You are on page 1of 20

The Role of Fault Interaction and Linkage in Controlling

Synrift Stratigraphic Sequences: Late Jurassic, Statfjord


East Area, Northern North Sea1

Nancye H. Dawers and John R. Underhill2

ABSTRACT imply that only from the perspective of fault


growth and linkage can the Late Jurassic structure
Examination of well-constrained three-dimen- and stratigraphy be fully understood.
sional seismic data demonstrates the role of fault
interaction and linkage in controlling the nature of
synrift sequences on the hanging wall of the INTRODUCTION
Statfjord East fault, a typical Late Jurassic structure
in the northern North Sea Brent province. Al- Field studies show that en echelon arrangement
though now a single fault, the Statfjord East fault of normal fault segments and the nature of the dis-
originally consisted of several en echelon seg- placement distribution along them result from seg-
ments, each of which defined individual subbasins. ment interaction and subsequent linkage (e.g.,
Structural and stratigraphic evidence, both along Peacock and Sanderson, 1991; Trudgill and
and across fault strike, indicates that the fault Cartwright, 1994; Dawers and Anders, 1995;
resulted from segment propagation, interaction, Cartwright et al., 1995, 1996). A number of model-
and linkage. Facies architecture, thickness varia- ing studies show that these spatial patterns result
tions, and the internal character of synrift forma- from stress perturbations due to fault slip (Segall
tions are temporally and spatially related to the sub- and Pollard, 1980; Bürgmann et al., 1994; Willemse
basin geometry. Variations in displacement along et al., 1996; Willemse, 1997; Crider and Pollard,
the fault segments exhibit characteristics of inter- 1998). Recent modeling results also show that fault
acting en echelon faults, including anomalous dis- interaction and linkage lead to marked temporal
placement gradients in regions of segment overlap. variability within an evolving fault array (Cowie,
We attribute the observed shifts in depocenters to 1998). Temporal variability in activity along seg-
local enhancement of displacement rates, resulting mented normal faults has been well documented in
from the interaction of neighboring fault segments. studies of neotectonic faults, i.e., faults active dur-
The results have far-reaching consequences for syn- ing the Quaternary, such as the Wasatch fault in
rift plays in the northern North Sea because they Utah (Machette et al., 1991). The behavior of seg-
mented faults over longer time scales, however, is
less well known. Because normal faults control the
©Copyright 2000. The American Association of Petroleum Geologists. All creation of accommodation space for syntectonic
rights reserved.
1Manuscript received August 3, 1998; revised manuscript received April deposition in rift basins (Leeder and Gawthorpe,
22, 1999; final acceptance June 30, 1999. 1987; Schlische, 1991; Gawthorpe et al., 1994), the
2Department of Geology and Geophysics, University of Edinburgh, Grant
Institute, The King’s Buildings, West Mains Road, Edinburgh EH9 3JW,
displacement history of normal faults should be
Scotland, United Kingdom; e-mail: Nancye.Dawers@glg.ed.ac.uk; recorded in the synrift stratigraphy.
John.Underhill@glg.ed.ac.uk In many continental areas, synrift stratigraphy is
This work was funded by the UK Natural Environment Research Council
(NERC), under the Realising Our Potential Award (ROPA) scheme (no.
poorly preserved or incompletely exposed due to
GR3/R9521), and Norsk Hydro Research Centre. Seismic interpretation was burial. Subsurface data, especially three-dimensional
undertaken using Schlumberger™ GeoQuest IESX software at facilities in (3-D) seismic surveys, thus provide an important
Edinburgh supported by the Centre for Marine and Petroleum Technology, Norsk
Hydro, Shell Expro, and Esso (UK) Ltd. We thank Gunn Mangerud and Pål Skott source of detailed information on the long-term
for their support and efforts in releasing data; Randi Jordan and Rolf Helland for structural evolution of normal fault systems and on
help with well data; and Anker Berge, Kjell-Owe Häger, Cai Puigdefàbregas, Arvid the development of rift basins; however, little atten-
Nøttvedt, Tom Dreyer, Roald Færseth, Anne Otelie Eide, Gunn Mangerud, Pål Skott,
Sarah Prosser, Paul Milner, Patience Cowie, Sarah Davies, Sanjeev Gupta, Aileen tion has been given to integrating the synrift strati-
McLeod, and Jon Turner for discussions. Bulletin reviewers Bruce Trudgill, Manuel graphic patterns with data on displacement varia-
Willemse, and Al Lacazette provided very helpful reviews; Juan Contreras, Patience
Cowie, and Simon Kattenhorn provided additional helpful comments. Gerard White
tions along the basin-bounding normal faults. A
assisted with figures and Chung-Lun Lau provided computer support. number of subsurface studies in the Upper Jurassic

AAPG Bulletin, V. 84, No. 1 (January 2000), P. 45–64. 45


46 Faults and Synrift Sequences

of the central and northern North Sea have de- 1°E Hydrocarbon field
0 50 km
scribed stratigraphic patterns that indicate spatial 2°E ZETA
and temporal variability of activity on normal faults
(e.g., Underhill, 1991a, 1991b; Prosser, 1993; D

E N
SNORRE
T N
Nøttvedt et al., 1995; Ravnås et al., 1997; Ravnås and S
A Area Shown
A
Bondevik, 1997). In this paper, we discuss the rela- E L in Fig. 3

G R A B
tionship between synrift stratigraphic patterns and T
E
normal fault development observed in a 3-D subsur- H VISUND

S
face data set to better understand processes of nor- STATFJORD
EAST
mal fault growth and rift basin development in time GULLFAKS

and space. Our results have implications for tectonic- I N


STATFJORD

stratigraphic models of rift basins, and show that A


S
B
temporal variability of stratigraphy in rift zones on a

G
BRENT

time scale of a few million years can be explained by STRATHSPEY I N 61°N


processes of fault interaction and linkage. K

V I
Location
GEOLOGICAL SETTING

Norway

N O R T H
The study area is located in the Brent hydrocarbon
province of the East Shetland Basin, on the western
flank of the Late Jurassic Viking Graben of the north-
ern North Sea (Figure 1). The East Shetland Basin
formed as a result of distributed continental exten-

U.K.
sion, which is estimated to be approximately 15%
100 km
(Yielding, 1990; Roberts et al., 1993a). The Middle
Jurassic (Bajocian) through Lower Cretaceous
(Ryazanian) extension is accommodated predomi- Figure 1—Location map of the East Shetland Basin and
study area. The Statfjord East fault lies along a large array,
nantly on large, east-dipping normal faults (e.g., Lee including faults bounding the Brent and Statfjord fields.
and Hwang, 1993). Most of the significant hydrocar-
bon fields within the Brent province are situated in
the uplifted footwalls (Figure 1). The Statfjord East
area lies along an alignment of fault-controlled hydro- Tarbert Formation and overlying Viking Group mud-
carbon fields, including Strathspey, Brent, and stones (Graue et al., 1987; Helland-Hansen et al.,
Statfjord (Figure 1). The Statfjord East area is some- 1992). Transgression continued with the deposition of
what unique in that the hanging wall also has been Bathonian–Oxfordian mudstones of the Heather
explored. There are several exploration wells in Formation and the subsequent shale-dominated
hanging-wall locations, and 3-D seismic coverage Draupne Formation (Vollset and Doré, 1984) [which is
extends across all of the hanging-wall basin. synonymous with the Kimmeridge Clay Formation in
UK nomenclature of Deegan and Scull (1977)]. The
Draupne Formation is coeval with numerous late
STRATIGRAPHY Oxfordian–Ryazanian sedimentary wedges throughout
the North Sea. A thin carbonate at the base of the
Figure 2 summarizes the relevant stratigraphy of Cromer Knoll Group marks the transition from active
the East Shetland Basin. The Upper Triassic–Lower extensional faulting to a phase of postrift thermal sub-
Jurassic Banks Group (Richards et al., 1993) [or sidence (Glennie and Underhill, 1998). Transgression
Statfjord Formation of Deegan and Scull (1977) and during the Late Jurassic in the northern North Sea
Vollset and Doré (1984)] and the Lower Jurassic enabled sea level to reach its maximum height region-
Dunlin Group were deposited in a thermally subsid- ally during the middle Kimmeridgian (sensu anglico)
ing basin following Permian–Triassic rifting (see and early Volgian (Hallam, 1969; Sneider et al., 1995).
Underhill, 1998). The overlying Middle Jurassic Brent Sea level then fell through the remainder of the
Group, except for the Tarbert Formation, represents Volgian and Early Cretaceous (Hallam, 1969; Rawson
a major deltaic complex, which advanced and and Riley, 1982).
retreated in response to thermal doming and subse- In this paper we focus on the seismic stratigra-
quent deflation (Underhill and Partington, 1993). phy of the synrift interval extending from the top
The onset of extension during the Middle Jurassic of the Brent Group through the base of the Cromer
(Bajocian) is marked by the drowning of the Brent Knoll Group. The base of the Cromer Knoll Group
delta and subsequent deposition of the shallow-marine generally is referred to as the base Cretaceous
Dawers and Underhill 47

