Hoved Opp Gave

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 120

Alles Vergängliche

Ist nur ein Gleichnis;


Das Unzulängliche,
Hier wird’s Ereignis;
Das Unbeschreibliche,
Hier ist’s getan;
Das Ewig-Weibliche
Zieht uns hinan.
Goethe (1833)
THE ARCTIC PLATE BOUNDARY
Cand. Scient. Thesis

Øyvind Engen
Department of Geology, University of Oslo, P.O. Box 1047 Blindern, N-0316 Oslo, Norway
oyvind.engen@geologi.uio.no

PREFACE

This thesis was carried out at the Department of Geology, University of Oslo, with Professor
Olav Eldholm as principal advisor and Professor Hilmar Bungum as co-advisor. Some of the
research was also done at NORSAR, Kjeller.
I am greatly indebted to my advisors - to Olav for his friendly support and utterly
professional cooperation in every difficult topic, and to Hilmar for his thorough reviews and
for sharing his profound knowledge.
Many more persons have contributed to the final result. First of all, I would like to
thank Filippos Tsikalas for his almost 24-hour computer support. The staff and students at
NORSAR are acknowledged, in particular Conrad Lindholm and Erik Hicks, for their help
and comments. Also, my special thanks go to Professors Jan Inge Faleide and Annik M.
Myhre for their arrangements and teaching; Steve Gibbons for his computer programming;
Shicun Ren for her encouragement and Chinese cooking; Ellen Sigmond for her fruitful
criticism on the Fram Strait; and the staff at Hydro EPI for their colleagueness.
The students at the Department of Geology, my friends and family have done their
best to make my study years a pleasant and vigorous experience. Thanks a lot!

Oslo, 20 June 2001,

Øyvind Engen

Notes:
• Cover illustration: Topography of the Eurasia Basin and adjacent areas, viewed from the Greenland Sea
towards the east. Data from IBCAO (Jakobsson et al., 2000a).
• The manuscript was written in Adobe FrameMaker 5.5, and figures were made with GMT 3.3.2 (Wessel and
Smith, 1991), Plates 4.1, IslandDraw, Adobe Photoshop 5.0 and Adobe Illustrator 8.0.
• Computer programs were written in FORTRAN 77 and nawk.

3
5

CONTENTS

1 INTRODUCTION ________________________________________________________________ 7
2 GEOLOGICAL FRAMEWORK _____________________________________________________ 11
2.1 Geometric Constraints ........................................................................................................... 11
2.2 Plate Tectonic Setting ............................................................................................................ 12
2.3 Mid-Oceanic Ridges .............................................................................................................. 14
2.4 Basins and Sub-Basins ........................................................................................................... 15
2.4.1 Greenland Sea ............................................................................................................... 15
2.4.2 Eurasia Basin ................................................................................................................. 15
2.5 Plateaus and Submarine Ridges .......................................................................................... 16
2.5.1 Yermak Plateau and Morris-Jesup Rise ..................................................................... 16
2.5.2 Lomonosov Ridge ......................................................................................................... 16
2.6 Svalbard-Laptev Sea Continental Margin .......................................................................... 17
3 GEOPHYSICAL DATA __________________________________________________________ 23
3.1 Display of Gridded Data ....................................................................................................... 23
3.2 Regional Data Sets ................................................................................................................. 25
3.2.1 Bathymetry .................................................................................................................... 26
3.2.2 Gravity ............................................................................................................................ 27
3.2.3 Magnetics ....................................................................................................................... 28
3.2.4 Sediment Thickness ...................................................................................................... 30
3.3 Local Data Sets ....................................................................................................................... 32
4 SEISMICITY ___________________________________________________________________ 37
4.1 Data Sources ........................................................................................................................... 38
4.2 Epicentre Processing .............................................................................................................. 41
4.2.1 Elimination of Redundancy ........................................................................................ 43
4.2.2 Quality Sorting .............................................................................................................. 44
4.3 Arctic Seismicity Map and Catalogue ................................................................................. 47
5 INTERPRETATION PROCEDURE __________________________________________________ 51
5.1 Geophysical Plate Boundary Indicators ............................................................................. 51
5.2 Course of Work ...................................................................................................................... 53
6 THE ARCTIC PLATE BOUNDARY _________________________________________________ 55
6.1 Spitsbergen Transform System ............................................................................................ 55
6.2 Gakkel Ridge .......................................................................................................................... 60
6.2.1 West Gakkel Ridge ....................................................................................................... 62
6.2.2 East Gakkel Ridge ......................................................................................................... 65
6.3 Laptev Sea ............................................................................................................................... 66
7 DISCUSSION __________________________________________________________________ 71
7.1 Seismicity of the Arctic Plate Boundary ............................................................................. 71
7.2 Intraplate Seismicity .............................................................................................................. 76
7.3 Crustal Accretion and Plate Boundary Segmentation ...................................................... 78
7.4 The Laptev Sea Continent-Ocean Transition ..................................................................... 81
7.5 Opening of Fram Strait .......................................................................................................... 86
8 SUMMARY AND CONCLUSIONS __________________________________________________ 91

APPENDIX ____________________________________________________________________ 95
A1 Abbreviations ......................................................................................................................... 95
A2 Making an Arctic Gravity Grid with GMT ........................................................................ 96
A3 Arctic Catalogue Focal Mechanisms ................................................................................... 98
A4 Main Output Data (CD-ROM) ........................................................................................... 105
REFERENCES _________________________________________________________________ 107
7

1 INTRODUCTION

The Arctic Ocean comprises two deep ocean basins, the Cenozoic Eurasia Basin and the
Mesozoic Amerasia Basin (Figure 1.1). These are separated by the transpolar Lomonosov
Ridge and flanked by one of the world’s broadest continental shelves (Kristoffersen, 1990a).
Spreading between the Eurasian and the North American plate has governed the regional
tectonic style since the early Albian (e.g., Lawver et al., 1990). The present plate boundary
encompasses two large mid-oceanic ridges, the Knipovich Ridge in the Greenland Sea and the
Gakkel Ridge in the Eurasia Basin. The Spitsbergen Transform System, a complex ridge-
transform fault region, connects the two mid-oceanic ridges, hence linking the Eurasia Basin
to the Norwegian-Greenland Sea and the Atlantic. This regional plate geometry reflects two
Cretaceous-Palaeocene1 rift systems that broke up at the Palaeocene-Eocene transition and
were offset by the mega-shear De Geer Zone (Faleide et al., 1993).
The plate boundary north of the Knipovich Ridge, for simplicity named the Arctic
plate boundary, is characteristic in two respects. First, the ultra-slow sea-floor spreading, at
0.26-0.75 cm/yr half-rate2, is the slowest known in the global mid-oceanic ridge (MOR)
system. Second, the plate boundary continues into continental lithosphere in the Laptev Sea,
and the present rotation pole is located in northern Siberia about 500 km further south (e.g.,
Talwani and Eldholm, 1977; Franke et al., 2000). Thus, the outer Laptev Sea margin marks the
transition from a narrow zone of slowly accreting oceanic crust to a broad continental rift
system (Fujita et al., 1990a; Drachev et al., 1998).
The permanent ice cover and the prevailing climatic conditions pose logistical,
technical and economic challenges to geophysical data acquisition in the high northern
latitudes. Thus, the Arctic plate boundary is still poorly studied and has commonly been
defined from regional bathymetry and seismicity (Perry et al., 1985; Eldholm et al., 1990a;
Fujita et al., 1990b). However, refined acquisition methods and the declassification of military
data have significantly increased the available geophysical data bank during the past 5-10
years. The new data include bathymetric maps (HDNO-VNIIOkeangeologia, 1999; Jakobsson
et al., 2000a) and an improved earthquake data base. Moreover, in 1998 and 1999 the US
Naval Research Laboratory (NRL) and the University of Oslo (UiO) conducted airborne
gravity and magnetic surveys in a wide corridor between Svalbard and the Lomonosov Ridge
(Figure 1.1) (Childers et al., 2001). This was part of the joint US-Norway Arctic
Aerogeophysics Program (UNAP) to investigate the evolution of the Eurasia Basin and its
continental margins.

1. Epochs and ages in this study refer to the timescales of Cande and Kent (1992, 1995) and
Gradstein and Ogg (1996).
2. In the following, ‘spreading rate’ will refer to half-rate, i.e. half the plate separation rate.
8

160˚

150˚


140
180

13
nT

na
1200 SIBERIA

Le
East Siberian Sea 200
EAST
150
1 SIBERIAN
100
SEA
75
70

50 New Siberian
3 75˚ Islands
AMERASIA 25
3


0
3

12
D
SIA
E BASIN
IN -25
G
A

-50
EUR

E
BAS

E LAPTEV
R -75 SEA
Z -100
O
N
-90 E Barents -150
-1000
Sea

70

-200
GS 80˚ -1200 11
BB

KR

1
GREENLAND 1
.
R

4
n
oh
1

Makarov
M

3
Basin

2
10
AY
NORW

60
Severnaja

4
e

Zemlya
dg

85˚
60 90˚
1
i
0

sin
v R

3
3
4

Figure 1.1: Main physiographic and


Ba
so

in
structural elements in the Arctic
Ridge

3
80˚

2
no

1
Bas KARA
4

plate boundary region. Bathymetry


mo

2 SEA
from Jakobsson et al. (2000a),
Lo

contour interval 1 km. The magnetic North


Pole 70˚
anomalies show the coverage of the Franz Josef
en

UNAP survey (Childers et al., 2001). Land


ds

The four plate boundary sub-


60˚
un

el

provinces (Chapter 6) are shown


kk
Am

within boxes: [1] Spitsbergen


en
Ga

4
Transform System; [2] West Gakkel
2 Nans

3
2

50˚
1

Ridge; [3] East Gakkel Ridge; [4] 85˚


MJR
Laptev Sea. The Lena Trough is the
4

bathymetric depression in the Fram


Strait between Svalbard and 40
BARENTS ˚
Greenland. BB, Baffin Bay; GS,
Le

YP
SEA
n
aT

Greenland Sea; HR, Hovgård Ridge;


ro
GREENLAND

ug

KR, Knipovich Ridge; MFZ, Molloy SFZ


h

Fracture Zone; MJR, Morris-Jesup MR


SVALBARD
1

80˚ 30
˚
1

Rise; MR, Molloy Ridge; SFZ, MFZ


2

Spitsbergen Fracture Zone; YP, HR 300 km


3
Yermak Plateau.
KR

20
-20˚

10

-10˚

˚
˚

In this study I integrate new and previously published data sets in a regional
analysis of the Arctic plate boundary. The main objectives are: 1) to improve our models and
understanding of the plate boundary location, segmentation and geodynamic character; and
2) to study the Arctic seismicity and its correlation with morphological elements. In
particular, the work includes:
9

• The compilation of Arctic regional geophysical maps and earthquake data;


