You are on page 1of 14

Distribution of subsurface uid-ow systems in the SW Barents Sea

Sunil Vadakkepuliyambatta
a,
*
, Stefan Bnz
a
, Jrgen Mienert
a
, Shyam Chand
b
a
Department of Geology, University of Troms, Dramsveien-201, 9037 Troms, Norway
b
Geological Survey of Norway (NGU), P.O. Box 6315 Sluppen, 7491 Trondheim, Norway
a r t i c l e i n f o
Article history:
Received 28 July 2012
Received in revised form
28 January 2013
Accepted 1 February 2013
Available online 26 February 2013
Keywords:
Fluid ow
Gas chimneys
Gas hydrates
Barents Sea
Seismic
a b s t r a c t
The SW Barents Sea is a large hydrocarbon-prone epi-continental Sea of the Norwegian Arctic region. A
signicant portion of the hydrocarbon gases generated in deep source rocks has leaked or migrated into
the shallow subsurface and is now trapped in gas hydrate and shallow gas reservoirs. The evolution of
sedimentary basins of this region has controlled the leakage of these uids through marine sediments.
Understanding the distribution of various uid-ow systems may enhance our knowledge of the evo-
lution of different basins in the SW Barents Sea and could help nd potential targets for future hydro-
carbon exploration. We analyze approximately 3000 2D multi-channel seismic proles and data from 60
wells covering the entire SW Barents Sea, to identify and classify uid-ow features, and study their
relationship to tectonic elements and geological history. Gas chimneys are the most abundant feature
among various other uid-ow features such as uid leakage along faults and fractures, seepage pipes,
and high amplitude anomalies potentially indicating trapped uids. Large uid-ow features, covering
areas as large as 600 km
2
, occur close to known hydrocarbon elds such as Snhvit, Skrugard, and Havis.
The uid-ow features occur above major deep-seated faults in the area suggesting a close relation to it.
The number of uid-ow features in the western part of the study area is signicantly higher than in the
eastern part. The amount of net erosion in the study area shows no direct control over the distribution of
uid-ow features, suggesting that the faults and distribution of mature source rocks control the uid
ow. The strong correlation between the locations of uid-ow features and structural elements in-
dicates that extensional tectonics, uplift and glaciations could have played major roles in the timing and
activity of the uid leakage, although erosion might have had an added effect.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
The vertical ow of uids through marine sediments is a
widespread and dynamic geological process that occurs on passive
and active continental margins worldwide. Fluid migration is
associated with excess pore-uid pressure, and is attributed to
temporally and spatially varying processes, such as rapid sediment
loading (e.g. Dugan and Flemings, 2000), uplift and erosion (e.g.
Dor and Jensen, 1996), dissociation of gas hydrate (e.g. Mienert
et al., 2005), polygonal faulting (e.g. Cartwright et al., 2007), and
hydrocarbon generation and leakage fromdeep and shallowsource
rocks and reservoirs (e.g. Heggland, 1998; Solheim and Elverhi,
1985; Hovland and Judd, 1988). Presence of shallow gas accumu-
lations associated with uid leakage are of interest for several
reasons: (1) Shallow gas accumulations may reduce the shear
strength of sediments, and pose a hazard to hydrocarbon explora-
tion and development (Andreassen et al., 2007a), (2) the occur-
rence of shallowgas and indications of uid owunderlying it may
point toward deeper prospective reservoirs (e.g. Heggland, 1998)
and (3) shallow gas accumulations could be of commercial interest
in the future (Carstens, 2005).
Vertical migration of gas through subsurface strata can cause
widely distributed acoustic low-velocity zones. These low-velocity
zones can deteriorate the seismic signal and create regions of
chaotic seismic signals. The nature and shape of this chaotic region
of acoustic signals can vary depending on the process of formation
of these zones. Chaotic regions in seismic data can also result from
mud mobilization triggered by vertical migration of uids (Lseth
et al., 2003). Mud mobilization can modify the structure of sedi-
ments to a disrupted succession and form a low-density sediment-
uid mixture. Vertical uid-ow features and chaotic seismic
reection zones are commonly observed with most types of sedi-
ment mobilizations (Graue, 2000; Hurst et al., 2003; Lseth et al.,
2003; Jackson and Stoddart, 2005). Fluid ow can also alter the
* Corresponding author. Tel.: 47 77623290.
E-mail address: sunil.vadakkepuliyambatta@uit.no (S. Vadakkepuliyambatta).
Contents lists available at SciVerse ScienceDirect
Marine and Petroleum Geology
j ournal homepage: www. el sevi er. com/ l ocat e/ marpet geo
0264-8172/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.marpetgeo.2013.02.007
Marine and Petroleum Geology 43 (2013) 208e221
formation, cause local sediment remobilization, and appear as
chaotic reections in the seismic prole (Ligtenberg, 2007). Ge-
ometry of subsurface uid-ow systems is hard to constrain by
direct observations (Talukder, 2012) and characterization of themis
difcult because ow can be highly transient and can vary in time
and space along complex and changing conduit systems (Hornbach
et al., 2007).
The SW Barents Sea is a part of the Arctic Ocean located north of
Norway (Fig. 1). Occurrence of shallow gas, gas hydrates and sea-
oor expulsion features has been reported fromseveral areas of the
SW Barents Sea (e.g. Andreassen et al., 1990; Perez-Garcia et al.,
2009; Chand et al., 2009, 2012; Ostanin et al., 2012). Migration of
uids into shallow sediments and seepage into the ocean through
the seaoor was probably the result of spillage and migration of
hydrocarbons in response to uplift and erosion processes in the
Cenozoic (e.g. Dor, 1995; Dor and Jensen, 1996; Henriksen et al.,
2011).
Uplift and erosion is known to have affected the SW Barents
Sea during Cenozoic times (Faleide et al., 1996; Dimakis et al., 1998;
Anell et al., 2009). This process is thought to have a very strong
impact on petroleum systems (e.g. Dor and Jensen, 1996; Bjrkum
et al., 2001; Cavanagh et al., 2006; Ohmet al., 2008; Henriksen et al.,
2011). The negative effects of uplift and associated erosion on hy-
drocarbon systems include tilting and opening of hydrocarbon-lled
traps resulting in spillage of oil and gas (Kjemperud and Fjeldskaar,
1992), gas expansion and ex-solution from oil (Skagen, 1993; Dor
and Jensen, 1996; Cavanagh et al., 2006), seal failure (e.g., Sales,
1993), reduction in temperature due to uplift resulting in imma-
ture source rocks (Dor et al., 2000; Ohm et al., 2008), and porosity
and permeability reduction due to diagenetic processes (Berglund
et al., 1986; Walderhaug, 1992).
Whereas, positive effects include the occurrence of thermally
matured source rocks at shallow levels (Bjrkum et al., 2001),
liberation of dissolved methane from formation water (Dor and
Jensen, 1996) due to decreases in pressure (e.g. Maximov et al.,
1984; Nesterov et al., 1990), ex-solution of light oil or condensate
fromgas (e.g. Piggott and Lines, 1991), fracture enhancement of less
permeable reservoirs (e.g. Aguilera, 1995) and remigration to
shallower structures (Waylett and Embry, 1992). In addition, local
re-deposition under a heavy overburden associated with erosion
can result in rapid maturation of source rocks and generation of
hydrocarbons (e.g. Dahl and Augustson, 1993).
Numerous glaciations also affected the SW Barents Sea region
during the late Cenozoic. Rapid buildup and removal of ice load, as
occurred in the SW Barents Sea, may have less impact on the
evolution of the basin (Lerche et al., 1997). However, distortions to
the thermal regime of sub-ice sediments caused by spatial and
temporal variations of ice thickness inuence the generation,
migration and present-day accumulation of hydrocarbons (Lerche
et al., 1997; Cavanagh et al., 2006). Ice loading can cause struc-
tural distortions leading to a redistribution of gas and oil in the
reservoir and spill of hydrocarbons (Lerche et al., 1997). Major ice-
loading effects are present in the Hammerfest Basin; where many
drilled wells found only a small amount of residual oil in rotated
and tilted structures (Rasmussen et al., 1993; Nyland et al., 1992).