AGE STRATIGRAPHY TECTONIC EPISODE conclusions. Differential compaction between foot-


Cretac- Valanginian Cromer Knoll Group “BCE” Post-Rift Subsidence walls and hanging walls of large faults also may
eous Ryazanian effect throw patterns; however, the throws here

Banks Dunlin Brent Viking or Humber


Volgian Draupne or
Kimmeridge are not large enough for this to be a cause of great
Late

Group
Kimmeridgian Clay Fm
Rifting concern.
Oxfordian
Callovian Heather Formation
Jurassic
Middle

Bathonian
Tarbert Fm STATFJORD EAST FAULT
Group Group Group
Bajocian Ness Fm Thermal Doming
Rannoch Fm Etive Fm
Aalenian Broom Fm & Deflation
“MCU”
Toarcian Drake Fm The Statfjord East fault is an approximately
Cook Fm
Pliensbachian 18-km-long, north-northeast–striking fault (Figure
Early

Burton Fm
Amundsen Fm Post-Rift Thermal
Sinemurian
Nansen Formation
Subsidence 4). Figure 4 shows the geometry of several regional
Hettangian Statfjord Formation layers that are offset by the fault. On the level of
Triassic Rhaetian the BCE, the fault has an overall continuous trace
with two left-stepping bends (Figure 4A). At the
Figure 2—Stratigraphic framework and tectonic events
of the East Shetland Basin. BCE = base Cretaceous event
northernmost bend, a fault splay extends basin-
(see text), MCU = middle Cimmerian unconformity. ward from the principal fault, i.e., the fault surface
where most of the displacement is taken up (Figure
4A). On the deeper horizons, such as the top of the
Heather Formation (Figure 4B) and the top of the
unconformity despite its Early Cretaceous age and Brent Group (Figure 4C), a series of northeast- to
the fact that in some structurally deep settings north-northeast–striking intrabasinal splays extend
there may be no stratigraphic omission (Rattey and from the principal fault surface into the hanging-
Hayward, 1993). This prominent, regional seismic wall basin. At a more regional scale, the Statfjord
marker results from a sharp acoustic impedance East fault is partially overlapped by the Statfjord
contrast between carbonates of the Lower Cre- fault (Figures 3, 4A), which becomes the more
taceous Cromer Knoll Group and underlying forma- dominant fault toward the south, outside our study
tions. We refer to it here as the base Cretaceous area (Figure 1).
event (BCE) because no stratigraphy is missing in
the Statfjord East hanging wall.
The top of the Banks Group has been mapped as Footwall
a prerift marker horizon for fault displacement anal-
ysis. Although the Jurassic rifting may have reacti- The uplifted footwall of the Statfjord East fault is
vated some earlier Permian–Triassic structures (see relatively simple in profile (Figure 5). A thin cap-
Færseth 1996), the Statfjord East fault is interpreted ping of Heather Formation occurs along the central
as being Late Jurassic–Early Cretaceous in age portion of the footwall. In well 33/9-12, the Tarbert
(Nyberg, 1987; Dahl and Solli, 1993; Færseth, Formation also is present, as is a thin drape of
1996). upper Volgian Draupne Formation shale. South of
this well, the Draupne Formation thickens across a
significantly eroded part of the Statfjord East foot-
DATA SETS wall, along which the entire Brent succession and
the upper part of the Dunlin Group are missing
The data sets include four overlapping 3-D seis- (Figure 5B). The hanging-wall geometry, described
mic surveys and data from 32 exploration wells in the following paragraphs, is more complex.
located within Norwegian Block 34/7 and the adja-
cent part of Block 33/9 (Figure 3). Biostratigraphic
zonation, well correlation, and facies analysis of Hanging-Wall Structure and Stratigraphy
cored intervals within the Statfjord East hanging
wall are described in Häger and Smelror (1997), The intrabasinal faults, which extend from the
Nøttvedt et al. (in press), and Dawers et al. (1999), principal fault surface into the hanging-wall basin
and are used here to place temporal constraints on (Figure 4), define several subbasins within the over-
the structural evolution of the area. all half graben. The subbasins are evident in the
For the purposes of this paper, the seismic interpre- geometry of the horizons shown in Figure 4 and are
tation has not been depth converted. Although veloci- illustrated in the along-strike seismic section shown
ty variations in detail may affect the patterns observed in Figure 6. The along-strike section shows that the
in synrift thickness and fault throw, well control is suf- subbasin bounding faults cut progressively younger
ficiently closely spaced to suggest that these varia- strata toward the north (Figure 6). Only the
tions will not change our general observations and southernmost intrabasinal fault is associated with
48 Faults and Synrift Sequences

Figure 3—Structural lt 61 30’


framework of the study 3-D Surveys au
o nf 34/7-4
area; faults shown are s
ge83 chi
Middle Jurassic and Mur re 34/7-1 N
younger. Sections AA′ sg8431
n or
S

Block 33/9
Block 34/7
and BB′ are shown in e86
Figures 5 and 6, and 34/7-7
sections aa′ and bb′ are sg9201 34/7-6

shown in Figure 7. 34/7-3


Also shown are the wells
used to constrain the
seismic interpretation, 34/7-10

Inner Snorre fault


outline of area shown in 34/7-15S
Area Shown 33/9-15 34/7-13
Figure 4, oil fields, and
in Fig. 4
outlines of three- A’
33/9-16 B’34/7-19
dimensional surveys.

lt
34/7-16

fau
Survey line, CDP (common Data
depth point) spacings are Gap

st
34/7-8
e86: 18.75 m, 12.5 m;

Ea
is
sg9201: 12.5 m, 12.5 m;
fjord 34/7-5 Vigd
sg8431: 25 m, 12.5 m; at
St
ge83: 50 m, 25 m. 33/9-5
Statfjord
34/7-17
East
tfjord
34/7-23A,S
Statfjord fault

33/9-4
b 33/9-7
Sta

33/9-12
34/7-24S 34/7-18
33/9-3
34/7-21 34/7-2

34/7-21A s
a b’ di
33/9-9 or
A 34/7-12 T 34/7-22
B
34/7-14
a’ 34/7-25

Hydrocarbon
Field
2 00’

61 15’