• Addressing data processing and display problems;
• Describing the plate boundary morphology and location; and
• Discussing plate-boundary-related research topics, such as: 1) properties of the Arctic
earthquake distribution; 2) the effect of ultra-slow spreading on the crustal accretion
process; 3) the transition from sea-floor spreading to continental rifting in the Laptev Sea;
and 4) the opening of the Fram Strait deep-water gateway.
The plate boundary is related to the formation of oceanic crust. Hence, the study area
includes both the Gakkel Ridge and its adjacent basins, as well as the complex plate boundary
extensions to the northern Knipovich Ridge and the Siberian shelf. In terms of structural
setting and geophysical character the studied area is subdivided into four plate boundary
provinces, the Spitsbergen Transform System; the West and East Gakkel Ridge; and the
Laptev Sea (Figure 1.1).
This study is part of the UiO Geophysics Research Group studies of high-latitude
basins and margins, plate tectonics and geodynamics. It is also part of the group’s ongoing
science cooperation with the NRL in the UNAP project, and with NORSAR.
Some of the results are previously presented at the 17th Norwegian Geological
Society Winter Conference in Oslo (Engen et al., 2001). Also, a draft research paper for
submission to an international journal during 2001 has been completed.
Below I first present the geological framework of the studied region, describe the
geophysical data base and show the compilation and properties of an Arctic earthquake
catalogue. The framework and data are then applied in a separate description of each plate
boundary province. Finally, the observations and results are discussed in terms of plate
tectonics and geodynamics. Some data-processing methods are described in the Appendix,
and the Arctic earthquake catalogue and other main output data are available on an enclosed
CD-ROM for easy access.
11

2 GEOLOGICAL FRAMEWORK

This chapter aims at stating the current knowledge of the Arctic plate boundary, i.e. the
background for the subsequent analysis. Intraplate structural elements are also included
because they either originate from the sea-floor spreading or were pre-break-up structures
influencing the plate boundary formation. Since the regional geology depends on the plate
configuration through time, I first establish the plate tectonic controls on the studied area in
space and time.

2.1 Geometric Constraints


The motion of lithospheric plates is constrained to the Earth’s approximately spherical
surface. Thus, Euler’s fixed-point theorem implies that every plate motion is a rotation about
a central axis (Le Pichon et al., 1973). The intersection between the axis and the surface is the
rotation pole, about which flowlines of plate motion form small circles (Figure 2.1). The
present Eurasia-North America rotation pole has been located in northern Siberia by means
of GPS measurements (DeMets et al., 1990) and earthquake stress (e.g., Franke et al., 2000).
However, in this study I apply the anomaly 1-5 stage pole of Pitman and Talwani (1972) and
Talwani and Eldholm (1977) at 68˚N 137˚E, obtained from fracture zone trends and
reconstructed magnetic anomalies in the North Atlantic1. Although this assumes a constant
spreading rate during the past 9.5 m.y. and is a boundary condition from outside the studied
area, it describes the Eurasia Basin plate motion robustly (Reksnes and Vågnes, 1985).

A B

Figure 2.1: Description of plate motion by a rotation ω about an axis. (After Fowler, 1990.)
A. Flowlines (latitudes of rotation) are small circles about the rotation pole.
B. Transform faults are parallel to flowlines, and the relative motion between plates A and B
increases with the distance to the rotation pole. Also, the rotation pole marks the change from
extensional to compressional plate motion along the same boundary.

1. In the following, Arctic spreading rates and flowlines are constructed from the stage poles
and angles of Talwani and Eldholm (1977).
12

The proximity to the rotation pole causes spreading at only 0.26 cm/yr at the Laptev
Sea margin, increasing to 0.75 cm/yr in the Molloy Fracture Zone. This classifies the present
Arctic plate boundary as ultra-slow spreading (Table 2.1). Hence, little space is available for
the accretion of new oceanic crust, and the melt generation process is probably slow (e.g.,
Reid and Jackson, 1981).

Typical topography
Spreading rate Example Reference
of axial regiona

cm/yr

0.3-1 ultra-slow deep axial valley Gakkel Ridge Edwards et al. (2001)
high-relief flanks Southwest Indian Ridge Cannat et al. (1999)
1-2 slow axial low Mid-Atlantic Ridge Macdonald (1982)
rugged flanks

2-4 intermediate (variable) Pacific-Antarctic Ridge Marks and Stock (1994)


4-10 fast axial high East Pacific Rise Macdonald (1982)
smooth flanks

a. The transition from an axial valley to an axial high is commonly diagnostic for the change from slow
to fast spreading. However, variations occur, e.g., the slow-spreading Reykjanes Ridge has an axial
high.

Table 2.1: Classification of mid-oceanic ridges after their spreading rate. The spreading rate affects the
generation of melt in the sea-floor spreading process and hence the relief and geochemistry of
the axial crust (Macdonald, 1982; Bown and White, 1994).

2.2 Plate Tectonic Setting


The onshore geology comprises a number of foldbelts in northern Siberia and Greenland
(Figure 2.4), revealing Precambrian accretion of terranes and the Palaeozoic Uralian and
Caledonian orogenies (Nalivkin, 1973; Dawes and Peel, 1981). Rifting episodes back to the
Devonian and Carboniferous are documented both in Svalbard and the Norwegian-
Greenland Sea (Eldholm et al., 1990b; Harland, 1997). The Amerasia Basin probably
developed from a Mesozoic rift system during the Cretaceous (Embry, 1990). In the
Norwegian-Greenland Sea and Eurasia Basin, the main plate tectonic episodes are (Figure
2.3):
• The opening of the North Atlantic commenced in the Late Cretaceous (~90 Ma) and
propagated northwards. The plate spreading opened the Labrador Sea and Baffin Bay
whereas the proto-Eurasia Basin and Norwegian-Greenland Sea were subject to
continental rifting (Kristoffersen, 1978);
• Sea-floor spreading in the two latter basins commenced at the Palaeocene-Eocene
transition (~55 Ma) and was connected by transform motion along the De Geer Zone.
13

Greenland became a separate plate, forming triple junctions in the Labrador Sea and the
western Eurasia Basin (Kristoffersen and Talwani, 1977; Srivastava, 1985);
• A change in the direction of spreading occurred at the Eocene-Oligocene transition (~33
Ma). Sea-floor spreading in Baffin Bay terminated and the plate tectonic setting changed
from three-plate to two-plate. Continent-continent translation along the De Geer Zone was
replaced by lithospheric extension and subsequent opening of the southern Greenland Sea
(Eldholm et al., 1987; Faleide et al., 1993);
• Sea-floor spreading in the northern Greenland Sea commenced in the Middle or Late
Miocene (Kristoffersen, 1990b; Eldholm et al., 1994).
There is evidence of voluminous, episodic magmatism during the Cretaceous and Palaeocene
both in Svalbard (Maher, 2001), Franz Josef Land (Dibner, 1998) and Greenland (Soper and
Higgins, 1991). Also, the Eurasia-Greenland-North America triple junction was associated
with extensive outpouring of melt (e.g., Feden et al., 1979). Whether this stems from
decompression melting during rifting or reveals an Arctic Large Igneous Province (LIP) is
disputed (Eldholm et al., in press).

Feature Age Length/Width Depth C.T. Reference

Ma km km km

Knipovich Ridge <33 550/ 120 0.6- 2.5 <15 Eldholm et al. (1990a)
Hovgård Ridge <33 170/ 40 1.2- 2.4 20 Myhre and Eldholm (1988)

Gakkel Ridge 55 1800/ 100-200 0.4- 3.4 1-4 Eldholm et al. (1990a)

Amundsen Basin 55 1800/ 250-450 4.3 4-10 Kristoffersen (1990a)


Nansen Basin 55 1800/ 200-500 3.7- 4.1 4-7 Kristoffersen (1990a)

Yermak Plateau 33-41 300/ 150 0.4- 1.4 20 Sundvor and Austegard (1990)

Morris-Jesup Rise 33-41 200/ 100 0.8- 1.2 18 Weigelt (1998)


Lomonosov Ridge >55 1700/ 50-80 0.5- 2.0 28 Jokat et al. (1995a)

Table 2.2: Main structural elements in the Eurasia Basin and northern Greenland Sea. Ages, from
interpreted magnetic anomalies (Vogt et al., 1981), refer to the age of the feature as a separate
tectonic element. Ages of the mid-oceanic ridges refer to the duration of the present axial
geometry. Lengths and widths are estimated averages from the IBCAO bathymetry (Jakobsson et
al., 2000a). Depth ranges also from the IBCAO. Approximate crustal thickness (C.T.) on the
Knipovich Ridge from Austegard and Sundvor (1991); Hovgård Ridge from Karlberg (1995);
Gakkel Ridge from Coakley and Cochran (1998); and Weigelt (1998) elsewhere.
14

2.3 Mid-Oceanic Ridges


The high-relief Knipovich Ridge (Table 2.2) is continuous and spreads at ~60˚ angle with the
axis. It is asymmetrically situated in the eastern Greenland Sea and approaches the Svalbard
continental margin at a low angle, indicating that the axis has migrated during the Cenozoic
(Sundvor and Eldholm, 1979; Eldholm and Sundvor, 1980). Probably, a post-Eocene excursion
of the plate boundary split the Hovgård Ridge microcontinent off Eurasia (Myhre and
Eldholm, 1988; Crane et al., 1988).
The Spitsbergen Transform System translates the plate boundary ~450 km to the
north and is composed of short, oblique-spreading ridge segments connected by dextral
transform faults. Its heat-flow distribution, tight segmentation and steep topography suggests
that axial jumps have occurred (Crane et al., 1982; Eldholm et al., 1990a).
The ultra-slow spreading of the Gakkel Ridge is nearly orthogonal to the axis. The
Eurasian flank spreads 0.06 cm/yr faster than the North American flank, indicating a
persistent asymmetry during most of the Cenozoic (Vogt et al., 1979). The unusually steep
and faulted ridge flanks (Weigelt and Jokat, 2001) are at normal depth except for anomalously
deep areas at either end of the ridge (Figure 2.2) (Vogt et al., 1979). The narrow and almost flat

Figure 2.2: Variation of geophysical parameters along the Eurasia-North America plate boundary north
of Iceland. From top to bottom: Distance to the anomaly 1-5 rotation pole at 68˚N 137˚E (Talwani
and Eldholm, 1977), anomaly 1-5 spreading rate, angle between the ridge axis and the anomaly
1-5 flowline, and maximum plate boundary depth. (After Eldholm et al., 1990a.)
15

axial valley is one of the deepest known (Grachev and Naryshkin, 1978). Most fracture zones
have offsets less than the width of the axial valley, and are magnetically indistinct (Eldholm et
al., 1990a). Gravity modelling indicates that the oceanic crust is locally less than 2 km thick
(Coakley and Cochran, 1998; Weigelt, 1998). Along with core samples this suggests a very low
degree of mantle melting below the axis, causing crust of nearly undepleted mantle peridotite
composition (Mühe et al., 1993, 1997; Hellebrand et al., in press). Recent tectono-magmatic
activity on the East Gakkel Ridge has been documented from axial magmatic constructions
probably related to a large earthquake swarm in 1999 (Müller and Jokat, 2000; Edwards et al.,
2001).

