ArticlePDF Available

Tectonic evolution of southern Baffin Bay and Davis Strait: Results from a seismic refraction transect between Canada and Greenland

Authors:

Abstract and Figures

Wide-angle reflection/refraction seismic data were acquired on a 450-km-long transect in southern Baffin Bay extending from Baffin Island to Greenland. Dense air gun shots were recorded on 22 ocean bottom seismometers. A P wave velocity model was developed from forward and inverse modeling of the observed travel times. Beneath the Baffin Island shelf, a three-layered continental crust is observed with velocities of 5.5 to 6.9km/s. Typical for transform margins, there is a sharp transition between continental and oceanic crust. Off Baffin Island, 7-km-thick oceanic crust is interpreted to lie in a major transform fault identified on the gravity map. Beneath the deep Baffin Bay basin, 9-km-thick oceanic crust is encountered but thins to 6km within an assumed fracture zone. The thicker than normal oceanic crust indicates an ample magma supply, possibly related to melt extracted from a mantle plume. Seaward of the Greenland continental crust, 20-km-thick igneous crust (6.3 to 7.3km/s) is encountered in a 25-km-wide zone interpreted as a leaky transform fault that can be correlated southward through Davis Strait. The igneous crust is bounded by a 20-km wide basin to the west, underlain by 4-km-thick crust of unknown affinity. This structure is probably associated with transform movements. A high-velocity lower crustal layer (7.1km/s) of 8km thickness is indicated beneath the Greenland crust and can be correlated into the adjacent thick igneous crust. Both the thick igneous and Greenland crust are covered by up to 4km of Paleogene volcanics (5.2 to 5.7km/s).
Content may be subject to copyright.
Tectonic evolution of southern Baffin Bay and Davis Strait:
Results from a seismic refraction transect between Canada
and Greenland
Thomas Funck,
1
Karsten Gohl,
2
Volkmar Damm,
3
and Ingo Heyde
3
Received 19 December 2011; revised 2 March 2012; accepted 8 March 2012; published 26 April 2012.
[1]Wide-angle reflection/refraction seismic data were acquired on a 450-km-long
transect in southern Baffin Bay extending from Baffin Island to Greenland. Dense air
gun shots were recorded on 22 ocean bottom seismometers. A Pwave velocity model
was developed from forward and inverse modeling of the observed travel times.
Beneath the Baffin Island shelf, a three-layered continental crust is observed with
velocities of 5.5 to 6.9 km/s. Typical for transform margins, there is a sharp transition
between continental and oceanic crust. Off Baffin Island, 7-km-thick oceanic crust is
interpreted to lie in a major transform fault identified on the gravity map. Beneath the
deep Baffin Bay basin, 9-km-thick oceanic crust is encountered but thins to 6 km within
an assumed fracture zone. The thicker than normal oceanic crust indicates an ample
magma supply, possibly related to melt extracted from a mantle plume. Seaward of the
Greenland continental crust, 20-km-thick igneous crust (6.3 to 7.3 km/s) is
encountered in a 25-km-wide zone interpreted as a leaky transform fault that can be
correlated southward through Davis Strait. The igneous crust is bounded by a 20-km
wide basin to the west, underlain by 4-km-thick crust of unknown affinity. This
structure is probably associated with transform movements. A high-velocity lower
crustal layer (7.1 km/s) of 8 km thickness is indicated beneath the Greenland crust and
can be correlated into the adjacent thick igneous crust. Both the thick igneous and
Greenland crust are covered by up to 4 km of Paleogene volcanics (5.2 to 5.7 km/s).
Citation: Funck, T., K. Gohl, V. Damm, and I. Heyde (2012), Tectonic evolution of southern Baffin Bay and Davis Strait:
Results from a seismic refraction transect between Canada and Greenland, J. Geophys. Res.,117, B04107,
doi:10.1029/2011JB009110.
1. Introduction
[2] The separation between Greenland and Canada has
remained largely unsolved because of an insufficient
understanding of the crustal nature in Baffin Bay and Davis
Strait and the tectonic relationship of both areas to the
Labrador Sea. In order to address this problem, the German
research vessel Maria S. Merian carried out a geophysical
survey in Davis Strait and southern Baffin Bay (Figures 1
and 2). The expedition was a component of the lead proj-
ect Plate Tectonics and Polar Gateways in the Earth Sys-
tem (PLATES and GATES)of the International Polar Year
(IPY 2007/08). The overall objective of the cruise was the
tectonic and sedimentary reconstruction of the opening of
Davis Strait and southern Baffin Bay. The study of this
paper covers the tectonic development and architecture of
the southern Baffin Bay and its transition to Davis Strait,
while the spreading history of southern and central Baffin
Bay is described in Suckro et al. [2012].
[3] No modern seismic refraction data are available for
southern Baffin Bay. Older sonobuoy profiles are interpreted
to indicate the presence of oceanic crust [Keen and Barrett,
1972]. However, the obtained velocity functions show a
substantial scatter and are not always compatible with
average oceanic crust [White et al., 1992]. Furthermore, no
linear magnetic anomalies typical for seafloor spreading
could be identified in Baffin Bay yet [Chalmers and
Pulvertaft, 2001]. The new data set will allow the nature of
the crust in southern Baffin Bay to be determined, whether it
consists of oceanic crust or of unroofed and partially ser-
pentinized mantle as suggested for northern Baffin Bay
[Reid and Jackson, 1997]. Despite the lack of unequivocal
evidence for oceanic crust in Baffin Bay, a prominent
gravity low in the center of the bay is interpreted as the
location of the extinct spreading axis [Whittaker et al.,
1997]. Other north-south striking anomalies are interpreted
as the expression of transform faults [Whittaker et al., 1997;
Chalmers and Pulvertaft, 2001]. The seismic refraction
1
Geological Survey of Denmark and Greenland, Copenhagen, Denmark.
2
Alfred Wegener Institute for Polar and Marine Research, Bremerhaven,
Germany.
3
Federal Institute for Geosciences and Natural Resources, Hanover,
Germany.
Published in 2012 by the American Geophysical Union.
JOURNAL OF GEOPHYSICAL RESEARCH, VOL. 117, B04107, doi:10.1029/2011JB009110, 2012
B04107 1of24
component of the project was designed to map the crustal
character and thickness in Baffin Bay, and to check for the
seismic expression of the gravity lows. This will set the stage
for the discussion of the tectonic development of the area.
[4] Paleogene volcanism in the Davis Strait and southern
Baffin Bay is associated with a mantle plume [Storey et al.,
1998]. The seismic refraction data will provide information
on the distribution and volume of this volcanism, both near
the surface and as underplated layers at deep-crustal levels.
This will also allow the interaction between the plume and
the transform-rifted margin in the Davis Strait [Funck et al.,
2007] to be investigated.
2. Geological Setting
2.1. General Setting
[5] Davis Strait is a bathymetric high between Baffin
Island to the west and Greenland to the east that separates
Labrador Sea from Baffin Bay (Figure 1). The oldest
undisputed seafloor spreading magnetic anomaly in the
Labrador Sea is magnetic chron 27n (62.061.7 Ma using
the time scale of Gradstein et al. [2005]) [Chalmers and
Laursen, 1995]. Other authors suggest that seafloor spread-
ing started during magnetic chron 33 (8473.6 Ma) [Roest
and Srivastava, 1989] or between magnetic chrons 29 and
31 (71.064.4 Ma) [Chian et al., 1995a]. A reorientation of
the spreading axis took place during magnetic chron 24r
(56.753.8 Ma), at the same time that seafloor spreading
initiated between Greenland and Europe. Spreading in the
Labrador Sea ceased by magnetic chron 13 (34.833.3 Ma)
[Srivastava, 1978]. The continental stretching and rifting in
the Labrador Sea and Davis Strait occurred over an extended
period of time as indicated by dyke swarms in West
Greenland. Based on the ages of these dykes, Larsen et al.
[2009] identified several phases of tectonomagmatic activ-
ity: initial stretching from the Late Triassic to Late Jurassic
(223 to 150 Ma), stretching and thinning in the Late Jurassic
(150 Ma), further stretching and thinning in the Early Cre-
taceous (140 to 133 Ma) followed by Paleogene thinning
and break-up (63 to 50 Ma).
[6] Baffin Bay is the northwest extension of the Labrador
Sea spreading system. The transform margin in Davis Strait
that links these two areas is characterized by the Ungava
transform fault, a name that was first introduced by Kerr
[1967]. Later, the term Ungava fault zone (UFZ) became
more commonly used (Figure 2). The position of the UFZ is
taken to be along the SE side of a line of positive gravity
anomalies. Davis Strait and the northern Labrador Sea are
bounded by volcanic margins [Keen et al., 2012; Gerlings
et al., 2009; Funck et al., 2007; Chalmers, 1997].
Onshore, Palaeogene volcanics crop out on either side of the
strait in a short narrow belt near Cape Dyer on Baffin Island
and in a wider zone in the Disko-Svartenhuk area of West
Greenland (Figure 2). Storey et al. [1998] identified two main
pulses of volcanism in West Greenland: one between 60.7
and 59.4 Ma and one between 54.8 and 53.6 Ma. The first
pulse is probably related to the arrival of the Iceland plume
beneath Greenland [Lawver and Müller, 1994; Storey et al.,
1998]. However, the assumed plume position at that time is
based on models and not on direct observations. Larsen and
Saunders [1998] explain the almost simultaneous volcanism
from 62 to 60 Ma in West Greenland [Storey et al., 1998],
East Greenland [Larsen and Saunders, 1998], and on the
British Isles [Pearson et al., 1996] by rapid lateral flow of a
small plume head that impinged on the continental litho-
sphere. Continental break-up of East Greenland from NW
Europe occurred at 56 Ma [Larsen and Saunders, 1998] and
caused a reorientation of the spreading axis in the Labrador
Sea [Srivastava, 1978]. Storey et al. [1998] suggested that the
Figure 1. Bathymetric map Baffin Bay and Labrador Sea.
Data source: General Bathymetric Chart of the Oceans
(GEBCO) 08 grid (http://www.gebco.net). Seismic refrac-
tion lines of the Maria S. Merian cruise MSM09/3 are
shown in red [Gohl et al., 2009]. Solid lines indicate pub-
lished seismic refraction lines in Baffin Bay and Davis Strait
as well as selected profiles in the Labrador Sea (lines 911
and 913[Jackson and Reid, 1994]; lines 912 and 914
[Reid and Jackson, 1997]; line 20013[Funck et al.,
2006]; line GR89-WA [Gohl and Smithson, 1993]; line
NUGGET 1 [Funck et al., 2007]; line NUGGET 2
[Gerlings et al., 2009]; line 88R2 [Chian and Louden,
1994]; line 90R1[ Chian et al., 1995b]). Dashed lines mark
the location of sonobuoy lines [Keen and Barrett, 1972;
Srivastava et al., 1982]. Abbreviations are FB, Foxe Basin;
LS, Lancaster Sound; NS, Nares Strait.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
2of24
second pulse of volcanism in West Greenland could be
related to this reorientation of the spreading axis, during
which remnants of the plume could have generated melts
along the UFZ. Volcanic layers, lava flows and seaward
dipping reflectors (SDR) are mapped in large areas of Davis
Strait [Skaarup et al., 2006; Chalmers and Laursen, 1995;
Chalmers, 1997]. In addition, volcanic rocks were drilled in a
number of wells (Hekja O-71 and Gjoa G-37 [Klose et al.,
1982]; Ralegh N-18 (BASIN database, Geological Survey
of Canada, Dartmouth, Nova Scotia, Canada); Hellefisk-1
Figure 2. Geologic map of the southern Baffin Bay, the Davis Strait region and the northernmost Lab-
rador Sea [after Chalmers and Oakey, 2007]. Red lines indicate the location of the MSM09/3 seismic
refraction lines (AWI20080[500/600/700]) with white circles showing the position of ocean bottom seism-
ometers (OBS) along line 600. The station number of every 5th OBS is labeled. Well locations are indi-
cated by yellow circles. The blue line marks the location of NUGGET line 1 [Funck et al., 2007].
Abbreviation SDR is seaward dipping reflections.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
3of24
and Nukik-2 [Rolle, 1985]). Laser Argon dating of basaltic
rocks from the Gjoa well (Figure 2) yielded ages of 59.5 Ma
[Williamson et al., 2001], which relates those rock samples to
the first pulse of volcanism.
[7] Baffin Bay forms a sedimentary basin that extends
from Davis Strait in the south to Nares Strait in the north.
The sediment thickness is up to 12 km in northern Baffin
Bay [Reid and Jackson, 1997]. The nature of the crust in
Baffin Bay is disputed. Keen et al. [1974] provide evidence
that the Baffin Bay crust was created by seafloor spreading,
which is supported by velocity models that are compatible
with oceanic crust [Keen and Barrett, 1972]. In contrast,
Reid and Jackson [1997] interpret their velocity model as
evidence for serpentinised mantle in northern Baffin Bay in
support for amagmatic continental rifting and separation.
Serpentinised mantle is often observed in ocean-continent
transition zones of non-volcanic continental margins [e.g.,
Funck et al., 2003, 2004] or at ultra-slow spreading ridges
[Jokat et al., 2003]. Chalmers and Pulvertaft [2001] point
out that there are no unequivocal magnetic anomalies in
Baffin Bay that can be related to seafloor spreading.
Absence of these anomalies could be related to oblique
spreading or to a blanketing effect of up to 12 km of sedi-
ments. The position of the extinct spreading axis and of the
transform faults in Figure 2 is determined from gravity and
magnetic anomaly maps [Chalmers and Pulvertaft, 2001;
Chalmers and Oakey, 2007].
2.2. Previous Geophysical Studies
[8] The coverage by refraction and regional deep seis-
mic reflection data is variable along the ocean basins west
of Greenland. The most detailed picture is available for the
Labrador Sea, where a number of refraction profiles define
the non-volcanic nature of the conjugate continental mar-
gins of Labrador and SW Greenland [e.g., Chian and
Louden, 1994; Chian et al., 1995b]. In addition, long
regional seismic reflection lines were acquired by the Federal
Institute for Geosciences and Natural Resources (BGR), the
Geological Survey of Canada (GSC) and LITHOPROBE
[e.g., Chalmers and Laursen, 1995; Chian et al., 1995a;
Hall et al., 2002].