2 20’
0 5 10 km

substantial thickening in the Heather Formation, Formation (Dawers et al., 1999; Nøttvedt et al., in
suggesting that the intrabasinal faults to the north press). Several reflectors onlap the top Kimmer-
were much less active during the deposition of the idgian reflector (Figure 7A). The erosion of the
Heather Formation, or perhaps had not yet formed. Statfjord East footwall here (see also Figure 5) sug-
The geometry created by the intrabasinal faults pro- gests that these internal reflectors, observed only in
duces apparent fold structures transverse to the this part of the basin, are related to scarp degrada-
trend of the overall hanging-wall basin. tion. Overlying these are more continuous reflec-
Figure 7 shows two seismic sections across the tors that overstep the eroded footwall (Figure 7A).
basin. In the deepest part of the basin, adjacent to These, in turn, are overlain by discontinuous inter-
the southernmost part of the Statfjord East fault, nal reflections that are truncated by the BCE.
both the Heather Formation and Draupne Formation The thin drape of upper Volgian Draupne
are distinctly wedge shaped in cross section (Figure Formation shale in well 33/9-12, located just north
7A). The internal nature of the Draupne Formation of the most intensely eroded footwall area, suggests
consists of an upper shallow-marine sequence and a that this portion of the Statfjord East fault had
lower shale sequence separated by a prominent, become relatively inactive by the late Volgian.
top Kimmeridgian reflector that onlaps the Heather Note, however, that the Statfjord fault, which lies
Dawers and Underhill 49

Figure 4—Map-view geometry (color-contoured in two-


way traveltime) of horizons offset by the Statfjord East
fault. (A) Base Cretaceous horizon, (B) top of the
Heather Formation, and (C) top of the Brent Group.

west of and overlaps the Statfjord East fault, shows a


small but clear displacement of the BCE (Figure 7A),
suggesting that displacement slowed or ceased on
this portion of the Statfjord East fault in the latest
Jurassic, and by the Early Cretaceous displacement
was being taken up on the main Statfjord fault.
Figure 7B is a traverse through wells 33/9-12,
34/7-24S, 34/7-21A, and 34/7-21 (see Figure 3).
Both the Heather Formation and Draupne For-
mation are wedge shaped (Figure 7B), although in
this subbasin the Heather Formation does not
thicken as drastically as it does farther south (see
Figure 7A). Shown in parentheses on Figure 7B is
the thickness of the Tarbert Formation encoun-
tered in wells 33/9-12, 33/9-21, and 33/9-21A but
not completely drilled through in well 33/9-24S.
The hanging-wall thickness of the Tarbert For-
mation is roughly double that of the footwall. This
ratio is consistent with previous observations sug-
gesting that the Tarbert is an initial synrift unit
(e.g., Yielding et al., 1992; Ravnås et al., 1997;
Underhill et al., 1997).

Overlaps Between the Statfjord East Fault and


Intrabasinal Splays

Figure 8 shows the cross sectional geometry of


the zones of overlap between the principal fault
surface, i.e., the fault with the greater displace-
ment, and the intrabasinal fault splays (Figure 8A).
Both the principal and intrabasinal faults clearly dis-
place the BCE at the northern overlap zone (Figure
8B). The middle intrabasinal splay does not dis-
place the BCE but does displace the top of the
Heather Formation (Figure 8C). The southern intra-
basinal does not displace the BCE and only locally
displaces the top of the Heather Formation (Figure
8D, also see Figure 4B). The southern and middle
intrabasinal faults appear to have been active early
relative to the principal fault surface, manifested by
synrift thickness variations. Across the southern
zone, thickening is observed in the Brent Group
and Heather Formation across the intrabasinal splay
(Figure 8D), whereas thickening is observed in the
Heather and Draupne formations across the middle
intrabasinal fault (Figure 8C). Note also that the
Heather Formation here is on average thinner than
that farther south, whereas the Draupne Formation
is overall of similar thickness here and toward the
south. Across the northern overlap zone, the only
synrift unit thickening toward the faults is the
50
(A)
SSW NNE
A A’
Faults and Synrift Sequences

(B) 33/9-12 34/7-5

2250 Footwall Erosion Heather Fm


e taceous
ous Brent Group base Cr
ce
ta
2500 re Dunlin Group
base C
Draupne Fm Banks Group +
2750
older units

Depth (msec TWTT)


3000

0 2 km

Intersection with a-a’


(Figure 7a)
Intersection with b-b’
(Figure 7b)

3250

Figure 5—(A) Along-strike seismic traverse AA′ (see Figure 3 for location) and (B) line drawing of the seismic section of the Statfjord East
footwall. Note the simple footwall structure and area of footwall erosion in the south. TWTT = two-way traveltime.
(A)
SSW NNE
B B’

(B) 34/7-24S 34/7-23A


-23S
2250 Draupne Fm
e Cr
bas et

Intersection with a-a’


(Figure 7a)
Intersection with b-b’
(Figure 7b)
aceous Heather Fm
2500 Draupne Fm Brent Group
ctor
fle
idgian re
m er
Kim
p
to Heather Fm
2750 Draupne Fm Brent Group Dunlin Group
Banks Group +
Heather Fm older units

Depth (msec TWTT)


3000 Dunlin Group
BanksGroup+
Brent Group older units
3250
Dawers and Underhill

Dunlin Group 0 2 km

Figure 6—(A) Along-strike hanging-wall traverse BB′ (see Figure 3 for location) and (B) line drawing of the seismic section. Intrabasi-
nal highs are associated with the hanging-wall splays. Thickness variations suggest the southernmost splay was most active during
51

Heather deposition, whereas those toward the north were more important during Draupne deposition. TWTT = two-way traveltime.
52
W E
a a’
Faults and Synrift Sequences

2250

BCE
2500 BCE

Intersection with A- A’
(Figure 5)
Intersection with B-B’
(Figure 6)
top Kimmeridgian
reflector
t up
Bren Gro
nt
2750 Bre
Group Draupne Fm
Dunlin
p+ Fm
rou
n ther
a ks Gnits
B er u Hea
3000 old p+
ou
Gr
p a nks units
Statfjord rou B er
nlin G old

Depth (msec TWTT)


fault Statfjord Du
3250 East fault
0 2 km
(A)
Figure 7—(A) Seismic line aa′ (see Figure 3 for location) and line drawing of the seismic section across the southern part of the
Statfjord East fault showing the eroded footwall block between the Statfjord East fault and the main Statfjord fault to the west. (B)
Traverse bb′ (see Figure 3) and line drawing across wells 33/9-12, 34/7-24S, 34/7-21A, and 34/7-21. TWTT = two-way traveltime.
W E
b b’

33/9-12 34/7-24S 34/7-21A 34/7-21

Intersection with A- A’
(Figure 5)
Thin Draupne &
2250 Heather;
Tarbert (~23 m)

Intersection with B-B’ (Figure 6)


Bend in Section
top Kimmeridgian
m reflector
BCE rF Fm )
e Fm the
2500 Draupn a Draupne 5m
He BCE (~5
er F m m)
Heath rt ~63
up (Tarbe
t Gro
Bren
2750 roup
lin G
Dun p oup
Statfjord rou n Gr
ntG Dunli
p+ Bre
rou East fault

Depth (msec TWTT)


a n ks Gnits p up +
B er u rou s Gro
3000 old lin G
Dawers and Underhill

Bankr units
Dun olde

0 2 km
(B)
53

Figure 7—Contiued.
54 Faults and Synrift Sequences

2000
0 2 km x x’
N
Northern Northern
x Overlap

Depth (msec TWTT)


2250 Overlap
Zone Zone Draupne Fm
BCE
la y e r Fm
Heath
e

sp 2500 up
p t Gro
fac

al rou Bren
sin in G
ba D unl
Sur

x’
ra
int

2750 Banks Group +


older units
y Middle
Overlap (B)
Zone 3000
y

2000
la

breaching
sp

y y’
ult

segments
l
ina

Middle
Fa

z
as

Overlap

Depth (msec TWTT)


2250
y’
rab

Zone
BCE
int

up
Southern nt Gro
Bre Draupne Fm m
pal

Overlap 2500
er F
splay Zone ath
asinal He
nci

b up
ra- Gro
int nlin
ri

2750
P Banks Group + Du
older units
z’ (A) (C)
3000
2000
z z’
0 1 km

Southern
Depth (msec TWTT)

2250

BCE Overlap
Zone
2500

Draupne Fm
Fm
ther
Hea
2750
Banks Group + oup
older units t Gr
Br e n
roup
3000
u n lin G
D
(D)
Figure 8—(A) Fault structure (at top of the Heather Formation) and section locations for parts B–D; (B) northern
overlap zone; (C) middle overlap zone; (D) southern overlap zone. The principal fault surface accommodates most
of the displacement across all three overlap zones. Intrabasinal fault activity youngs progressively from south to
north, as evidenced by thickness variations and lack of offset of top Heather (D) and top Draupne (C). TWTT = two-
way traveltime.