2.4 Basins and Sub-Basins

2.4.1 Greenland Sea


The evolution of the Greenland Sea is strongly influenced by the change in relative plate
motion at the Eocene-Oligocene transition (~33 Ma) (Figure 2.3). During the Late Palaeocene
to Eocene, left-lateral continent-continent translation formed a pull-apart setting in the south
and transpression in the north, initiating the Spitsbergen orogeny in Svalbard and Greenland
(Harland, 1969; Faleide et al., 1993). The change in relative plate motion replaced this setting
by crustal extension and sea-floor spreading (Eldholm et al., 1987). However, the oldest
oceanic crust in the north Greenland Sea is difficult to date because of poorly resolved
magnetic anomalies (Vogt et al., 1981). Further north, the formation of deep-water conditions
was delayed until 7.5-5 Ma (Lawver et al., 1990). This passage, the Fram Strait, is still the only
deep-water connection between the Arctic and the world oceans. Its opening must therefore
have had palaeoclimatic implications (Eldholm et al., 1994), which have been investigated on
legs 151 and 162 of the Ocean Drilling Program (Thiede et al., 1996; Raymo et al., 1999).

2.4.2 Eurasia Basin


The 1800-km-long Gakkel Ridge divides the Eurasia Basin into the Nansen Basin in the south
and the Amundsen Basin in the north. Magnetic anomaly 24B on both sides of the ridge
shows that the basin opening at ~55 Ma was abrupt (Vogt et al., 1979). However, a broad belt
of unknown crust lies between anomaly 24B and the continental shelf. The Eurasia Basin may
therefore have opened even earlier, perhaps before the Norwegian-Greenland Sea (Sundvor
and Austegard, 1990). Sediments only partially cover the ridge axis but increases to ~2 km
average thickness towards the Lomonosov Ridge (Jokat et al., 1995a). However, the Nansen
Basin has received more sediments, reaching 3-6 km thickness in fans that extend basinward
from large troughs in the Barents-Kara margin (Ostenso, 1974; Kristoffersen, 1990a; Vågnes,
1996). The different sediment input on the two flanks is reflected by the bathymetry, showing
16

an almost flat ocean floor in the Amundsen Basin (Jokat et al., 1995a) while the Nansen Basin
shoals towards the sediment sources on the margin (Kristoffersen, 1990a).

2.5 Plateaus and Submarine Ridges

2.5.1 Yermak Plateau and Morris-Jesup Rise


The Yermak Plateau and Morris-Jesup Rise are conjugate marginal plateaus with similar
geophysical characteristics (Table 2.2). They formed as a single feature at the Eurasia-
Greenland-North America triple junction during the Middle Eocene to Oligocene (41-33 Ma),
possibly as a result of hotspot activity (Feden et al., 1979; Jackson et al., 1984).
The Yermak Plateau has been studied by seismic profiling of good to variable
coverage (Sundvor et al., 1982; Kristoffersen and Husebye, 1985; Jokat et al., 1995b;
Andreassen, 1997). Its location at the transition between the western shear margin and the
northern shear-rift margin of Svalbard is reflected in the internal structure of its southern part.
Here, considerably stretched and block-faulted crust (Sundvor and Austegard, 1990), possibly
associated with volcanism (Andreassen, 1997), indicating extensive crustal thinning. Elevated
topography compared with age (Kristoffersen and Husebye, 1985), velocity structure (Chan
and Mitchell, 1982; Weigelt, 1998), and a smooth magnetic field (Jackson et al., 1984) also
favour a continental affinity. The northern part of the plateau may be of oceanic origin
(Jackson et al., 1984), but the continent-ocean transition is only approximately delineated
(Sundvor and Austegard, 1990). The overlying sediments are 1-2 km thick and ice-eroded
(Vogt et al., 1994; Weigelt, 1998).
The Morris-Jesup Rise is less studied than its southern conjugate. However,
individual seismic profiles indicate approximately 1 km thin, eroded sediments overlying
strongly deformed crustal rocks (Jokat et al., 1995b). Gravity modelling suggests that the crust
is of oceanic composition (Weigelt, 1998). The Oligocene to recent sea-floor spreading has
separated the Morris-Jesup Rise and the Yermak Plateau to their current positions ~200 km to
either side of the Gakkel Ridge axis (Vogt et al., 1979).

2.5.2 Lomonosov Ridge


The Lomonosov Ridge rises 3 km above the Amundsen Basin as an elongate, shallow
structure between the Greenland and East Siberian margins (Table 2.2) (Weber and Sweeney,
1990). Its geometric outline resembles that of the Gakkel Ridge (Figure 1.1), and suggests it is
a continental sliver torn off the Eurasian continent by the opening of the Eurasia Basin (Vogt
et al., 1979). Whether the Lomonosov Ridge was a separate plate or attached to Greenland or
North America is unclear (Reksnes and Vågnes, 1985), but plate tectonic reconstructions
indicate a North American affinity since ~80 Ma (Rowley and Lottes, 1988; Lawver et al.,
17

1990). Its aseismic character implies a stable, intraplate environment (Weber and Sweeney,
1990). The ridge is flat-topped (Ostenso and Wold, 1977; Jokat et al., 1992) and may have
experienced erosion by ice-grounding down to about 1 km water depth (Jakobsson, 1999).
The sediment infill pattern on its flanks indicates it was subaerial until about chron 22 (~49
Ma) (Jokat et al., 1995a). Below the 500 m thick marine sediment cover, faults and half-grabens
with terrestrial sediments appear on the Eurasian side (Jokat et al., 1995a; Weigelt, 1998). The
Amerasian side displays prograding, probably terrigenous sediments which support its pre-
rift history as a developing, Eurasian continental margin (Weigelt, 1998).

2.6 Svalbard-Laptev Sea Continental Margin


The margin between Svalbard and Severnaya Zemlya is up to 1400 km wide and comprises
the epeiric Barents and Kara seas. Its crustal structure has been studied only fragmentarily
near Svalbard (Sundvor et al., 1977, 1978; Baturin et al., 1994) and Franz Josef Land (Baturin,
1987), and in the Laptev Sea (Drachev et al., 1998; Franke et al., in press). In the Barents Sea, a
very thin, Tertiary and Quarternary sediment cover overlies progressively younger Late
Palaeozoic to Tertiary strata to the northwest, implying that the Barents Shelf has been tilted
and eroded (e.g., Ziegler et al., 1986). Rasmussen and Fjeldskaar (1996) calculated an Early
Tertiary flank uplift ranging from 0.5 km off Norway to 2 km in Svalbard, associated with the
Arctic rifting and sea-floor spreading. Erosion followed in the Eocene-Miocene. A second
erosional episode occurred during the glacial period, when lowered sea level and flow of a
regional glacier eroded the shelf by ~1 km (Elverhøi et al., 1995; Nilsen, 1996). This resulted in
wide north-south-trending troughs that downcut the margin and develop into the thick fans
in the proximal Nansen Basin (Ostenso, 1974; Kristoffersen, 1990a). Trough and fan systems of
this kind are typical for glacial margins (Vorren et al., 1991; Eidvin et al., 1993).
The Laptev and New Siberian shelf areas constitute a >500 km wide continental rift
system that bridges the Gakkel Ridge and the east Siberian Cherskii Mountains, which
delineate the continental extension of the plate boundary (Fujita and Cook, 1990). The rift is
covered by Maastrichtian to recent sediments whose thickness reach ~10 km in the western
grabens and 6.6 km in the east (Drachev et al., 1998). About 70% of the shelf areas are
shallower than 20 m, and hence low sea-levels during the glacial period may have influenced
the shelf development considerably (Jakobsson, 2000).
19

Figure 2.3: Plate tectonic reconstruction of the


Present Chron 13n (33 Ma) Arctic and North Atlantic, keeping
Chron 5n (9.5 Ma) Eurasia fixed. The reconstruction satisfies
constraints from onshore geological data
but relies on a poorly constrained
continent-ocean transition (COT).
Therefore, the extent of oceanic crust, in
particular in the Greenland Sea, should be
considered approximate. Dates represent
reconstructions to regionally identified
magnetic anomalies and show the
approximate timing of the following
events:
• 135-120 Ma: Opening of the Amerasia
Basin.
• 84 Ma: Opening of the North Atlantic
south of the Charlie-Gibbs Fracture Zone,
propagating north between Greenland
and Canada.
• 53 Ma: After the opening of the
Norwegian-Greenland Sea and the
Chron 21n (47 Ma) Chron 24n (53 Ma) Chron 34n (84 Ma) Eurasia Basin; continent-continent
translation between Svalbard and
Greenland.
• 47 Ma: Minor change in relative plate
motion; more oblique spreading west of
Greenland.
• 33 Ma: Major change in relative motion;
cessation of sea-floor spreading west of
Greenland; onset of sea-floor spreading in
the southern Greenland Sea.
• 9.5 Ma: Minor change in relative motion;
sea-floor spreading commenced in the
northern Greenland Sea.
(After Rowley and Lottes, 1988.)

Chron M0r (120 Ma) 135 Ma


21

500 km

Figure 2.4: Geological map of the Eurasia Basin and adjacent provinces. Ocean basin morphology should be considered recent studies (Eldholm et al., 1990a). Oceanic basement ages from Gradstein and Ogg (1996). (Adapted from
approximate because of sparse data. In particular, note that the plate boundary is considerably changed in more Churkin and Johnson, 1976, in the UNESCO Geological World Atlas.)
23

3 GEOPHYSICAL DATA

Bathymetry, gravity and magnetic data were compiled for the purpose of establishing a base
of all published Arctic data, processed for optimal display of the plate boundary province
where possible. Three types of data were obtained: regional grids; regional and local line data;
and images. Grids constitute the bulk of the data, and are complemented by line data, which
are digital sample values along ship and aircraft tracks. Line data of regional coverage were
gridded to ease data management and display. The gridding and display process involved
some data programming and is described below, followed by a listing of the regional and
local data sets. Image data, which comprise printed maps and visualisations of yet
unpublished grids, are presented in their relevant context. The regional coverage of the new
Arctic data sets led me to investigate a method for calculating sediment thickness, hence this
is included as an additional topic.