[9] In the Baffin Bay, modern seismic refraction lines with
ocean bottom seismometers are limited to the entrance of the
Nares Strait [Jackson and Reid, 1994; Reid and Jackson,
1997], where Reid and Jackson [1997] interpret velocities
of 6.8 km/s as indication for serpentinized mantle rather than
oceanic crust. The remainder of the Baffin Bay is only
covered by short sonobuoy profiles (Figure 1), which show
velocities compatible with oceanic crust [Keen and Barrett,
1972].
[10] In the Davis Strait, two lines from the NUGGET
experiment [Funck et al., 2007; Gerlings et al., 2009] and a
line along the Greenland coast [Gohl and Smithson, 1993]
provide some information on the crustal structure in that
region (Figure 1). NUGGET line 1 shows that southern
Davis Strait is underlain by thinned continental crust with
the exception of an area around the UFZ, where crust of an
oceanic affinity was observed [Funck et al., 2007]. The
oceanic crust is associated with pronounced gravity and
magnetic anomalies that can be correlated through Davis
Strait. This suggests that the UFZ acted as a leaky transform
fault during phases of transtension. The 7 to 12-km-thick
continental crust on NUGGET line 1 is underlain by a
5 km thick high-velocity zone (HVZ) that Funck et al.
[2007] associate with southward flow of plume material
along lithospheric thin spots beneath the Davis Strait region,
similar to a model suggested earlier by Nielsen et al. [2002].
[11] While Funck et al. [2007] found no evidence for a
HVZ beneath the thicker continental crust close to Greenland,
Gohl and Smithson [1993] identified an up to 8-km-thick
HVZ at the base of the Greenlandic crust, where the Moho
depth varies between 30 to 42 km. Gohl and Smithson
[1993] interpret the HVZ to be associated with hot spot
magmatism in the Davis Strait and Baffin Bay region. This
magmatism is also observed along NUGGET line 2
(Figure 1), which extends from southern Davis Strait into
northern Labrador Sea. Here, Gerlings et al. [2009]
observed a 12-km-thick oceanic crust that is overlain by
an up to 2-km-thick sequence of Palaeogene basalts. This is
in the area of BGR line 776, where Chalmers and Laursen
[1995] identified seaward-dipping reflections that are
indicative of volcanic-style margins. Keen et al. [2012] also
report seaward-dipping reflectors off Labrador. However,
the volcanism is limited to the northern part of Labrador
Sea. Farther south, the continental margins have a non-
volcanic character with serpentinised mantle in the continent-
ocean transition zone [Chian et al., 1995b].
3. Wide-Angle Seismic Experiment
3.1. Data Acquisition and Processing
[12] Geophysical data were collected in September and
October 2008 onboard the German research vessel Maria
S. Merian [Gohl et al., 2009] including the three seis-
mic refraction lines AWI20080500, AWI20080600 and
AWI20080700. For simplicity, these lines are subsequently
referred to as lines 500, 600, and 700 (Figures 1 and 2).
This paper presents the results from line 600, a 450-km-
long transect (Figure 3) from the shelf off Baffin Island,
across the southern Baffin Bay and onto the shelf off West
Greenland. Close to the southeastern termination of line
600, the location of the 3201-m-deep Hellefisk-1 well was
crossed. Results from the crossing line 500 can be found in
the work of Suckro et al. [2012]. Figure 2 shows the loca-
tion of line 600 in a geological context. The characterization
of basement types is primarily based on seismic reflection
data, potential field data and plate reconstructions since
seismic refraction data are scarce in this area (see 2.2 Pre-
vious Geophysical Studies).
[13] A total of 25 ocean bottom seismometers (OBS) were
deployed at equal distances along the line (18.4 km spacing).
The instruments were provided by IFM-GEOMAR (Kiel,
Germany) and were equipped with a three-component 4.5-Hz
seismometer and a hydrophone [Gohl et al., 2009]. The
seismic source consisted of 16 G. Guns
TM
ranging in size
from 1.1 to 4 L (70 to 250 in
3
) with a total volume of 50 L
(3100 in
3
) operated at a pressure of 145 bar. In addition, two
Bolt guns with a volume of 32 L (1953 in
3
) each were used at
a pressure of 80 bar (1160 psi). The shot interval was one
minute, which resulted in an average shot spacing of 156 m at
a variable ship speed of 4 to 5 knots. For navigation, the
ships Global Positioning System (GPS) system was used
with an accuracy of 1020 m [Gohl et al., 2009].
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
4of24
[14] The shots were simultaneously recorded by a
3450-m-long digital streamer (SEAL system by Sercel) with
276 channels. The purpose of the coincident seismic
reflection line was to assist the velocity modeling of
the seismic refraction data by providing detailed informa-
tion on the geometry of the basement and sedimentary
layers.
[15] During the shooting of line 600, the vessel encoun-
tered an area with active fishing at the seaward limit of the
Greenland shelf. This happened at night during a heavy
snow storm with a visibility well below 50 m. For this rea-
son, the streamer had to be recovered and the ship detoured
to the north (Figure 3). This resulted in a kink in the line
with shots being offset by up to 23 km from the deployed
positions of OBS 616, 617 and 618. As discussed later, the
resolution of the velocity model is reduced in this zone.
[16] After recovery of the OBS, the data were dumped to
disk, corrected for OBS clock drift, converted to SEG-Y
format, and debiased. Travel time picks of the direct wave
were used to recalculate the position of the instruments at the
seafloor using a least squares method and velocity-depth
profiles for the water column obtained from CTD (conduc-
tivity, temperature, depth) measurements. Shot-receiver
ranges were calculated from the new OBS locations. Record
sections are displayed with a band-pass filter from 8 to
24 Hz (Figures 48). Deconvolution improved the recogni-
tion of seismic phases to some extent but was less successful
in the shallow water where the large Bolt guns produced
significant reverberations. Trace amplitudes in the record
sections are weighted by their distance to the OBS to
increase amplitudes for large offsets.
3.2. Methodology
[17] The goal of the analysis of the seismic refraction data
was to obtain a two-dimensional velocity model for the
sediments, volcanics, crust, and uppermost mantle. Line 600
was shot along a great circle arc with the exception of the
off-line shots due to fishing vessels. This arc forms the
baseline for the model with the origin (x = 0 km) defined by
the northwestern-most shot.
[18] The Pwave velocity model was developed using the
program RAYINVR [Zelt and Smith, 1992; Zelt and
Figure 3. Location map of line 600 (full name: AWI20080600) of the MSM09/3 expedition. The red
lines mark the shot locations along lines 600 and 500. White circles indicate the location of ocean bottom
seismometers (OBS) on line 600, gray circles mark OBS with no data. The OBS station numbers are anno-
tated (red annotations refer to OBS that are shown in Figures 48). Blue lines indicate segments of line
600 along which coincident seismic reflection data were collected [Gohl et al., 2009]. The bathymetry
(GEBCO 08 grid) is plotted with solid lines; the contour interval is 200 m.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
5of24
Forsyth, 1994], which allows both for forward and inverse
modeling. Initially, a forward model was developed from top
to bottom (seafloor to mantle) by fitting the observed travel
times. Layer boundaries within the sediments and the base-
ment geometry were taken from the coincident seismic
reflection record whenever the data quality allowed for this.
However, the low fold of the seismic reflection data, the
location of the seabed multiple, and the gap in the reflection
data due to the deviation around fishing vessels, prevent a
clear definition of the basement along most portions of line
600. After the development of the initial forward model, both
layer boundaries and velocities within layers were optimized
by using the inversion algorithm in RAYINVR. Gravity data
were used to help defining the deep structures at the edges of
the line without ray coverage (see section 4.3).
3.3. Seismic Data
[19] Three of the 25 OBS had technical problems, which
prevented the recording of data (OBS 605, 606 and 621).
Most records exhibit a high signal-to-noise ratio that allow
for the identification of seismic phases down to the mantle
(mantle refraction P
n
) as for example on OBS 609 (Figure 4).
The Moho reflection (P
m
P) can be recognized on most
records, in particular for OBSs on crust with oceanic affinity
(Figure 4) but also in the continental domain (Figure 5). Most
instruments between OBS 604 and 619 show a prominent
refraction that is approximately horizontal when plotted with
Figure 4. Seismic record, raypaths and travel times for OBS 609. (top) Record section of hydrophone
component with calculated travel times (red lines), (middle) calculated travel time (black lines) with
observed travel times (red lines), and (bottom) raypaths through the velocity model (same color scheme
as in Figure 9). The uncertainty of the observed travel times is given by the height of the vertical bars.
Travel times are displayed with a reduction velocity of 6.8 km/s. The horizontal scale is distance along
the velocity model. A red triangle marks the OBS position. For phase names see also Table 1.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
6of24
a reduction velocity of 6.8 km/s (e.g., Figure 4). This indi-
cates the presence of a crustal layer with a velocity around
6.8 km/s, which is very typical for layer 3 in oceanic crust
[White et al., 1992].
[20] Three important record sections are discussed here in
more detail. OBS 625 (Figure 6) at the SE end of line 600
shows a high-amplitude mid-crustal reflection (P
c1
P). Such
mid-crustal reflections are generally not observed in oceanic
crust as there is a rather smooth transition from the upper to
lower crustal velocities [cf. White et al., 1992]. In contrast,
mid-crustal reflections are very typical for continental crust
[Mooney and Brocher, 1987; Funck and Louden, 1999;
Funck et al., 2000b, 2003]. Hence, this reflection is regarded
as an indication for the presence of continental crust at this
part of the line. Further evidence is provided by the observed
velocities (see section 4.1).
[21] A second feature to be discussed for OBS 625
(Figure 6) is the occurrence of a low-velocity zone that is
revealed by a time delay (jump) in the first arrivals some
30 km to the NW of the OBS (at a distance of 417 km). At
this distance, the refraction within the volcanic layer (labeled
P
B
and inferred from correlation with the Hellefisk-1 well)
disappears and some 200 ms later another refraction (P
c1
,
within the upper crust) becomes the new first arrival. Such a
time delay can only be generated by a layer between the
volcanic layer and the upper crust with a velocity lower than
in the volcanic layer. Velocities within such a low-velocity
zone (LVZ) cannot be constrained due to the lack of
refractions (see also section 4.2).
[22] One OBS that was very difficult to model was OBS
620 (Figure 7). This was due to a very complex geometry
close to that OBS with a deep basin that caused problems
with phase identification. A strong reflection at a distance of
Figure 5. Seismic record, raypaths and travel times for OBS 624 (hydrophone component). For details
see caption for Figure 4. For phase names see also Table 1. Abbreviations are HVZ, high-velocity zone;
L2, oceanic layer 2; LVZ, low-velocity zone.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
7of24
330 to 335 km was initially thought to be the Moho reflec-
tion P
m
P. However, later it was determined to be a reflection
from the base of the basin. The most important phase on this
record is the P
2
refraction - a phase that is very difficult to
spot due to its low amplitude. The velocity (5.5 km/s) is
similar to that observed in the upper crust between 317 and
330 km. However, using a velocity of 5.5 km/s for the entire
basin cannot explain the time delay for deep crustal phases
Figure 6. (top) Part of record section of OBS 625 (hydro-
phone component) displayed with a reduction velocity of
5.4 km/s. The horizontal scale is distance along the velocity
model. (middle) The same record shown together with calcu-
lated travel times (red lines) for some selected phases. (bot-
tom) The corresponding velocity model (same color
scheme as in Figure 9) with the raypaths of the selected
phases. The red triangle marks the position of the OBS.
For phase names see also Table 1. Abbreviations are LVZ,
low-velocity zone; MCB mid-crustal boundary.
Figure 7. (top) Part of record section of OBS 620 (hydro-
phone component) displayed with a reduction velocity of
5.9 km/s. The horizontal scale is distance along the velocity
model. Yellow arrows indicate the zone where the weak P
2
phase can be correlated. (middle) The same record shown
together with calculated travel times (red lines) for some
selected phases. (bottom) The corresponding velocity model
(same color scheme as in Figure 9) with the raypaths of the
selected phases. The red triangle marks the position of the
OBS. For phase names see also Table 1.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
8of24
observed on instruments farther to the SE. This is why
velocities <5.5 km/s had to be used for the basin. To match
the observed P
2
phase at OBS 620 (Figure 7), a thin layer
with a velocity of 5.5 km/s was introduced at the top of the
basin. Overall, the raypaths were very sensitive to even
slight changes in the basin like velocities, steepness of the
flanks, and thickness of the top layer. This made it difficult
to obtain an acceptable fit for all the observed seismic
phases.
[23] OBS 604 (Figure 8) is located at the position where
an abrupt lateral change in the crustal velocities had to be
introduced. To the southeast, an oceanic velocity structure is
encountered with the P
2
and P
3
refractions. Crustal refrac-
tions in the northwest (P
c1
and P
c2
) arrive earlier due to the
shallowing of the basement. The phase velocity of the P
c2
refraction is higher than that of the P
3
phase in oceanic layer
3. However, this apparent higher velocity is also related to
the northwestward shallowing of the basement. The actual
modeled velocity in the mid-crustal layer through which the
P
c2
is propagating, is only 6.1 km/s. This compares to
6.8 km/s in oceanic layer 3. The change from oceanic to
continental crust around 60 km is also indicated by the
presence of a mid-crustal reflection (P
c2
P) to the northwest
that is not observed to the southeast.
[24] Off Greenland, the P
c2
Pphase (Figure 5) marks the
top of a high-velocity lower crustal layer. In order to obtain a
visible P
c2
Preflection, velocities have to increase from the
constrained 6.8 km/s at the base of the lower crust to
values >7.0 km/s below. As there are no refractions within
the high-velocity zone, there are no direct constraints on its
Figure 8. Seismic record, raypaths and travel times for OBS 604 (hydrophone component). For details
see caption for Figure 4. For phase names see also Table 1. Abbreviations are L2, oceanic layer 2;
L3, oceanic layer 3; MC, mid-crustal layer; UC, upper crust.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
9of24
velocity. However, the presence of high velocities as
deduced from the P
c2
Preflections is consistent with the
better constrained results from NUGGET line 1 in southern
Davis Strait [Funck et al., 2007].
4. Results
[25] Below, the Pwave velocity model for line 600 is
presented. First, the model is described followed by a
detailed account of the model resolution and uncertainties.
Finally, two-dimensional gravity modeling is employed to
check for consistency of the velocity model with the
observed gravity data.