Often, large gas anomalies overlie hydrocarbon discoveries in
SW Barents Sea indicating leakage of gas from the deeper forma-
tions (Chand et al., 2008, 2009, 2012; Heggland, 1998, 2004;
Meldahl et al., 2001). A recent study from the Loppa High (Chand
et al., 2012) reported seepage of gas into the water column indi-
cating that the gas migration is still active through open faults. The
locations of these anomalies may indicate possible targets for
Figure 1. General bathymetric map of SW Barents Sea with major basins and hydrocarbon discoveries. The location of the study area (red star) and seismic lines in the following
gures are also shown.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 209
future hydrocarbon exploration. Therefore, it is important to un-
derstand the distribution of uid-ow anomalies and their gov-
erning controls. Here, we analyze 2D seismic data covering the
whole SW Barents Sea (Fig. 2) in order to delineate and classify
various uid-owfeatures, to characterize their distribution, and to
better understand their relation to the structural setting, glacia-
tions, uplift, and erosion during the Cenozoic.
2. Geological setting
2.1. Geological history of the area
The Barents Sea is a large epi-continental sea, bound to the west
and north by young passive margins that developed during the
Cenozoic opening of the NorwegianeGreenland Sea and Eurasia
basin, respectively (Faleide et al., 1993). The SW Barents Sea con-
tains some of the deepest sedimentary basins worldwide with
sediment cover exceeding 10 km at places (e.g. Nordkapp Basin,
Srvestsnaget Basin) (Smelror et al., 2009). These basins formed in
response to several phases of regional tectonism within the North
AtlanticeArctic region, culminating with continental separation of
Eurasia and Greenland, and accretion of oceanic crust, in the early
Tertiary (Faleide et al., 1993).
The western part of the Barents Sea shows much more complex
tectonic development than the eastern part, with a mosaic of ba-
sins, platforms, and structural highs (Faleide et al., 1984; Gabrielsen
et al., 1990; Gudlaugsson et al., 1998). The SW Barents shelf formed
a central part of the northern Pangean margin from the late
Devonian (Worsley, 2008) and is underlain by early Devonian
metamorphic basement formed during the Caledonian Orogeny
(Smelror et al., 2009). Extensional tectonic movements during
earlyemiddle Devonian, Carboniferous, Permian, Triassic and late
Jurassiceearly Cretaceous (Johansen et al., 1993) dominate the late
Paleozoic and Mesozoic tectonic history of the SW Barents Sea.
Extensional faulting also affected the region during late Cretaceous
and Paleogene (Faleide et al., 1993). In late Devonian to early
Carboniferous times, most of the Barents Sea was affected by crustal
extension (Gjelberg, 1981, 1987; Faleide et al., 1993). These rifting
events caused the formation of Hammerfest Basin as a NE-trending
half graben (Cavanagh et al., 2006). Evaporites were deposited in
the Nordkapp Basin and possibly also in the Troms Basin (Worsley,
2008). Extensional faulting affected the Loppa and the Stappen
Highs during late Carboniferous to early Permian period (Brekke
and Riis, 1987), while northeastern part of Bjarmeland Platform
and Nordkapp Basin were stable (Riis et al., 1986).
The western part of the Barents Sea has been the most tecton-
ically active sector throughout Mesozoic and Cenozoic times;
although Triassic to early Jurassic was relatively quiet tectonically
(Gabrielsen et al., 1990). However, reactivated rifting in the
Permian-early Triassic led to tilting in the Loppa High and Stappen
high region (Fig. 1) (Gudlaugsson et al., 1998). The early Triassic was
characterized by regional subsidence in eastern areas and sediment
inux into the regional sag basin from the East (Gudlaugsson et al.,
1998; Worsley, 2008). Salt tectonics inuenced the depositional
patterns in the Nordkapp and Maud Basins (Fig. 1) (Gabrielsen et al.,
1990).
Rifting and associated block (extensional) faulting started again
in the Mid-Jurassic, related to the opening of Central Atlantic
(Ziegler, 1988), increased during the period from late Jurassic into
early Cretaceous, with the opening of southern North Atlantic
(Faleide et al., 1993), and terminating with the formation of the
now well-known major basins and highs (Fig. 1). Structural
Figure 2. Coverage of 2D seismic data (thin black lines) used for the study with major basins and boundaries (thick black lines) in the area.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 210
development in this period was complicated. On one hand,
extreme rates of subsidence are seen in the Troms Basin and the
western part of the Bjrnya Basin during the early Cretaceous
(Aptian to Albian). On the other hand, indications of local early
Cretaceous inversion are found, e.g. along the Ringvassy-Loppa
Fault Complex (RLFC) and its junction with the Asterias Fault
Complex (Fig. 1) (Gabrielsen et al., 1990). During the early Creta-
ceous, Harstad, Troms, and Bjrnya basins underwent large-
scale subsidence and become major depocenters (Faleide
et al., 1993). Uplift and erosion in the northeastern Barents Sea
brought more sediments into these depocenters (Smelror et al.,
2009). Throughout the Cretaceous, rifting prevailed and caused
formation of pull-apart basins such as Srvestsnaget Basin and
Vestbakken Volcanic Province (Fig. 1) (Smelror et al., 2009). In the
western part of the area, especially in the Vestbakken Volcanic
Province, there was abundant magmatic activity, probably in the
Paleocene and Eocene (Jebsen and Faleide, 1998; Faleide et al.,
2008).
The opening of the North Atlantic was associated with uplift
during the Paleocene-Eocene (Eldholm et al., 1987; Faleide et al.,
1996). Inversion and folding reached a maximum in Eocene to
Oligocene times (Talwani and Eldholm, 1977; Myhre et al., 1982;
Eldholm et al., 1987).
The Barents Sea was affected by extensive erosion, related to
deglaciation and uplift in the late PlioceneePleistocene (Vorren
et al., 1991; Nyland et al., 1992). During late PlioceneePleistocene,
the entire Barents Shelf was uplifted and eroded, and large
amounts of sediment were deposited along the western margin
(Nyland et al., 1992). Particularly large accumulations are found in
the Bjrnya trough mouth fan (Fig. 1), with up to 6 km-thick
packages of glacigenic sediments, a main depocenter during the
late Cenozoic (Hjelstuen et al., 2004; Vorren et al., 1991). In the
southern part of the SW Barents Sea, the Hammerfest Basin, the
Nordkapp Basin and the Loppa High suffered lesser amounts of
uplift and erosion (Smelror et al., 2009). The Upper Regional Un-
conformity (URU), formed during the late PlioceneePleistocene at
about 2.5 Ma, separates the glacigenic sediments of the Barents Sea
from the deeper pre-glacial sedimentary rocks (Tertiary and older)
and is a major seismic reector in the study area (Solheim and
Kristoffersen, 1984; Eidvin et al., 1993). Criss-crossing glacial line-
ations and iceberg plough marks are a major feature of the Barents
Sea seabed and paleo seabeds affected by glacial erosion (Elverhi
and Solheim, 1983; Andreassen et al., 2007b).
2.2. Major source and reservoir rocks
Numerous source-rock formations are present in the SW
Barents Sea (Fig. 3) (Dor, 1995). The amount of total organic carbon
(TOC), hydrocarbon generative potential, and hydrogen index of
various source rocks in the Norwegian Barents Sea are discussed in
detail by Ohm et al. (2008). The most widely distributed and most
prolic source rock is the Hekkingen Formation of late Jurassic age,
which consists of dark, organic rich shales (Dalland et al., 1988).
Despite being widespread over most of the southern Barents Sea,
these shales have not realized their full hydrocarbon-generation
potential because of varying depth of burial and thereby maturity.
The unit is thought to be mature for oil and gas generation in a
narrow belt at the western margin of the Hammerfest Basin and
along the western fringe of the Loppa High (Dor, 1995). Hekkingen
shale also has the highest TOC, highest hydrocarbon generative
potential, and hydrogen index among the source rocks in the
Norwegian Barents Sea (Ohm et al., 2008).
Triassic shales such as, Snadd Formation and Havert Formation,
are also shown to have potential for generating hydrocarbons
(Bjory et al., 2009). The cumulative generation potential from the
thick Triassic (and possibly Lower Jurassic) sedimentary pile is
large, and it is widely assumed that the major gas discoveries of the
South Barents Basin emanate from this source (Johansen et al.,
1993). Permian and Carboniferous shales are the major source of
hydrocarbons in the Finnmark Platform (Ohm et al., 2008).