3.1 Display of Gridded Data


Grids are constructed from line data by
assigning representative values to equi-
dimensional cells after their distance to the
original sample points. After gridding, the
data represent a digitised surface instead of
the variable coverage of line data. Thus, the
data volume in densely sampled areas is
reduced, but areas with few data points
will contain mostly interpolated values
(Figure 3.1). The total data volume is
reduced further because matrix entries are
stored in the computer instead of irregular,
geographic positions. Gridding of large Figure 3.1: Gridding of bathymetry data off the
coast of Prins Karls Forland, Svalbard for
data sets takes some computing time, but
the IBCAO grid (Jakobsson et al., 2000a).
eases advanced visualisation and math The mesh represents the IBCAO grid with
cell spacing 2.5x2.5 km. The value of each
operations on one or more data sets, e.g., grid cell is determined from the underlying
volume calculations, filtering, colour source data. In areas of sparse data the grid
cell values are interpolated. (After
coding and artificial illumination. Jakobsson et al., 2000b.)
In this study all gridding was
performed with the GMT free mapping and display software (Wessel and Smith, 1991). In
general, GMT operated effectively, but problems emerged when displaying geographical
grids over the North Pole. Because geographical grids are registered in (longitude, latitude)
24

coordinates, the longitudinal grid density approaches infinity towards the poles and causes
distortion (Figure 3.2). To moderate this all geographical grids were projected into a cartesian
coordinate system, i.e., with rectangular grid cells in the plane. The projection involved
additional averaging, but on a regional scale the loss of detail was considered negligible. The
compilation and display of a new Arctic gravity grid with GMT is described in Appendix A2.
GMT includes rapid and robust routines for visualising grids, and comes with a
number of changeable colour tables for this purpose. However, the colour tables had to be
modified in order to emphasise subtle geophysical variations in as different areas as the
Spitsbergen Transform System and the Laptev Sea. I therefore made some efforts to develop
appropriate colour tables for bathymetry, gravity and magnetics, given in Appendix A4.

0˚ 0˚
18 18
˚

˚
86

86
0˚ 0˚
90

90
˚

Figure 3.2: Displaying GMT grids over the North Pole. The original geographic grid (left) is ˚
increasingly distorted towards the pole because the cell density approaches infinity. A
resampling to interpolated cartesian coordinates (right) unveils the geophysical trends of the
data. Data from the UNAP aeromagnetic survey (Childers et al., 2001).
25

3.2 Regional Data Sets


Table 3.1 lists recent maps and digital data sets that allow us to study the regional
morphological character and segmentation of the Arctic plate boundary. In addition,
individual bathymetry and gravity profiles from surface and submarine measurements
contribute locally (Section 3.3).

Data
acquisition
Map scale/

Submarine
Data Contour Digital version References

Airborne

Satellite
interval

Surface
Bathymetry
IBCAO 1 : 8 795 800 Grid 2.5 x 2.5 km x x Jakobsson (2000)
500 m Jakobsson et al. (2000a)
Russian 1 : 5 000 000 x x HDNO-VNII-
200 m Okeangeologia (1999)
NRL 1 : 4 704 075 x x Perry et al. (1985)
500 m
GEBCO 1 : 6 000 000 500 m contours x Johnson et al. (1979)
500 m Jones et al. (1994)
Free-air gravity
UNAP gravity 1 : 1 200 000 Track values x Childers et al. (2001)
2 mGal Grid 2.5 x 2.5 km This study
KMS99 - Grid 2 x 2’ x Andersen and Knudsen (1998)
-
Laxon/McAdoo - Grid 4 x 4’ x Laxon and McAdoo (1998)
-
AGP - Grid 5 x 5’ x x x x Forsberg and Kenyon (1999)
(preliminary) 10 mGal
Magnetics
UNAP magnetics 1 : 1 200 000 Track values x Childers et al. (2001)
25 nT Grid 2.5 x 2.5 km This study
GSC - Grid 5 x 5 km x x Verhoef et al. (1996)
-
DNAG 1 : 6 000 000 x x Kovacs et al. (1990)
Track values

Table 3.1: Key regional bathymetric, gravity and magnetic data sets used to delineate the Arctic plate
boundary. Contour interval refers to the plate boundary province. AGP, Arctic Gravity Project;
DNAG, Decade of North American Geology; GEBCO, General Bathymetric Chart of the Oceans;
GSC, Geological Survey of Canada; HDNO, Head Department of Navigation and Oceanography
(Russia); IBCAO, International Bathymetric Chart of the Arctic Ocean; KMS, Kort- og
Matrikelstyrelsen (Denmark); NRL, US Naval Research Laboratory; UNAP, US-Norway Arctic
Aerogeophysics Program.
26

3.2.1 Bathymetry
The bathymetry data base includes the General Bathymetric Chart of the Oceans (GEBCO)
(Johnson et al., 1979; Jones et al., 1994) whose plotted ship tracks show a relatively good data
coverage in ice-free areas but sparse and unsystematic acquisition in the high Arctic. The
Perry et al. (1985) map added new tracks and previously unpublished maps to the existing
data base, but its data coverage is not shown. Both maps portray the Arctic ocean floor and
MOR province with a resolution of >10 km.
A new Russian map (HDNO-VNIIOkeangeologia, 1999) is contoured from 5-15-km-
spaced depth soundings and hence represents a leap forward in level of detail. Although their
original data points are not available, several Russian contours have been incorporated in the
International Bathymetric Chart of the Arctic Ocean (IBCAO) (Figure 3.9a), which has been

Figure 3.3: Gakkel Ridge portrayed by the IBCAO (Jakobsson et al., 2000a) to the left and the GEBCO
(Jones et al., 1994) to the right. Note that the plate boundary in profile A-B, interpreted as the
deepest point in the axial valley, is mapped 15 km apart in the two data sets. (After Jakobsson,
2000.)
27

published as a test version digital grid (Jakobsson et al., 2000a). The IBCAO is based on the
past 40 years of sea and land surveying, including a comprehensive nuclear submarine data
set. Hence, a combined on- and offshore digital terrain model of 2.5 km resolution has been
achieved, providing the basis for a much improved regional understanding of the plate
boundary and adjacent areas. A comparison between the GEBCO and IBCAO maps is shown
in Figure 3.3.
Due to an erroneous survey line near
the eastern flank of the Yermak Plateau, the 83˚

IBCAO is not correct here (Figure 3.4). However,

3
L
3
no other errors have been reported (M. H
YERMAK H

4
PLATEAU
Jakobsson, pers. comm., 2000). Navigation
uncertainties vary because the submarines have 82˚
H
used inertial navigation, but are nonetheless
3
small for the purposes of this study. For instance, 2
1
uncertainties of ± 600 m in position and ± 0.5 % in
depth are given for the Russian map (Grikurov, N SVALBARD MARGIN 81˚

1999).
20˚

25˚
Minor discrepancies between the Figure 3.4: Non-existing seamount chain off
northern Svalbard in the IBCAO
IBCAO and the Russian map show that their bathymetry (Jakobsson et al., 2000a)
data coverages differ locally. However, both due to a corrupt survey line (M.
Jakobsson, pers. comm., 2000).
maps are highly relevant for the present regional Contour interval 1 km.
analysis and have served as primary bathymetry
sources.

3.2.2 Gravity
Both shipboard, airborne and satellite-derived gravity data were compiled. Shipboard and
airborne surveys measure the total gravity field, which must be corrected by a reference
gravity formula for the effects of the earth’s rotation and polar flattening. The residual field is
then corrected for the motion of the survey vessel and reduced to its corresponding value at
sea-level, yielding a free-air gravity field (Kearey and Brooks, 1991). Satellites apply precise
radar altimetry during repeated orbits to measure the height of the sea surface. When
corrected for dynamic topography due to ocean circulation, the sea surface represents an
equipotential surface of the gravity field, the geoid. The free-air field is computed by Fourier
analysis of the vertical geoid deflection after a filtering sequence to remove the directional
effects of satellite motion (Sandwell and Smith, 1997).
The free-air anomaly is assumed to represent density contrasts in the subsurface.
Because it contains a minimum of assumptions and because marine surveys are carried out at
28

a constant elevation, it is the commonly applied anomaly type in marine studies (Dehlinger,
1978).
Arctic gravity data prior to the mid-1980s were published by Sobczak et al. (1990) as
a regional gridded map. However, its coverage was incomplete and only included the
westernmost part of the Arctic plate boundary. More recently published data are given in
Table 3.1, but still no complete circum-Arctic gravity map is available because of inadequate
satellite orbits and the ice cover which inhibits surface ship operation. In addition, the
irregular ice surface disturbs standard satellite altimetry by scattering radar echoes. Laxon
and McAdoo (1994, 1998) examined repeated swipes during ice-covered and ice-free periods
and computed an ice correction filter that enhanced the readability of the northernmost
satellite soundings. Their gravity grid covers the plate boundary north to 82˚N. However, my
main data source (Table 3.1) has been a gridded version of the UNAP aerogravity data
(Childers et al., 2001), merged with the most recent satellite gravity grid, KMS99 (Andersen
and Knudsen, 1998). The combined data set is shown in Figure 3.9b. The UNAP data were
recorded with GPS navigation and cover about 1000 km of the plate boundary east of the
Spitsbergen Transform System (Figure 1.1). KMS99, which is based on both Geosat and ERS-1
data, has the same geographic coverage as the Laxon and McAdoo (1998) grid, but its
resolution is improved (Table 3.1). It also includes onshore gravity. The two data sets will be
part of the Arctic Gravity Project uniform grid, which is scheduled for publication in
December 2001 and contains a large amount of new data. A preliminary colour image
(Forsberg and Kenyon, 1999) has been consulted on regional trends in the eastern plate
boundary provinces (Figure 3.9b’).

3.2.3 Magnetics
The objective of a magnetic survey is to measure the magnetic effects of the subsurface. To
achieve this both the diurnal variation, the geomagnetic reference field and the elevation of
the vessel must be corrected for (e.g., Kearey and Brooks, 1991). A good correction is only
obtained if the survey is not conducted during magnetic storms, which have high impacts in
the Arctic due to the proximity to the magnetic north pole.
The DNAG summary of Arctic magnetic profiles (Kovacs et al., 1990) shows that the
survey line spacing is relatively good in the western Eurasia Basin but poorer farther east.
These and other profiles have been incorporated in the GSC regional grid (Figure 3.9c), which
was the first grid to cover the entire Arctic plate boundary province (Verhoef et al., 1996). Still,
however, it is not very detailed east of 80˚E, where incorrect navigation in the contributing
Russian data sets is evident. The GSC and UNAP gridded data sets (Childers et al., 2001) have
been my primary data sources. The UNAP data (Figure 3.5) offer better resolution than the
GSC grid, a fact that may be attributed to the accurate GPS navigation of the survey. A
29

merging of the two grids was unsuccessful because of a static shift in the western plate
boundary provinces, and an elaborated correction is beyond the scope of this study.