4.1. Velocity Model
[26] The Pwave velocity model for line 600 is shown in
Figure 9. Although some interpretation of the observed
velocity structure is given here, a more detailed account can
be found in the next section (section 5). The upper part of the
model consists of a series of sediment layers. In the central
part of the line in the deep basin of Baffin Bay, velocities
within the sedimentary layers vary from 1.6 to 4.2 km/s and
the total thickness reaches 6 km. However, the velocities of
1.6 km/s close to the seafloor are not constrained by any
refractions. In the NW beneath the Baffin Island shelf, the
lowermost sediment layer is modeled with a velocity of 4.3
to 4.4 km/s. On the Greenland side velocities are not as high;
sediments above the volcanic basement reach a velocity of
3.8 km/s. However, the maximum sediment thickness on
line 600 is slightly higher off Greenland than off Baffin
Island with values of 6 and 5 km, respectively.
[27] Volcanic rocks were drilled in the lower section of the
Hellefisk-1 well. The velocity model (Figure 9) shows
velocities of 5.2 to 5.7 km/s that are consistent with a vol-
canic layer with a maximum layer thickness of 4 km. This
thickness is quite variable as the volcanic sequence consists
of two mounded structures. The minimum thickness is as
little as 1 km. To the northwest, the layer pinches out at
365 km. Between 333 and 357 km another layer appears
with velocities of 4.7 to 4.9 km/s (poorly constrained) and a
thickness of 5 km. This layer forms a basin. At the SE end of
the line, a low-velocity zone (LVZ) is observed beneath the
basalts. The modeled velocity is 4.1 km/s within the LVZ,
but it should be emphasized here that the velocity within the
LVZ is not constrained due to the absence of refractions;
even the depth of the top of the LVZ is undefined. However,
the seismic reflection data indicate the disappearance of
high-amplitude reflections at that depth level. The velocity
in this layer is <5.6 km/s but the composition remains
unclear (basalts or non-volcanic sediments). Other, local
LVZs are observed within the sedimentary column.
[28] The crust shows large lateral variations both in
velocity and thickness (Figure 9). Between 0 and 60 km, the
crust is modeled with three layers with velocities of 5.5
5.9 km/s in the upper crust, 6.1 to 6.3 km/s at mid-crustal
levels and 6.6 to 6.9 km/s within the lower crust. This
velocity distribution and the occurrence of mid-crustal
reflections are consistent with continental crust [Mooney and
Brocher, 1987]. Rough basement topography with three
basement highs and a seaward shallowing of the Moho
indicates that the crust has been faulted and thinned. The
Moho depth is only constrained between 40 and 60 km
where it shallows from 15 to 13 km. The landward deepen-
ing of the Moho between 0 and 40 km is inferred from the
gravity modeling (see section 4.3.).
[29] As discussed in section 3.3, a rather abrupt velocity
change occurs at 60 km. Between 60 and 333 km, the crust
is divided into an upper and lower crustal layer interpreted
as oceanic layers 2 and 3 (Figure 9). Velocities within layer
3 are fairly constant and around 6.8 km/s at the top and 7.0
to 7.2 km/s at the base as is typical for oceanic crust [White
et al., 1992]. Layer 2 velocities can be divided laterally into
three segments: between 60 and 135 km velocities range
from 5.5 to 6.0 km/s, between 135 and 227 km there is an
increase to 6.1 to 6.5 km/s, and between 227 and 333 km
velocities decrease again to 5.5 to 6.0 km/s. The thickness
of layer 2 is generally around 2 km, but locally exceeds
3 km. The total oceanic crustal thickness varies from 5 to
9 km and the Moho discontinuity is located at depths
between 13 and 16 km. The previously discussed basin
between 333 and 357 km is partly covered by a thin layer
with a velocity of 5.5 km/s, which is similar to the adjacent
oceanic layer 2. The exact thickness of this layer, which is
assumed to be basaltic, is not determined seismically. The
model shows a termination of this layer at 352 km as there
was no seismic evidence for a continuation to the SE.
[30] Crustal velocities beneath the basin (333 to 357 km)
are not constrained seismically. However, reflections from
the base of the basin and from the Moho give a thickness
estimate of 4 km for the crust in this zone (Figure 9). Farther
to the SE, the crust thickens substantially; the Moho deepens
from 15 km beneath the basin to 26 km between 380 and
400 km. In this zone, two crustal layers can be distinguished
with velocities of 6.3 to 6.6 km/s and 6.8 to 7.3 km/s in the
upper and lower layer, respectively. This crust is interpreted
as thick igneous crust. Farther to the SE toward Greenland,
crustal velocities decrease and this, together with the obser-
vations of mid-crustal reflections, suggest a continental
character of the crust. In the upper continental crust, veloc-
ities range from 5.8 to 6.2 km/s while the lower crust exhi-
bits velocities from 6.5 to 6.8 km/s. The upper crust is
between 3 and 4 km thick, while the lower crustal thickness
decreases from 10 km in the NW to 6 km in the SE. Beneath
the continental crust, another layer was introduced into the
model to explain a set of reflections (P
c2
P). These phases
originate from a depth level that is shallower than the Moho
in the zone that is adjacent to the thick igneous crust. This
observation was accommodated by continuing the high
velocities of the igneous crust beneath the continental crust
and is interpreted as magmatic underplating. Velocities in
this high-velocity zone (HVZ) are not constrained due the
lack of refractions from this layer. However, the velocity
change in the model from 6.8 km/s at the base of the lower
continental crust to 7.1 km/s at the top of the HVZ yields in
an impedance contrast that is sufficient to create detectable
reflections. Funck et al. [2007] found a similar and well-
resolved underplated layer in the southern Davis Strait,
where the velocity was 7.4 km/s. This gives an idea of the
possible range of velocities in the HVZ. Mantle velocities
within the oceanic domain were modeled between 7.8 and
7.9 km/s. Beneath the continental crust off Baffin Island,
mantle velocities of 8.0 km/s were used in the model.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
10 of 24
Figure 9. (top) Magnetic anomalies, (middle) Pwave velocities with interpretation, and (bottom) diag-
onal values of the resolution matrix of the Pwave velocity model along line 600. Data source for the mag-
netic data: Verhoef et al. [1996]. The outer parts of the velocity model and of the resolution matrix with no
ray coverage are shown in white, velocities are specified in km/s, the velocity contour interval is 0.1 km/s.
Red circles show the location of OBSs that were used for the modeling, the black circles refer to non-
functional OBSs. The location of the Hellefisk-1 well and the cross-point with refraction line 500 are indi-
cated by a red and blue line, respectively. Abbreviations are BAS, basalts/volcanic rocks; FZ, fault zone;
MCB, mid-crustal boundary; L2, oceanic layer 2; LC, lower crust; LVZ, low-velocity zone; MC, mid-
crustal layer; MCB, mid-crustal boundary; UC, upper crust.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
11 of 24
4.2. Model Resolution and Uncertainty
[31] To assess the resolution and uncertainty of the
velocity model, a formal error analysis for individual phases
is given next to plots showing the ray coverage. In addition,
the values of the diagonal of the resolution matrix for the
velocity nodes are calculated. The resolution matrix is a
good indicator that distinguishes between poor and well
resolved parts of a model. Ideally, values of the resolution
matrix are 1 but values >0.5 indicate reasonably well
resolved model parameters [Lutter and Nowack, 1990].
[32] The formal error analysis for individual phases on line
600 is summarized in Table 1. The normalized c
2
is based
on assigned pick uncertainties of 40200 ms depending on
the quality of each individual travel time pick. Pick uncer-
tainties are graphically indicated in Figure 10. The model is
generally well constrained with a total RMS misfit of 80 ms
between calculated and picked travel times. This is close to
the average travel time uncertainty of 83 ms. The normalized
c
2
of 0.68 is below the optimum value of 1 when travel
times are fitted within the given pick uncertainty. This could
indicate that either the model was fit to noise in the data or
the pick uncertainties were overestimated.
[33] The diagonal values of the resolution matrix of the
velocity nodes are shown in a gridded version in Figure 9
and show a close correlation with the ray coverage
(Figure 10). That is, areas with good and reversed ray cov-
erage are characterized by high resolution values. Within the
sedimentary column, velocities are generally well resolved,
although there are some exceptions. In the area where the
ship deviated from the track line (ca. 265 to 335 km), there
are zones with reduced resolution that are caused by lack of
observations (absence of close offset shots) or by complex-
ities associated with the three-dimensionality introduced by
the off-line shots. During the modeling, large lateral varia-
tions were avoided within that area. In this zone, the
observed seismic phases are generally consistent with the
velocities in the adjacent areas with online shots. However,
an offset in time is often noticed, which indicates a change in
the sedimentary thickness perpendicular to the line. In these
cases, the offset was accepted because no consistent two-
dimensional model could be found that would explain all the
observations within this zone. Thus, the velocity model is
representative for the location of the OBS rather than for the
off-line shot locations. Phases that displayed these problems
were not included in the error analysis. Another zone with
poor resolution is within the sediments on the Baffin Island
shelf between 0 and 30 km, where low-velocity zones occur
and where not all raypaths are reversed.
[34] The low-velocity zone beneath the basalts at the
southeastern end of the profile is not well resolved. How-
ever, this is the nature of low-velocity zones in which no
refractions can be observed which would allow for a direct
determination of the velocity. The existence of the low-
velocity zone is well constrained as seen by the characteristic
delay of refractions (Figure 6). Similar, the low resolution in
the basin between 333 and 357 km is related to the com-
plexities in the velocity structure there. The data cannot be
explained without a basin at that location that is covered by a
thin layer with a higher velocity than in the underlying
sediments. The reduced resolution reflects the uncertainty of
the thickness of the overlying layer and the velocity within
the basin.
[35] Velocities within the oceanic crust are well resolved
(Figure 9), but as before, the resolution decreases in the
region where the ship deviated from the line during the
shooting. Nevertheless, some P
3
rays sampled the lower
crust in this region and are consistent with layer 3 velocities
farther to the NW. Also the few P
2
observations are con-
sistent with layer 2 velocities outside this zone. However, no
P
n
or P
m
Pphases are observed between 280 and 335 km,
which could determine the Moho depth. As mentioned
above, velocities within low-velocities zones cannot be
resolved and a resolution value of 0.3 in the basin around
340 km indicates uncertainty in the velocities there. Reduced
resolution within the crust close to the outer limit of the
model coincides with a decrease of the number of reversed
raypaths in these areas. The lower crust around 390 km has
well resolved velocities at the top (resolution values of 0.7
0.8), but at the bottom and within the adjacent HVZ,
velocities are only poorly resolved. Mantle velocities display
resolution values up to 0.9 in areas where P
n
phases are
observed. The plot with the ray coverage (Figure 10) also
indicates the segments of layer boundaries that are con-
strained by wide-angle reflections.
[36] Absolute errors of the model can be estimated by
perturbation of single nodes (velocity and boundary nodes)
in the model and examination of the sensitivity of the travel
times to these perturbations. Such a test was performed
within the basin between 333 to 357 km (Figure 9). It was
found that velocities that deviate 0.4 km/s from the mod-
eled 4.7 to 4.9 km/s can still explain most of the travel time
observations, although it becomes increasingly difficult to fit
all observations once the velocities are varied by more
than 0.2 km/s.
[37] Line 600 offers another opportunity to check the
accuracy of the velocity model since the line crosses
the Hellefisk-1 well. Figure 11 shows a comparison of
the velocity model with the sonic log from the well
(obtained from GEUS data repository). To reduce the scatter
in the sonic log data, a median filter with a filter length of
20 m was used. It should also be noted that the log data
were not used to adjust the velocity model; only the wide-
angle reflections and refractions as well as the coincident
seismic reflection data were used for the modeling. There is
Table 1. Number of Observations, n, RMS Misfit Between
Calculated and Picked Travel Times, t
rms
, and Normalized c
2
for Individual Phases on Line 600
Phase nt
rms
(s) c
2
Direct wave 1735 0.026 0.379
P
S
(all sediment refractions) 3129 0.045 0.354
P
S
P(all sediment reflections) 1337 0.096 1.683
P
B
(refraction in basalt) 585 0.048 0.441
P
B
P(reflection base basalt) 45 0.046 0.214
P
2
(refraction in oceanic layer 2) 1532 0.132 0.913
P
3
(refraction in oceanic layer 3) 3518 0.068 0.575
P
c1
(refraction in upper crust) 747 0.066 0.724
P
c1
P(reflection base upper crust) 102 0.045 0.372
P
c2
(refraction in mid-crustal layer) 882 0.090 0.859
P
c2
P(reflection base mid-crustal layer) 255 0.094 0.521
P
c3
(refraction in lower crust) 30 0.035 0.197
P
m
P(Moho reflection) 1670 0.077 0.619
P
n
(refraction in mantle) 2177 0.113 0.876
All phases 17744 0.080 0.683
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
12 of 24
Figure 10
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
13 of 24
a general agreement between the two velocity curves but
there are also deviations. For example, the velocity model
has different velocity gradients within layers, since the
refractions primarily constrain velocities at the top of each
layer. The layer at a depth of 1500 m and a sonic velocity of
4.5 km/s is completely missing in the velocity model
because no refractions could be detected in the seismic data
and, hence, such a layer is not required by the data. By
missing this layer and having too low a velocity gradient in
the underlying layer, the average velocity down to basement
was underestimated. Thus, the top of the volcanic unit
(velocities >5 km/s) is modeled 400 m too deep. However,
in two-way travel time, the velocity model still fits the
travel time to the top of the volcanic layer as can be seen
when compared with the coincident seismic reflection data
(Figure 12). Velocities within the basalts compare fairly
well (around 5.5 km/s). In summary, the comparison shows
that the refraction data are unable to resolve the detailed
velocity structure within the sediments but still match the
long-wavelength features. It can also be seen that the
absolute depth error of layer boundaries at the basement
level is on the order of 400 to 500 m.
4.3. Gravity Modeling
[38] Gravity modeling can provide additional constraints
on velocity models, in particular in areas where the
Figure 10. Ray coverage along line 600 obtained from point to point(shot to OBS) ray-tracing for all travel time picks
used for the modeling (only phases below basement are shown). The figure is divided into four panels for different depth
levels. The lower part of each panel shows the layer boundaries of the velocity model together with the raypaths; the upper
part visualizes the calculated travel times (solid lines) and the associated travel time picks (vertical bars with a height
corresponding to the estimated pick uncertainty). Not every travel time pick and ray is plotted, a distance of 2 km between
picks is used to keep the plot readable. Triangles mark the positions of the OBSs. For phase names see also Table 1. Abbre-
viations are HVZ, high-velocity zone; L2, oceanic layer 2; L3, oceanic layer 3; LC, lower crust; MC, mid-crustal layer; UC,
upper crust.