Figure 3 provides a summary by geological age of all major
source rocks and reservoirs, proven and postulated, in the Nor-
wegian Barents Sea. The most signicant proportion of the hydro-
carbon resources proven to date in both the Norwegian and Russian
Barents Sea are trapped within strata of Jurassic age (Dor, 1995).
The major discoveries in the SW Barents Sea have principal reser-
voirs consisting of loweremiddle Jurassic sandstone called the St
Formation (Dalland et al., 1988). Larsen et al. (1993) estimate that
about 85% of the Norwegian Barents Sea resources lay within this
formation. Minor Triassic gas accumulations have been found in the
Hammerfest Basin and on the margins of the Nordkapp Basin.
3. Data used
The uid-owfeatures were mapped using approximately 3000
2D multi-channel seismic proles covering the SW Barents Sea
(Fig. 2). In addition to the seismic data, well logs with formation
tops from 60 wells in the SW Barents Sea aided the interpretation.
The 2D lines were from different surveys with different horizontal
and vertical resolutions.
4. Observation and interpretation of uid-ow anomalies
Multi-channel seismic data from different basins of the SW
Barents Sea are used for mapping various types of uid-leakage
anomalies. Mapping of the source of uid-ow anomalies was
often difcult because of varying survey specications in resolution
and seismic-signal quality. The uid-ow features interpreted on
seismic data were subdivided into three main categories; 1) uid/
gas chimneys; 2) leakage along faults, and 3) all other features, not
included in the rst two categories.
4.1. Fluid/gas chimneys
Many of the seismic proles from the study area show zones of
chaotic acoustic signals and/or acoustic masking (Fig. 4). The size
and shape of these acoustically anomalous zones varies widely.
In most parts, these anomalous zones terminated at different
shallow stratigraphic levels showing high-amplitude anomalies,
suggesting the presence of gas. Such features are interpreted as
gas chimneys, which can be dened as a region of distorted
seismic signals resulting from irregularly distributed low-velocity
gas-charged zones, formed due to an upward migration of gas/
uids (Meldahl et al., 2001; Lseth et al., 2002; Heggland, 2004;
Judd and Hovland, 2007; Arntsen et al., 2007; Connolly et al.,
2008). Figure 4(a) shows gas chimneys located in the Bjrnya
Basin west of the Loppa High. They show well-dened zones of
chaotic seismic signals and terminate at the Top Cretaceous level.
Patchy high-amplitude reections were observed above both
chimneys. The small depressions on the seaoor were inconclu-
sive for pockmarks, since the area was glaciated in the past and
thus the depressions may be glacial plough marks. The gas
chimneys are located in the Bjrnyrenna Fault Complex (BFC)
area (Fig. 1), and the small gas chimney is located right above a
fault (Fig. 4a). We hypothesize that the gas migrates upward
through this fault. Seismic data from the Polheim Sub-platform
west of the Loppa High shows an exceptionally large gas chimney
(Fig. 4b). A very large zone, approximately 12 km wide, of acoustic
masking characterizes it. Shallow high-amplitude reections mark
the upper termination of the chimney. High-amplitude anomalies
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 211
are patchy, polarity-reversed and crosscut some parts of the
lithology. The seismic signal is highly chaotic inside the chimney,
making it difcult to identify the source of gas leakage. However,
the area is part of the BFC and the uid migration could be along
these faults.
Vertical and lateral migration of uids is visible in Srvestsnaget
Basin close to Veslemy High (Fig. 4c). High-amplitude anomalies
located at the Plio-Pleistocene boundary (URU) mark the upper
termination of a columnar zone of uid migration. The high-
amplitude reverse polarity reections at this stratigraphic level
suggest accumulation of gas. This feature is interpreted as a gas
chimney. High-amplitude reections are also visible on other
stratigraphic boundaries adjacent to the gas chimney and away
from the termination point of the chimney, suggesting lateral
migration in the up-dip direction along the strata.
The seismic line located on the southeast part of Loppa High,
close to the junction of Asterias FC and RLFC (Fig. 1), shows another
example of a gas chimney (Fig. 4d). The region of uid migration is
marked by a zone, approximately 1-km wide, of seismic reectors
that are chaotic and deteriorated in amplitude. Faults may be
Figure 3. Some major proven and potential reservoir and source rocks in the Barents Sea (modied after Dor, 1995; Ohm et al., 2008).
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 212
present inside the chimney, but are difcult to identify in the
seismic prole due to loss of seismic energy inside the chimney
zone. A very strong, polarity reversed reector at the terminating
point of the chimney just beneath the URU reector suggests
accumulation of gas. The Snadd Formation, one of the major source
rocks in the area (Fig. 3, Section 2.2) and other deeper source rocks
could be the source of uids here; however, it is very difcult to
identify it accurately from the seismic data.
Figure 4. a) Showuid ow in the southeast part of Bjrnya Basin. The leakage zone of large uid ow feature is w8 km at its widest region. Patchy high-amplitude anomalies are
present on top of the leakage zone. The leakage zone covers an area of w36 km
2
. The small chimney sits on top of a fault and shows bright spots on top of the termination. b) A
12 km wide uid migration feature SWof the Loppa High. It covers an area of w130 km
2
. The shallow high amplitude reections show reverse polarity with respect to the seaoor
indicating presence of gas. They also cross cuts lithological reectors suggesting formation of gas hydrates. c) Seismic prole from Srvestsnaget Basin close to Veslemy High shows
chaotic acoustic signals, which terminates in Late Tertiary sediments leaving bright spots. This feature is interpreted as a gas chimney, and the high-amplitude anomalies indicate
the presence of gas. d) Shows a gas chimney from the SW part of Loppa High. Highly chaotic seismic reectors dene the zone of uid ow, and a very bright reector is visible on
the termination point of the chimney, suggesting accumulation of gas.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 213
4.2. Leakage along faults
2D seismic data show chaotic reections and high-amplitude
anomalies associated with faults in many parts of the SW Barents
Sea (Fig. 5). A seismic prole from the northern part of the Loppa
High shows two major faults, cutting through the Hekkingen For-
mation and extending up to the URU, associated with several minor
faults (Fig. 5a). Highly chaotic and low-amplitude reections occur
close to the root of the major faults, suggesting uid migration from
a much deeper source and branching of uid migration along the
Figure 5. a) Seismic prole from northern part of the Loppa High shows leakage along faults in the Hekkingen formation. The two major faults act as main pathways for leakage.
Accompanying minor faults also shows signs of uid ow with high amplitude anomalies on their anks. The uid ow feature is narrowat the base (w2 km) and spreads as it goes
up covering an area of w290 km
2
. b) Shows uid leakage along faults from the southern part of the Loppa High. The source of the leakage could be early Triassic Havert Formation or
the underlying Paleozoic strata. c) Fluid leakage along faults from the northwest part of the Hammerfest Basin. Two faults, one of them cutting through the Jurassic Hekkingen
formation, showhigh-amplitude anomalies along them and at their termination in the shallowstrata, indicating gas migration and accumulation. Hekkingen formation could be the
source of the upward migrating gas. d) Seismic prole across the Samson Dome, NE of the Hammerfest Basin, show w7 km wide region of uid migration along faults. The leakage
zone is on the southern part of the dome, covering an area of w141 km
2
.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 214
faults. High-amplitude anomalies are visible along the fault planes
close to the termination of the faults. Occurrences of high-
amplitude anomalies in seismic data along the faults indicate that
gas is present (Lseth et al., 2009; Cartwright et al., 2007;
Ligtenberg, 2005). Minor faults also show high-amplitude re-
ections along fault planes. No high-amplitude reections are
observed above the URU, suggesting trapping of upward migrating
gas at or beneath URU. Hekkingen formation could be the source of
shallow gas in this area, although deeper formations could also
contribute to the gas leakage (See Section 2.2).
Figure 5(b) shows a heavily faulted region from the southern
part of the Loppa High. Noisy seismic reections are visible close to
and along the faults. It also shows high-amplitude reections on
the fault planes close to the termination point of faults, suggesting
presence of gas. A few faults extend almost all the way to the
seaoor and their termination coincides with the occurrence of
small depressions on the seaoor. Based on the 2D seismic data we
cannot resolve whether these depressions are pockmarks or not.