Draupne Formation (Figure 8B), but it is possible In the case of the southern and middle splays, the
that the wedge-shaped character of the Draupne cessation of activity on the intrabasinal splay was
Formation here in part may be the result of erosion. coupled with displacement localization on the over-
The relationships described in this paper and lapping principal fault surface toward the west
illustrated in Figure 8 indicate that, initially, displace- (Figure 8B, C). This scenario is inferred from the syn-
ment occurred along the intrabasinal faults, first rift thickness variations and the lack of offset of the
along the southern fault (Figure 8D), then along the upper synrift layers on the southern and middle
middle fault (Figure 8C), and finally on the northern- splays, which were noted in the discussion of
most fault (Figure 8B). We observe a pattern of activ- Figures 4 and 6. These relationships suggest that the
ity and then cessation of activity on the intrabasinal principal fault surface evolved through the breach-
faults that progresses from south to north with time. ing of the overlap zones (Figure 8A).
Dawers and Underhill 55

(A) Figure 9—(A) Along-strike throw


profile for the top of the (prerift)
Throw - top Banks Group Banks Group. (B) Map view of the
800
fault structure on this datum.
700 Principal Surface Summing in the throws of the
splays reduces the magnitude of
Hanging wall Splays
600 local throw minima along the
Throw (msec TWTT)

Footwall Splay principal fault surface, a finding


500
that is consistent with the
Summed Throw hanging-wall splays representing
400
remnants of en echelon segments
300 that became inactive as breaching
structures formed in the footwall.
200 TWTT = two-way traveltime.
100

0
0 2 4 6 8 10 12 14 16 18
Distance (km)

(B)
rd
tfjo
Sta

Principal Fault Surface


East
Statfjord
Footwall
Splay

01
sg92

e86
N

Intrabasinal
Fault Splays
0 2 km

ALONG-STRIKE DISPLACEMENT PROFILES Figure 9B shows the map-view fault geometry on


the top Banks Group, which illustrates the varia-
Figures 9 and 10 show along-strike displacement tion in the horizontal component, or heave, along
profiles for the Statfjord East fault. Figure 10 also the fault segments. Throw along principal seg-
includes data from the overlapping portion of the ments of the fault, i.e., those segments with the
Statfjord fault to the west. Displacement data are greatest displacement, is plotted as filled squares
limited because of partial erosion of the Heather (Figure 9A). The maximum throw occurs about 2
Formation and Brent and Dunlin groups on the km north of the southern edge of the e86 survey
southern portion of the footwall; consequently, we and tapers off gradually northward. Note that local
show curves for the top Banks Group (Figures 9, displacement minima correlate to areas where
10) and BCE horizons (Figure 10). The displace- hanging-wall splays are observed in map view
ment measure used is throw, i.e., the vertical com- (Figure 9B). Throw across four hanging-wall splays
ponent of the total displacement vector. Throw and one footwall splay also is plotted in Figure 9A.
was measured on every tenth line in the e86 data Summing the throw across the splays with that
set, which has lines spaced every 18.75 m and ori- measured along the principal fault trace reduces
ented parallel to the east-southeast extension direc- the magnitude of the local throw minima, hence
tion. The northern tip of the Statfjord East fault lies producing a relatively smooth and continuous
north of the e86 survey; in the sg9201 survey throw profile (Figure 9A).
(Figure 10), throw was measured using similarly The pattern of displacement variation observed
spaced traverses parallel to the e86 lines. in Figure 9A is typical of faults that evolve from
Variation in the throw of the top Banks Group mechanically interacting en echelon segments
along the Statfjord East fault is shown in Figure 9A. (Dawers and Anders, 1995). For initially unlinked
56 Faults and Synrift Sequences

interacting fault segments, the throw profiles are mimics that of the Banks and Brent groups, which
asymmetric with high displacement gradients have throw patterns here consistent with that
toward the region of fault overlap (e.g., Peacock expected for a prefaulting datum; however, the BCE
and Sanderson, 1991; Willemse et al., 1996). throw pattern along this splay is consistent with a
Following the breaching of the overlap zone (typi- synfaulting interpretation. The conclusion drawn
cally across the part of the overlap zone closest to from Figure 11 is that activity localized on the mid-
the mutual footwall, as observed here), displace- dle splay during Heather deposition, whereas signif-
ment accrues on the developing fault surface, icant activity did not localize on the northern splay
whereas activity on the previously active overlap- until during Draupne deposition. This pattern is
ping intrabasinal fault diminishes (Childs et al., consistent with northward growth of the Statfjord
1995; Cartwright et al., 1996). The observation that East fault by segment linkage; furthermore, this pat-
throw on the intrabasinal splays spatially coincides tern confirms the interpretation of northward
with throw anomalies of similar magnitude along migration of activity on the intrabasinal faults based
the principal segment indicates that the fault seg- on the cross sectional geometries of the fault over-
ments are geometrically coherent (Walsh and lap zones illustrated in Figure 8.
Watterson, 1991). In other words, the intrabasinal The data shown in Figure 11 do not rule out a
splays together with the principal fault surface component of vertical propagation in addition to
form a smoothly varying displacement profile, as lateral propagation through linkage. The basinward
would be expected for a single fault (e.g., Dawers dip of hanging-wall reflectors immediately adjacent
and Anders, 1995), and thus act on the large scale to the fault segments shown in Figures 7 and 8 is
as a single fault. consistent with vertical propagation from depth.
Figure 10A compares the total throw profile of This geometry is similar to observations made
the top Banks Group on the Statfjord East and along synsedimentary normal faults in other parts
Statfjord faults (upper curves) with that observed of the North Sea and in the Gulf of Suez rift, where
for the BCE (lower curves). The difference in the basinward-dipping strata have been shown to rep-
character of the curves for top Banks Group and resent the limbs of monoclines formed initially by
the BCE horizons is that the BCE horizon repre- vertical propagation of blind normal faults
sents the youngest synfaulting datum, and thus (Withjack et al., 1990; Gawthorpe et al., 1997;
records only a very small amount of displacement. Gupta et al., 1999); however, vertical propagation
The important point in Figure 10A is that there is alone would not explain the difference in the
less throw of the Banks Group on the Statfjord fault throw data shown in Figure 11.
relative to the Statfjord East fault, whereas the
reverse is true of the BCE. At BCE level, only the
northern part of the Statfjord East fault exhibits DISCUSSION
throws significantly greater than seismic resolution
(approximately 15 ms). This pattern is consistent The Statfjord East fault is similar to normal faults
with the stratigraphic interpretation that late in the described from other areas. Studies in the Neogene
rifting displacement became localized on the Basin and Range province in the western United
Statfjord fault and the overlapping portion of the States and Mesozoic rift basins in the eastern
Statfjord East fault became relatively inactive United States illustrate hanging-wall subbasins
(whereas the northern part of the Statfjord East separated by intrabasinal highs or transverse folds
remained active). The previously active hanging- adjacent to fault segment overlaps (Anders and
wall splays, particularly those in the south, do not Schlische, 1994; Schlische, 1995; Schlische and
show clear displacement at the BCE level (Figure Anders, 1996), similar to the structures observed here
10B, see also Figure 4A). (Figure 6). Along a Basin and Range fault in Utah,
It is not feasible to construct throw profiles using Wu and Bruhn (1994) observed that segments step-
synrift horizons below the BCE for the principal ping toward the footwall became progressively
Statfjord East segments because of either erosion or linked from the center of the structure outward. In
nondeposition on the footwall; however, sufficient east Africa, dated volcanic units along Lake Malawi
data are available for the two northern intrabasinal (Ebinger, 1989) and 2-D (two-dimensional) seismic
splays and are shown in Figure 11. Note that for the data (Contreras, 1999) indicate that segments along
middle splay, the throw pattern for the top of the the lake’s western fault system were initially individ-
Brent Group is similar to that for the prerift Banks ual subbasins. Sedimentological data from the Suez
Group, whereas the throw profile for the top of the rift system in Sinai, Egypt, demonstrate that segment
Heather Formation records less accumulated dis- interaction controlled Miocene sedimentary disper-
placement, consistent with this being a synfaulting sal systems there (Gupta et al., 1999). Although
horizon. Along the northern intrabasinal splay, the these studies illustrated the importance of the fault
throw gradient for the top of the Heather Formation linkage process, the paucity of hanging-wall data
Dawers and Underhill 57