Figure 11. Comparison of the sonic log velocities from the
well Hellefisk-1 (GEUS database) with the velocities
obtained from the modeling of the seismic refraction data
along line 600.
Figure 12. Part of seismic reflection record on line
BGR08315 around the Hellefisk-1 well. The well location
is marked by a blue line. (top) Migrated seismic record,
and (bottom) the same record shown together with velocity
model (Figure 9) converted to two-way travel time. Trian-
gles indicated the location of OBS 624 and 625. Abbrevia-
tion LVZ is low-velocity zone.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
14 of 24
Figure 13. Gravity modeling along line 600. (top) Models A through C explore the Moho geometry at
the outer limits of the model with no constraints from the seismic refraction data (compare with ray-
coverage in Figure 10). (bottom) Models A, D and E test for lateral variations of the density in the mantle
and within the basin around 350 km. The observed gravity is the free-air gravity measured during the
MSM09/3 expedition. The shaded area indicates the range of gravity values in a 20-km-wide zone to
either side of line 600 (extracted from DTU10 1-min grid [Andersen, 2010]). All densities in the model
are given in kg m
3
. For details see text. Abbreviations are HVZ, high-velocity zone; L2, oceanic layer 2;
LC, lower crust; MC, mid-crustal layer; UC, upper crust.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
15 of 24
Figure 14. Free-air gravity map with interpretation of crustal character along line 600 and along other
available seismic refraction lines. Shading by artificial illumination from the east. Data source: Satellite
altimetry, DTU10 1-min grid [Andersen, 2010]. Blue lines refer to the MSM09/3 seismic refraction lines
(AWI20080[500/600/700]) [Gohl et al., 2009]; pink lines show the location of other seismic refraction
experiments: GR89-WA [Gohl and Smithson, 1993], NUGGET 1 [Funck et al., 2007], and NUGGET 2
[Gerlings et al., 2009]; dashed pink lines indicate sonobuoy experiments: C1 to C10, B1 and D1 [Keen
and Barrett, 1972], D3 [Srivastava et al., 1982]. The distribution of crustal domains and structural ele-
ments is indicated by black and gray lines. White circles mark the location of OBS on line 600. Yellow
circles indicate the positions of wells (for well names compare with Figure 2). Abbreviation UFZ is
Ungava fault zone.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
16 of 24
velocities are poorly resolved. In addition, gravity modeling
can be used to verify how consistent the velocity model is
with the gravity data. Gravity values are derived from the
KSS31M sea gravimeter that was installed on the vessel
Maria S. Merian. In this study, two-dimensional gravity
modeling was performed along line 600 using the algorithm
of Talwani et al. [1959]). The observed gravity in the
models (Figure 13) are free-air gravity anomalies measured
along the main line, i.e., not along the local deviation during
the shooting. Densities were obtained from conversion of the
Pwave velocities (Figure 9) using the curve shown by
Ludwig et al. [1970], which is approximated by
r¼2:83v4þ70:4v3598v2þ2230v700
where ris the density in kg/m
3
and vis the Pwave velocity
in km/s. Densities of 1.03 and 3.30 g/cm
3
were used for the
seawater and the mantle, respectively. There is no fixed ref-
erence gravity value. Instead, a reference is calculated from
minimizing the misfit between observed and calculated grav-
ity. Some caution has to be applied in the two-dimensional
modeling approach as some gravity anomalies terminate very
close to line 600 (Figure 14). Figure 13 illustrates that the
gravity variations can be up to 70 mgal in a 20-km-wide
zone to either side of line 600. Hence, the modeling and
discussion focuses on general trends in the gravity data.
[39] In the starting model (Model A; Figure 13, top), the
Moho is continued horizontally to the outer boundaries
where there is no ray coverage. At the NW end of the line
(between 0 and 40 km) the Moho depth was kept at 15 km.
This depth is not compatible with the gravity data as the
observed and calculated gravity deviate by up to 190 mgal.
To compensate for that, the Moho depth was increased to
30 km at that end of the line (model B) and this resulted in a
greatly reduced misfit. Full thickness continental crust
beneath Baffin Island is likely to be >30 km if one uses the
measurements in the nearby Precambrian crust in northern
Labrador as an indicator for the thickness [Funck and
Louden, 1998, 1999; Funck et al., 2000a]. Receiver func-
tions on southern Baffin Island also indicate a crustal
thickness of >40 km [Darbyshire, 2003]. However, the NW
extension of line 600 runs for another 100 km on the shelf
(Figure 2) where the crustal thickness is less than onshore.
[40] At the SE end of the line, a decrease of Moho depth
by 4 km (model C) can reduce the misfit, although the cal-
culated gravity is still too low. Further shallowing of the
Moho in that area, however, started to affect the overall
shape of the calculated gravity and was therefore not pur-
sued. Instead, the mantle density was increased from 3300 to
3350 kg m
3
at the SE end of the line (model D; Figure 13,
bottom), assuming that the lithosphere beneath the conti-
nental crust might be colder and denser than in the oceanic
domain. However, this assumption is questionable as this
zone is affected by the mantle plume as indicated by the
HVZ beneath the continental crust that is interpreted as
magmatic underplating. Decreasing the mantle density
beneath the oceanic crust has a similar effect. This could be
explained by high degrees of melting that strip out the
magnesium and iron, which leaves a low-density restite
behind. In any case, a lateral density change in the mantle
greatly reduces the misfit between observed and calculated
gravity in the southeast. Alternatively, a density increase in
the HVZ can produce a similar effect as a lateral density
variation within the mantle. When the density of the conti-
nental mantle beneath Baffin Island is increased to
3350 kg m
3
as well, the Moho there has to be adjusted to a
depth of 31 to 32 km to keep the fit with the observed
gravity. Between 70 and 110 km, the pressure at the base of
the model decreases by 10 MPa. This indicates some iso-
static imbalance that is evidenced by the flat Moho in this
zone while the basement shallows toward Baffin Island. This
imbalance can be compensated for by a gradual landward
increase of the density in the mantle. Such a lateral density
increase would be compatible with a cooling of the litho-
sphere from the oceanic crust into the region that is inter-
preted as transform fault [Chalmers and Pulvertaft, 2001].
At the same time, such a density increase would reduce the
misfit between observed and calculated gravity in this
region.
[41] In a last modeling step, the misfit in the area of the
basin at 340 km was investigated. As mentioned above,
velocities within the basin are not well resolved and the
velocity uncertainty may be up to 0.2 to 0.4 km/s. To obtain
a good fit, the density had to be reduced from 2510 to
2310 kg m
3
(model E; Figure 13, bottom). Although this
decreases the misfit in the area of the basin, it seems unlikely
that the density within the basin is that low. A reduction by
200 kg m
3
compared to the empirical formula by Ludwig
et al. [1970] corresponds to a velocity decrease from 4.8
to 3.1 km/s. This is far too low to explain the observed
travel times in the seismic refraction data. In addition, both
density and velocity would be lower than in the overlying
sedimentary layers. Alternatively, the sedimentary basin
could be deeper than the model indicates, as the seismic
constraints on this are not very strong. However, the crust
underneath the basin is already rather thin (4 km).
[42] In summary, even with some density adjustments,
there is no perfect fit to the observed gravity data. Many of
the deviations are likely due to three-dimensional effects that
cannot be treated properly in a two-dimensional model. One
example is the 4-km-deep sedimentary basin around 20 km
(Figure 13) that is required to model the observed delays of
crustal phases in that zone. The basin has a pronounced
signature in the calculated gravity, but cannot be seen in the
observed gravity. This indicates that the basin has to be a
local feature that violates the assumption of two-dimensionality
in the gravity model. Figure 14 shows a free-air gravity map
of the study area. The map indicates that most features cross
line 600 at an oblique angle. The most prominent departure
from two-dimensionality occurs around the pronounced
gravity high around OBS 621. This high can be correlated
southward all the way through the Davis Strait but dis-
appears just to the north of the line.
5. Discussion and Interpretation
[43] In section 4.1, the model for line 600 was presented
with a description of the velocity distribution in the sedi-
ments, the crust, and the mantle. In addition, a first inter-
pretation of the crustal structure was given. However, a more
thorough discussion of the model is necessary and is pre-
sented in this section. The model is discussed in context of
the regional geology as well as compared to studies in sim-
ilar tectonic and geologic settings. The discussion starts in
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
17 of 24
the NW of the profile off Baffin Island and then moves
toward the SE following the segmentation of the crust indi-
cated in Figure 9.
5.1. Continental Crust (060 km)
[44] The segment between 0 and 60 km is interpreted as
continental crust (Figure 9). The velocity structure is divided
into three layers with velocities that are not typical of oce-
anic crust [White et al., 1992]. In addition, the occurrence of
mid-crustal reflections is a strong indicator for continental
crust [Mooney and Brocher, 1987]. Velocities of 5.5 to 6.3
and 6.6 to 6.9 km/s for the upper and lower crust, respec-
tively, compare well with the range of velocities found
elsewhere in West Greenland and Eastern Canada along the
shores of the Labrador Sea and the Baffin Bay. For this
comparison, the mid-crustal layer is considered as part of the
upper crust. The Archean Nain Province in Labrador, where
Funck and Louden [1998] report velocities of 5.8 to 6.5 and
6.6 to 6.9 km/s in the upper and lower crust, respectively, is
very close to the study area. In the Nain Province, the
uppermost crust was interpreted to have a gneissic compo-
sition based on Pwave velocities and a Poissons ratio of
0.20 to 0.24. To the SE of Baffin Island along NUGGET line
1, upper and lower crustal velocities of 5.8 to 6.1 and 6.4 to
6.6 km/s were reported [Funck et al., 2007]. However, upper
crustal velocities there are only poorly resolved. Farther to
the east on that line (off Greenland), upper crustal velocities
were as low as 5.4 km/s. This is very close to the observed
5.5 km/s on line 600. Funck et al. [2007] suggested a
granitic composition of the upper crust on NUGGET line 1.
Intense faulting on line 600 as indicated by the rough base-
ment topography between 0 and 60 km, may result in a
reduction of the Pwave velocity. Laboratory measurements
of rock samples measured at pressures compatible with mid-
crustal levels (300500 MPa) indicate Pwave velocities of
6.07 0.22 km/s for granite [Holbrook et al., 1992], which
is higher than what is observed on line 600. However, on a
seismic refraction line off SW Greenland, Chian et al.
[1995b] reported upper crustal velocities of 5.4 km/s in an
area that is characterized by widespread granitic intrusions.
[45] One other interesting aspect of the continental crust at
the NW end of line 600 is the abrupt transition to oceanic
crust at around 60 km (Figure 9), without any indication for
a transitional crust. Transitional crust at non-volcanic con-
tinental margins is often characterized by velocities that
differ from the adjacent normal continental and oceanic crust
and that are often difficult to interpret. The lack of such a
transition zone is interpreted to be related to the location of
line 600, which is parallel to the assumed extinct spreading
axis [Chalmers and Pulvertaft, 2001; Chalmers and Oakey,
2007] (Figure 2). The boundary from continental to oce-
anic crust crosses a N-S striking gravity low (Figure 14) that
is interpreted as transform margin [Chalmers and Pulvertaft,
2001]. Crustal thinning across transform margins is charac-
terized by a thinning of the crust over short distance. One
well-studied transform margin is the margin off Ghana in the
Equatorial Atlantic. Edwards et al. [1997] show that the
continental crust thins from 23 to 4 km over a distance of
only 20 km when initial oceanic crust is encountered. Such a
thinning is comparable to the one observed on line 600,
although it is only constrained by gravity modeling. Another
example for rapid thinning of the continental crust with no
transition zone can be found at the transform margin south of
Svalbard [Breivik et al., 2003] or off South Africa where the
transition occurs over a 50-km-wide zone [Parsiegla et al.,
2009].
[46] The thickness of the continental crust at the transition
to the oceanic crust is 7 km (Figure 9). This value is com-
patible with the 8 km found at the seaward limit of the Baffin
Island crust on NUGGET line 1 [Funck et al., 2007]. There
are no seismic refraction data on Baffin Island to determine
the thickness of the unstretched crust. However, receiver
functions indicate a crustal thickness between 40 and 45 km
beneath southern Baffin Island [Darbyshire, 2003]. This
would indicate a stretching factor of six.
5.2. Oceanic Crust (60333 km)
[47] The zone between 60 and 333 km is interpreted as
oceanic crust but displays both lateral thickness and velocity
variations. Velocity variations are restricted to the upper
crustal layer (oceanic layer 2). Figure 15 shows a compari-
son with average oceanic crust in the North Atlantic with a
crustal age between 59 and 127 Ma as compiled by White
et al. [1992]. The comparison shows that the layer 2
velocities of 6.1 to 6.5 km/s between 135 and 227 km are
slightly higher than average. Otherwise the velocities fall
well into the range of typical oceanic crust both in terms of
the absolute velocities and velocity gradients. With respect to
crustal thickness, the values on line 600 are typically 6 km or
9 km, which are either slightly lower or slightly higher when
Figure 15. Comparison of velocity-depth functions from
line 600 with average oceanic crust. The gray area shows
the range of typical oceanic crust encountered in the Atlantic
Ocean with a crustal age between 59 and 127 Ma [White
et al., 1992]. The color lines show the velocities from line
600; the position (km) refers to the distance along the
velocity model (Figure 9).
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
18 of 24
compared with average crust. White et al. [1992] found a
global average thickness of 7.1 0.8 km.