The uid migration is along two major faults in the deeper part, and
branches out through many smaller faults in the shallowstrata. Top
Havert formation of early Triassic age could be the major source of
hydrocarbons. Deeper formations could also be contributing to the
leakage (See Section 2.2). Seismic prole fromnorthwestern part of
the Hammerfest Basin shows a perfect example of uid leakage
along a fault (Fig. 5c). High-amplitude reections occur along the
stratigraphic horizons over close distances away from the faults.
The uid leakage is mainly along the bigger fault until Top Creta-
ceous level and then branches along another fault. High-amplitude
reections are visible close to the terminating point of the faults,
suggesting shallow gas accumulation. The main fault cuts across
the Top of Hekkingen formation, which could be the source of
leaking gas (Dor, 1995). The Samson Dome area in the Bjarmeland
Platform, northeast of the Hammerfest Basin (Fig. 1), shows uid
leakage and shallow gas accumulation over an anticlinal structure
(Fig. 5d). Noisy zones of seismic signals and high-amplitude
anomalies close to the seaoor suggest upward migration of
uids. The top of the anticlinal structure marked by the late Jurassic
Hekkingen Formation and formations below show many small
faults, formed probably due to extensional forces. These faults may
be acting as migration pathway for hydrocarbons. Much larger
faults are also present on the anks of the anticlinal structure and
may also play a role in uid migration. The Hekkingen formation is
a unit with hydrocarbon potential, but is not mature in this area
(Dor, 1995). Much deeper Triassic formations, such as Havert for-
mation, could be the source of upward migrating gas (Fig. 3).
4.3. Other features
Many seismic expressions associated with uid ow other than
gas chimneys and leakages along faults are present in the SW
Barents Sea. These features include seepage pipes, pockmark-like
depressions, acoustic pull downs related to uid migration, and
bottom-simulating reectors (BSR). They are rare and widely
distributed in the study area. Seepage pipes are columnar zones of
disrupted reections with localized amplitude anomalies
(Cartwright et al., 2007). Pockmarks are depressions on the seabed
resulting from the seepage of gas and pore uids in soft sediments
(Judd and Hovland, 2007). Pockmarks of different sizes and den-
sities are common in some parts of the Barents Sea (Solheim and
Elverhi, 1985; Hovland and Judd, 1988). The central Barents Sea
shows seabed depressions up to 1000 m in diameter (Solheim and
Elverhi, 1993; Lammers et al., 1995). BSRs are high-amplitude
reverse-polarity reections resulting from abrupt change in
acoustic impedance at the boundary between a hydrate-bearing
layer and underlying gassy sediments (Sloan and Koh, 2007; Judd
and Hovland, 2007). Figure 6(a) shows a vertical, narrow zone of
acoustic disturbance from the southern part of the Bjarmeland
Platform north of Norsel High. The zone of chaotic seismic signals
terminates in shallowstrata. However, faults (one of which extends
to the seaoor) are visible above the termination point and these
faults may represent pathways for uid migration upwards through
the strata. High-amplitude reections are visible above the termi-
nation of the vertical zone of uid leakage. The zone is approxi-
mately 800-m wide and such regions of narrow, vertical uid
migration, and is interpreted as a seepage pipe. The source of uid
leakage could be Triassic or older formations, since the leakage
zone go deeper than the Havert formation representing early
Triassic. A seismic prole from central part of the Finnmark Plat-
form shows chaotic zones of seismic reectors and high-amplitude
anomalies indicating migration of uids (Fig. 6b). The reectors
show acoustic pull down suggesting presence of low-velocity ma-
terial, possibly gas. The source of uid owcould be early Permian/
late Carboniferous rn formation or much deeper Carboniferous
shales (Fig. 3, Ohm et al., 2008). Also visible is a crater-like
depression, 1.2 km wide, which can be interpreted as a paleo
channel.
The Bjrnyrenna Fault Complex, west of the Loppa High,
shows a gas chimney, 6 km-wide, with high-amplitude reections
at its shallow terminating point (Fig. 6c). The chaotic region in the
seismic data suggests upward migration of uids. High-amplitude
reections observed above the chimney are polarity reversed with
respect to the seaoor, suggesting accumulation of gas beneath it.
These enhanced reections are also discontinuous and crosscut
stratigraphic boundaries near it. This crosscutting event can be
interpreted as a gas-hydrate-related BSR (Holbrook et al., 1996;
Kvenvolden, 1998). Higher-order hydrocarbon gases could be
migrating along with other uids through the chimney and can
form gas hydrates in this part of the Barents Sea (Chand et al.,
2008). BSRs have been observed in seismic data earlier in this
area and are suggested to have gas compositionwith fewpercents of
ethane and propane along with methane (Laberg and Andreassen,
1996).
Although no high-resolution 3D-seismic multibeam data were
used for this study, we were able to identify possible pockmark-like
depressions in some areas. It should be kept in mind that the SW
Barents Sea was heavily glaciated in the Cenozoic and glacial
plough marks are very common in the study area. This makes
identication of surface uid-expulsion features difcult with 2D
data. However, seismic proles from the Bjrnya fan showed
pockmark-like depressions on the seaoor, approx. 600 m wide
(Fig. 6d). These depressions showed vertically stacked high-
amplitude reections underneath them. Reectors underneath
the channel showed acoustic pull down. The high-amplitude re-
ections suggest the presence of gas and uid ow. The gas
escaping to the seaoor may be attributed to the formation of these
depressions. It is also possible that these are erosional channels as
their axis is along the slope of Bjrnya fan.
5. Distribution of uid-ow features
Fluid-ow features, shallow gas accumulations, and associated
gas hydrates are characterized from different parts of the SW
Barents Sea (e.g. Andreassen et al., 1990; Laberg and Andreassen,
1996). Fluid-ow features, especially large gas chimneys, are hy-
drocarbon migration pathways and the location of these features is
important for the hydrocarbon industry (Heggland, 1998) since
these chimneys could pose a drilling hazard due to high pore-uid
pressures that may be present (Lseth et al., 2002). Our study
shows that the uid ow is abundant and widespread in the SW
Barents Sea (Fig. 7).
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 215
Gas chimneys are the most common uid-ow feature in the
study area and appear in most parts with various sizes and shapes.
The Ringvassy-Loppa and Bjrnyrenna Fault Complexes, northern
part of the Troms Basin, the Polheim Sub-platform and the
Veslemy High show a distinctly higher density of gas chimneys
compared to other regions in the SW Barents Sea. Fewer gas
chimneys are present in the Troms Basin and the easternpart of the
study area, especially the Finnmark Platform, Nordkapp Basin and
the Bjarmeland Platform. The Loppa High and the heavily faulted
areas around it also show a high concentration of gas chimneys. In
Figure 6. a) A vertical, narrow zone of uid leakage from the southern part of the Bjarmeland Platform north of the Nyslepp Fault Complex. The leakage zone is approximately
800 m wide, which is interpreted as a seepage pipe. b) Shows chaotic seismic reectors, and high amplitude reector from the central part of the Finnmark Platform, suggesting
migration of uids. A depression, approximately 600 m wide, is also visible above the uid migration feature. The downward bending reectors indicate acoustic pull-down due to
low-velocity material, probably gas. c) Cross-cutting high-amplitude reectors associated with the gas chimney from the Bjrnyrenna Fault Complex. Huge region of masked
acoustic reections and polarity reversed high-amplitude anomalies show uid ow and gas accumulation. The patchy, crosscutting high-amplitude reector mimics the seaoor
and is interpreted as the base of gas hydrate stability zone. d) Seismic prole from the Bjrnya fan showing pockmark-like depressions and associated high amplitude feeders. The
depressions were few hundred meters across. Acoustic pull-down and presence of bright spots suggest the presence of gas.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 216
the Hammerfest Basin, gas chimneys are present close to the
Snhvit reservoir. Many of the gas chimneys were located right
above major faults in the study area (Fig. 7). Srvestsnaget Basin
shows many gas chimneys, which are not related to any major faults.
However, they are small and widely distributed.
Leakages along faults are present in almost all parts of the SW
Barents Sea. Their density, however, was higher in the western part,
especially in the Loppa high and areas surrounding it. RLFC and BFC
show high concentrations of leakage along faults. The major fault
boundaries between different basins also show numerous fault
leakages (Fig. 7). Our observations indicate that the uid leakages
are mainly related to major faults in the area.
