfordian. Sedimentation rates increased by an another was progressive and gradual, with for example, move-
order of magnitude relative to the Callovian (Fig. 11) and water ment on N–S faults decreasing as movement started to occur
depths increased, with open marine shales of the Heather on NE–SW faults.
Formation deposited across the whole of the Inner Moray Firth There are several models for the kinematic evolution of the
during Late Oxfordian times. On seismic data from the Inner North Sea that either require a single extension direction or
Moray Firth there is evidence of Middle to Late Oxfordian and have a changing extension direction during the Jurassic.
younger Late Jurassic sediments thickening into the hanging Given that there remains a lack of evidence for strike-slip
walls of the NE–SW orientated faults that dominate this area and oblique-slip movement (Underhill 1991a; Thomson &
(e.g. Fig. 5 and Underhill 1991a). Similarly, in the western Outer Underhill 1993; Erratt et al. 1999), our model, in common with
Moray Firth (Ettrick area), water depths, sedimentation and Roberts et al. (1990) and Erratt et al. (1999), considers most of
subsidence rates all increased during the Oxfordian consistent the faults to be normal, dip-slip faults which accommodated
with the effects of active faulting (Fig. 8). The Western Outer extension orthogonally to their strike. Thus, sequential faulting
Moray Firth is dominated by NE–SW and ENE–WSW to E–W was superimposed upon the trilete junction with intra-rift faults
orientated faults, with thickening of the Oxfordian succession responding to the prevailing far-field stresses. We propose
visible on seismic data (Fig. 8). A NE–SW orientated thick of that the Bathonian to Callovian was dominated by an overall
early Callovian to Early Kimmeridgian age is preserved in the E–W extension direction, during the Oxfordian to Early
Outer Moray Firth and the Fisher Bank Basin (X on Fig. 13b). Kimmeridgian the prevailing extension direction was NW–SE
Whilst water depths increased from non-marine to shallow and lastly during the Late Kimmeridgian to Volgian the
marine conditions during the Oxfordian in the Witch Ground extension direction had a NE–SW orientation. This is equiva-
Graben and triple junction there is no sign of the major lent to a rotation of extension direction of approximately 135
acceleration in subsidence rates recognized elsewhere during in 10 Ma.
384 R. J. Davies et al.

Fig. 13. (a) Map showing the distribution of the Bathonian to lower Callovian; (b–d) isopach maps based on thicknesses penetrated by wells
of (b) the lower Callovian to Middle Oxfordian, (c) the Middle Oxfordian to lower Kimmeridgian and (d) lower Kimmeridgian to lower Volgian.
The stratigraphic intervals are very similar to those used in the quantitative analysis we present (Fig. 12). The maps demonstrate that
sedimentation accumulation during the Middle and Late Jurassic underwent systematic changes. A N–S orientated depocentre formed during the
Bathonian and Early Callovian within the Viking and Central grabens. During the Oxfordian a NE–SW orientated depocentre developed within
the Moray Firth extending into the Fisher Bank Basin (marked X). During the Kimmeridgian and the Volgian a NW–SE orientated rift zone
developed within the Central Graben (marked Y). The complexities of salt-induced variations in accommodation space are only mapped in
Quadrant 21 in (d), however, the development of salt withdrawal basins probably took place throughout the Jurassic. This suggests that the
observations made on selected seismic lines and from subsidence recorded within particular wells can be extrapolated over the triple junction
as a whole.

Changes in the extension direction did not lead to the were orthogonal to the new extension direction (Fig. 14). This
development of a series of oblique-slip faults but instead led to has resulted in a characteristic rhombohedral fault pattern in
the development of a new population of dip-slip faults that plan view within the North Viking Graben (Faerseth et al. 1997)
Jurassic rifting of the North Sea 385