[48] Figure 9 shows that the lateral velocity variations in
layer 2 correlate with changes in the observed magnetic
anomalies. The zone with the higher velocities (6.1
6.5 km/s) displays positive anomalies, while the lower
velocities (5.56.0 km/s) are generally associated with neg-
ative magnetic anomalies. Changes of magnetic polarity in
oceanic crust mirror the reversals of the Earths magnetic
field over time and, hence, the lateral changes observed
along line 600 could indicate that the segments of oceanic
crust have a different age. However, there is currently no
consensus that there are seafloor-spreading magnetic
anomalies in Baffin Bay (Figure 16). Nevertheless, the
positive magnetic anomalies between OBS 609 and 612
show up as a linear feature that is parallel with the spreading
axis, which is itself parallel to line 600. The location of the
spreading axis is assumed to coincide with a gravity low
[Whittaker et al., 1997] (Figure 14), which is seismically
confirmed by Suckro et al. [2012]. If the magnetic anomaly
between OBS 609 and 612 is indeed related to seafloor-
spreading, it may correspond to polarity chron 21n (45.3 to
47.2 Ma) as this was the last clear spreading anomaly in the
Labrador Sea [Roest and Srivastava, 1989] and has a similar
distance to the extinct spreading axis. The southwestern and
northeastern limitation of the magnetic anomaly fits with the
location of a transform fault and an assumed fracture zone
(cf. Figure 14).
[49] On the free-air gravity map (Figure 14), the extinct
spreading axis in Baffin Bay stands out as a gravity low
parallel to line 600. In addition, another gravity low can be
seen that continues northward from the extinct spreading
axis, roughly between OBS 604 and 608. This gravity low
correlates with a proposed transform fault along which the
spreading axis is displaced to the north [Chalmers and
Oakey, 2007] (Figure 2). It is in this zone that the oceanic
crustal thickness is reduced to 6 km. A second zone with
reduced crustal thickness is observed near OBS 612
(Figure 9). Northward of this zone, the spreading axis shows
a minor displacement (Figure 14) interpreted as the location
of a fracture zone [Chalmers and Oakey, 2007] (Figure 2).
There is no clear signal in the gravity that would connect the
displacement in the spreading axis with the zone of reduced
crustal thickness on line 600. However, such a trend would
be consistent with the northward orientation of the transform
fault in the west and the UFZ in the east. Variations in the
thickness of oceanic crust can have various reasons. In the
study region, thickness variations may for example relate to
the variable nature of volcanic activity in the Davis Strait
region. The crust in transform faults and fracture zones is
often anomalously thin and the observed velocities can
deviate from normal oceanic crust [Detrick et al., 1993].
Detrick et al. [1993] show that the crust in North Atlantic
fracture zones is frequently quite thin (<12 km thick) and is
characterized by low Pwave velocities and the absence of a
normal seismic layer 3. These anomalies are explained by a
reduced magma supply at the end of spreading segments. In
contrast, the velocity structure within the transform fault and
the assumed fracture zone on line 600 is much closer to
normal oceanic crust even though it is thinner than the local
average. This probably relates to the influence of the Iceland
plume that provided ample magma supply in southern Baffin
Bay. This extra magma supply generated the 9-km-thick
oceanic crust along the line, some 2 km thicker than average
oceanic crust. Although the thickness decreases by 3 km
within the transform fault and the assumed fracture zone, the
resulting thickness of 6 km is still close to normal oceanic
crust.
[50]Storey et al. [1998] identified two main pulses of
volcanism in West Greenland: one between 60.7 and
59.4 Ma and one between 54.8 and 53.6 Ma. On line 600,
the distance to the extinct spreading axis is only 40 km and
an oceanic crustal thickness of 9 km is observed, indicating
that the plume must have influenced the region well after
53.6 Ma to create the thicker than normal oceanic crust.
While the age of the oceanic crust along line 600 is not
known, some estimates can be made. Srivastava and Keen
[1995], in an interpretation of a seismic reflection line
across the extinct spreading axis in the Labrador Sea, assume
a half-spreading rate of 3 mm/yr between magnetic anoma-
lies 18 and 13. At anomaly 13 (~33 Ma), Greenland and
North America are thought to have moved as a single plate.
Under the assumption that these parameters are applicable to
the Baffin Bay as well, a 40 km distance to the axis at a half-
spreading rate of 3 mm/yr would translate to 13 million
years prior to the cessation of spreading. This yields an age
estimate of 46 Ma for the oceanic crust on line 600, some 7
millions of years after the last main pulse of volcanism
sampled onshore West Greenland.
[51] The area between 285 and 335 km is seismically
poorly resolved due to the deviation from the line. However,
the most convincing fit to the observed travel times was
achieved with a velocity structure that is compatible with
oceanic crust. This zone is within the area that Gregersen
and Bidstrup [2008] define as the Aasiaat Basin and
Aasiaat Structural Trend. Some deep reflections in this zone
are interpreted to have an Early to mid-Cretaceous age
[Gregersen and Bidstrup, 2008] but lack a direct well con-
trol. With that age, the sediments would post-date the initi-
ation of seafloor spreading and the underlying basement
should be of continental character. This would contradict our
interpretation as oceanic crust. In contrast, G. Oakey and
J. Chalmers (A new model for the Paleogene motion of
Greenland relative to North America: Plate reconstructions
of the Davis Strait and Nares Strait regions between
Canada and Greenland, submitted to Journal of Geophysi-
cal Research, 2012) infer from plate reconstructions that
the Aasiaat Basin is most likely underlain by Paleocene
oceanic crust.
5.3. Transition Zone (333400 km)
[52] The region between 333 and 400 km is named tran-
sition zone because its seismic character is different from
both the adjacent 9-km-thick oceanic crust to the NW and
the continental crust to the SE (Figure 9). The transition zone
is divided into two very distinct segments. The area that is
labeled leaky transform fault is characterized by 20-km-thick
igneous crust, while the basin to the NW is underlain by only
4-km-thick crust of unknown nature.
5.3.1. Leaky Transform Fault
[53] The 20-km-thick igneous crust between 375 and
400 km lies within the area referred to as Ungava fault zone
(UFZ) a transform fault system that links the spreading axis
in the northern Labrador Sea to that in the southern Baffin
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
19 of 24
Figure 16. Magnetic anomaly map. Shading by artificial illumination from the southeast. Data source:
Verhoef et al. [1996] and Oakey [2001]. Blue lines refer to the MSM09/3 seismic refraction lines
(AWI20080[500/600/700]) [Gohl et al., 2009]; thin solid lines show the location of other seismic refrac-
tion experiments: GR89-WA [Gohl and Smithson, 1993], NUGGET 1 [Funck et al., 2007], and NUGGET
2[Gerlings et al., 2009]; dashed lines indicate sonobuoy experiments: C1 to C10, B1 and D1 [Keen and
Barrett, 1972], D3 [Srivastava et al., 1982]. The bold dashed lines indicate the location of the extinct
spreading axis (taken from the gravity low shown in Figure 14). White circles mark the location of
OBS on line 600 (every 5th station number is annotated). Yellow circles indicate the positions of wells.
The outline of the leaky transform fault is taken from the gravity data (Figure 14).
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
20 of 24
Bay. Storey et al. [1998] suggested that the Davis Strait went
through transtensional phases that created gaps between the
Canadian and Greenlandic continental crust. These gaps
were then filled with melt that formed thick igneous crust. In
this sense the UFZ acted as a leaky transform fault. Storey
et al. [1998] link the Early Eocene volcanism in West
Greenland to net extension on the UFZ, caused by a change
in plate kinematics as the North Atlantic opened at
55 Ma. Funck et al. [2007] identified 16-km-thick igne-
ous crust within the UFZ in southern Davis Strait on
NUGGET Line 1. They favor a Paleocene age of the UFZ
based on radiometric dates of the basalts (59.5 Ma)
[Williamson et al., 2001] in the nearby Gjoa G-37 well.
[54] The igneous crust within the UFZ correlates with a
gravity high that can be correlated through the entire Davis
Strait (Figures 14). With the addition of this study, the oce-
anic or igneous character of this gravity high is confirmed at
both ends (line 600 and NUGGET line 1) and there can be
little doubt that this feature is composed of igneous crust
along its entire length. In addition, the UFZ stands out as a
prominent feature on the magnetic map (Figure 16),
although the correlation is not as good as in the gravity data.
In particular, the northern part of the UFZ appears less
continuous in the magnetic data.
[55] Comparison of the velocity models of the UFZ in the
north (line 600) and in the south (NUGGET line 1) indicates
some lateral changes (Figure 17). In general, the crustal
velocities on these two lines are very similar and any dif-
ferences are close to the limit of seismic resolution. More
significant seems to be the division into a distinct HVZ on
NUGGET line 1 that is overlain by two crustal layers that
are similar to normal oceanic crust with a total thickness of
8 km. The distinct HVZ could indicate a multiphase devel-
opment. First, transtension or extension between Greenland
and Canada produced a gap that was filled with new igneous
crust with a velocity structure similar to normal oceanic
crust. Afterwards, this new crust, as well as the adjacent
thinned continental crust, was magmatically underplated,
creating the HVZ. This magmatic underplating is consistent
with geodynamic models [e.g., Nielsen et al., 2002] that
suggest southward flow of plume material through Davis
Strait along lithospheric thin spots. In the north, no such
division into an upper and lower unit could be detected.
However, the zone where thick igneous crust is observed is
much narrower on line 600 than on NUGGET line 1 (30 km
versus 60 km width). The detection of a potentially weak
reflection within the igneous crust from the top of a HVZ
would be significantly more difficult in such a narrow zone.
The thickness decrease of the igneous crust from north to
south seems to be compatible with the previously mentioned
southward channelling of plume material through Davis
Strait. Including the overlying basalts, the igneous thickness
is 23 km in the north compared to 17 km in the south. The
volcanism also extends into the Baffin Bay, where wide-
spread Paleogene basalts are observed off Disko and Svar-
tenhuk (Figure 2). A volcanic margin extends as far north as
the northern end of line 500 where there is indication for a
high-velocity lower crustal layer [Suckro et al., 2012].
5.3.2. Basin/Transform Fault
[56] Between 333 and 357 km, the velocity model
(Figure 9) indicates the presence of a 5-km-deep basin. As
discussed earlier, velocities within the basin are poorly
resolved but are probably within 0.2 km/s of the modeled
4.7 to 4.9 km/s. The crust immediately below the basin was
not sampled by refracted waves and, hence, the velocities are
unknown.
[57] There are two possible explanations for the basin.
Either this segment is a thin sliver of stranded continental
crust left between the leaky transform fault and the oceanic
crust or it is a transform fault, possibly associated with the
UFZ. If the basin is located on continental crust, the infill
could contain Cretaceous or older sediments. However, as
the igneous crust of the leaky transform fault can be corre-
lated directly to the SE flank of the basin, it appears more
likely that the bulk of the infill consists of volcanic material,
possibly interbedded with some non-volcanic sediments.
The velocities of 4.7 to 4.9 km/s can be easily explained with
volcanic rocks. Sonic log measurements in the Lopra-1
borehole on the Faroe Islands show velocities between 3.3
and 6.6 km/s in the volcanic sequence, which consists of
basalts and hyaloclastites [Christie et al., 2006].
[58] On the free-air gravity map (Figure 14), the basin
shows up as a north-south striking linear feature bordered by
two gravity highs. The western gravity high (around OBS
618 or at 320 km in Figure 13) follows the shelf break
northward into the Baffin Bay with a variable amplitude.
Watts [1988] note that such a free-air gravity high near the
Figure 17. Comparison of velocity functions from within
the Ungava fault zone. Simplified models from NUGGET
line 1 (between 200 and 230 km) [Funck et al., 2006] and
line 600 (between 380 and 395 km, this study). For location
of the lines see Figure 1. Note that velocities below a depth
of 12 km are not well resolved for this segment of line 600.
Abbreviation HVLC is high-velocity lower crust.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
21 of 24
continental shelf break is typical for Atlantic continental
margins.
[59] Seafloor spreading in a north-south direction requires
a transform fault between the oceanic crust in the southern-
most Baffin Bay and the continental crust observed at the SE
end of line 600. Based on the gravity signature, the leaky
transform fault disappears just to the north of line 600
(Figure 14). Hence, the basin could easily form this trans-
form fault if extended northward up to the extinct spreading
axis. The interpretation as a transform fault can also explain
the low crustal thickness of 4 km as is commonly observed
in fracture zones and transform faults and discussed above
[Detrick et al., 1993]. In this context it is interesting to note
that a similar basin was observed on NUGGET line 1, where
the leaky transform fault (UFZ) is bounded by a graben
structure to the west [Funck et al., 2007]. Although the
velocity structure within the graben is not very well
resolved, the model for NUGGET line 1 indicates a volcanic
fill (velocities of 5.4 km/s) and the crust is potentially only
3 to 4 km thick (not including the underlying HVZ)
[Funck et al., 2007]. However, the 5.4-km/s layer could
also represent oceanic layer 2, in which case the crust
would be 7 km thick; but still thinner than the crust within
the UFZ. In summary, the UFZ seems to be bounded by a
narrow zone with thinned crust to the west. However,
more work is needed to determine if this feature is con-
tinuous through Davis Strait.
5.4. Underplated Continental Crust (400450 km)
[60] Finally, the continental crust at the SE end of line 600
is discussed. Velocities in the upper and lower crust between
400 and 450 km are well-resolved (Figure 9) and are very
distinct from the oceanic and igneous crust observed else-
where along the line. Hence, there is little doubt that this
crust has a continental character with velocities of 5.9 to
6.2 km/s and 6.5 to 6.8 km/s in the upper and lower layer,
respectively. Further evidence for a continental affinity
comes from the observed mid-crustal reflections. While the
lower crustal velocities are comparable to the ones found off
Baffin Island, upper crustal velocities off Greenland are
slightly lower than off Baffin Island. However, the NW and
SE ends of line 600 are not conjugate. In a plate recon-
struction [Skaarup et al., 2006], the SE end of line 600
would move southward to a position some 100 to 200 km to
the north of NUGGET line 1 (for location see Figure 1).
Hence, the Baffin Island continental crust on NUGGET line
1 would come close to the continental crust off Greenland on
line 600. Indeed, velocities on NUGGET line 1 are 5.8 to
6.1 km/s and 6.4 to 6.6 km/s in the upper and lower crust,
respectively [Funck et al., 2007], close to the values found
off Greenland on line 600 (Figure 9).
[61] The crust at the SE end of line 600 has a continental
affinity but was later modified by the Iceland plume that
created the basaltic layer at the top (as drilled in the Helle-
fisk-1 well) and the HVZ at the base of the crust. The HVZ is
interpreted as magmatic underplating. A similar layer is
observed on NUGGET line 1 beneath the thinned portions of
the continental crust but disappears to the east where the
Greenlandic continental crust thickens [Funck et al., 2007].