-150˚

˚
-12

180
˚


84˚
82
84˚

82
˚
0 ˚
-90 3 15
˚

3
3

1
82˚

3
1

3 3

3
120˚

-60˚
1

3
82˚

3
3
1

90˚

1
3
GREENLAND 3

˚
82
3

˚
-30
1
80
˚

3
3
1

˚
1
80
1

78 FRANZ
60

˚ 1 JOSEF
˚

LAND
3
˚
78

SVALBARD
200 km
30˚

Figure 3.5: Ungridded version of the UNAP aeromagnetic data (Childers et al., 2001), collected by NRL
and UiO in 1998 and 99. Bathymetry, in km, from Jakobsson et al. (2000a).
30

3.2.4 Sediment Thickness


Published sediment thickness maps have been constructed from a relatively small number of
seismic profiles. Kristoffersen (1990a) integrated earlier compilations with seismic reflection
data from the LOREX 79 and FRAM experiments (1979-82) and obtained the sediment
distribution map in Figure 3.6. Likewise, Jackson et al. (1990) compiled all previously
published seismic lines from the Arctic Ocean and arrived at a map of similar resolution.
Vågnes (1996), assuming isostasy and the North Atlantic depth/age relationship for oceanic
basement, constructed a sediment thickness map showing the same trends as Kristoffersen
(1990a) and Jackson et al. (1990). However, this map was not calibrated with seismic data.

Figure 3.6: Distribution of Cenozoic sediments in the Eurasia Basin and the Laptev Sea. Large numbers
annotate isopachs, whereas small numbers annotate bathymetric contours. In the shaded areas
off Greenland the estimated depth to magnetic basement is >5 km. (After Kristoffersen, 1990a.)

Computing a new sediment map. Figure 3.6 shows a larger sediment infill in the Nansen
Basin than in the Amundsen Basin. To quantify the effects of this asymmetric sediment load,
e.g., in terms of isostatic compensation, sediment maps of comparable resolution to the new
bathymetry and potential field data are needed. Unfortunately, published seismic profiles
across the Eurasia Basin are still too sparse to significantly update the present sediment maps.
However, I note that bathymetry may be inverted from dense gravity data using sparse
bathymetry soundings as calibration points (Smith and Sandwell, 1994), Similarly, from
sediment thickness at selected fixed points and the new high-quality bathymetry and gravity
data, it is possible to invert the gravity field for sediment thickness over oceanic basement.
31

The observed free-air anomaly is a combined effect of several density contrasts in the
subsurface, and the largest contrast exists at the ocean floor. Still, the sediment-basement
contrast is significant, in particular on oceanic crust, and should affect the free-air anomaly
accordingly. Now the concept is to correct the gravity field for all factors but the effect of the
sediment-basement interface, and then invert this residual field for the depth to basement.
Practical two-dimensional inversion methods have been presented by Oldenburg (1974) and
Granser (1987) and were recently used by Géli and Blanc (1998) in the Barents Sea. A similar
but presumably improved sediment inversion for the Eurasia Basin forms part of my
proposed Dr. Scient. research project. Anyway, I present here the current status of the project,
describing the preliminary steps of the inversion procedure (Figure 3.7):
1. Assigning densities to crustal blocks. A four-layer model with mantle, oceanic basement,
sediments and water is assumed. The oceanic basement is assigned a constant density and
the sediments are assumed to obey an exponential density-depth function due to
compaction, ρ = ρ 0 e – λz , where ρ0 is the surface density, z the burial depth and λ an
empirical compaction constant that has been published for several regions and sediment
types (e.g., Telford et al., 1990).
2. Processing of the gravity field. Deep bodies will produce longer-wavelength anomalies
than shallow bodies (e.g., Nettleton, 1976), and accordingly the effects of Moho
topography can be removed by a low-cut filtering of the gravity field. On the other hand, a
high-cut filtering is necessary to remove noise and because the sediment inversion routine
(step 5) evaluates a Fourier series that only converges for low frequencies (Granser, 1987).

1 2 3 4 5
mGal mGal mGal

WATER

mGal
SEDIMENTS
? ? ? ? ? ? ? ? ? ? ? ? ? ? ? ?

BASEMENT

ho
Mo
MANTLE

Correct for Downward


Band-pass filter Invert
water layer continue

Figure 3.7: A method for calculating sediment thickness from bathymetry and gravity data, assuming a
simple crustal density model. The steps are based on Oldenburg (1974) and Granser (1987).
32

3. Calculating the gravity attraction of the water layer. The problem is forward in nature
because the volume and density of the water layer is known (Parker, 1972). The calculated
effect is subtracted from the processed gravity field. The corrected, residual field is
assumed to reflect only the sediment-basement contrast.
4. Downward continuation. The inversion series converges faster if the basement relief is
calculated relative to the median basement depth (Oldenburg, 1974). Hence, the residual
field is downward continued to this educatedly guessed depth.
5. Inversion. The inversion routine iteratively calculates the basement relief relative to the
median depth. Basically, this is a solution of the inverse problem, i.e., several sediment-
basement interfaces give rise to identical gravity anomalies. However, if the gravity field is
appropriately filtered and a density-depth distribution is assumed as in steps 1-2, the
solution is unique (Granser, 1987).
At some locations the depth to basement is known from seismic profiles, and the
density variation with depth can be estimated from velocity profiles (Ludwig et al., 1970).
These points serve as calibration points where the calculated and observed basement depths
can be correlated. The correlation depends mainly on the precision of the density model. The
individual correlation values may be gridded or contoured to yield a smooth image of the
regional density variation, which may be interpreted further in terms of depositional pattern
and history.

3.3 Local Data Sets


Table 3.2 shows individual surface-ship and submarine measurements that have
complemented the regional grids where the interpolated grid values provide meagre
information on the plate boundary geometry. Some seismic profiles from the plate boundary
province have been published (e.g., Jackson et al., 1990; Weigelt, 1998). However, most of the
local data sets are high-resolution bathymetry and side-scan sonar data from the past 15 years
of surveying. As an example, a detailed survey of the Molloy Deep is shown in Figure 3.8a
(Thiede et al., 1990). Particularly important contributions have come from the SCICEX
programme, in which the Gakkel Ridge has been systematically mapped with submarine-
mounted equipment. The SCICEX bathymetry data have been incorporated in the IBCAO
grid (Jakobsson et al., 2000a), but displays finer details than the 2.5-km IBCAO resolution
allows for. Because the data were available as colour images only (Figure 3.8b), they have
served mainly as a reference during the later stages of the study. The SCICEX data also
include continuous gravity measurements east of the UNAP survey area (Childers et al.,
2001).
33

Figure 3.8: Detailed bathymetry surveys of the Arctic plate boundary.


A. SeaBeam sonar survey of the Molloy Deep (Thiede et al., 1990).
B. SCAMP side-scan and bathymetry recordings from the West Gakkel Ridge, viewed from the
west (yellow arrow). The data were collected during the 1998-99 SCICEX cruises and processed
and displayed by the Hawaii Mapping Resource Group (http://www.soest.hawaii.edu/hmrg/).
34

Province Data type

Bathymetry

Gravity
Seismic
Reference

STS

GR

LS
Coakley and Cochran (1998) x x x

Crane et al. (1995) x x

Drachev et al. (1998) x x


Eldholm and Windisch (1974) x x x x

Franke et al. (in press) x x

Heidland et al. (1995) x x

Jackson et al. (1982) x x

Jokat et al. (1995b) x x x


Myhre and Eldholm (1988) x x

Sundvor et al. (1977) x x x

Talwani and Eldholm (1977) x x x x


Thiede (1988); Thiede et al. (1990) x x x

Weigelt (1998) x x x

Table 3.2: Local, ungridded profiles from the Arctic plate boundary and its rift extension in the Laptev
Sea. Additional profiles off the axis are listed in Jackson et al. (1990). LS, Laptev Sea; GR, Gakkel
Ridge; STS, Spitsbergen Transform System.

Figure 3.9: Geophysical grids and one colour image (B’) processed for plate boundary analysis during
this study. All grids are resampled to a cartesian grid of 2.5x2.5 km cell spacing for smooth
display (~800000 grid cells). Coastlines from Jakobsson et al. (2000a). Polar stereographic
projection. The GMT colour tables applied are given in Appendix A4.
A. IBCAO digital bathymetry and terrain model (Jakobsson et al., 2000a).
B. Free-air gravity anomalies from the UNAP survey (Childers et al., 2001), complemented by the
KMS99 digital gravity model between 70˚ and 82˚N. (Andersen and Knudsen, 1998). The
construction of the grid is demonstrated in Appendix A2. Areas were data are absent are shown
in white. B’. Colour image of the Arctic Gravity Project preliminary free-air grid (Forsberg and
Kenyon, 1999).
C. GSC magnetic anomalies (Verhoef et al., 1996).
35

160˚

160˚
150˚

150˚
160˚

150˚
A B C

˚
˚


140

140

140
13

13
13
m mGal nT
3239 236 3034

50 200
0 150
40
100
30
-500 75
75˚ 75˚ 20 75˚
50


10

˚
-1000 25

0
12

12
12
0 0
-2000 -10 -25
-50
-20
-3000 -75
-30
-100
80˚ 0˚ 0 ˚
80˚ 0˚
-4000 11 80˚ -40 11 -150 11
-50 -200
-5534 -293 -1602

0˚ 0˚ 0˚
10 10 10

85˚ 85˚ 85˚


90˚ 90˚ 90˚

80˚ 80˚ 80˚

70˚ 70˚ 70˚


B’
60˚ 60˚ 60˚

50˚ 50˚ 50˚


85˚ 85˚ 85˚

40 40 40
˚ ˚ ˚

80˚ 30 30 80˚ 30
˚ 80˚ ˚ ˚

200 km
20

20
-20˚

-20˚
10

10


-10˚

-10˚
20
-20˚

10

-10˚
˚

˚
˚
˚

˚
˚
37

4 SEISMICITY

The dynamics and spatial distribution of Arctic earthquakes are less determined than in most
ocean basins due to magnitudes close to the detection limit and a poor circum-Arctic
seismograph network (Fujita et al., 1990b). However, the installation of several permanent
and temporary recording stations as part of the World-Wide Standardised Seismograph
Network (WWSSN) allowed Sykes (1965) to publish a set of significantly improved Arctic
relocations with error estimates. The epicentre distribution outlined new ridge and transform
plate boundary segments. Subsequently, the Arctic seismicity has been updated by Barazangi
and Dorman (1970), Wetmiller and Forsyth (1978), and Fujita et al. (1990b). Moreover,
fundamental contributions are offered by the routine global determinations from the
International Seismological Centre (ISC) and the US Geological Survey (PDE locations).
During the past 10 years several broadband and array stations have been established,
improving the data base significantly. The coverage of Arctic stations is shown in Figure 4.1.