Fig. 15. Comparison of published rotation of extension direction


in the Afro-Arabian rift system and the North Sea triple junction:
(a) this study; (b) from various studies (Lyberis 1988; Ring 1994;
Fig. 14. Intra-rift faults superimposed upon a simple trilete junction Bosworth & Taviani 1996; Bonini et al. 1997). In both rifts there is
developed during three progressive stages during (a) the late a systematic rotation in the extension direction and stress field.
Bathonian to Early Callovian, (b) the Oxfordian and lastly (c) during
the Kimmeridgian and Volgian respectively. In each case faults are
considered to be dominantly dip-slip in origin, with an extension
direction that was orthogonal to their strike.
is c. 105 and takes place over c. 25 Ma. Bosworth et al. (1992)
identify a 45 rotation in the extension direction within the
and in parts of the Central Graben (Bartholomew et al. Kenyan rift, which took place over 600 ka. This gives a rate of
1993). rotation of 75 Ma1. This is thought to be the result of
There is little certainty regarding the possible causes of the far-field changes in plate boundary constellation (Streckler et al.
rotational stress field. In the Afro-Arabian rift system there has 1990). The evidence from the Afro-Arabian rift is consistent
been a rotation in the extension direction from NE–SW during with our findings in the North Sea, with far-field stresses being
the Miocene (25–16.5 Ma) to ENE–WSW (16.5–10 Ma) and the most likely causal mechanism. Thomas & Coward (1995)
then E–W (5 Ma) (Fig. 15). This rotation of extension direction have proposed that a complex interaction between extension
386 R. J. Davies et al.

Fig. 16. Cartoon illustrating the lithospheric and asthenospheric changes that generated a thermal dome in the North Sea region during the
Middle Jurassic. We speculate that doming led to the development of three zones of lithospheric weakness that took on a trilete geometry.

Fig. 17. Cartoon showing that after Middle Jurassic doming the three resultant rift zones were exploited as zones of weakness during subsequent
Middle–Late Jurassic rifting. Active faulting during this rift phase varied through time in both orientation and location depending upon the
prevailing far-field least principal stress direction.

directions of the Arctic and Atlantic oceans, with possibly an needing the least energy to form in homogeneous materials. As
influence from rift zones emanating from the Neo-Tethyan such, it is analogous to the formation of triple junctions in
system in Central and Eastern Europe may have caused stress young oceanic crust (McKenzie & Morgan 1969; Patriat &
field rotation. The exact mechanism for generating a rotation in Courtillot 1984). Unlike their oceanic counterparts which con-
the stress field remains speculative. tinue to develop an individual identity in response to any
variability in geometry and stability of their respective, com-
ponent arms (e.g. McKenzie & Morgan 1969), the individual
Impact of thermal doming rift arms of the North Sea are likely to have experienced less
The North Sea rift shows a clear trilete configuration but the direct effect of the doming as the feature declined during
movement history and distribution of Middle and Late Jurassic deflation. Although the initial formation of a simple trilete
intra-rift faults made little contribution to its development, in junction may have been the result of doming, subsequent
many cases developing obliquely to the rift arms themselves. extension during the Middle and Late Jurassic led to the
Instead the triple junction is an older feature that immediately development of a series of intra-rift fault sets that did not
pre-dates the Middle and Late Jurassic faulting. develop synchronously but during a 10 Ma rift phase.
The likelihood that the triple junction formed central to a
Middle Jurassic thermal dome (Underhill & Partington 1993,
1994) suggests that initial development of the triple junction CONCLUSIONS
may have been as a result of the dome rise and decay. We + It has proved popular to try and place Middle and Late
suggest that the trilete system originally developed in response Jurassic deformation within one simple overall stress regime.
to the thermal doming and deflation of the Central North Sea In doing so, several workers have had to invoke consider-
resulting from the development and evolution of a warm, able amounts of strike- or oblique-slip deformation on faults
diffuse and transient plume dome (Underhill & Partington within the rift arm that are not ideally orientated for dip-slip
1993, 1994). The dome rise and decay would have provided a motion.
zone of weakened lithosphere that, during later rifting, recorded + In this study we provide evidence that supports an interpret-
the rotation of the ambient stress field (Figs 16 & 17). ation of predominantly dip-slip faulting during rifting. The
The development of the three rift arms is consistent with the faulting was sequential, with each fault set having a different
formation of three equidistant (i.e. 120 spaced) fractures movement history. This negates the need for a complex
Jurassic rifting of the North Sea 387