Although the distribution of seismic data used for the study is
uneven, the western part of the study area clearly shows a higher
concentration of uid-ow features. This can be attributed to
various reasons. Presence of high number of faults in the western
part increases the odds of uid ow since fault reactivation during
uplift and glacial cycles could negatively affect the sealing ability.
The upper Jurassic, Triassic and Permian/Carboniferous source
rocks are oil mature in the western part and can exsolve gas during
uplift and erosion (Ohmet al., 2008). Triassic source rocks are over-
mature or gas-mature in the western part of SW Barents Sea (Ohm
et al., 2008). In addition, evaporites in the eastern part of the study
area might have negatively affected hydrocarbon migration to the
shallow strata (Dor and Jensen, 1996).
6. Large uid-ow features
The high density of 2D seismic proles in some areas allowed us
to map the aerial extent of some of the uid-owfeatures in the SW
Barents Sea. The size and shape of these features are approximate,
since they are derived from 2D data. Areas such as the Loppa High,
northern part of Hammerfest Basin, northern part of Troms Basin
and the RLFC have good coverage of 2Dseismic data, which allowed
the mapping of large features in these locations.
In all, 93 large uid-ow features were mapped (Fig. 8). They
have peculiar shapes and cover huge areas. Most of them terminate
in the shallow strata, with high amplitude reections at their crest,
suggesting shallow gas accumulation. The area covered by these
large uid-ow features varied from 1 to 600 km
2
. Of the 93 large
features mapped, 81% were comparatively smaller in area, varying
from 1 to 50 km
2
. The largest mapped uid-ow feature is a gas
chimney, located on the RLFC and northern part of Troms Basin,
covering approximately 600 km
2
. The total area covered by these
large uid-ow features is approximately 3000 km
2
, which is
approximately one percent of the SW Barents Sea. The shape of
these features varies widely from circular to elongate. Elongated
uid-ow features are seen located above major faults, while cir-
cular uid-ow features are present on all parts.
The Ringvassy-Loppa and Bjrnyrenna Fault Complexes, the
northern part of the Troms Basin, the Polheim Sub-platform, and
the Veslemy High contain most of the large uid-ow features.
Srvestsnaget Basin and Finnmark Platform show few small uid-
ow features. The Veslemy High has a denser distribution of
uid-ow features. These features are also larger than in other
areas and mostly occur above major fault boundaries. The northern
part of Troms Basin, devoid of salt domes, also shows large uid-
ow features especially along the boundaries with Veslemy High,
Loppa High and Polheim Sub-Platform. The Hammerfest Basin
shows large uid-ow features close to the Snhvit reservoir and
on its boundary with the Loppa High. The Bjrnya Basin shows
two of the largest uid-ow features, covering 260 km
2
and
Figure 7. Map showing the distribution of gas chimneys and leakage along faults in the SW Barents Sea. The black lines are the major fault boundaries. Most of the features were
located right on top of major faults in the area.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 217
200 km
2
, above the Bjornyrenna Fault Complex. The Loppa High
shows a very large gas chimney on the northern part, which covers
an area of 290 km
2
. The largest uid-owfeature in the eastern part
of the study area is located above the Samson Dome and covers
150 km
2
. The Nordkapp Basin shows no large uid-ow features.
7. Discussion
We observed subsurface uid-ow systems on all parts of the
SW Barents Sea. The observed features were of various types,
interpreted as gas chimneys, leakage along faults and fractures and
other related features. However, gas chimneys were dominant in
most parts of the study area (Fig. 7). A wide variety of uid-ow
expressions probably indicates the variable response to glacial
cycles, uplift and erosion in the different basins. Fluid-ow features
are spatially related to the structural elements of the study area,
especially with faults (Fig. 7) separating the major basins and plat-
form in the SW Barents Sea. Most of the uid-ow features were
located above major deep-seated faults located in hydrocarbon-rich
source/reservoir rocks in the study area. Fluid ow along faults is
also widespread in the study area (Figs. 5 and 7). Faults and fractures
can act as migration pathways for uids since they are generally
more permeable than the surrounding rock (Berndt et al., 2003;
Cartwright et al., 2007). Removal of sedimentary overburden during
uplift and erosion, and the consequent decrease in pressure, will
cause the gas in a gas accumulation to expand. The effect of this
expansion, assuming that the pre-existing structure was lled to
spill, will be the expulsion of hydrocarbons from the closure. This
will result in shorter oil legs and/or a less dense, and hence smaller,
gas accumulation (c.f. Nyland et al., 1992). The same effect could
occur during repeated loading and unloading of glaciers (Lerche
et al., 1997). Thus glacial cycles, isostatic uplift, and erosion during
the Cenozoic can be responsible for the majority of the uid leakage
observed throughout the SW Barents Sea.
Near the Loppa High, where large uid-ow features are
concentrated, it is evident that the leakage is along the rotated fault
blocks (Fig. 4a), which might have been reactivated during glacial
cycles. These rotated fault blocks contain one of the prolic reser-
voir rocks of the Barents Sea, the St formation. The Snadd for-
mation with shale and sandstone inter-bedding could also be a
source for hydrocarbons in this area. Fracturing of the cap rock can
be a major cause of uid leakage (Arntsen et al., 2007), which could
be the reason for uid leakage in the Samson Dome (Fig. 4c).
Sudden release of pressure from a highly pressurized uid accu-
mulation could have resulted in the formation of seepage pipe in
the Bjarmeland Platform (Fig. 6a) (Lseth et al., 2011; Cartwright
et al., 2007). Structural traps, especially faults, are thought to be
highly sensitive to the effects of uplift and erosion (Henriksen et al.,
2011). Extensional fault-dominated areas such as the Hammerfest
Basin, Loppa High and Veslemy High showed a high density of
uid-ow features. Salt-related structural traps, which are
considered stable against considerable amounts of uplift and
erosion, present in the Nordkapp and Troms basins showed rela-
tively less uid-ow features. However, numerous small gas
chimneys occur along the southern and western rims of the
Nordkapp Basin (Fig. 7). Thus, the structural evolution of SW
Barents Sea might have played a major role in the uid owhistory
of the area.
The large uid-ow features represent signicant areas of hy-
drocarbon leakage. Many of the largest ones were located close to
the two major fault complexes in the area, RLFC and BFC. Large
uid-ow zones suggest unfocussed migration of gas from the
Figure 8. Distribution of large uid ow features in the area. They vary in area fromw1 km to w600 km
2
. Histogram plot of area covered by these large features show 81% of them
falling in the category with an area of 1e50 km
2
.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 218
rotated fault blocks, which could provide multiple points of hy-
drocarbon spillage. Such extensive areas of leakage suggest the
presence of signicant hydrocarbon reservoirs in the subsurface.
Some of the uids from these reservoirs have clearly leaked from
traps. There still, however, may be economic quantities of hydro-
carbon beneath these uid-leakage areas, as evidenced by discov-
eries in the Snhvit and Goliat elds in Hammerfest Basin (Dor,
1995).
The source of the observed uid-ow features was difcult to
identify from seismic data due to their deep origin and acoustic
masking caused by upward migrating gas. The tentative maturity of
various source rocks in the Norwegian Barents Sea suggested by
Ohm et al. (2008) shows oil mature upper Jurassic Hekkingen
Formation in the Hammerfest basin and the northern part of the
Loppa High. Triassic and Permian/Carboniferous formations (Snadd,
Havert) are oil mature in major basins in the SW Barents Sea
including the Hammerfest Basin. The presence of multiple source
rocks, especially in the western part of study area, increases the
chance of hydrocarbongenerationanddissolutionof gas fromoil and
can also contribute to mixing of hydrocarbons from different strat-
igraphic intervals, as observed insome wells (e.g. 7120/2-1, 7125/1-1,
Goliat oil well 7122/7-3) in the Barents Sea (Ohm et al., 2008).