(A) Figure 10—(A) Comparison of


800
Throw - top Banks Group throw profile for the base
Statfjord East fault
Cretaceous event (BCE) with top
700 Statfjord fault Banks Group for Statfjord East and
main Statfjord faults. Despite less
600 Throw - base Cretaceous throw of the top Banks Group on
Throw (msec TWTT)

500
Statfjord East fault the Statfjord fault relative to the
Statfjord fault Statfjord East fault, the throw of the
400 BCE is greater on the Statfjord fault
than on the Statfjord East fault.
300 At BCE level, only the northern
part of the Statfjord East fault
200
has throws greater than
100 Throw approximately 15 ms. Thus late
Resolution
ca. 15 msec in rifting deformation localized
0 onto the Statfjord fault with the
0 2 4 6 8 10 12 14 16 18 southern portion of the Statfjord
Distance (km) East fault becoming inactive; the
northern part of the fault remained
active. (B) Interpretation of faults
(B) from a BCE dip-magnitude map.
rd The previously active southern and
tfjo
Sta middle hanging-wall splays do not
show clear displacements at BCE
level. TWTT = two-way traveltime.
East
Statfjord

e86

horizon
taceous
base Cre
nt
poorly re
solved pme
e Escar
Snorr
N

0 2 km

01
s g 92

and the erosion of footwall marker horizons in (e.g., Peacock and Sanderson, 1991; Trudgill and
most subaerial settings limits our understanding of Cartwright, 1994; Cartwright et al., 1995, 1996; Dawers
the temporal evolution of faults. Because of the 3-D and Anders, 1995; Childs et al., 1995; Willemse et
nature of the subsurface data set and the submarine al., 1996; Crider and Pollard, 1998;). The relatively
setting, this study provides important insights into smooth throw profile, produced by summing the
the temporal and spatial evolution of normal faults Banks Group throws across the intrabasinal splays
over the time scale of the entire rifting episode and with that observed along the principal fault surface,
has implications for the tectonic-stratigraphic clearly demonstrates that the intrabasinal faults are
development of rift basins. associated with the evolution of the Statfjord East
fault. Structural and stratigraphic evidence suggests
an overall northward migration of fault activity,
Implications for Evolution of Normal Faults which is further illustrated in Figure 12. Figure 12
shows isopachs (in two-way traveltime) of the
Displacement variations observed along the Heather Formation and the upper Oxfordian–
Statfjord East fault (e.g., Figure 9) are consistent with Kimmeridgian Draupne Formation shale sequence.
previous studies of fault interaction and linkage During the Bathonian–late Oxfordian, sediment
58 Faults and Synrift Sequences

250
Throw (msec TWTT)

Middle Intrabasinal Fault


200
Northern Intrabasinal Fault
150

100

50
M ono
cline
0
400 450 500 550 600 650 700 750 800
e86 Inline Number

top Banks Group top Heather Formation 0 1 km


top Brent Group top Draupne Formation/BCE
Figure 11—Throw profiles for middle and northern intrabasinal faults. On the middle fault, the throw pattern for
top Brent is similar to that for (prerift) top Banks Group, whereas the top Heather profile records less accumulated
displacement, consistent with this being a synfaulting horizon. For the northern intrabasinal splay, all horizons
except for the BCE (base Cretaceous event) have gradients expected for a prefaulting datum. Activity occurred on
the middle splay during Heather deposition, but significant activity did not occur in the north until Draupne depo-
sition, consistent with northward growth of the Statfjord East fault by segment linkage. TWTT = two-way traveltime.

accumulation was most pronounced along the the south. The northern segments, i.e., toward
southernmost part of the fault (Figure 12A). With the tip, experienced enhanced activity and even-
time, the locus of activity shifted toward the north- tually became incorporated into the growing
east. Isopachs of the late Oxfordian–Kimmeridgian structure by linkage. The migration of the linkage
lower Draupne Formation shale indicate that con- events is producing the overall northward propa-
tinued subsidence occurred along the southern gation of the fault.
part of the fault and the development of a new This study has not attempted to address the early
depocenter along a segment toward the northeast patterns of fault initiation. A recent sedimentologi-
(Figure 12B). We suggest that deposition of the cal study of the earliest synrift unit in this area, the
Heather Formation and the lower part of the Tarbert Formation, shows, however, that the intra-
Draupne Formation occurred during a phase of basinal faults controlled small depocenters during
ongoing linkage of fault segments. the Bajocian (Davies et al., 1998). Thus the intra-
The structural relationships and the depocenter basinal faults probably initiated early as a simple en
patterns we observe are consistent with the pat- echelon array along which the segments later
tern of stress redistribution due to slip on a nearby became progressively more active and linked to
interacting fault (Segall and Pollard, 1980; Bürgmann form the Statfjord East fault. Lateral growth of the
et al., 1994; Willemse et al., 1996; Willemse, 1997). overall fault structure by linkage was on the order
Fault segments in certain positions relative to of 5–10 km in approximately 20 m.y. The time
active segments develop increased displacement scale of activity increasing on an individual seg-
rates as a result of frequent stress redistribution ment and that segment becoming incorporated
following episodes of slip (Cowie, 1998; Gupta into the large fault structure probably is on the
et al., 1998). In the case of normal faults, en ech- order of 5 m.y. It is difficult to infer the slip rates for
elon or co-planar segments experience enhanced the Statfjord East fault segments because of the
activity, whereas faults in the immediate hanging absence of synrift units on the footwall. In the case
wall and footwall are inhibited from further of the intrabasinal faults, the data in Figure 11
activity. In the case discussed here, the data sug- imply slip rates on the order of a few millimeters
gest that the process of fault growth involved per 1000 yr (assuming reasonable interval veloci-
increased activity on one or more segments in ties for the Viking Group).
Dawers and Underhill 59

Heather Fm Lower Draupne Fm


Bathonian- Late Oxfordian -
Late Oxfordian Kimmeridgian

34/7-5 34/7-5

34/7-17 34/7-17

34/7-23A 34/7-23A
33/9-7 33/9-7
34/7-23S
34/7-23S
Limit of Seismically
100

33/9-12 Main 33/9-12


Main 34/7-24S Depo- 34/7-24S
Depo- 250
50 centers
centers
Re
so
lve

25
d Heather

34/7-21 34/7-21
250 50
300 200 60 40
30
350
34/7-21A 35 34/7-21A
S 400
S 25
10
15

er
150
ta

ta

40

h
t
tf

ea
tf

20
jo

jo

45
H
rd

s
rd

p
edg

la
fa

fa

e of

on
ul

stud
ul

y ar N
t

ea

(A) 0 2 km (B)
Figure 12—Thickness variation (in two-way traveltime) of shale successions. (A) Heather Formation and (B) lower
Draupne Formation. Fault segments that control thickness variation are shown as bold lines; faults that were much less
active or had not yet formed are shown as thinner lines. Modified from Dawers et al. (1999) and Gupta et al. (1998).