This observation was interpreted by Funck et al. [2007] to be
consistent with the channelling model of Sleep [1997] and
Nielsen et al. [2002]. In this model, buoyant plume material
ponds at the base of the lithosphere and then flows laterally
and preferentially along thin lithosphere. With a thickness
between 9 and 13 km (not including the volcanics and the
HVZ), the continental crust around the Hellefisk-1 well can
be considered as thinned when compared with full-thickness
continental crust in this area. Gohl and Smithson [1993]
report a thickness of up to 40 km on a line in Davis Strait
close to the coast of Greenland (for location see Figure 14).
6. Conclusions
[62] The data from line 600 provide new insight into the
crustal affinity and geodynamic evolution in the southern
Baffin Bay. The central part of the line is characterized by
oceanic crust that is divided into several segments based on
crustal thickness and on velocities within layer 2 (Figure 9).
In the NW near Baffin Island, the line crosses a transform
margin with a sharp transition from continental into oceanic
crust. The oceanic crust at this transition is 6 to 7-km thick,
which is approximately the same as normal oceanic crust
(7 km) [White et al., 1992]. However, it is thinner than along
most of the remainder of the line. Thus, the segment is
interpreted to be part of the transform fault that can be cor-
related northward in the gravity data (Figure 14). An offset
in the extinct spreading axis correlates with another segment
of thinned oceanic crust (6 km) that is assumed to lie within
an oceanic fracture zone. Similar to the transform fault in the
west, this assumed fracture zone cuts the spreading axis at an
oblique angle, which is further confirmation for oblique
seafloor spreading in Baffin Bay as suggested by Chalmers
and Pulvertaft [2001].
[63] A third zone with a reduced crustal thickness (4 km)
is interpreted as a transform fault with a probable oceanic
composition that separates oceanic crust in the southern
Baffin Bay from 20-km-thick igneous crust and continental
crust farther landward. The igneous crust coincides with a
gravity high that can be correlated through the Davis Strait
and into the Labrador Sea, where similar igneous crust is
observed on NUGGET line 1 [Funck et al., 2007]. This
igneous crust is thought to form a leaky transform fault
(Ungava fault zone) where material from the Paleogene
Iceland plume filled the gaps that developed during phases
of transtension. The plume also created the high-velocity
zone beneath the continental crust on the Greenland side of
the line. Based on its distance to the extinct spreading axis
(40 km), the oceanic crust on line 600 may have an
approximate age of 46 Ma. Given that the oceanic crust here
is thicker than the global average (9 km outside the fracture
zones and transform faults compared to 7 km for typical or
average crust), the plume seems to have influenced the
spreading system for a long time after the initial impact of
the plume. This impact is thought to coincide with the first
major volcanic pulse in West Greenland at around 60.7 to
59.4 Ma [Storey et al., 1998]. The thicker than normal oce-
anic crust in the southern Baffin Bay is in contrast to
northern Baffin Bay, where Reid and Jackson [1997] find
serpentinized mantle indicative of a low magma supply.
Hence, somewhere in Baffin Bay, a transition from volcanic
to non-volcanic continental margins is expected. This would
be similar to what is observed in the Labrador Sea [Keen
et al., 2012].
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
22 of 24
[64] The continental crust at the SE end of the line is
overlain by volcanic rocks that were drilled in the Hellefisk-1
well. The velocity model indicates a maximum thickness of
4 km for the basalts. A low-velocity zone observed beneath
some of these basalts may indicate the presence of older
sediments deposited prior to the Paleocene volcanics.
However, as there are no velocity constraints at all for this
LVZ, the layer may also represent basalts with a lower
velocity.
[65]Acknowledgments. We thank all personnel onboard the RV
Maria S. Merian who helped to collect the seismic data. We are grateful
to Jim Chalmers and Gordon Oakey who made their geological map avail-
able in digital format. John Hopper read an earlier version of the manu-
script. Reviews by Charlotte Keen, Keith Louden, and the Associate
Editor greatly improved the manuscript. Funding for the cruise was pro-
vided by Deutsche Forschungsgemeinschaft. Various parts of the project
were funded by AWI, BGR, and GEUS with support by the Danish
National Research Council. IfM-Geomar provided ocean bottom seis-
mometers through an EU grant (contract RITA-CT-2004-505322). Support
for the data analysis was provided by Capricorn Greenland, Dong E&P
Grønland A/S, ExxonMobil, and Husky Energy. The paper is published
with the permission of the Geological Survey of Denmark and Greenland.
References
Andersen, O. B. (2010), The DTU10 Gravity field and Mean sea surface,
paper presented at Second International Symposium of the Gravity Field
of the Earth (IGFS2), Univ. of Alaska Fairbanks, Fairbanks.
Breivik, A. J., R. Mjelde, P. Grogan, H. Shimamura, Y. Murai, and
Y. Nishimura (2003), Crustal structure and transform margin development
south of Svalbard based on ocean bottom seismometer data, Tectonophy-
sics,369,3770, doi:10.1016/S0040-1951(03)00131-8.
Chalmers, J. A. (1997), The continental margin off southern Greenland:
Along-strike transition from an amagmatic to a volcanic margin, J. Geol.
Soc.,154, 571576, doi:10.1144/gsjgs.154.3.0571.
Chalmers, J. A., and K. H. Laursen (1995), Labrador Sea: The extent of
continental and oceanic crust and the timing of the onset of seafloor
spreading, Mar. Pet. Geol.,12, 205217, doi:10.1016/0264-8172(95)
92840-S.
Chalmers, J. A., and G. N. Oakey (2007), Cretaceous-Palaeogene develop-
ment of Labrador Sea and Davis Strait, Geophys. Res. Abstr.,9, 01638.
Chalmers, J. A., and T. C. R. Pulvertaft (2001), Development of the conti-
nental margins of the Labrador Sea: A review, in Non-Volcanic Rifting of
Continental Margins: A Comparison of Evidence From Land and Sea,
edited by R. C. L. Wilson et al., Geol. Soc. Spec. Publ.,187,77105,
doi:10.1144/GSL.SP.2001.187.01.05.
Chian, D., and K. E. Louden (1994), The continent-ocean crustal transi-
tion across the southwest Greenland margin, J. Geophys. Res.,99,
91179135, doi:10.1029/93JB03404.
Chian, D., C. Keen, I. Reid, and K. E. Louden (1995a), Evolution of non-
volcanic rifted margins: New results from the conjugate margins of the
Labrador Sea, Geology,23, 589592, doi:10.1130/0091-7613(1995)
023<0589:EONRMN>2.3.CO;2.
Chian, D., K. E. Louden, and I. Reid (1995b), Crustal structure of the
Labrador Sea conjugate margin and implications for the formation of non-
volcanic continental margins, J. Geophys. Res.,100, 24,23924,253,
doi:10.1029/95JB02162.
Christie, P., I. Gollifer, and D. Cowper (2006), Borehole seismic studies of
a volcanic succession from the Lopra-1/1A borehole in the Faroe Islands,
northern North Atlantic, Geol. Surv. Den. Greenl. Bull.,9,2340.
Darbyshire, F. A. (2003), Crustal structure across the Canadian High Arctic
region from teleseismic receiver function analysis, Geophys. J. Int.,152,
372391, doi:10.1046/j.1365-246X.2003.01840.x.
Detrick, R. S., R. S. White, and G. M. Purdy (1993), Crustal structure of
North Atlantic fracture zones, Rev. Geophys.,31, 439458, doi:10.1029/
93RG01952.
Edwards, R. A., R. B. Whitmarsh, and R. A. Scrutton (1997), The crustal
structure across the transform continental margin off Ghana, eastern equa-
torial Atlantic, J. Geophys. Res.,102, 747772, doi:10.1029/96JB02098.
Funck, T., and K. E. Louden (1998), Wide-angle seismic imaging of pris-
tine Archean crust in the Nain Province, Labrador, Can. J. Earth Sci.,
35, 672685, doi:10.1139/e98-019.
Funck, T., and K. E. Louden (1999), Wide-angle seismic transect across the
Torngat Orogen, northern Labrador: Evidence for a Proterozoic crustal
root, J. Geophys. Res.,104, 74637480, doi:10.1029/1999JB900010.
Funck, T., K. E. Louden, R. J. Wardle, J. Hall, J. W. Hobro, M. H. Salisbury,
and A. Muzzatti (2000a), Three-dimensional structure of the Torngat
Orogen (NE Canada) from active seismic tomography, J. Geophys.
Res.,105, 23,40323,420, doi:10.1029/2000JB900228.
Funck, T., K. E. Louden, and I. D. Reid (2000b), Wide-angle seismic imag-
ing of a Mesoproterozoic anorthosite complex: The Nain Plutonic Suite in
Labrador, Canada, J. Geophys. Res.,105, 25,69325,707, doi:10.1029/
2000JB900237.
Funck, T., J. R. Hopper, H. C. Larsen, K. E. Louden, B. E. Tucholke, and
W. S. Holbrook (2003), Crustal structure of the ocean-continent transition
at Flemish Cap: Seismic refraction results, J. Geophys. Res.,108(B11),
2531, doi:10.1029/2003JB002434.
Funck, T., H. R. Jackson, K. E. Louden, S. A. Dehler, and Y. Wu (2004),
Crustal structure of the northern Nova Scotia rifted continental mar-
gin (Eastern Canada), J. Geophys. Res.,109, B09102, doi:10.1029/
2004JB003008.
Funck, T., H. R. Jackson, S. A. Dehler, and I. D. Reid (2006), A refraction
seismic transect from Greenland to Ellesmere Island, Canada: The crustal
structure in southern Nares Strait, Polarforschung,74,97112.
Funck, T., H. R. Jackson, K. E. Louden, and F. Klingelhöfer (2007), Seis-
mic study of the transform rifted margin in Davis Strait between Baffin
Island (Canada) and Greenland: What happens when a plume meets a
transform, J. Geophys. Res.,112, B04402, doi:10.1029/2006JB004308.
Gerlings, J., T. Funck, H. R. Jackson, K. E. Louden, and F. Klingelhöfer
(2009), Seismic evidence for plume-derived volcanism of the continental
margin in southern Davis Strait and northern Labrador Sea, Geophys. J.
Int.,176, 980994, doi:10.1111/j.1365-246X.2008.04021.x.
Gohl, K., and S. B. Smithson (1993), Structure of Archean crust and passive
margin of southwest Greenland from seismic wide-angle data, J. Geo-
phys. Res.,98, 66236638, doi:10.1029/93JB00016.
Gohl, K., B. Schreckenberger, and T. Funck (2009), The expedition of the
research vessel Maria S. Merianto the Davis Strait and Baffin Bay in
2008(MSM09/3), Rep. Polar and Mar. Res. 587, 99 pp., Alfred Wegener
Inst. for Polar and Mar. Res., Bremerhaven, Germany.
Gradstein, F. M., J. G. Ogg, and A. G. Smith (2005), A Geologic Time Scale
2004, 610 pp., Cambridge Univ. Press, Cambridge, U. K.
Gregersen, U., and T. Bidstrup (2008), Structures and hydrocarbon prospec-
tivity in the northern Davis Strait area, offshore West Greenland, Petrol.
Geosci.,14, 151166, doi:10.1144/1354-079308-752.
Hall, J., K. E. Louden, T. Funck, and S. Deemer (2002), Geophysical char-
acteristics of the continental crust along the Lithoprobe Eastern Canadian
Shield Onshore-Offshore Transect (ECSOOT): A review, Can. J. Earth
Sci.,39, 569587, doi:10.1139/e02-005.
Holbrook, W. S., W. D. Mooney, and N. I. Christensen (1992), The seismic
velocity structure of the deep continental crust, in Continental Lower
Crust, edited by D. M. Fountain et al., pp. 143, Elsevier, New York.
Jackson, H. R., and I. Reid (1994), Crustal thickness variations between the
Greenland and Ellesmere Island margins determined from seismic refrac-
tion, Can. J. Earth Sci.,31, 14071418, doi:10.1139/e94-124.
Jokat, W., O. Ritzmann, M. C. Schmidt-Aursch, S. Drachev, S. Gauger, and
J. Snow (2003), Geophysical evidence for reduced melt production on the
Arctic ultraslow Gakkel mid-ocean ridge, Nature,423, 962965,
doi:10.1038/nature01706.
Keen, C. E., and D. L. Barrett (1972), Seismic refraction studies in Baffin
Bay: An example of a developing ocean basin, Geophys. J. R. Astron.
Soc.,30, 253271, doi:10.1111/j.1365-246X.1972.tb05812.x.
Keen, C. E., M. J. Keen, D. J. Ross, and M. Lack (1974), Baffin Bay: Small
ocean basin formed by sea-floor spreading, AAPG Bull.,58, 10891108.
Keen, C. E., K. Dickie, and S. A. Dehler (2012), The volcanic margins of
the northern Labrador Sea: Insights to the rifting process, Tectonics,31,
TC1011, doi:10.1029/2011TC002985.
Kerr, J. W. (1967), A submerged continental remnant of the Labrador Sea,
Earth Planet. Sci. Lett.,2, 283289, doi:10.1016/0012-821X(67)90143-4.
Klose, G. W., E. Malterre, N. J. McMillan, and C. G. Zinkan (1982), Petro-
leum exploration offshore southern Baffin Island, northern Labrador Sea,
Canada, in Arctic Geology and Geophysics, edited by A. F. Embry and
H. R. Balkwill, Mem. Can. Soc. Pet. Geol.,8, 233244.
Larsen, H. C., and A. D. Saunders (1998), Tectonism and volcanism at the
southeast Greenland rifted margin: A record of plume impact and later
continental rupture, in Proc. Ocean Drill. Program Sci. Results,152,
503533.
Larsen, L. M., L. M. Heaman, R. A. Creaser, R. A. Duncan, R. Frei, and
M. Hutchinson (2009), Tectonomagmatic events during stretching and
basin formation in the Labrador Sea and the Davis Strait: Evidence
from age and composition of Mesozoic to Palaeogene dyke swarms
in West Greenland, J. Geol. Soc.,166, 9991012, doi:10.1144/0016-
76492009-038.
Lawver, L. A., and R. D. Müller (1994), Iceland hotspot track, Geology,22,
311314, doi:10.1130/0091-7613(1994)022<0311:IHT>2.3.CO;2.