Isostatic uplift and associated erosion is considered to have
affected the hydrocarbon accumulation and migration in the SW
Barents Sea. Many workers (Vorren et al., 1991; Riis, 1992; Nyland
et al., 1992; Henriksen et al., 2011) have discussed net erosion in
the Barents Sea. Their studies suggest net erosion values of
approximately 900e2000 m in the SW Barents Sea and conclude
that different amounts of uplift and erosion in the area could affect
the petroleum systems. We correlated the net erosion of the
SW Barents Sea (Henriksen et al., 2011) with the distribution of
uid-ow features. We could not nd any relation between the
amount of net erosion and the distribution of uid-ow features
(Fig. 9). The areas with a dense distribution of uid-owfeatures in
the western half of the study area show an erosion of approxi-
mately 1000 m, whereas the eastern half, with a net erosion of
around 2000 m, shows comparatively less uid-ow features.
This suggests that the uid leakage in the study area is primarily
controlled by the presence of mature source rocks and structural
traps, especially faults. Tectonic faulting and fracturing is often
suggested as the controlling mechanism for the location and dis-
tribution of uid-ow observations worldwide (Gay et al., 2007;
Barnes et al., 2010; Talukder et al., 2003). In the western part of the
study area, most of the uid-ow features terminate in younger
sediments (late Tertiary-Quaternary) (Figs. 4, 5c, 6c, and d) which
suggests that they might have been triggered by uplift and glacia-
tions during the late PlioceneePleistocene.
Gas chimneys terminate at different stratigraphic levels in the
study area (Figs. 4e6). This suggests that the timing of uid ow
could be different from basin to basin. However, regional tectonic
events, such as rifting, uplift or glaciations would affect the whole
region and could trigger uid leakage at the same time regionally.
Difference in the lithology of shallow sediments in the area could
cause different termination point for chimneys, even though the
same tectonic event triggered them. Since the structural evolution
of the SW Barents Sea is complex and different from basin to basin,
a more focused study on individual basins may be needed to
accurately analyze the origin and timing of these uid-ow
features.
8. Conclusions
Fluid ow is abundant and widespread in the SW Barents Sea
with some very large gas chimneys covering approximately one
Figure 9. The amount of net erosion (black lines, after Henriksen et al., 2011) does not seem to affect the distribution of uid ow (in pink).
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 219
percent of the study area. Different types of uid-ow features
occur in the area, the most dominant being gas chimneys. The
distribution of uid-ow features showed a direct relationship
with the structural setting of the SW Barents Sea. This could be
due to the inability of deep-seated faults to effectively seal the
trapped hydrocarbons. Fault traps could be highly sensitive to
the effects of uplift, erosion and glaciations. The uid-ow fea-
tures were highly dense in the western part of the SW Barents
Sea. The difference in distribution could be due to a) presence of
multiple source rocks, which are over-mature or gas mature in
the western part, b) availability of structural traps, especially
faults and c) presence of evaporites in the eastern part. The
amount of net erosion in the study area did not seem to be
controlling the distribution of uid leakage features. Given the
relation of uid ow to the structural setting, extensional tec-
tonics, uplift and glaciations in the Plio-Pleistocene time could
have played major roles in the timing and activity of the uid
leakage, although erosion might have had an added effect.
However, due to the structural complexity of the study area and
the number of uid-ow features observed, a more focused
study on the timing and origin of uid-ow features in indi-
vidual basins is necessary.
Acknowledgments
The authors are thankful to TGS-NOPEC Geophysical Company
for providing some of the seismic data. We thank Schlumberger for
providing the Petrel Interpretation software. We are grateful to
Andrew Smith for correcting the English language. Reviews by
Martin Hovland and Herald Ligtenberg signicantly improved the
manuscript.
References
Aguilera, R., 1995. Naturally Fractured Reservoirs, second ed. PennWell, Tulsa, OK.
Andreassen, K., Hogstad, K., Berteussen, K.A., 1990. Gas hydrate in the southern
Barents Sea, indicated by a shallow seismic anomaly. First Break 8, 235e245.
Andreassen, K., Nilssen, E., degaard, C., 2007a. Analysis of shallow gas and uid
migration within the Plio-Pleistocene sedimentary succession of the SW
Barents Sea continental margin using 3D seismic data. Geo-Marine Letters 27,
155e171.
Andreassen, K., degaard, C.M., Rafaelsen, B., 2007b. Imprints of former ice streams
imaged and interpreted using industry three-dimensional seismic data from
the south-western Barents Sea. In: Geological Society, London, Special Publi-
cations, vol. 277, pp. 151e169.
Anell, I., Thybo, H., Artemieva, I.M., 2009. Cenozoic uplift and subsidence in
the North Atlantic region: geological evidence revisited. Tectonophysics 474,
78e105.
Arntsen, B., Wensaas, L., Loseth, H., Hermanrud, C., 2007. Seismic modeling of gas
chimneys. Geophysics 72, Sm251eSm259.
Berglund, L.T., Augustson, J., Frseth, R., Gjelberg, I., Ramberg-Moe, H., 1986. The
evolution of the Hammerfest Basin. In: Spencer, A.M., et al. (Eds.), Habitat of
Hydrocarbons on the Norwegian Continental Shelf: Proceedings of an Inter-
national Conference (Habitat of Hydrocarbons, Norwegian Oil and Gas Finds).
Graham & Trotman for the Norwegian Petrolium Society, London, pp. 319e338.
Barnes, P.M., Lamarche, G., Bialas, J., Henrys, S., Pecher, I., Netzeband, G.L.,
Greinert, J., Mountjoy, J.J., Pedley, K., Crutchley, G., 2010. Tectonic and geological
framework for gas hydrates and cold seeps on the Hikurangi subduction
margin, New Zealand. Marine Geology 272, 26e48.
Berndt, C., Bnz, S., Mienert, J., 2003. Polygonal fault systems on the mid-Norwegian
margin: a long-term source for uid ow. In: Geological Society, London,
Special Publications, vol. 216, pp. 283e290.
Bjory, M., Hall, P.B., Ferriday, I.L., Mrk, A., 2009. Triassic Source Rocks of the
Barents Sea and Svalbard. AAPG Convention, Denver, Colorado, USA.
Bjrkum, P.A., Walderhaug, O., Nadeau, P.H., 2001. Thermally driven porosity
reduction: impact on basin subsidence. In: Geological Society, London, Special
Publications, vol. 188, pp. 385e392.
Brekke, H., Riis, F., 1987. Tectonics and Basin Evolution of the Norwegian Shelf
Between 62-Degrees-N and 72-Degrees-N. In: Norsk Geol Tidsskr, vol. 67.
pp. 295e321.
Carstens, H., 2005. Gas found in glacial, shallow sands. GeoExPro, Geoscience and
Technology Explained 4, 24e25.
Cartwright, J., Huuse, M., Aplin, A., 2007. Seal bypass systems. Aapg Bulletin 91,
1141e1166.
Cavanagh, A.J., Di Primio, R., Scheck-Wenderoth, M., Horseld, B., 2006. Severity and
timing of Cenozoic exhumation in the southwestern Barents Sea. Journal of the
Geological Society, London 163, 761e774.
Chand, S., Mienert, J., Andreassen, K., Knies, J., Plassen, L., Fotland, B., 2008. Gas
hydrate stability zone modelling in areas of salt tectonics and pockmarks of the
Barents Sea suggests an active hydrocarbon venting system. Marine and
Petroleum Geology 25, 625e636.
Chand, S., Rise, L., Ottesen, D., Dolan, M.F.J., Bellec, V., Be, R., 2009. Pockmark-like
depressions near the Goliat hydrocarbon eld, Barents Sea: morphology and
genesis. Marine and Petroleum Geology 26, 1035e1042.
Chand, S., Thorsnes, T., Rise, L., Brunstad, H., Stoddart, D., Be, R., Lgstad, P.,
Svolsbru, T., 2012. Multiple episodes of uid ow in the SW Barents Sea (Loppa
High) evidenced by gas ares, pockmarks and gas hydrate accumulation. Earth
and Planetary Science Letters 331e332, 305e314.
Connolly, D.L., Brouwer, F., Walraven, D., 2008. Detecting fault-related hydrocarbon
migration pathways in seismic data: implications for fault-seal, pressure, and
charge prediction. Gulf Coast Association of Geological Societies Transactions
58, 191e203.
Dahl, B., Augustson, J.H., 1993. The inuence of Tertiary and Quaternary sedimen-
tation and erosion on hydrocarbon generation in the Norwegian offshore
basins. In: Dor, A.G., et al. (Eds.), Basin Modeling: Advances and Applications.
NPF Special Publications 3. Elsevier, Amsterdam, pp. 419e432.