Figure 4.1: Seismological stations that operated for one or more years during the Arctic Catalogue time
span (1955-99). Stations presently operating are indicated by solid dots, whereas closed stations
are represented by open circles. Seismograph networks are shown by shaded areas. Initial year
of operation are given in parentheses. For closed stations the terminal year is also provided. If
several stations existed in the same location, their combined operating years and the name of the
longest operating station are given. Names and locations in northeast Siberia are approximate.
SFZ, Spitsbergen Transform System. (Adapted from Fujita et al., 1990b.)
38

4.1 Data Sources


In this study I have examined earthquakes north of 72˚N from January, 1955 through
December, 1999, with the objective of elucidating plate boundary characteristics. I have
processed available earthquake reports from the NORSAR data bases, attempting to minimise
location errors, and compiled focal mechanisms. The result is an updated and improved
seismicity data base, the Arctic Catalogue (Appendix A3-A4).
An earthquake location contains information on epicentre position and depth
(Figure 4.2), origin time, magnitude, and commonly recording parameters and error
estimates. The NORSAR epicentre data base contains mainly teleseismic events, but also
locally recorded earthquakes in the Svalbard region and Arctic Canada. The major
contributions come from the International Seismological Centre (ISC), the University of
Bergen (BER), Engdahl et al. (1998) and PDE catalogues (Table 4.1). The PDE contains rapidly
published preliminary locations based on a large number of stations. With a somewhat longer
delay, more precise locations are published in the ISC catalogue, which includes readings
from an even larger number of stations. Engdahl et al. (1998) have further refined and
relocated about 100 000 global ISC locations, analysing the seismic waveform. Finally, the BER
catalogue contains records from about 20 stations in mainland Norway, Svalbard, Jan Mayen
and the North Sea region.

Figure 4.2: Earthquake location terminology. Shown here is a large earthquake rupturing the entire
depth of brittle crust, the schizosphere. The hypocentre, or focus, is the nucleation point of the
earthquake, while the epicentre is the point on the surface directly above the focus. The moment
centroid is the torque axis of the total moment release. (Adapted from Scholz, 1990.)

A focal mechanism, or fault-plane solution, is based on waveform analyses from


stations surrounding the earthquake (Figure 4.3). With a large number of stations the mode of
faulting that caused the earthquake can be determined; whether it is normal, reverse, strike-
slip, or intermediate between these end members. Also, the directional components of the
moment release can be estimated, yielding the moment tensor (e.g., Scholz, 1990). The torques
39

of the tensor must balance, and hence the standard earthquake model involves a double
couple of forces (Figure 4.3) (Nakano, 1923). Because reliable focal mechanism solutions need
ample amounts of data, most of the Arctic mechanisms are from relatively large, MS>5.5,
earthquakes (Figure 4.4).
The solutions were divided into manual and the routinely computed centroid-
moment tensor (CMT) solutions, and then grouped according to the predominant mode of
faulting. Most manual solutions, based on analyses of teleseismic and/or local records, have
been compiled by Savostin and Karasik (1981), Jemsek et al. (1986), Fujita et al. (1990a, b),
Franke et al. (2000) and Hicks et al. (2000). The automated CMT solutions are from the
Harvard University (Harvard CMT), complemented by one US Geological Survey solution
from 1997 (Sipkin et al., 1999). Harvard CMT solutions are computed for Mw>5.5 earthquakes
and published four times a year. The CMT method (Dziewonski et al., 1981; Dziewonski and

1 2 3

B C

Figure 4.3. Focal mechanisms. Compiled from Bullen and Bolt (1985), Fowler (1990) and Scholz (1990).
A. Obtaining the focal mechanism of a strike-slip earthquake. The significance of first motions at
stations A-F [1] is applied to divide the source region into compressional (+) and dilatational (-)
quadrants [2]. The quadrants are separated by the fault plane and the auxiliary plane, but which
is which must be determined from geological knowledge, aftershock distribution and
consistency with nearby mechanisms. The double couple of forces is represented by arrows. P,
pressure axis; t, tensional axis. In the final mechanism [3], black conventionally denotes
compression and white dilatation.
B. Model and focal mechanism of a normal earthquake.
C. Reverse earthquake.
40

Woodhouse, 1983; Woodhouse and Dziewonski, 1984) allows for non-double-couple


mechanisms, which imply a finite torque and thus may indicate anelastic deformation at an
accreting plate boundary (Miller et al., 1998).
The composite focal mechanism catalogue (Appendix A3) contains up to six different
solutions for each event, and these may differ significantly (Fujita et al., 1990b). In such cases I
sorted the mechanisms by solution method and year (Table 4.1) and selected the highest-
ranked solution for the analysis in Chapter 6.

Events
Agency/reference Period
Total Filtered
Epicentre locations 2707 1295 1955-99
1. Engdahl et al. (1998) 612 518 1964-95
Sykes (1965) 45 42 1955-62
2. ISC 962 484 1964-96
3. PDE 408 244 1962-99
4. BER 604 2 1980-99
EID 25 1 1995-96
NAO 43 2 1983-99
Misc. 8 2 1980-92
Focal mechanisms 137 1959-99
1. Franke et al. (2000) 19 1983-96
2. Chung and Gao (1997) 1 1987
3. Fejerskov et al. (1996) 2 1991-96
4. Cook (1988) 5 1960-80
5. Jemsek et al. (1986) 5 1964-76
6. Wetmiller and Forsyth (1982) 1 1978
7. Savostin and Karasik (1981) 13 1964-76
8. Stein et al. (1979) 1 1976
9. Sykes and Sbar (1974) 1 1971
10. Horsfield and Maton (1970) 2 1967
11. Lazareva and Misharina (1965) 2 1959
12. Harvard CMT 84 1977-99
Sipkin et al. (1999) 1 1997

Table 4.1: Contributors to the Arctic Catalogue (Appendix A3-A4), ranked after their presumed quality.
The filtered ‘plate boundary catalogue’ (Section 4.2.2), contains only M S ≥ 3.0 events recorded
by at least 12 stations. Focal mechanisms [1]-[11] are manual solutions, while [12] are automated
agency solutions. Agencies with the same priority are listed alphabetically. Agency abbreviations
in Appendix A1.
41

70
40 A 39 B 59
60
Number of solutions

Number of solutions
50
30 28 45
25
40

20
30

12
10 20 16
10 9 13
7
5 10
2 4

0 0
1960 1970 1980 1990 2000 5.0 5.5 6.0 6.5
Year Magnitude (MS)

Normal Reverse Strike-slip Composite

Figure 4.4: Focal mechanism statistics.


A. Time distribution. Note the dominance of strike-slip and composite mechanisms, although most
of the plate boundary is composed of ridge segments (Figure 1.1). The high number of solutions
from 1995-99 is reflects an earthquake swarm from January-August 1999, from which 22 Harvard
CMT solutions are available (Müller and Jokat, 2000).
B. Magnitude distribution. Focal mechanisms are commonly computed for large earthquakes only.

4.2 Epicentre Processing


The NORSAR epicentre catalogue contains 9829 locations for the region and time-period of
interest. However, the many reporting agencies cause a large redundancy. Therefore, the
catalogue had to meet two requirements to facilitate seismological plate boundary
interpretation:
1. Each location must represent only one event; and
2. The location uncertainties must be sufficiently small. Low location quality yields imprecise
events, while high quality yields few events to interpret.
A sequence of processing routines developed and refined at NORSAR (e.g., Lindholm and
Bungum, 2000) addresses these problems by first eliminating the redundancy and then
sorting the events by parameters of interest. I expanded and tailored the routines for this
study during a dedicated one-month stay at NORSAR. The added functionalities include
routines for number-of-stations sorting and display of the individual processing steps, and
resulted in the two-part processing sequence shown in Figure 4.4.
42

N=9829 N=9829 N=4359


1. Elimination of redundancy

Group Pick

1 2 3

Ms/NST
info?
N=1295 N=1307 N=2707
2. Quality sorting

Ms≥3.0 NST≥12

6 5 4
Engdahl et al. (1998) ISC BER NAO
Engdahl et al. (1998) ISC BER NAO
Sykes
Sykes(1965)
(1965) PDE
PDE EID
EID Other
Other
Magnitude (MS): 6.0 5.0 4.0 3.0 2.0
43

Figure 4.4: Action of the two-part epicentre processing sequence, illustrated by the epicentre locations
in the Spitsbergen Transform System. The processing steps are:
1. Compilation of all epicentre locations north of 72˚N between January, 1955 and December,
1999
2. Grouping of locations according to the criteria in Table 4.2
3. Selection of the presumed best-constrained location in each group (Table 4.1)
4. Establishment of the Arctic Catalogue by keeping locations that can be quality evaluated
5. Filtering by number of stations to reduce mean location error (Figure 4.8)
6. Magnitude cut-off to discriminate against artificial sources, resulting in a filtered ‘plate
boundary catalogue’ (Table 4.1).
Numbers of epicentre locations refer to the entire region north of 72˚N. Agency abbreviations in
Appendix A1.

4.2.1 Elimination of Redundancy


The first part of the sequence groups multiple reports from the same event and evaluates the
quality of each report. The grouping procedure takes into account empirical location
uncertainties for each recording year (Table 4.2). Because the uncertainty estimates are aimed
at obtaining a duplicate-free catalogue, it cannot be excluded that some real events may have
been grouped together. The reports in each group are then sorted after their presumed
location quality (Table 4.1). The ranking gives priority to reports with listed magnitudes, focal
depths and location errors, and large numbers of recording stations. Finally, the highest-
ranked report in each group is selected. The grouping extracted 4359 events, yielding a
reporting redundancy of 2.25.
Of the catalogues in Table 4.1, Engdahl et al. (1998) is given priority because of a
refined processing that has lead to systematic differences in event location compared to the
ISC. Their processing involved waveform analysis of regional as well as teleseismic, ∆ > 30˚,
records with azimuthal coverages greater than 180˚. Whereas the ISC locations depend almost
entirely on the first-P and -S arrivals, Engdahl et al. (1998) reanalysed later phases, in
particular with respect to multiple ray paths and effects of plate boundaries. Furthermore,
they corrected travel times with the ak135 reference earth model (Kennett et al., 1995) rather
than the standard Jeffreys and Bullen (1940) model. Although the locations are fewer, they are
considered more precise than the corresponding ISC locations. The Sykes (1965) locations,

Same agency Different agency


Period
s km s km
1955-1969 130 200 4000 400
1970-1979 70 150 700 400
1980-1998 40 100 80 300
1999 20 100 40 200

Table 4.2: Empirical grouping parameters applied to process Arctic Catalogue epicentre locations. If
two locations are within the same time (s) and distance (km) window, they are grouped under
the same event. Even locations from the same agency may be grouped (Figure 4.4).
44

listing location errors and number of recording stations, were assigned the same priority as
Engdahl et al. (1998) because there is no temporal redundancy between them. Despite the
different reporting periods, the location standard deviations are comparable, i.e., ~10 km and
6.0 km, respectively. Finally, the ISC and the PDE locations, based on a larger number of
stations, take precedence over catalogues with a regional or local focus. Most of the PDE
reports are from 1997-99, a period from which ISC reports were not yet available.