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
23 of 24
Ludwig, W. J., J. E. Nafe, and C. L. Drake (1970), Seismic refraction, in
The Sea, vol. 4, New Concepts of Sea Floor Evolution, Part I, edited
by A. E. Maxwell, pp. 584, Wiley-Intersci., Hoboken, N. J.
Lutter, W. J., and R. L. Nowack (1990), Inversion for crustal structure using
reflections from the PASSCAL Ouachita experiment, J. Geophys. Res.,
95, 46334646, doi:10.1029/JB095iB04p04633.
Mooney, W. D., and T. M. Brocher (1987), Coincident seismic reflection/
refraction studies of the continental lithosphere: A global review, Rev.
Geophys.,25, 723742, doi:10.1029/RG025i004p00723.
Nielsen, T. K., H. C. Larsen, and J. R. Hopper (2002), Contrasting rifted
margin styles south of Greenland: Implications for mantle plume dynam-
ics, Earth Planet. Sci. Lett.,200, 271286, doi:10.1016/S0012-821X(02)
00616-7.
Oakey, G. N. (2001), Magnetic anomaly, Davis Strait region, Canadian and
Greenland Arctic, Open File Rep. 39358, Geol. Surv. of Can., Ottawa,
Ont., Canada.
Parsiegla, N., J. Stankiewicz, K. Gohl, T. Ryberg, and G. Uenzelmann-
Neben (2009), Southern African continental margin: Dynamic processes
of a transform margin, Geochem. Geophys. Geosyst.,10, Q03007,
doi:10.1029/2008GC002196.
Pearson, D. G., C. H. Emeleus, and S. P. Kelley (1996), Precise
40
Ar/
39
Ar
age for the initiation of Palaeogene volcanism in the Inner Hebrides and
its regional significance, J. Geol. Soc.,153, 815818, doi:10.1144/
gsjgs.153.6.0815.
Reid, I., and H. R. Jackson (1997), Crustal structure of the northern Baffin
Bay: Seismic refraction results and tectonic implications, J. Geophys.
Res.,102, 523542, doi:10.1029/96JB02656.
Roest, W. R., and S. P. Srivastava (1989), Sea-floor spreading in the
Labrador Sea: A new re-construction, Geology,17, 10001003,
doi:10.1130/0091-7613(1989)017<1000:SFSITL>2.3.CO;2.
Rolle, F. (1985), Late Cretaceous-Tertiary sediments offshore central West
Greenland: Lithostratigraphy, sedimentary evolution, and petroleum
potential, Can. J. Earth Sci.,22, 10011019, doi:10.1139/e85-105.
Skaarup, N., H. R. H. H. Jackson, and G. Oakey (2006), Margin segmenta-
tion of Baffin Bay/Davis Strait, eastern Canada, based on seismic reflec-
tion and potential field data, Mar. Pet. Geol.,23, 127144, doi:10.1016/j.
marpetgeo.2005.06.002.
Sleep, N. H. (1997), Lateral flow and ponding of starting plume material,
J. Geophys. Res.,102, 10,00110,012, doi:10.1029/97JB00551.
Srivastava, S. P. (1978), Evolution of the Labrador Sea and its bearing on
the early evolution of the North Atlantic, Geophys. J. R. Astron. Soc.,
52, 313357, doi:10.1111/j.1365-246X.1978.tb04235.x.
Srivastava, S. P., and C. E. Keen (1995), A deep seismic reflection profile
across the extinct Mid-Labrador Sea spreading center, Tectonics,14,
372389, doi:10.1029/94TC02453.
Srivastava, S. P., B. MacLean, R. F. Macnab, and H. R. Jackson (1982),
Davis Strait: Structure and evolution as obtained from a systematic
geophysical survey, in Arctic Geology and Geophysics, edited by A. F.
Embry and H. R. Balkwill, Mem. Can. Soc. Pet. Geol.,8, 267278.
Storey, M., R. A. Duncan, A. K. Pedersen, L. M. Larsen, and H. C. Larsen
(1998),
40
Ar/
39
Ar geochronology of the West Greenland Tertiary volca-
nic province, Earth Planet. Sci. Lett.,160, 569586, doi:10.1016/
S0012-821X(98)00112-5.
Suckro, S., et al. (2012), The crustal structure of southern Baffin Bay:
Implications from a seismic refraction experiment, Geophys. J. Int.,
in press.
Talwani, M., J. L. Worzel, and M. Landisman (1959), Rapid gravity com-
putations for two-dimensional bodies with application to the Mendocino
submarine fracture zone, J. Geophys. Res.,64,4959, doi:10.1029/
JZ064i001p00049.
Verhoef, J., W. R. Roest, R. Macnab, and J. Arkani-Hamed, and the Mem-
bers of the Project Team (1996), Magnetic anomalies of the Arctic and
North Atlantic oceans and adjacent land areas, Open File Rep. 3125, Geol
Surv. of Can., Ottawa, Ont., Canada.
Watts, A. B. (1988), Gravity anomalies, crustal structure and flexure of the
lithosphere at the Baltimore Canyon Trough, Earth Planet. Sci. Lett.,89,
221238, doi:10.1016/0012-821X(88)90174-4.
White, R. S., D. McKenzie, and K. ONions (1992), Oceanic crustal thick-
ness from seismic measurements and rare earth element inversions,
J. Geophys. Res.,97, 19,68319,715, doi:10.1029/92JB01749.
Whittaker, R. C., N. E. Hamann, and T. C. R. Pulvertaft (1997), A new
frontier province offshore northwest Greenland: Structure, basin develop-
ment, and petroleum potential of the Melville Bay area, AAPG Bull.,81,
978998.
Williamson, M.-C., M. E. Villeneuve, L. M. Larsen, H. R. Jackson, G. N.
Oakey, and B. Maclean (2001), Age and petrology of offshore basalts
from the Southeast Baffin Island Shelf, Davis Strait, and the Western
Greenland continental margin, paper presented at Joint Annual Meeting,
Geol. and Mineral. Assoc. of Can., St. Johns, Newfoundland, Canada.
Zelt, C. A., and D. A. Forsyth (1994), Modeling wide-angle seismic data for
crustal structure: Southeastern Grenville Province, J. Geophys. Res.,99,
11,68711,704, doi:10.1029/93JB02764.
Zelt, C. A., and R. B. Smith (1992), Seismic traveltime inversion for 2-D
crustal velocity structure, Geophys. J. Int.,108,1634, doi:10.1111/
j.1365-246X.1992.tb00836.x.
V. Damm and I. Heyde, Federal Institute for Geosciences and Natural
Resources, Stilleweg 2, D-30655 Hanover, Germany. (volkmar.damm@bgr.
de; ingo.heyde@bgr.de)
T. Funck, Geological Survey of Denmark and Greenland, Øster Voldgade
10, DK-1350 Copenhagen, Denmark. (tf@geus.dk)
K. Gohl, Alfred Wegener Institute for Polar and Marine Research, Am
Alten Hafen 26, D-27568 Bremerhaven, Germany. (karsten.gohl@awi.de)
FUNCK ET AL.: SEISMIC REFRACTION SOUTHERN BAFFIN BAY B04107B04107
24 of 24
... Two German expeditions led by the Alfred Wegener Institute with the research vessels RV Maria S. Merian and RV Polarstern were conducted in 2008 and 2010, respectively. The 2008 data (AWI-2008 lines;Funck et al., 2012a;Suckro et al., 2012Suckro et al., , 2013 comprise three transects from Baffin Island to Greenland using between 12 and 24 ocean-bottom seismometers. The 2010 lines cover the Melville Bay area and the entrance to Nares Location of released multichannel seismic reflection data and wells in the Labrador-Baffin Seaway (select unreleased data is also shown for the Labrador margin and released data up to 2018 is shown for the West Greenland margin). ...
... In the central part of Davis Strait, Suckro et al. (2013) reported a zone 60 km wide with increased crustal velocities in the area of the Ungava Fault Zone (Fig. 10d), which was interpreted as intruded continental crust or new igneous crust that can be correlated with the igneous crust on NUGGET line 1 in southern Davis Strait (Fig. 10c). Similarly, Funck et al. (2012a) observed a 25 km wide zone with igneous crust in northern Davis Strait on seismic line AWI-20080600 (Fig. 10e). Here the crust is up to 23 km thick, including a sequence of volcanic rocks up to 3 km thick. ...
... P-wave velocity models for the Labrador Sea and Davis Strait: a) conjugate seismic refraction profiles 90-R1 and 88-R2 across the Labrador and southwestern Greenland continental margins, respectively (velocity models after Chian and Louden (1994) and Chian et al. (1995a); red triangles mark location of interpreted magnetic anomalies after Roest and Srivastava (1989)), and conjugate profiles merged at magnetic chron C26; b) seismic-refraction profile 4 of the SIGNAL experiment across the extinct spreading axis in the Labrador Sea (after Delescluse et al., 2015); c) line 1 of the NUGGET experiment (Funck et al., 2007) The southernmost seismic line (AWI-20080600 in Fig. 10e) in Baffin Bay was discussed above in reference to Davis Strait, but the line crosses southern Baffin Bay parallel to the interpreted location of the extinct spreading axis (Oakey and Chalmers, 2012) at a distance of 40 km. Funck et al. (2012a) observed oceanic crust with a variable thickness along the line. In the southeast, the oceanic crust is 9 km thick, which is more than the global average of 7 km (White et al., 1992), indicating an ample magma supply. ...
... The NCRM CTSE occupies parts of the east Canadian continental margin that are underlain by continental crust (Fig. 3). The latter can be up to c. 15-25 km thick below the pre-rift and syn-rift TSEs (Funck et al., 2007;Funck et al., 2012;Keen et al., 2018), and thins toward the oceanic domain (Fig. 4). The northern CTSE limit corresponds to the southern boundary of the Eurekan contractional deformation near 77° N, and the southern limit for the purpose of this volume is defined at 64° N. ...
Article
Full-text available
The Northeast Canada Rifted Margin (NCRM) Composite Tectono-Sedimentary Element (CTSE) developed during a long and complex history that produced two tectono-sedimentary elements (TSEs): (1) the pre-rift TSE of pre-Cretaceous age; and (2) the syn-rift TSE of Early Cretaceous-Paleocene age. The pre-rift TSE includes the oldest and most poorly known offshore sedimentary accumulations which mainly evolved in a cratonic setting. In contrast, Cretaceous-lower Paleocene sedimentary basins of the syn-rift TSE are known from several wells, seismic data, outcrops, and seabed samples, and their extent and distribution are mapped in most parts of the margin. The syn-rift TSE is the most prospective part of the margin and hydrocarbon shows have been documented in some wells and offshore seeps studies. This review provides insights into the Paleozoic-Cenozoic evolution of the NE Canada rifted margin in the Labrador Sea, Davis Strait, and Baffin Bay. In this context, we discuss structural inheritance and rift development, and account for confirmed and potential hydrocarbon systems and plays.
... Seafloor spreading ended by chron C13 in the Late Eocene (Oakey and Chalmers, 2012). As a result of this complex rift history, Baffin Bay is ringed by magma-rich margins in the south, and magma-poor margins in the north (Reid and Jackson, 1997;Skaarup et al., 2006;Funck et al., 2012;Keen et al., this volume). ...
Chapter
Full-text available
Sedimentary basins within the Labrador-Baffin Seaway are the product of rifting between Greenland and the paleo-North American Plate. Rifting started in the Early Cretaceous, with seafloor spreading initiated in the Paleocene and ending near the Eocene-Oligocene boundary. A change in the spreading direction in the latest Paleocene resulted in transform offsets in the Davis Strait and along fracture zones in Baffin Bay, with deformation in northern Baffin Bay during the Eurekan Orogeny. Since the stratigraphy of western Baffin Bay is poorly constrained, analogues are used from the well studied Labrador and West Greenland margins and exposures on nearby Bylot Island. The generally northwest-trending basement structures are infilled with Cretaceous strata, which are overlain by a seaward-thickening wedge of post-rift Paleocene to Middle Miocene sedimentary rocks. Finally, a thick Middle Miocene and younger interval blankets the deep water and oceanic crust, with clinoforms locally developed on the shelf.
... In the large volcanic rock-covered area offshore central West Greenland (Fig. 2), prevolcanic successions have not been drilled, but subvolcanic basins ((?)Danian-Cretaceous or older) are likely present. This may be supported by a 'low-velocity zone' below the volcanic rocks at Hellefisk-1 well from a refraction seismic study (Fig. 25b;Funck et al., 2012) and a zone with 'lower sediments' from a gravity modelling study ( Fig. 25c; Skaarup, 2001Skaarup, , 2002. Farther north, along the northern West Greenland margin, the Upper Cretaceous includes thick mudstone-dominated successions deposited in large basins with periodically active boundary faults. ...
Chapter
Full-text available
The chapter describes the stratigraphy and geological evolution of West Greenland continental margin, including the tectonostratigraphy, the seismic stratigraphy, the lithostratigraphy, and the biostratigraphy, with ties to all offshore wells on the margin, and with descriptions from selected onshore wells and outcrops.
... The thickness of the basalts and dykes varies significantly, which also results in a high variability of the P-wave velocities. Funck et al. (2012) divided oceanic layer 2 in the southern Baffin Bay into segments with different velocities (5.5-6.0 km/s and 6.1-6.5 km/s) and thicknesses. These segments are probably separated by fracture zones and were formed at different ages . ...
Article
Full-text available
We present results derived from a seismic refraction experiment and gravity measurements about the crustal structure of southern Sri Lanka and the adjacent Indian Ocean. A P-wave velocity model was derived using forward modelling of the observed travel times along a 509 km long, N-S trending profile at 81°E longitude. Our results show that the continental crust below southern Sri Lanka is up to 38 km thick. A ~ 65 km wide transition zone, which thins seawards to ~7 km thickness, divides stretched continental from oceanic crust. The adjacent, 4.7 to 7 km thick normal oceanic crust is covered by up to 4 km thick sediments. The oceanic crust is characterized by intra-crustal reflections and displays P-wave velocity variations, especially in oceanic layer 2, along our profile. In the central part of the profile, the uppermost mantle layer is characterized by normal P-wave mantle velocities of 8.0–8.1 km/s. At the southern end of the profile, unusual low upper mantle seismic velocities, ranging from 7.5 to 7.6 km/s only, characterize the uppermost mantle layer. These low upper mantle velocities are probably caused by partially serpentinized upper mantle. At even greater depths the upper mantle layer is characterized by velocities of 8.3 km/s on average. The type of margin along our profile is difficult to identify, since it is characterized by features typical for different types of margins.