Dalland, A., Worsley, D., Ofstad, K., 1988. A Lithostratigraphical Scheme for the
Mesozoic and Cenozoic Succession Offshore Mid- and Northern Norway.
Norwegian Petroleum Directorate Bulletin 4.
Dimakis, P., Braathen, B.I., Faleide, J.I., Elverhi, A., Gudlaugsson, S.T., 1998. Cenozoic
erosion and the preglacial uplift of the SvalbardeBarents Sea region. Tectono-
physics 300, 311e327.
Dor, A.G., 1995. Barents Sea geology, petroleum resources and commercial
potential. Arctic 48, 207e221.
Dor, A.G., Jensen, L.N., 1996. The impact of late Cenozoic uplift and erosion on
hydrocarbon exploration: offshore Norway and some other uplifted basins.
Global and Planetary Change 12, 415e436.
Dor, A.G., Scotchman, I.C., Corcoran, D., 2000. Cenozoic exhumation and prediction
of the hydrocarbon system on the NW European margin. Journal of
Geochemical Exploration 69e70, 615e618.
Dugan, B., Flemings, P.B., 2000. Overpressure and uid ow in the New Jersey
continental slope: implications for slope failure and cold seeps. Science 289,
288e291.
Eidvin, T., Jansen, E., Riis, F., 1993. Chronology of Tertiary fan deposits off the
western Barents Sea: implications for the uplift and erosion history of the
Barents Shelf. Marine Geology 112, 109e131.
Eldholm, O., Faleide, J.I., Myhre, A.M., 1987. Continent-ocean transition at the
Western Barents Sea Svalbarrd continental-margin. Geology 15, 1118e1122.
Elverhi, A., Solheim, A., 1983. The physical environment western Barents Sea, 1:
1,500,000, sheet A, surface sediment distribution. Norsk Polar Institutt Rap-
portserie 179A, 23.
Faleide, J.I., Solheim, A., Fiedler, A., Hjelstuen, B.O., Andersen, E.S., Vanneste, K., 1996.
Late Cenozoic evolution of the western Barents Sea-Svalbard continental
margin. Global and Planetary Change 12, 53e74.
Faleide, J.I., Tsikalas, F., Breivik, A.J., Mjelde, R., Ritzmann, O., Engen, O., Wilson, J.,
Eldholm, O., 2008. Structure and evolution of the continental margin of Norway
and Barents Sea. Episodes 3, 82e91.
Faleide, J.I., Gudlaugsson, S.T., Jacquart, G., 1984. Evolution of the western Barents
Sea. Marine and Petroleum Geology 1, 123e150.
Faleide, J.I., Vgnes, E., Gudlaugsson, S.T., 1993. Late Mesozoic-Cenozoic evolution of
the south-western Barents Sea in a regional rift-shear tectonic setting. Marine
and Petroleum Geology 10, 186e214.
Gabrielsen, R.H., Frseth, R.B., Jensen, L.N., Kalheim, J.E., Riis, F., 1990. Structural
Elements of the Norwegian Continental Shelf Part I: the Barents Sea Region.
Norwegian Petroleum Directorate Bulletin 6.
Gay, A., Lopez, M., Berndt, C., Sranne, M., 2007. Geological controls on focused uid
ow associated with seaoor seeps in the Lower Congo Basin. Marine Geology
244, 68e92.
Gjelberg, J.G., 1981. Upper Devonian (Famennian) e Middle Carboniferous succes-
sion of Bjrnya: a study of ancient alluvial and coastal marine sedimentation.
Norsk Polarinstitutt Skrifter 174, 1e67.
Gjelberg, J.G., 1987. Early Carboniferous graben style and sedimentation response,
Svalbard. In: Miller, J., Adams, A.E., Wright, V.P. (Eds.), European Dinantian
Environments. John Wiley & Sons, Chichester, pp. 93e113.
Graue, K., 2000. Mud volcanoes in deepwater Nigeria. Marine and Petroleum
Geology 17, 959e974.
Gudlaugsson, S.T., Faleide, J.I., Johansen, S.E., Breivik, A.J., 1998. Late Palaeozoic
structural development of the South-western Barents Sea. Marine and Petro-
leum Geology 15, 73e102.
Heggland, R., 1998. Gas seepage as an indicator of deeper prospective reservoirs. A
study based on exploration 3D seismic data. Marine and Petroleum Geology 15,
1e9.
Heggland, R., 2004. Denition of geohazards in exploration 3-D seismic data using
attributes and neural-network analysis. Aapg Bulletin 88, 857e868.
Henriksen, E., Bjrnseth, H.M., Hals, T.K., Heide, T., Kiryukhina, T., Klvjan, O.S.,
Larssen, G.B., Ryseth, A.E., Rnning, K., Sollid, K., Stoupakova, A., 2011. Chapter
17 Uplift and erosion of the greater Barents Sea: impact on prospectivity
and petroleum systems. In: Geological Society, London, Memoirs, vol. 35,
pp. 271e281.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 220
Hjelstuen, B.O., Sejrup, H.P., Haidason, H., Berg, K., Bryn, P., 2004. Neogene and
Quaternary depositional environments on the Norwegian continental margin,
62

Ne68

N. Marine Geology 213, 257e276.


Holbrook, W.S., Hoskins, H., Wood, W.T., Stephen, R.A., Lizarralde, D., 1996. Methane
hydrate and free gas on the Blake Ridge from vertical seismic proling. Science
273, 1840e1843.
Hornbach, M.J., Ruppel, C., Van Dover, C.L., 2007. Three-dimensional structure of
uid conduits sustaining an active deep marine cold seep. Geophysical Research
Letters 34.
Hovland, M., Judd, A.G., 1988. Seabed Pockmarks and Seepages. Graham and Trot-
man, London.
Hurst, A., Cartwright, J., Huuse, M., Jonk, R., Schwab, A., Duranti, D., Cronin, B., 2003.
Signicance of large-scale sand injectites as long-term uid conduits: evidence
from seismic data. Geouids 3, 263e274.
Jackson, C., Stoddart, D., 2005. Temporal constraints on the growth and decay of
large-scale mobilized mud masses and implications for uid ow mapping in
sedimentary basins. Terra Nova 17, 580e585.
Jebsen, C., Faleide, J.I., 1998. Tertiary rifting and magmatism at the western Barents
Sea margin (Vestbakken volcanic province). In: 3rd International Conference on
Arctic Margins, p. 92.
Johansen, S.E., Ostisty, B.K., Birkeland, ., Federovsky, Y.F., Martirosjan, V.N., Bruun
Cristensen, O., Cheredeev, S.I., Ignatenko, E.A., Margulis, L.S., 1993. Hydrocarbon
potential in the Barents Sea region: play distribution and potential. In:
Vorren, T.O., Bergsager, E., Dahl-Stamnes, .A., Holter, E., Johansen, B., Lie, E.,
Lund, T. (Eds.), Arctic Geology and Petroleum Potential. Elsevier, Amsterdam,
pp. 273e320.
Judd, A., Hovland, M., 2007. Seabed Fluid Flow. Cambridge University Press,
Cambridge.
Kjemperud, A., Fjeldskaar, W., 1992. Pleistocene glacial isostasy-implications for
petroleum geology. In: Larsen, R.M., Brekke, H., Larsen, B.T., Talleraas, E. (Eds.),
Structural and Tectonic Modeling and its Application to Petroleum Geology. NPF
Special Publications 1. Elsevier, Amsterdam, pp. 187e195.
Kvenvolden, K.A., 1998. A primer on the geological occurrence of gas hydrate. In:
Geological Society, London, Special Publications, vol. 137, pp. 9e30.
Laberg, J.S., Andreassen, K., 1996. Gas hydrate and free gas indications within the
Cenozoic succession of the Bjornoya Basin, western Barents Sea. Marine and
Petroleum Geology 13, 921e940.
Lammers, S., Suess, E., Hovland, M., 1995. A large methane plume east of Bear Island
(Barents Sea): implications for the marine methane cycle. Geologische Run-
dschau 84, 59e66.
Larsen, R.M., Fjran, T., Skarpnes, O., 1993. Hydrocarbon potential of the Norwegian
Barents Sea based on recent well results. In: Vorren, T.O., Bergsager, E., Dahl-
Stamnes, .A., Holter, E., Johansen, B., Lie, E., Lund, T. (Eds.), Arctic Geology and
Petroleum Potential. Elsevier, Amsterdam, pp. 321e331.