4.2.2 Quality Sorting


To obtain a common magnitude reference all local (ML), body-wave (mb), and moment (MW)
magnitudes are converted into surface-wave magnitude (MS). Because a reliable magnitude
conversion table for the Arctic has not yet been published, I applied global linear regression
relationships (Figure 4.5).
MS is a stable measure for Arctic earthquakes, which are generally shallow,
moderately large and teleseismic events. However, surface waves travelling across the COT
may be considerably affected. For instance, the Lg wave (Press and Ewing, 1952), which is a
guided wave in continental crust, is blocked at continental margins (e.g., Mendi et al., 1997).
Surface-wave reflections and refractions is common, arising from the different acoustic
properties of continental and oceanic lithosphere (Bungum and Capon, 1974) (Figure 4.6).

ML = 0.7MW + 1.0

mb = 0.6MW + 2.0

MS = 1.3MW - 2.0

Figure 4.5: Global relationship between earthquake magnitude scales. MW is the only magnitude scale
that describes earthquake size correctly, because it can be related to physical properties; all other
scales saturate at high magnitudes. For the Arctic earthquakes, which plot left of the vertical line,
there is a linear relationship between the magnitude scales. mb, body-wave-; MJMA, Japan
Meteorological Agency-; MS, surface-wave-; ML, local (Richter)-; MW, moment magnitude. (After
Idriss, 1985.)
45

Also, network magnitudes (mb) may suffer a sizeable bias when the number of stations is
reduced (Ringdal, 1976). Consequently, magnitudes may be under-estimated for many Arctic
events. In this study, however, reliable epicentre locations are considered more important
than accurate magnitudes
A common quality reference is
established by keeping only locations
with information on both magnitude and
number of stations. The remaining 2707
locations constitute the Arctic Catalogue
(Table 4.1) and were grouped by number
of stations before display in Figure 4.7.
From this procedure it is evident that a
larger number of stations decreases the
location error. Much of the adjacent,
apparently intraplate, seismicity may
thus result from inaccurate recording by
few stations. The number of stations is Figure 4.6: Refraction of surface waves from two
teleseismic earthquakes recorded at
probably the best diagnostic criterion for NORSAR. The rays are reflected and
refracted at continental margins, reducing
location accuracy.
the amplitudes of the seismic waveform.
Figure 4.7 shows that a regional (After Bungum and Capon, 1974.)

analysis of the plate boundary requires a


better resolution than offered by the 2707-event Arctic Catalogue. I therefore quality
evaluated the events further, intending to keep only events constrained by a 12-15 km
location error, comparable to the width of the axial valley (Jokat et al., 1995b). Accordingly,
the Engdahl et al. (1998) locations require a minimum of 30 stations (Figure 4.8). However, at
this threshold level only 309 events remain, which are too few to aid in the location of the
plate boundary. The number is reduced further because locations with lower priority (Table
4.1) have greater or even no error estimates and consequently require more stations. Besides,
older locations are typically determined only to tenths of degrees, corresponding to an 11x2
km grid at 80˚N. It may be objected that my grouping criteria (Table 4.2) reflect the
uncertainty of the Arctic Catalogue: All events separated by less than 100 km in position and
20 s in origin time is treated as the same event, even when reported by the same agency. In
this respect, Figure 4.4 shows that the grouped locations are often scattered by more than 20
km. However, the higher-ranked locations possess the smallest errors, and thus the grouping
actually improves the accuracy of the catalogue.
46

160˚

150˚


140

13
Number of stations
1-6
7-11
12-538
6 5 4 3 MS

75˚


12
1


80˚ 11


3 10
3

85˚
90˚
3
1

80˚

70˚

60˚

3 50˚
85˚ Figure 4.7: Events recorded
between 1955 and 1999 by
at least one station. For
3

1 40 clarity, only M S ≥ 3.0


˚ events are shown. The
confinement of epicentres
to the plate boundary
improves as the number of
80˚ 30 stations increases. Note the
˚ defocusing of seismicity at
the transition from the
Eurasia Basin to the Laptev
Sea. Bathymetry, in km,
from Jakobsson et al.
20
-20˚

10

-10˚

(2000a).
˚
47

The balancing of a sufficient number 30


N= 600
of stations and a reasonable location error
was approached by trial and error. I

Location error (km)


eventually arrived at the following quality 20

criteria:
1. Retention of events recorded by at least 12
stations, i.e., a mean location error of 10 10

km for Engdahl et al. (1998) locations


(Figure 4.8).
0
2. MS 3.0 high-pass filtering to discriminate
0 100 200 300
against artificial sources. Number of stations
The 12-station threshold for well-constrained
Figure 4.8: Standard error of epicentre
epicentres compares with the 10-station locations as a function of number of
stations for Engdahl et al. (1998)
levels of Sykes (1965) and Engdahl et al. locations north of 72˚N. 12 locations plot
(1998). The resulting ‘plate boundary outside the diagram.

catalogue’ contains 1295 events (Table 4.1)


and is plotted in the Chapter 6 figures.

4.3 Arctic Seismicity Map and Catalogue


It is apparent from Figure 4.7 that seismicity is an excellent tool for locating the plate
boundary in the poorly accessible Eurasia Basin. Plate-boundary-related earthquakes
dominate the Arctic seismicity, but the catalogue is also influenced by locally recorded, small
intraplate earthquakes in Svalbard and Canada (Table 4.3; Figure 4.9a-c). The recently
improved data acquisition is demonstrated by about half of the events being from 1990-99.
Most earthquakes are moderately large, MS 4-5. However, earthquake distributions
are generally self-similar, i.e., the number of earthquakes drops exponentially over all
magnitudes. This is stated in the Gutenberg-Richter relationship,
log N = a – bM , (4.1)
where N is the number of earthquakes greater than or equal to magnitude M, and a and b are
constants (Gutenberg and Richter, 1949). The Arctic Catalogue shows this relationship, being
log-linearly distributed between MS 4.3 and MS ~6.2 (Figure 4.9d). The lower break-down in
log-linearity, MS 4.3, is due to recording limits and is accordingly the completeness threshold
of the catalogue. This magnitude agrees with the MS 4.1-4.3 threshold of Kristoffersen (1982).
However, Figure 4.9c shows that the lower detection limit has decreased with time. Hence,
subsets containing only the most recent events are expectedly complete down to even lower
magnitudes. The upper break-down either indicates that the recording period is too short to
48

sample the population of large earthquakes on the ultra-slow-spreading ridge, or that a


different scaling relation applies here. The b-value of the Gutenberg-Richter relationship
shows the relative abundance of large earthquakes and consequently the amount of stress
build-up before rupture of the source region (e.g., Wyss, 1973). The b-value of the Arctic
Catalogue is 1.1 (Figure 4.9d), which is intermediate between b-values for ridges and
transforms on the Mid-Atlantic Ridge (Einarsson, 1986).

1200 250
1114
A B
1000
200
Number of events

Number of events
800
150

600

100
400 370
294
218 222 240
50
200
141
78
30
0 0
1960 1970 1980 1990 2000 5 10 15 20 25 30 35 40 45 50

Year Number of stations

7
C D
1000
6
Number of events
Magnitude (MS)

5 logN=7.7-1.1MS
100
4

3
10

1 1
1960 1970 1980 1990 2000 3 4 5 6 7

Year Magnitude (MS)

Figure 4.9: Arctic Catalogue statistics. The 1295 events of the filtered catalogue are shown in red and
the additional 1412 events before filtering in blue.
A. Number of recorded events as a function of time, showing the improved recording performance
rather than normal fluctuations in earthquake activity.
B. Number of events as a function of number of stations. 318 events plot outside the diagram.
C. MS as a function of time. The horizontal line marks the completeness threshold found from (D).
Note the lower magnitude limit decreasing in a stepwise fashion with time.
D. Cumulative number of events as a function of MS, showing the Gutenberg-Richter relationship
between the MS 4.3 completeness threshold (vertical line) and MS~6.2.
49

Events
Province
Unfiltered Filtered

Interplate 1799 1019

Arctic plate boundary 1017 628

Mid-Atlantic Ridge 782 391

Intraplate 908 276

Svalbard 361 44

Arctic Canada 221 85


Barents-Kara margin 199 92

Greenland 89 42

North Atlantic 9 3

Other (North American plate) 29 10

Total 2707 1295

Table 4.3: Plate tectonic distribution of events in the Arctic Catalogue. Plate boundary events
predominate, whereas a large number of small, locally recorded intraplate events do not pass the
number-of-stations filtering.
51

5 INTERPRETATION PROCEDURE

Based on the optimised regional and local data sets, the detailed geophysical interpretation
aimed at locating Arctic plate boundary segments with the resolution of the IBCAO and
UNAP data sets. Figures 3.9 and 4.7 obviously outline the oceanic plate boundary in terms of
morphology, crustal magnetisation and seismicity, while the continental rift extension is more
diffuse. The Arctic plate boundary is insufficiently analysed, but has been discussed by
Kristoffersen (1982), Eldholm et al. (1990a) and Coakley and Cochran (1998), among others.
Therefore, I used other slow- to ultra-slow-spreading segments of the MOR system as
analogues (Table 5.1). As a starting point, I assume that the geophysical indicators of the Mid-
Atlantic Ridge (1.5 cm/yr spreading rate) and the Southwest Indian Ridge (0.8 cm/yr) also
apply to the Arctic plate boundary.
The interpretation followed Vogt (1986b) in defining a morphological, magnetic and
seismological plate boundary, which were integrated in a final plate boundary location.
Special efforts were made to evaluate location errors and the capabilities of each data set to
record recent plate boundary shifts.