Article
Full-text available
Wegener's Continental Drift Theory has laid the foundations of modern plate tectonics. However, despite decades of work and studies around the globe, modern plate tectonics still does not explain all the datasets acquired up to now and is well overdue for a major update. We propose a new theory, the orogenic bridge theory, which partly builds on the Continental Drift Theory and modern plate tectonics and reconciles them with the idea put forward by a competing theory, the Land Bridge Theory (or Isthmian Links). The orogenic bridge theory states that the style of continental rifting is directly controlled by preexisting orogenic structures. On the one hand, preexisting orogens trending parallel to an opening rift facilitate breakup and rift propagation and control the strike and geometry of rift-related structures. This endmember has already been broadly studied worldwide. On the other hand, orogens oriented orthogonal (or highly oblique) to the opening rift will act as strong barriers forcing the rift to step, therefore delaying or impeding breakup and rift propagation and localizing the formation of major-offset transform faults. In the present contribution, we review the evidence in favor of a correlation between rift-orthogonal orogens and major transforms and discuss some of the main implications of the orogenic bridge theory.
Thesis
The Icelandic plume, a major convective upwelling, has had a considerable influence on the geological evolution of the North Atlantic region. Direct manifestations of this major convective upwelling include positive residual depth anomalies and long wavelength free-air gravity anomalies, both of which reach from Baffin Island to Norway and from Newfoundland to Svalbard. Signifi cant shear wave velocity anomalies, observed in full-waveform tomographic models between 100 km and 200 km depth, show the Icelandic plume has a complex, irregular planform. These anomalies suggest about fi ve horizontal fi ngers radiate away from the central plume conduit. The best imaged fingers lie beneath the British Isles, southern Scandinavia and Greenland, extending ~1,000 km from the Icelandic plume. It is proposed that these radial miscible fi ngers develop due to the Saffman-Taylor instability, a fluid dynamical phenomenon which occurs when a less viscous fluid is injected into a more viscous fluid. Mobility ratio (i.e. the ratio of fluid viscosities), Peclet number (i.e. the ratio of advective and diffusive transport rates) and thickness of the horizontal layer into which the fluid is injected, together control the presence of fi ngering due to the Saffman-Taylor instability. Estimates for the Icelandic plume suggest the mobility ratio is at least 15, the Peclet number is ~ 2 x 10⁴, and the asthenospheric channel thickness is 100 ± 50 km. Appropriately scaled laboratory experiments play a key role in developing a quantitative understanding of the spatial and temporal evolution of mantle plume planforms. My results prove that the presence or absence of radial miscible fi ngering due to the Saffman-Taylor instability is controlled by changes in mobility ratio, Peclet number and horizontal layer thickness. At large horizontal thicknesses, gravity has an increasingly important influence and acts to damp the production of radial viscous miscible fi ngers. Observed values from the Icelandic plume suggest the fluid dynamics may be more complex than the Saffman-Taylor instability alone. Additional processes, such as interaction with the base of the lithospheric plate, along with the Saffman-Taylor instability, may be the origin of the fi ngers.
Article
Full-text available
The Palaeogene volcanic succession in the northern part of the Nuussuaq Basin in West Greenland comprises three formations: the Vaigat and Svartenhuk Formations of Paleocene age (61–58 Ma) and the Naqerloq Formation of Eocene age (57–54 Ma). In this study, we formalise and describe the volcanic stratigraphy on Svartenhuk Halvø and the areas with lavas that flowed across the basin boundary onto the adjoining basement areas in the north and east. The Vaigat Formation comprises three members. The Kakilisaat and Nerutusoq Members are of minor volume and consist of, respectively, crustally contaminated basalts and chemically enriched basalts with relatively high contents of incompatible trace elements. They are overlain by the voluminous Nunavik Member of tholeiitic picrites (MgO ≥12 wt%) and subordinate magnesian basalts. The oldest volcanic deposits are commonly foreset-bedded hyaloclastites, and the overlying subaerial lavas are mainly thin, grey, crumbling flows. Eruption sites were mainly within the basin, with depocentres in the south and hyaloclastite and lava transport directions towards the north. Thicknesses vary from up to at least 2000 m in the south to ≥380 m in the northernmost exposures close to 72°N. The Svartenhuk Formation comprises four members. The lowest, Kuugaartorfik Member, is up to 100 m thick and consists partly of quartzofeldspathic and partly volcanogenic sediments; it is restricted to northern Svartenhuk Halvø and the Innerit peninsula. The overlying volcanic Tunuarsuk, Nuuit and Skalø Members are voluminous and widespread, with a combined thickness of up to 1800 m. They consist of tholeiitic basalts with similar chemical compositions but with correlatable stratigraphic variation patterns. The Tunuarsuk Member consists of interspersed flow groups of thin, grey flows and massive, brown flows; the Nuuit Member comprises mainly massive brown flows, and the Skalø Member is dominated by light grey flows. The Svartenhuk Formation oversteps the Vaigat Formation on the basement in the north and east. In these distal areas the Tunuarsuk and Nuuit Members constitute the major volumes, and preserved thicknesses are up to 1400 m. In northern and eastern Svartenhuk Halvø and also farther to the north and east, foreset-bedded hyaloclastites indicate transport directions towards the north and possibly east from eruption sites within the basin. The Naqerloq Formation comprises one member, the Arfertuarsuk Member, consisting of flows of brown basalt with relatively enriched chemistry and a single trachyte flow. The member is only found in western Svartenhuk Halvø and on Skalø, where it conformably overlies the older lavas with up to 350 m thickness preserved after erosion. Dykes of all three formations are present. The distribution of dykes of the Naqerloq Formation suggests that this originally extended much farther east. Picrites and basalts of the Vaigat and Svartenhuk Formations are geochemically related; the picritic lavas represent erupted primitive magmas, whereas the basaltic lavas represent fractionated melts formed in deep magma chambers. The melts formed from a geochemically depleted but heterogeneous mantle; in addition melts from enriched sources were occasionally incorporated. The enriched basalts of the Naqerloq Formation arose from another mantle source. Low contents of V, Cu and Ni in some crustally contaminated lavas indicate that accumulation of these elements may be present at depth.
Chapter
Full-text available
The papers contained in this bulletin provide a comprehensive summary and updated understanding of the onshore geology and evolution of Baffin Island, the Labrador-Baffin Seaway, and surrounding onshore regions. This introductory paper summarizes and links the geological and tectonic events that took place to develop the craton and subsequent Proterozoic to Cenozoic sedimentary basins. Specifically, the Precambrian and Paleozoic geology of Baffin Island and localized occurrences underlying the adjacent Labrador-Baffin Seaway, the Mesozoic to Cenozoic stratigraphy and rift history that records the opening and evolution of the Labrador-Baffin Seaway, the seismicity of the region, as well as both the mineral (Baffin Island) and hydrocarbon (onshore and offshore) resource potential are discussed.
Article
Full-text available
Extruded basalt flows overlying sedimentary sequences present a challenge to hydrocarbon exploration using reflection seismic techniques. The Lopra-1/1A re-entry well on the Faroese island of Suduroy allowed us to study the seismic characteristics of a thick sequence of basalt flows from well logs and borehole seismic recordings. Data acquired during the deepening operation in 1996 are presented here. The re-entry well found that the seismic event at 2340 m, prognosed from the pre-drill Vertical Seismic Profile (VSP) as a decrease in impedance, was not base basalt and the deepened well remained within the lower series basalts. Nonetheless, compressional and shear sonic logs and a density log were recorded over the full open hole interval. These allowed a firm tie to be made with the reflected wavefield from a new VSP. The sonic logs show a compressional to shear wavespeed ratio of 1.84 which is almost constant with depth. Sonic compressional wavespeeds are 3% higher than seismic velocities, suggesting dispersion in the basalt flows. Azimuthal anisotropy was weakly indicated by the shear sonic log but its orientation is consistent with the directions of mapped master joints in the vicinity of the well. The VSP downgoing compressional wavelet shows good persistence, retaining a dominant period of 28 ms at 3510 m depth. Average vertical velocity is 5248 m/s, higher than previously reported. Attenuation can largely be modelled by geometrical spreading and scattering loss, consistent with other studies. Within the piled flows, the effective Q from scattering is about 35. Elastic layered medium modelling shows some hope that a mode-converted shear wave may be observed at moderate offsets. Like its predecessor, the 1996 VSP indicates a decrease in impedance below the final depth of the well. However, it is unlikely to be basement or sediment and is probably an event within the volcanic sequence.
Article
Full-text available
Dynamic processes at sheared margins associated with the formation of sedimentary basins and marginal ridges are poorly understood. The southern African margin provides an excellent opportunity to investigate the deep crustal structure of a transform margin and to characterize processes acting at these margins by studying the Agulhas-Falkland Fracture Zone, the Outeniqua Basin, and the Diaz Marginal Ridge. To do this, we present the results of the combined seismic land-sea experiments of the Agulhas-Karoo Geoscience Transect. Detailed velocity-depth models show crustal thicknesses varying from ˜42 km beneath the Cape Fold Belt to ˜28 km beneath the shelf. The Agulhas-Falkland Fracture Zone is embedded in a 50 km wide transitional zone between continental and oceanic crust. The oceanic crust farther south exhibits relatively low average crustal velocities (˜6.0 km/s), which can possibly be attributed to transform-ridge intersection processes and the thermal effects of the adjacent continental crust during its formation. Crustal stretching factors derived from the velocity-depth models imply that extension in the Outeniqua Basin acted on regional as well as more local scales. We highlight evidence for two episodes of crustal stretching. The first, with a stretching factor β of 1.6, is interpreted to have influenced the entire Outeniqua Basin. The stresses possibly originated from the beginning breakup between Africa and Antarctica (˜169-155 Ma). The second episode can be associated with a transtensional component of the shear motion along the Agulhas-Falkland Transform from ˜136 Ma. This episode caused additional crustal stretching with β = 1.3 and is established to only have affected the southern parts of the basin. Crustal velocities directly beneath the Outeniqua Basin are consistent with the interpretation of Cape Supergroup rocks underlying most parts of the basin and the Diaz Marginal Ridge. We propose that the formation of this ridge can be either attributed to a transpressional episode along the Agulhas-Falkland Transform or, more likely, to thermal uplift accompanying the passage of a spreading ridge to the south.
Article
Full-text available
Baffin Bay represents the northern extension of the extinct rift system in the Labrador Sea. While the extent of oceanic crust and magnetic spreading anomalies are well constrained in the Labrador Sea, no magnetic spreading anomalies have yet been identified in Baffin Bay. Thus, the nature and evolution of the Baffin Bay crust remain uncertain. To clearly characterize the crust in southern Baffin Bay, 42 ocean bottom seismographs were deployed along a 710-km-long seismic refraction line, from Baffin Island to Greenland. Multichannel seismic reflection, gravity and magnetic anomaly data were recorded along the same transect. Using forward modelling and inversion of observed traveltimes from dense airgun shots, a P-wave velocity model was obtained. The detailed morphology of the basement was constrained using the seismic reflection data. A 2-D density model supports and complements the P-wave modelling. Sediments of up to 6 km in thickness with P-wave velocities of 1.8-4.0 km s-1 are imaged in the centre of Baffin Bay. Oceanic crust underlies at least 305 km of the profile. The oceanic crust is 7.5 km thick on average and is modelled as three layers. Oceanic layer 2 ranges in P-wave velocity from 4.8 to 6.4 km s-1 and is divided into basalts and dykes. Oceanic layer 3 displays P-wave velocities of 6.4-7.2 km s-1. The Greenland continental crust is up to 25 km thick along the line and divided into an upper, middle and lower crust with P-wave velocities from 5.3 to 7.0 km s-1. The upper and middle continental crust thin over a 120-km-wide continent-ocean transition zone. We classify this margin as a volcanic continental margin as seaward dipping reflectors are imaged from the seismic reflection data and mafic intrusions in the lower crust can be inferred from the seismic refraction data. The profile did not reach continental crust on the Baffin Island margin, which implies a transition zone of 150 km length at most. The new information on the extent of oceanic crust is used with published poles of rotation to develop a new kinematic model of the evolution of oceanic crust in southern Baffin Bay.
Article
A refraction and wide-angle reflection seismic study was carried out in southern Nares Strait (northernmost Baffin Bay) on a 378 km long profile running from Pituffik/Thule Air Base on Greenland into Makinson Inlet on Ellesmere Island, Canada. Eight ocean bottom seismometers and eight land stations were deployed to record the airgun shots along the line. A crustal velocity model was developed by forward and inverse modeling techniques. The Proterozoic Thule Basin can be correlated across Nares Strait as a continuous structure with a total thickness of 4-5 km. The basin is divided into three units. The upper unit has velocities of 4.5-5.0 km s-1 and a Poisson's ratio (σ) of 0.30, indicating a high content of carbonates. The middle unit is characterized by high velocities (6.1 km s-1) and a Poisson's ratio of 0.28. This unit is interpreted to correlate with the basaltic sills found in the Cape Combermere formation. The lower unit is a low-velocity zone and, hence, its velocities are unconstrained. The underlying crust is divided into three layers, upper crust (6.0-6.2 km s-1, σ = 0.25), middle crust (6.1-6.3 km s-1, a = 0.26) and lower crust (6.7-6.9 km s-1, σ = 0.26). These properties are consistent with a granitic/gneissic composition in the upper/middle crust and with granulites in the lower crust. Moho depth on either side of Nares Strait is 36 km, with some shallowing to 33 km in a 100 km wide zone. The Moho shallowing is related to local uplift at the Carey Islands just to the south of the line based on correlation with the gravity data. With only minor lateral changes of crustal velocities and Moho depth across southern Nares Strait, the eastern Archean/Proterozoic part of Ellesmere Island appears to be part of the same plate as Greenland with Thule Basin as intracratonic feature. No structures consistent with a strike-slip boundary could be resolved in southern Nares Strait.
Article
A successor to A Geologic Time Scale 1989 (Cambridge, 1990), this volume introduces the theory and methodology behind the construction of the new time scale, before presenting the scale itself in extensive detail. An international team of over forty stratigraphic experts develops the most up-to-date international stratigraphic framework for the Precambrian and Phanerozoic eras. A large wallchart (not available for eBook) summarizing the time scale at the back of the book completes this invaluable reference for researchers and students.