Implications for Tectonic-Stratigraphic shale-prone sequences with aggradational stratal


Models of Rifts geometries indicates high subsidence rates, i.e., dis-
placement rates on faults are creating accommoda-
Previous studies in the North Sea and elsewhere tion space faster than it can be filled with sediment.
illustrate several tectonic-stratigraphic features that Prosser (1993) referred to these patterns as the rift
are common to half-graben basins and that can be initiation, when displacement rates are relatively
linked to spatial and temporal variability of normal low, and the rift climax, when displacement rates
fault activity. The facies distribution in shallow- are relatively high. The rift-basin shale sequences
marine and lacustrine rift settings shows broadly commonly are overlain by a sand-prone sequence
similar patterns consisting of shallow-marine or flu- as faulting slows and eventually ceases.
vial sediments during the initiation of the basin The stratigraphic framework of the Statfjord East
overlain by deeper water shales (e.g., Schlische, area consists of a similar rift-basin succession, i.e.,
1991; Prosser, 1993). The frequent association of Tarbert Formation sandstone overlain by shales of
60
Southern 2 Segments Segments Interact,
34/7-5
Most Active During Focusing
34/7-5
Heather Deposition Shale Deposition
New Depocenter
33/9-7 34/7-17
34/7-23A
Develops 34/7-17
33/9-7
34/7-23S
33/9-12 34/7-23A
34/7-23S
34/7-24S Slight Northward 33/9-12
Shift of Depocenter 34/7-24S

34/7-21 Bathonian - Late Oxfordian -


34/7-21A 34/7-21A
n its its Kimmeridgian
Late Oxfordian n
up e rU 34/7-21 e rU
ro ld up
tG +O Heather Fm
ro Old
en tG +
Br o up en up
Br o
Gr Gr
in in
nl (A) nl (B)
Du N Du N
0 0
2 km 2 km
Faults and Synrift Sequences

Activity Localizes on Further Northward Migration


Stafjord Fault & Diminishes of Faulting and
on Southernmost ? Shoreline Migration
Statfjord East Segment 34/7-5
?
33/9-5 34/7-5
34/7-17 33/9-5
33/9-7
? 34/7-23A
Continued Draupne 33/9-7
Shale Deposition 34/7-23S 34/7-17
Thin Draupne 33/9-12
34/7-23A
Shale Drape 34/7-23S
? 33/9-12
?
34/7-24S Inactive Portion
34/7-24S
e of En Echelon
f ac
34/7-21 re Segments
rp
ca 34/7-21A Volgian ho Ryazanian
S S
d
de
ra ce n its 34/7-21A 34/7-21
eg a ts
D r ef e rU ni
ld p U
ho up O (C) ou lder (D)
S ro p + Gr
Heather Fm tG u ent +O
Statfjord ren Gro Heather Fm Br roup
Fault B lin
n G
Du lin
D un
Statfjord East N N
0 0
2 km 2 km

Figure 13—Evolution for the Statfjord East area from the time of (A) Heather deposition, (B) lower Draupne shale deposition, (C) upper Draupne
deposition during the Volgian, and (D) earliest Cretaceous.
Dawers and Underhill 61

the Heather and Draupne formations, followed by activity along the southern part of the Statfjord East
discontinuous shoreface sandstones in the upper- fault appears to be associated with the transfer of
most part of the Draupne Formation (Figure 2) (see activity onto the Statfjord fault, which overlaps the
also Dawers et al., 1999). The Tarbert Formation is Statfjord East structure. Structural evidence of inter-
known from well data to be widespread in this part action at this larger scale is the skewed nature of the
of the East Shetland Basin and thickest in structural throw distribution on the Statfjord East fault seen in
lows (e.g., Underhill et al., 1997). These characteris- Figure 9. If the Statfjord East structure as a whole
tics are consistent with the Tarbert Formation hav- were to be considered a single isolated fault, then we
ing been deposited across a region of distributed would expect the throw distribution to be symmetric
faulting with little total strain early in the rifting (e.g., Dawers and Anders, 1995). This view implies
episode (Yielding et al., 1992). The deeper marine, that processes discussed here for the intrabasinal
more focused nature and wedge-shaped geometry of faults, and hence temporal variability, should be
the Heather Formation and lower Draupne observed along the whole of the Strathspey-Brent-
Formation shale depocenters observed here (Figure Statfjord-Statfjord East fault system (Figure 1), and
12), and on the larger scale as well (Lee and Hwang, probably on similar time scales. Some evidence of this
1993), may be attributed to the enhancement of dis- view exists. The Statfjord fault within our study area
placement rates, and hence hanging-wall subsidence appears to have been most active relatively late in the
rates, as individual fault segments interacted. During rifting episode, whereas high rates of displacement
deposition of the sand-prone upper Draupne For- are inferred to have occurred earlier (approximately
mation (Figure 13C, D), despite the regional fall in late Oxfordian) for the part of the Statfjord segment
sea level, evidence for local shoreline retrogradation south of our study area (Roberts et al., 1993b; A. E.
suggests that fault growth led to a relative sea level McLeod, 1999, personal communication).
rise within our study area.
Our observations suggest that the structural char-
acter of the rift climax is that of a still highly seg- CONCLUSIONS
mented fault (Figures 12; 13A, B). We interpret the
intrabasinal splays to represent the remnants of en Results of integrating seismic stratigraphy with
echelon segments that show a pattern of increased fault analysis suggest that the Statfjord East fault of
activity and cessation that migrated northward with the northern North Sea evolved by fault segment
time. This process of growth by segment interaction linkage during Late Jurassic rifting. The intrabasinal
and linkage encompasses the time periods recorded fault splays in the Statfjord East hanging wall repre-
by the wedge-shaped shale successions of the sent inactive tips of earlier formed en echelon seg-
Heather Formation and lower Draupne Formation ments; thus they represent an integral part of over-
(Figures 12; 13A, B). This evolution contrasts with all fault development rather than being related to
the view that the border fault is a well-developed other tectonic events or discrete phases of rifting.
throughgoing structure during the period of high The synrift stratigraphy of the Viking Group, in par-
displacement rate (e.g., Prosser 1993). Rather than ticular the thickness variation in the Heather
rapid deepening and shale deposition occurring Formation and lower Draupne Formation shale
mainly after segment linkage (due to the newly sequences, suggests that the subbasins controlled
linked fault being “underdisplaced” and inhibiting by these splays young progressively toward the
propagation until a critical displacement-length ratio north. This conclusion is consistent with the seg-
is reached (see Schlische and Anders, 1996)), our ments becoming progressively more active and
observations suggest that rift-climax shale deposi- eventually linking in a south-to-north manner. We
tion is closely associated with the interaction of fault attribute this pattern of fault activity and depocen-
segments prior to the formation of a fully linked fault ter development to the interaction of normal fault
array in the latest Jurassic–earliest Cretaceous segments.
(Figure 13D). The results of this study have several implica-
An implication of this study is that diachroneity in tions for stratigraphic studies and hydrocarbon
rifting might be best explained through the process- exploration in rift basins. Because the spatial
es of fault propagation, interaction, and linkage. Our extent of the fault interaction is determined by the
results suggest that individual fault segments along scale of the fault segments, synrift sequences will
an array may show variable activity over a time scale vary spatially along fault systems. For example,
on the order of 5 m.y. For example, along the south- high displacement rates near segment centers
ern part of the Statfjord East fault, most of the activi- may promote rift-climax stratal patterns and facies
ty occurred in the early part of the rifting, whereas associations, whereas shallow-marine conditions
at its northern tip it appears that fault activity was may persist at fault tips and in overlap zones
confined mostly to the latest Jurassic–earliest between unlinked faults. Thus an improved under-
Cretaceous (see Figures 8, 10, 11). The cessation of standing of patterns of fault activity may aid in
62 Faults and Synrift Sequences