interaction between oblique-slip, strike-slip and dip-slip North Sea. In: Hardman, R. F. P. (ed.) Exploration Britain: Geological insights
faults controlled by a single stress regime. for the next decade. Geological Society, London, Special Publications, 67,
65–105.
+ Our results suggest that there was a shift in the extension Davies, R. J. 1995. The Stratigraphic Evolution of the Middle and Upper Jurassic of the
direction from E–W during the Bathonian and Callovian Outer Moray Firth, North Sea. PhD thesis, University of Edinburgh, UK.
through to NW–SE during the Oxfordian and early Davies, R. J., Stephen, K. S. & Underhill, J. R. 1996. A Re-evaluation of the
Kimmeridgian to NE–SW in the late Kimmeridgian and Middle and Upper Jurassic Stratigraphy and Flooding History of the Moray
Volgian. This apparent rotation in the extension direction of Firth Rift System, North Sea. In: Hurst, A. et al. (ed.) Geology of the Humber
Group: Central Graben and Moray Firth, UKCS. Geological Society, London,
135  in 10 Ma is similar in magnitude to changes in Special Publications, 114, 81–108.
extension direction identified within other, younger basins. Davies, R. J., O’Donnell, D., Bentham, P. N., Gibson, J. P. C., Curry, M. R.,
+ We speculate that the rise and fall of a thermal dome during Dunay, R. E. & Maynard, J. R. 1999. The origin and genesis of major
the Middle Jurassic was the fundamental precursor to rifting Jurassic unconformities within the North Sea triple junction area of the
as it provided the zones of lithospheric weakness within North Sea, UK. In: Fleet, A. J. & Boldy, S. A. R. (eds) Petroleum Geology of
Northwest Europe: Proceedings of the 5th Conference. Geological Society, London,
which sequential faulting and rifting could occur, responding 117–131.
to changes in the Middle and Late Jurassic intra-plate stress Davies, S. J., Dawers, N. H., McLeod, A. E. & Underhill, J. R. 2000. The
regime. Structural and Sedimentological evolution of early syn-rift successions: The
Middle Jurassic Tarbert Formation, North Sea. Basin Research, 12, 343–365.
Dawers, N. H. & Underhill, J. R. 2000. The role of fault interaction and
RJD would like to thank ExxonMobil International Ltd for giving linkage in controlling sun-rift stratigraphic sequences: Late Jurassic Stafjord
permission to publish this paper. John Gibson and James Maynard East Area, northern North Sea. American Association of Petroleum Geologist
are thanked for help with some of the seismic interpretation. RJD is Bulletin, 84, 45–64.
grateful to Ian Cloke for his encouragement. The Nevis Field Doré, A. G. & Gage, M. S. 1987. Crustal alignments and sedimentary
partners (Amerada Hess Ltd, Enterprise Oil Ltd, British Gas North domains in the evolution of the North Sea, North-East Atlantic Margin
and Barents shelf. In: Brooks, J. & Glennie, K. (eds) Petroleum Geology of
Sea Holdings and OMV UK Ltd and the Shearwater and Puffin Field North West Europe. Graham & Trotman, London, 1131–1148.
partners (Superior Oil, Esso, Shell and Arco) are thanked for giving
Eggink, J. W., Riegstra, D. E. & Suzanne, P. 1996. Using 3D seismic to
permission to use 3D seismic data. Peter Broad of Geoteam is understand the structural evolution of the UK Central Graben. Petroleum
thanked for permission to use Fig. 5 in this publication. Nopec Geoscience, 2, 83–96.
International are thanked for giving permission to use 2D seismic Erratt, D. 1993. Relationships between basement faulting, salt withdrawal and
data across the Ettrick Field (Fig. 8). JRU and JDT acknowledge Late Jurassic rifting, UK Central North Sea. In: Parker, J. R. (ed.) Petroleum
Schlumberger Geoquest for their permission to use their Geoframe Geology of Northwest Europe: Proceedings of the 4th Conference. Geological Society,
and IESX seismic interpretation software. We are grateful to Simon London, 1211–1219.
Stewart for discussion. Scot Fraser, Alain Mascle and Rodmar Erratt, D., Thomas, G. M. & Wall, G. R. T. 1999. The evolution of the
Ravnas all provided constructive reviews and Tony Spencer provided Central North Sea Rift. In: Fleet, A. J. & Boldy, S. A. R. (eds) Petroleum
very helpful editorial guidance. Geology of Northwest Europe:Proceedings of the 5th Conference. Geological Society,
London, 63–82.
Færseth, R. B. 1996. Interaction of Permo-Triassic and Jurassic extensional
REFERENCES fault blocks during the development of the northern North Sea. Journal of
the Geological Society, London, 153, 931–944.
Abbotts, I. L. (ed.) 1991. United Kingdom Oil and Gas Fields 25 Year Færseth, R. B., Knudsen, B. E., Liljedahl, T., Midboe, P. S. & Soderstrom, B.
Commemorative Volume. Geological Society, London, Memoir, 14. 1997. Oblique rifting and sequential faulting in the Jurassic development of