Lerche, I., Yu, Z., Trudbakken, B., Thomsen, R.O., 1997. Ice loading effects in sedi-
mentary basins with reference to the Barents Sea. Marine and Petroleum Ge-
ology 14, 277e338.
Ligtenberg, J.H., 2005. Detection of uid migration pathways in seismic data: im-
plications for fault seal analysis. Basin Research 17, 141e153.
Ligtenberg, J.H., 2007. Indications for Pressure Release from Zechstein Rafts during
Late Kimmerian e Implications for Reducing Drilling Risks, EAGE 69th Con-
ference & Exhibition, London.
Lseth, H., Wensaas, L., Arntsen, B., 2002. Gas Chimneys e Indication of Fractured
Cap Rocks, AAPG Hedberg Conference, Vancouver, BC, Canada.
Lseth, H., Gading, M., Wensaas, L., 2009. Hydrocarbon leakage interpreted on
seismic data. Marine and Petroleum Geology 26, 1304e1319.
Lseth, H., Wensaas, L., Arntsen, B., Hanken, N.M., Basire, C., Graue, K., 2011. 1000 m
long gas blow-out pipes. Marine and Petroleum Geology 28, 1047e1060.
Lseth, H., Wensaas, L., Arntsen, B., Hovland, M., 2003. Gas and uid injection
triggering shallow mud mobilization in the Hordaland Group, North Sea. In:
Geological Society, London, Special Publications, vol. 216, pp. 139e157.
Maximov, S.P., Zolotov, A.N., Lodzhevskaya, M.I., 1984. Tectonic conditions for oil
and gas generation and distribution on ancient platforms. Journal of Petroleum
Geology 7, 329e339.
Meldahl, P., Heggland, R., Bril, B., de Groot, P., 2001. Identifying faults and gas
chimneys using multiattributes and neural networks. The Leading Edge 20,
474e482.
Mienert, J., Vanneste, M., Bnz, S., Andreassen, K., Haidason, H., Sejrup, H.P., 2005.
Ocean warming and gas hydrate stability on the mid-Norwegian margin at the
Storegga slide. Marine and Petroleum Geology 22, 233e244.
Myhre, A.M., Eldholm, O., Sundvor, E., 1982. The margin between Senja and Spits-
bergen fracture zones: Implications from plate tectonics. Tectonophysics 89,
33e50.
Nesterov, I.I., Salvamanov, F.K., Kontorovich, A.E., Kulakhmetov, N.K., Surkov, V.S.,
Tromuk, A.A., Shpilman, V.I., 1990. West Siberian oil and gas superprovince. In:
Geological Society, London, Special Publications, vol. 50, pp. 491e502.
Nyland, B., Jensen, L.N., Skagen, J., Skarpnes, O., Vorren, T.O., 1992. Tertiary uplift and
erosion in the Barents Sea: magnitude, timing and consequences. In: Larsen, R.M.,
Brekke, H., Larsen, B.T., Talleraas, E. (Eds.), Structural and Tectonic Modeling and
its Application to Petroleum Geology. Elsevier, Amsterdam, pp. 153e162.
Ohm, S.E., Karlsen, D.A., Austin, T.J.F., 2008. Geochemically driven exploration
models in uplifted areas: examples from the Norwegian Barents Sea. Aapg
Bulletin 92, 1191e1223.
Ostanin, I., Anka, Z., di Primio, R., Bernal, A., 2012. Identication of a large Upper
Cretaceous polygonal fault network in the Hammerfest basin: implications on
the reactivation of regional faulting and gas leakage dynamics, SW Barents Sea.
Marine Geology 332e334, 109e125.
Perez-Garcia, C., Feseker, T., Mienert, J., Berndt, C., 2009. The Hkon Mosby mud
volcano: 330 000 years of focused uid ow activity at the SW Barents Sea
slope. Marine Geology 262, 105e115.
Piggott, N., Lines, M.D., 1991. A case study of migration from the West Canada Basin.
In: Geological Society, London, Special Publications, vol. 59, pp. 207e225.
Rasmussen, A., Kristensen, S.E., Van Veen, P.M., Stelan, T., Vail, P.R., 1993. Use of
sequence stratigraphy to dene a semi-stratigraphic play in Anisian sequences,
southwest Barents Sea. In: Vorren, T.O., Bergsager, E., Dahl-Stamnes, .A.,
Holter, E., Johansen, B., Lie, E., Lund, T. (Eds.), Arctic Geology and Petroleum
Potential. Elsevier, Amsterdam, pp. 439e455.
Riis, F., Vollset, J., Sand, M., 1986. Tectonic development of the western margin of the
Barents Sea and adjacent areas. In: Halbouty, M.T. (Ed.), Future Petroleum
Provinces of the World. AAPG Memoirs, pp. 661e676.
Riis, F., 1992. Dating and Measuring Uplift and Subsidence in Norway and the
Norwegian Shelf in Glacial Periods. In: Norsk Geol Tidsskr, vol. 72. pp. 325e331.
Sales, J.K., 1993. Closure vs. seal capacity-a fundamental control on the distribution
of oil and gas. In: Dor, A.G., et al. (Eds.), Basin Modeling: Advances and Ap-
plications. NPF Special Publications 3. Elsevier, Amsterdam, pp. 399e413.
Skagen, J.I., 1993. Effects of hydrocarbon potential caused by tertiary uplift and
erosion in the Barents Sea. In: Vorren, T.O., et al. (Eds.), Arctic Geology and Pe-
troleum Potential. NPF Special Publications 2. Elsevier, Amsterdam, pp. 711e719.
Sloan, E.D., Koh, C.A., 2007. Clathrate Hydrates of Natural Gases, third ed. CRC Press,
Boca Raton, FL.
Smelror, M., Petrov, O., Larssen, G.B., Werner, S., 2009. Geological History of the
Barents Sea. Geological Survey of Norway, Trondheim, Norway.
Solheim, A., Elverhi, A., 1985. A pockmark eld in the central Barents Sea; gas from
a petrogenic source? Polar Research 3, 11e19.
Solheim, A., Elverhi, A., 1993. Gas-related sea oor craters in the Barents Sea. Geo-
Marine Letters 13, 235e243.
Solheim, A., Kristoffersen, Y., 1984. Sediments above the upper regional unconfor-
mity; thickness, seismic stratigraphy and outline of the glacial history. Norsk
Polar Institutt Rapportserie 179B, 26.
Talukder, A.R., 2012. Review of submarine cold seep plumbing systems: leakage to
seepage and venting. Terra Nova 24, 255e272.
Talukder, A.R., Comas, M.C., Soto, J.I., 2003. Pliocene to recent mud diapirism and
related mud volcanoes in the Alboran Sea (Western Mediterranean). In:
Geological Society, London, Special Publications, vol. 216, pp. 443e459.
Talwani, M., Eldholm, O., 1977. Evolution of Norwegian Greenland Sea. Geological
Society of America Bulletin 88, 969e999.
Vorren, T.O., Richardsen, G., Knutsen, S.-M., Henriksen, E., 1991. Cenozoic erosion
and sedimentation in the western Barents Sea. Marine and Petroleum Geology
8, 317e340.
Walderhaug, O., 1992. Magnitude of Uplift of the St and Nordmela Formations in
the Hammerfest Basin e a Diagenetic Approach. In: Norsk Geol Tidsskr, vol. 72.
pp. 321e323.
Waylett, D.C., Embry, A.F., 1992. Hydrocarbon loss from oil and gas elds of the
Sverdrup Basin, Canadian Arctic Islands. In: Vorren, T.O., et al. (Eds.), Arctic
Geology and Petroleum Potential. NPF Special Publications 2. Elsevier,
Amsterdam, pp. 195e204.
Worsley, D., 2008. The post-Caledonian development of Svalbard and the western
Barents Sea. Polar Research 27, 298e317.
Ziegler, P.A., 1988. Evolution of the Arctic-North Atlantic and the Western Tethys.
Am. Assoc. Geol. Mem. No. 43, 198 pp.
S. Vadakkepuliyambatta et al. / Marine and Petroleum Geology 43 (2013) 208e221 221

You might also like