reconstructing synrift paleogeographies and may Ebinger, C., 1989, Geometric and kinematic development of
help to explain discrepancies between regional and border faults and accommodation zones, Kivu-Rusivi rift,
Africa: Tectonics, v. 8, p. 117–133.
local sea level changes. Færseth, R. B., 1996, Interaction of Permo–Triassic and Jurassic
As a result of the temporal changes associated extensional fault-blocks during the development of the
with fault interaction, the apparent duration of the northern North Sea: Journal of the Geological Society, London,
rift episode may vary if observations are made v. 153, p. 931–944.
Gawthorpe, R. L., A. J. Fraser, and R. E. Ll. Collier, 1994,
using a spatially limited data set. The duration of Sequence stratigraphy in active extensional basins:
the rifting thus may be significantly underrepre- implications for the interpretation of ancient basin fills: Marine
sented in the stratigraphic record. Moreover, and Petroleum Geology, v. 11, p. 642–658.
because stratal patterns are strongly influenced by Gawthorpe, R. L., I. Sharp, J. R. Underhill, and S. Gupta, 1997,
local fault interactions, variation in synrift stratal Linked sequence stratigraphic and structural evolution of
propagating normal faults: Geology, v. 25, p. 795–798.
geometry alone is not sufficient evidence of tempo- Glennie, K. W., and J. R. Underhill, 1998, Origin, development
ral variation in the regional strain rate or of multi- and evolution of structural styles, in K. W. Glennie, ed.,
ple rifting events. Petroleum geology of the North Sea: London, Blackwell
Science, p. 42–84.
Graue, E., W. Helland-Hansen, J. R. Johnsen, L. Lømo, A. Nøttvedt,
K. Rønning, A. Ryseth, and R. J. Steel, 1987, Advance and
REFERENCES CITED retreat of the Brent delta system, Norwegian North Sea, in
J. Brooks and K. Glennie, eds., Petroleum geology of
Anders, M. H., and R. W. Schlische, 1994, Overlapping faults, northwest Europe: London, Graham and Trotman, p. 915–937.
intrabasin highs, and the growth of normal faults: Journal of Gupta, S., P. A. Cowie, N. H. Dawers, and J. R. Underhill, 1998,
Geology, v. 102, p. 165–179. A mechanism to explain rift basin subsidence and stratigraphic
Bürgmann, R., D. D. Pollard, and S. J. Martel, 1994, Slip patterns through fault array evolution: Geology, v. 26,
distributions on faults: effects of stress gradients, inelastic p. 595–598.
deformation, heterogeneous host-rock stiffness, and fault Gupta, S., J. R. Underhill, R. L. Gawthorpe, and I. Sharp, 1999,
interaction: Journal of Structural Geology, v. 16, p. 1675–1690. Role of fault interaction in controlling syn-rift sediment
Cartwright, J. A., B. D. Trudgill, and C. S. Mansfield, 1995, Fault dispersal patterns: Miocene, Abu Alaqa Group, Suez rift, Sinai,
growth by segment linkage: an explanation for scatter in Egypt: Basin Research, v. 11, p. 167–189.
maximum displacement and trace length data from the Häger, K-O., and M. Smelror, 1997, Reworking of dinocysts as a
Canyonlands Grabens of SE Utah: Journal of Structural means of obtaining useful geological information: a study of
Geology, v. 17, p. 1319–1326. the Upper Jurassic Draupne Formation in the Tampen area,
Cartwright, J. A., C. Mansfield, and B. Trudgill, 1996, The growth Norwegian North Sea (abs.): American Association of
of faults by segment linkage, in P. G. Buchanan and D. A. Stratigraphic Palynologists, Annual Meeting Abstracts, p. 20.
Nieuwland, eds., Modern developments in structural Hallam, A., 1969, Tectonism and eustasy in the Jurassic: Earth
interpretation, validation and modelling: London, Geological Science Reviews, v. 5, p. 45–68.
Society Special Publication 99, p. 163–177. Helland-Hansen, W., M. Ashton, L. Lømo, and R. J. Steel, 1992,
Childs, C., J. Watterson, and J. J. Walsh, 1995, Fault overlap zones Advance and retreat of the Brent delta: recent contributions to
within developing normal fault systems: Journal of the the depositional model, in A. C. Morton, R. S. Haszeldine,
Geological Society, London, v. 152, p. 535–549. M. R. Giles, and S. Brown, eds., Geology of the Brent Group:
Contreras, J., 1999, Tectonic and stratigraphic modeling of the London, Geological Society Special Publication 61,
evolution of continental rift basins: Ph.D. thesis, Columbia p. 109–127.
University, New York, 109 p. Lee, M. J., and Y. J. Hwang, 1993, Tectonic evolution and
Cowie, P. A., 1998, A healing-reloading feedback control on the structural styles of the East Shetland Basin, in J. R. Parker, ed.,
growth rate of seismogenic faults: Journal of Structural Petroleum geology of northwest Europe: Proceedings of the
Geology, v. 20, p. 1075–1087. 4th Conference, London, The Geological Society,
Crider, J. G., and D. D. Pollard, 1998, Fault linkage: three- p. 1137–1149.
dimensional mechanical interaction between echelon Leeder, M. R., and R. L. Gawthorpe, 1987, Sedimentary models for
normal faults: Journal of Geophysical Research, v. 103, extensional tilt-block/half-graben basins, in M. P. Coward, J. F.
p. 24,373–24,391. Dewey, and P. L. Hancock, eds., Continental extensional
Dahl, N., and T. Solli, 1993, The structural evolution of the Snorre tectonics: London, Geological Society Special Publication 28,
field and surrounding areas, in J. R. Parker, ed., Petroleum p. 139–152.
geology of northwest Europe: Proceedings from the 4th Machette, M. N., S. F. Personius, A. R. Nelson, D. P. Schwartz, and
Conference, London, The Geological Society, p. 1159–1166. W. R. Lund, 1991, The Wasatch fault zone, Utah—
Davies, S. J., J. R. Underhill, N. H. Dawers, and A. E. McLeod, segmentation and history of Holocene earthquakes: Journal of
1998, The evolution of early syn-rift depositional systems Structural Geology, v. 13, p. 137–149.
(abs.): AAPG Hedberg Research Conference Abstracts, p. 33. Nøttvedt, A., R. H. Gabrielsen, and R. J. Steel, 1995,
Dawers, N. H., and M. H. Anders, 1995, Displacement-length Tectonostratigraphy and sedimentary architecture of rift
scaling and fault linkage: Journal of Structural Geology, v. 17, basins, with reference to the northern North Sea: Marine and
p. 607–614. Petroleum Geology, v. 12, p. 889–901.
Dawers, N. H., A. M. Berge, K.-O. Häger, C. Puigdefàbregas, and Nøttvedt, A., A. Berge, N. H. Dawers, R. B. Færseth, K.-O. Häger,
J. R. Underhill, 1999, Controls on Late Jurassic, subtle sand G. Mangerud, and C. Puigdefàbregas, in press, Evaluation of
distribution in the Tampen Spur area, northern North Sea, in the syn-rift play potential of the Snorre fault block, in
A. J. Fleet and S. A. R. Boldy, eds., Petroleum geology of NW A. Nøttvedt, ed., Integrated Basins Studies Special Publication:
Europe: Proceedings from the 5th Conference, London, The London, The Geological Society.
Geological Society, p. 827–838. Nyberg, I. T., 1987, Statfjord Øst, in A. M. Spencer et al., eds.,
Deegan, C. E., and B. J. Scull, 1977, A standard lithostratigraphic Geology of the Norwegian oil and gas fields: London, Graham
nomenclature for the central and northern North Sea: Bulletin and Trotman, p. 351–362.
of the Norwegian Petroleum Directorate, v. 1, 36 p. Peacock, D. C. P., and D. J. Sanderson, 1991, Displacements,
Dawers and Underhill 63