Andrews, I. J., Long, D., Richards, P. P., Thomson, A. R., Brown, S., the northern North Sea. Journal of Structural Geology, 19(10), 1285–1302.
Chesher, J. A. & McCormac, M. 1990. The Geology of the Moray Firth. British Frost, R. E. & Rose, J. F. 1996. Tectonic quiescence punctuated by strike-slip
Geological Survey UK Offshore Regional Report. HMSO, London. movement: influences on Upper Jurassic sedimentation in the Moray Firth
Argent, J. D., Stewart, S. A. & Underhill, J. R. 2000. Controls on the Lower and the North Sea Region. In: Hurst, A. et al. (ed.) Geology of the Humber
Cretaceous Punt Sandstone Member, a massive deep-water clastic depo- Group: Central Graben and Moray Firth, UKCS. Geological Society, London,
system, Inner Moray Firth, UK North Sea. Petroleum Geoscience, 6, 275–285. Special Publications, 114, 145–162.
Badley, M. E., Price, J. D., Rambech Dahl, C. & Agdestein, T. 1988. The Frostick, L. E., Linsey, T. K. & Reid, I. 1992. Tectonic and climatic control
structural evolution of the North Viking Graben and its bearing upon of the Triassic sedimentation in the Beryl Basin, northern North Sea.
extensional modes of basin formation. Journal of the Geological Society, London, Journal of the Geological Society, London, 149, 13–26.
145, 455–472. Harland, W. B., Armstrong, R. L., Cox, A. V., Craid, L. E., Smith, A. G. &
Bartholomew, I. D., Peters, J. M. & Powell, C. M. 1993. Regional structural Smith, D. G. 1990. A Geologic Timescale. Cambridge University Press.
evolution of the North Sea: oblique-slip and reactivation of basement Helgesen, D. E. 1999. Structural development and trap formation in the
lineaments. In: Parker, J. R. (ed.) Petroleum Geology of Northwest Europe. Central North Sea HP/HT play. In: Fleet, A. J. & Boldy, S. A. R.
Proceedings of the 4th Conference. Geological Society, London, 1109–1122. (eds) Petroleum Geology of Northwest Europe: Proceedings of the 5th Conference.
Boldy, S. A. R. & Brealey, S. 1990. Timing, nature and sedimentary result of Geological Society, London, 1029–1034.
Jurassic tectonism in the Outer Moray Firth. In: Hardman, R. F. P. & Hillis, P. R., Thomson, K. & Underhill, J. R. 1994. Quantification of Tertiary
Brooks, J. (eds) Tectonic events responsible for Britain’s oil and gas reserves. erosion in the Inner Moray Firth using sonic velocity data from the Chalk
Geological Society, London, Special Publications, 55, 259–279. and the Kimmeridge Clay. Marine and Petroleum Geology, 11, 283–289.
Bonini, M., Souriot, T., Boccaletti, M. & Brun, J. P. 1997. Successive Hodgson, N. A., Farnsworth, J. & Fraser, A. J. 1992. Salt related tectonics,
orthogonal and oblique extension episodes in a rift zone: Laboratory sedimentation and hydrocarbon plays in the Central Graben, North Sea,
experiments with application to the Ethiopian rift. Tectonics, 16, 347–362. UKCS. In: Hardman, R. F. P. (ed.) Exploration Britain: Geological insights for
Bosworth, W. & Taviani, M. 1996. Late Quaternary reorientation of stress the next decade. Geological Society, London, Special Publications, 67, 31–63.
field and extension direction in the southern Gulf of Suez, Egypt: Jackson, M. P. A. & Vendeville, B. C. 1992. Regional extension as a trigger for
Evidence from uplifted coral terraces, mesoscopic fault arrays, and diapirism. Geological Society of America Bulletin, 106, 57–73.
borehole breakouts. Tectonics, 15, 791–802. Jacquin, J., Dardeau, G., Durlet, C., De Graciansky, P. C. & Hantzpergue, P.
Bosworth, W., Streckler, M. P. & Blisniuk, P. M. 1992. Integration of East 1998. The North Sea Cycle: An Overview of 2nd-Order Transgressive/
Africcan paleostress and present-day stress data: implications for Regressive Facies Cycles in Western Europe. In: de Graciasky, P. C.,
continental stress field dynamics. Journal of Geophysical Research, 97(B8), Hardenbol, J., Jacquin, T., Farley, M. & Vail, P. R. (eds) Mesozoic and
11 851–11 865. Cenozoic Sequence Stratigraphy of European Basins. SEPM Special Publication,
Brun, J. P. & Tron, V. 1993. Development of the North Viking Graben: 60, 445–446.
inferences from laboratory modeling. Sedimentary Geology, 86, 31–51. Jeremiah, J. M. & Nicholson, P. H. 1999. Middle Oxfordian to Volgian
Cains, S. 2000. Controls on the Lower Cretaceous Punt Sandstone Member, sequence stratigraphy of the Greater Shearwater area. In: Fleet, A. J. &
a massive deep-water clastic deposystem, Inner Moray Firth, UK North Boldy, S. A. R. (eds) Petroleum Geology of Northwest Europe: Proceedings of the 5th
Sea. Discussion. Petroleum Geoscience, 6, 280–282. Conference. Geological Society, London, 153–170.
Cockings, J. H., Kessler, L. G., Mazza, T. A. & Riley, L. A. 1992. Bathonian Lyberis, N. 1988. Tectonic evolution of the Gulf of Suez and the Gulf of
to mid-Oxfordian sequence stratigraphy of the South Viking Graben, Aqaba. Tectonophysics, 153, 209–220.
388 R. J. Davies et al.

McKenzie, D. P. & Morgan, W. J. 1969. Evolution of Triple Junctions. Hardenbol, J., Jacquin, T., Farley, M. & Vail, P. R. (eds) Mesozoic and
Nature, 224, 125–133. Cenozoic Sequence Stratigraphy of European Basins. SEPM Special Publication,
McLeod, A. E., Dawers, N. H. & Underhill, J. R. in press. The propagation 60, 481–506.
and linkage of normal faults: insights from the Strathspey-Brent-Statfjord Stewart, S. A., Fraser, I. S., Cartwright, J. A., Clark, J. A. & Johnson, H. D.
array, Northern North Sea. Basin Research, 12. 1999. Controls upon Upper Jurassic sediment distribution in the Durward-
Patriat, P. & Courtillot, V. 1984. On the stability of triple junctions and its Dauntless area, UK blocks 21/11,21/16 UK. In: Fleet, A. J. & Boldy,
relation to episodicity in spreading. Tectonics, 3, 317–332. S. A. R. (eds) Petroleum Geology of Northwest Europe. Proceedings of the 5th
Penge, J., Taylor, B., Huckerby, J. A. & Munns, J. W. 1993. Extension and salt Conference. Geological Society, London, 879–896.
tectonics in the east Central Graben. In: Parker, J. R. (ed.) Petroleum Geology Streckler, M. R., Blisniuk, M. & Eisgbacher, G. H. 1990. Rotation in
of Northwest Europe, Proceedings of the 4th Conference. Geological Society, extension direction in the central Kenya Rift. Geology, 18, 299–302.
London, 1197–1209. Thomas, M. P. & Coward, M. P. 1995. Late Jurassic–Early Cretaceous
Prosser, S. D. 1993. Rift related linked depostional systems and their seismic inversion of the northern East Shetland Basin, northern North Sea. In:
expression. In: Williams, G. D. & Dobb, A. (eds) Tectonics and Seismic Buchanan, J. G. & Buchanan, P. G. (eds) Basin Inversion. Geological Society,
Sequence Stratigraphy. Geological Society, London. Special Publications, 71, London, Special Publications, 88, 285–306.
35–66. Thomson, K. & Underhill, J. R. 1993. Controls on the development and
Rattey, R. P. & Hayward, A. B. 1993. Sequence stratigraphy of a failed rift evolution of structural styles in the inner Moray Firth Basin. In: Parker,
system. The Middle Jurassic to Early Cretaceous basin evolution of the J. R. (ed.) Petroleum Geology of Northwest Europe. Proceedings of the 4th Conference.
Central and northern North Sea. In: Parker, J. R. (ed.) Petroleum Geology of Geological Society, London, 1167–1177.
Northwest Europe, Proceedings of the 4th Conference. Geological Society, London, Turner, J. D. 1997. The Subsidence of Sedimentary Basins. PhD thesis, University
215–249. of Edinburgh, UK.
Ring, U. 1994. The influence of preexisting structure on the evolution of the Underhill, J. R. 1991a. Implications of Mesozoic basin development in the
Cenozoic Malawi rift (East African rift system). Tectonics, 13, 313–326. western Inner Moray Firth, UK. Marine and Petroleum Geology, 8, 359–369.
Roberts, A. M. & Holdsworth, R. E. 1999. Linking the onshore and offshore
structures: Mesozoic extension in the Scottish Highlands. Journal of the Underhill, J. R. 1991b. Controls on Late Jurassic seismic sequences, Inner
Moray Firth, U.K. North Sea. A critical test of a key segment of Exxon’s
Geological Society, London, 156, 1061–1064.
original global cycle chart. Basin Research, 3, 79–98.
Roberts, A. M., Yielding, G. & Badley, M. E. 1990. A kinematic model for the
orthogonal opening of the late Jurassic North Sea rift system, Denmark- Underhill, J. R. & Partington, M. A. 1993. Jurassic thermal doming and