segment linkage and relay ramps in normal fault zones: Journal stratigraphy of the Middle to Upper Jurassic, Viking Graben,
of Structural Geology, v. 13, p. 721–733. North Sea, in R. J. Steel, ed., Sequence stratigraphy on the
Prosser, S., 1993, Rift-related linked depositional systems and their northwest European margin: Amsterdam, Elsevier, Norwegian
seismic expression, in G. D. Williams and A. Dobbs, eds., Petroleum Society Special Publication 5, p. 167–197.
Tectonics and seismic sequence stratigraphy: London, Trudgill, B., and J. Cartwright, 1994, Relay-ramp forms and
Geological Society Special Publication 71, p. 35–66. normal-fault linkages, Canyonlands National Park, Utah:
Rattey, R. P., and A. B. Hayward, 1993, Sequence stratigraphy of a Geological Society of America Bulletin, v. 106, p. 1143–1157.
failed rift system: the Middle Jurassic to Early Cretaceous basin Underhill, J. R., 1991a, Implications of Mesozoic–Recent basin
evolution of the central and northern North Sea, in J. R. development in the western Inner Moray Firth, UK: Marine
Parker, ed., Petroleum geology of northwest Europe: and Petroleum Geology, v. 8, p. 359–369.
Proceedings of the 4th Conference, London, The Geological Underhill, J. R., 1991b, Controls on Late Jurassic seismic
Society, p. 215–249. sequences, Inner Moray Firth, UK North Sea: a critical test of a
Ravnås, R., and K. Bondevik, 1997, Architecture and controls on key segment of Exxon’s original global cycle chart: Basin
Bathonian–Kimmeridgian shallow-marine synrift wedges of the Research, v. 3, p. 79–98.
Oseberg-Brage area, northern North Sea: Basin Research, v. 9, Underhill, J. R., 1998, Jurassic, in K. W. Glennie, ed., Petroleum
p. 197–226. geology of the North Sea: London, Blackwell Science,
Ravnås, R., K. Bondevik, W. Helland-Hansen, L. Lømo, A. Ryseth, p. 245–293.
and R. J. Steel, 1997, Sedimentation history as an indicator of Underhill, J. R., and M. A. Partington, 1993, Jurassic thermal
rift initiation and development: the late Bajocian–Bathonian doming and deflation in the North Sea: implications of the
evolution of the Oseberg-Brage area, northern North Sea: sequence stratigraphic evidence, in J. R. Parker, ed., Petroleum
Norsk Geologisk Tidsskrift, v. 77, p. 205–232. geology of northwest Europe: Proceedings from the 4th
Rawson, P. F., and L. A. Riley, 1982, Latest Jurassic–Early Conference, London, The Geological Society, p. 337–345.
Cretaceous events and the “late Cimmerian unconformity” in Underhill, J. R., M. J. Sawyer, P. Hodgson, M. D. Shallcross, and
the North Sea area: AAPG Bulletin, v. 66, p. 2628–2648. R. L. Gawthorpe, 1997, Implications of fault scarp degradation
Richards, P. C., G. K. Lott, H. Johnson, R. W. O’B. Knox, and J. B. for Brent Group prospectivity, Ninian field, northern North
Riding, 1993, Jurassic of the central and northern North Sea, Sea: AAPG Bulletin, v. 81, p. 999–1022.
in R. W. O’B. Knox and G. W. Cordey, eds., Lithostratigraphic Vollset, J., and A. G. Doré, 1984, A revised Triassic and Jurassic
nomenclature of the UK North Sea: Nottingham, British lithostratigraphic nomenclature for the Norwegian North Sea:
Geological Society, p. 219. Norwegian Petroleum Directorate Bulletin no. 3, 53 p.
Roberts, A., G. Yielding, N. J. Kusznir, I. Walker, and D. Dorn- Walsh, J. J., and J. Watterson, 1991, Geometric and kinematic
Lopez, 1993a, Mesozoic extension in the North Sea: coherence and scale effects in normal fault systems, in A. M.
constraints from flexural backstripping, forward modelling Roberts, G. Yielding, and B. Freeman, eds., The geometry of
and fault populations, in J. R. Parker ed., Petroleum geology of normal faults: London, Geological Society Special Publication
northwest Europe: Proceedings of the 4th Conference, 56, p. 193–203.
London, The Geological Society, p. 1123–1136. Willemse, E. J. M., 1997, Segmented normal faults: correspondence
Roberts, A. M., G. Yielding, and M. E. Badley, 1993b, Tectonic and between three-dimensional mechanical models and field data:
bathymetric controls on stratigraphic sequences within Journal of Geophysical Research, v. 102, p. 675–692.
evolving half-graben, in G. D. Williams and A. Dobbs, eds., Willemse, E. J. M., D. D. Pollard, and A. Aydin, 1996, 3D analyses
Tectonics and seismic sequence stratigraphy: London, of slip distributions on normal fault arrays with consequen-
Geological Society Special Publication 71, p. 87–121. ces for fault scaling: Journal of Structural Geology, v. 18,
Schlische, R., 1991, Half-graben basin filling models: new p. 295–309.
constraints on continental extensional basin development: Withjack, M. O., J. Olson, and E. Peterson, 1990, Experimental models
Basin Research, v. 3, p. 123–141. of extensional forced folds: AAPG Bulletin, v. 74, p. 1038–1054.
Schlische, R. W., 1995, Geometry and origin of fault-related folds Wu, D., and R. L. Bruhn, 1994, Geometry and kinematics of
in extensional settings: AAPG Bulletin, v. 79, p. 1661–1678. active normal faults, South Oquirrh Mountains, Utah:
Schlische, R. W., and M. H. Anders, 1996, Stratigraphic effects and implications for fault growth: Journal of Structural Geology,
tectonic implications of the growth of normal faults and v. 16, p. 1061–1075.
extensional basins, in K. K. Beratan, ed., Reconstructing the Yielding, G., 1990, Footwall uplift associated with Late Jurassic
history of Basin and Range extension using sedimentology and normal faulting in the northern North Sea: Journal of the
stratigraphy: Boulder, Geological Society of America Special Geological Society, London, v. 147, p. 219–222.
Paper 303, p. 183–203. Yielding, G., M. E. Badley, and A. M. Roberts, 1992, The structural
Segall, P., and D. D. Pollard, 1980, Mechanics of discontinuous evolution of the Brent province, in R. S. Haszeldine, M. R.
faults: Journal of Geophysical Research, v. 85, p. 4337–4350. Giles, and S. Brown, eds., Geology of the Brent Group:
Sneider, J. S., P. de Clarens, and P. R. Vail, 1995, Sequence London, Geological Society Special Publication 61, p. 27–43.
64 Faults and Synrift Sequences

ABOUT THE AUTHORS

Nancye H. Dawers John R. Underhill


Nancye Dawers holds a B.S. John Underhill graduated from
degree from the University of Ken- Bristol University in 1982 and was
tucky, an M.S. degree from the awarded his Ph.D. in 1985 from
University of Illinois at Urbana- the University of Wales. He subse-
Champaign, and an M.Ph. degree quently worked for Shell before mov-
and a Ph.D. from the Lamont- ing to Edinburgh, where he has
Doherty Earth Observatory of recently been promoted to the Chair
Columbia University. Since 1996 she of Stratigraphy. John has been award-
has been a research associate at the ed the European Association of Petro-
University of Edinburgh. Nancye’s leum Geoscientists’ Distinguished
research interests include fault Lecturer award twice, the AAPG
growth, displacement and length scaling in fault popula- Matson Award once, and was one of AAPG’s 1999 Distin-
tions, neotectonics, and basin analysis. guished Lecturers.

You might also like