Mid Norway. In: Blundell, D. J. & Gibbs, A. D. (eds) Tectonic Evolution of the deflation in the North Sea: implications of the sequence stratigraphic
North Sea Rifts. Clarendon Press, Oxford, 180–199. evidence. In: Parker, J. R. (ed.) Petroleum Geology of Northwest Europe:
Proceedings of the 4th Conference. Geological Society, London, 337–345.
Roberts, A. M., Kusznir, N. J., Walker, I. M. & Dorn-Lopez, D. 1995.
Quantitative analysis of Triassic extension in the northern Viking Graben. Underhill, J. R. & Partington, M. A. 1994. Use of maximum flooding surfaces
Journal of the Geological Society, London, 152, 15–26. in determining a regional control on the intra-Aalenian (Mid-Cimmerian)
Sclater, J. G. & Christie, P. A. F. 1980. Continental stretching: an sequence boundary: Implications for North Sea basin development and
explanantion of the post-mid-Cretaceous subsidence of the Central North Exxon’s sea-level chart. In: Posamentier, H. W. & Wiemer, P. J. (eds) Recent
Sea Basin. Journal of Geophysical Research, 85, 3711–3739. advances in siliciclastic sequence stratigraphy. American Association of Petroleum
Sears, R. A., Harbury, A. R., Protoy, A. J. G. & Stewart, D. J. 1993. Structural Geologists Memoir, 58, 449–484.
styles from the Central Graben in the UK and Norway. In: Parker, J. R. Underhill, J. R. & Brodie, J. A. 1993. Structural geology of the Easter Ross
(ed.) Petroleum Geology of Northwest Europe, Proceedings of the 4th Conference. Peninsula, Scotland: implications for the Great Glen Fault Zone. Journal of
Geological Society, London, 1231–1243. the Geological Society, London, 150, 515–527.
Speksnijder, A. 1987. The structural configuration of Cormarant Block IV in Wakefield, L. L., Droste, H., Giles, M. R. & Janssen, R. 1993. Late Jurassic
context of the northern Viking Graben structural framework. Geologie en plays along the western margin of the Central Graben. In: Parker, J. R.
Mijnbouw, 65, 357–379. (ed.) Petroleum Geology of Northwest Europe: Proceedings of the 4th Conference.
Steckler, M. S. & Watts, A. B. 1978. Subsidence of the Atlantic-type Geological Society, London, 459–468.
continental margin off New York. Earth and Planetary Science Letters, 41, Wood, M., Bingham, G. & Mitchell, P. 1999. Predicting sand body distribu-
1–13. tion and porosity for Callovian to late Oxfordian succession; Block 9/13
Stephen, K. J., Underhill, J. R., Partington, M. A. & Hedley, R. J. 1993. The UKCS: an integrated approach. In: Fleet, A. J. & Boldy, S. A. R. (eds)
genetic sequence stratigraphy of the Hettangian to Oxfordian succession, Petroleum Geology of Northwest Europe: Proceedings of the 5th Conference, Conference.
Inner Moray Firth. In: Parker, J. R. (ed.) Petroleum Geology of Northwest Geological Society, London, 1281–1287.
Europe, Proceedings of the 4th Conference. Geological Society, London, 485–505. Ziegler, P. A. 1990. Tectonic and palaeogeographic development of the
Stephen, K. J. & Davies, R. J. 1998. Documentation of Jurassic sedimentary North Sea rift system. In: Blundell, D. J. & Gibbs, A. D. (eds) Tectonic
cycles from the Moray Firth basin UK North Sea. In: de Graciasky, P. C., Evolution of North Sea Rifts. Clarendon Press, Oxford, 1–36.

Received 27 April 1999; revised typescript accepted 26 February 2001

You might also like