5.1 Geophysical Plate Boundary Indicators


The morphological plate boundary is commonly represented by a high-relief ridge with a
deep and wide axial valley (Kristoffersen, 1982). Being formed at the plate boundary, ridge
mounts and oceanic plateaus show axial symmetry. The axial valley is continuous on a
regional scale and may have an axial magmatic construction, or intrarift ridge, revealing the
most recent volcanic activity (Macdonald, 1982). Transform segments are outlined by a
distinct transform valley, which is generally more distinct with larger transform offsets and
slower spreading (Fox and Gallo, 1984). Paired ridge mounts or -chains commonly extend
basinward of the transform (Savostin and Karasik, 1981). At ridge-transform intersections
(Figure 5.1) there is a closed-contour nodal basin and high corners because of the special
dynamics of this plate boundary geometry (Fox and Gallo, 1984; Chen, 1992). Axial

Figure 5.1: Typical topography of a ridge-


transform intersection (RTI) on a slow
to ultra-slow spreading ridge (1 cm/
yr). Note the high ridge-transform
(RT) and ridge-nontransform (RN)
corners, and the nodal basin. (After
Fracture zone Transform fault Fracture zone Fox and Gallo, 1984.)
52

topography down to one year of age may be mapped by high-resolution bathymetry


surveying (Edwards et al., 2001), hence revealing minute plate boundary changes.
The free-air gravity field may be considered a representation of bathymetry with
short-wavelength, local features removed (e.g., Smith and Sandwell, 1994). Thus, bathymetric
trends and axial offsets are emphasised. A pronounced, negative anomaly results from the
combined effect of the axial valley and the partially molten, low-density mantle material
below (Dehlinger, 1978). Because fracture zones juxtapose oceanic crust of different age and
thickness, they represent lateral density contrasts and may give rise to basinwide, flowline-
striking gravity anomalies (Louden and Forsyth, 1976). However, recent, small-scale plate
boundary jumps are probably poorly resolved (Smith and Sandwell, 1994).
The magnetic plate boundary is defined by a large, positive axial anomaly arising
from the recently accreted oceanic crust, which is strongly magnetised, relatively unaltered

Data Ridge Transform Resolution Reference

yrs

Bathymetry • Axial valley • Ridge-normal scarps 1-10 Fox and Gallo (1984)
• Axial magmatic on the ocean floor Fowler (1990)
construction • Offset ridge Cannat et al. (1999)
• Elevated flanks segments
• Axial topographic • Flowline-striking
symmetry ridge mounts
• Transform valley
• RTI nodal basins
• High ridge-
transform corners
• Transpressive highs
• Pull-apart basins
(transtensive)

Free-air • Negative axial • Offset anomalies 106 Dehlinger (1978)


gravity anomaly • Edge anomaly
• Positive flank
anomalies

Magnetics • Positive axial • Offset anomalies 105-106 Vogt (1986a, b)


anomaly • Edge anomaly
• Axial symmetry
of anomalies

Seismicity • Epicentres • Epicentres, 10 Sykes (1970, 1972)


• Dip-slip or concentrated at RTI Engeln et al. (1986)
composite focal • Strike-slip or
mechanisms composite focal
mechanisms

Table 5.1: Main geophysical indicators of a slow to ultra-slow-spreading plate boundary. Due to
complex geometries in the Spitsbergen Transform System and small-offset transforms on the
Gakkel Ridge, only some of these indicators are observed along the Arctic plate boundary. In
addition, large variability is commonly observed. RTI, ridge-transform intersection.
53

and little covered by sediments (e.g., Vogt, 1979). Shifts in the symmetric pattern of sea-floor
spreading anomalies indicate offsets across fracture zones. The present magnetic chron 1n has
existed since 0.8 Ma (Cande and Kent, 1992), hence, axial shifts within this time interval may
not be resolved.
The seismological plate boundary is revealed by a concentration of epicentres. Focal
mechanisms indicate the direction of plate motion (Figure 4.3), but anomalous modes of
faulting are common, in particular at ridge-transform intersections (Engeln et al., 1986; Wolfe
et al., 1993). Epicentre location errors are generally larger than navigation errors in the
geophysical data. Because of the short recording period of high-quality earthquake
catalogues, the epicentres show the present sites of deformation, rendering seismicity a
precise tool for mapping recent axial shifts.

5.2 Course of Work


The detailed and integrated interpretation (Figure 5.2) generally yielded overlapping
morphological, magnetic and seismological expressions of the plate boundary and thus a
simple geological model. An example from the Gakkel Ridge is given in Figure 5.3. However,

Framework
- Regional geology
- Synthetic flowlines

Bathymetry
Epicentres

Magnetics Gravity
Focal
mechanisms
Local profiles

Magnetic Morphologic Seismological


plate boundary plate boundary plate boundary

Consistent?

Figure 5.2: Flowchart of the


Integrated plate boundary integrated geophysical
interpretation procedure.
54

local deviations had to be investigated by iterated interpretation. This involved adjusting


each plate boundary location within the uncertainty limits of the data. The few deviations that
remained after successive iterations were considered consistent with a model of recent axial
shifts. In these cases the morphological plate boundary was emphasised and the magnetic
and seismological boundaries plotted as alternate plate boundary locations.
The geological framework constrained the range of possible locations. First, onshore
foldbelts and lines of crustal weakness (Figure 2.4), several of which have been mapped
offshore (Drachev et al., 1998), have probably influenced the location of large fracture zones.
Second, the assumption of rigid plates implies that rotation poles and angles obtained from
North Atlantic magnetic anomalies (Pitman and Talwani, 1972; Talwani and Eldholm, 1977)
are valid also in the Eurasia Basin. Synthetic flowlines were constructed from the rotation
poles and angles and applied: 1) as a test of transform azimuths; and 2) as constraints on the
duration of relative motion along the transform faults.

BATHYMETRY MAGNETICS
AV

30˚
L
30˚ nT
0 L

L H
4

H
km

3k 0 nT
m
FST H
H FST 35˚ 35˚
25 HIC 25
˚ ˚ L

0 nT H
H

T
0n
L 100 nT
AH

10
AV

H
50 km
0'

0'

L
˚3

˚3

30 30
˚ ˚
86

86

L
0'

0'
0'

0'
˚0

˚0
˚3

˚3
86

86
85

85

GRAVITY SEISMICITY H
al

0
mG

m H
AL

30˚ L G 30˚
al
50

H L
35˚ 35˚
25 25
˚ ˚
50
mG H L
al
L
AL
0'

0'
˚3

˚3

30 30
˚ ˚
86

86
0'

0'
0'

0'
˚0

˚0
˚3

˚3
86

86
85

85

Figure 5.3: Geophysical signature of the 13-km-offset FZ 8 on the Gakkel Ridge (Figure 6.5). The
transform fault offsets the axial valley and the potential field anomalies, and is linked to
flowline-striking topography and concentrated earthquakes. The plate boundary is represented
by a bold line, and synthetic flowlines by a dashed line. Bathymetry from Jakobsson et al.
(2000a), contour interval 0.25 km; gravity and magnetics from Childers et al. (2001), contoured at
10 mGal and 25 nT intervals, respectively; and seismicity from the filtered Arctic Catalogue. The
map projection is oblique Mercator about the anomaly 1-5 rotation pole, making the present
spreading direction horizontal. AH, axial high; AL, axial low; AV, axial valley; FST, flowline-
striking topography; HIC, high inside corner.
55

6 THE ARCTIC PLATE BOUNDARY

The interpretation led to the division of the plate boundary into four morphologically distinct
provinces, the Spitsbergen Transform System, the West and East Gakkel Ridge, and the
Laptev Sea (Figure 1.1). The choice of plate boundary location is described separately for each
province, together with reference figures from previous studies. The first-order segmentation
is summarised in Table 6.1 on page 61.

6.1 Spitsbergen Transform System (Figure 6.2)


The plate boundary comprises a series of short ridge and dextral transform segments. In a
regional sense, the province is a ~400 km wide oceanic depression between the Svalbard and
Greenland margins, decreasing to ~100 km in the north. Hence, the ocean basins between the
ridge province and the adjacent margins are narrow. The Spitsbergen Transform System can
be subdivided into four main structural elements, the Molloy Fracture Zone; the Molloy
Ridge; the Spitsbergen Fracture Zone and the Lena Trough (Table 6.1).
A complex plate boundary has
previously been inferred in this region
(Husebye et al., 1975; Savostin and Karasik,
1981; Vogt et al., 1981). However, the
bathymetric map of Perry et al. (1985) outlined
the two southern transforms and a largely
continuous Lena Trough depression farther
north. Eldholm et al. (1990a) interpreted this
depression as the axial valley of a curved and
obliquely spreading ridge (Figure 6.1). The
plate boundary location in this study is largely
similar to that of Eldholm et al. (1990a) except
for the Lena Trough where two 24- and 10-km-
offset transforms are inferred.
The location and morphology of the
present Spitsbergen Transform System may Figure 6.1: The location of the Spitsbergen
Transform System by Eldholm et al.
suggest relatively young features. In this (1990a). Bathymetry from Perry et al.
(1980; 1985), contour interval 400 m.
respect, I note that: The plate boundary is shown by bold
• The complex ridge and basin morphology lines and synthetic anomaly 1-5
flowlines (Talwani and Eldholm,
may reflect steep and closely spaced 1977) by dashed lines. Epicentres are
represented by dots.
basement faults resulting from ephemeral
changes in the plate boundary geometry;
56

• The Spitsbergen Transform System is an area of oblique sea-floor spreading separating the
distinctly different azimuths of the Knipovich and Gakkel ridges. A similar, but continuous
change in azimuth occurs at the Mohn-Knipovich ridge transition farther south (Eldholm
et al., 1990a);
• Sea floor spreading was delayed with respect to the Eurasia and southern Greenland Sea
basins (Kristoffersen, 1990b; Eldholm et al., 1994).
The northern Knipovich Ridge is spreading obliquely at a rate of 0.8-0.9 cm/yr
without large-offset transforms (Eldholm et al., 1990a). The prominent 3.0-3.7 km deep and
~10 km wide axial valley approaches the Svalbard continental margin at a low angle and
becomes entirely buried by lower slope sediments before terminating at the Molloy Fracture
Zone (Eldholm and Windisch, 1974; Sundvor and Eldholm, 1979). The ridge mounts have
acted as a depositional barrier for sediments from the Svalbard margin (Figure 6.3). Hence a
considerable load is exerted on the eastern ridge flank, possibly explaining the greater relief
west of the axis (Desimon and Karasik, 1979) and that the earthquake distribution is skewed
to the east (cf. Byrkjeland et al., 2000). Earthquakes in the axial valley are unevenly
distributed.
The Molloy Fracture Zone is a north-deepening trough that strikes along the
anomaly 1-5 flowline. It reaches 5607 m in the Molloy Deep, the greatest depth in the North
Atlantic. Thiede et al. (1990) interpreted the topography of the Molloy Deep region as a ridge-
transform intersection with a nodal basin, a mantle protrusion and probably a high inside
corner. Moreover, they found that the bathymetric trend of the Molloy Fracture Zone
continues into a depression that may be a relict nodal basin. The fracture zone (FZ) is also
evident in the free-air gravity field as a flowline-striking, negative anomaly that continues
onto the Svalbard and Greenland continental margins, terminating at distinct offsets in the
margin morphology. The Molloy Deep is associated with a pronounced ‘bull’s eye’ gravity
low superimposed on the fracture zone anomaly. The seismicity is focused along the
transform, with somewhat increased activity towards the Molloy Deep. The abundant focal
mechanisms show nearly uniform, dextral strike-slip motion (Harvard CMT; Fejerskov et al.,
1996).

Figure 6.2: Spitsbergen Transform System. Location in Figure 1.1.


A. Bathymetry, contour interval 0.25 km, from Jakobsson et al. (2000a). Epicentres and focal
mechanisms from the plate boundary sorting of the Arctic Catalogue (Table 4.1). The synthetic
flowlines and anomaly 5 position assume the stage poles and rotation angles of Talwani and
Eldholm (1977). AB, Amundsen Basin; HD, Hayes Deep; HR, Hovgård Ridge; KR, Knipovich
Ridge; LV, Lena Valley; MD, Molloy Deep; MFZ, Molloy Fracture Zone; MR, Molloy Ridge; NB,
Nansen Basin; NFZ, ‘Nansen Fracture Zone’; NLTFZ, North Lena Trough Fracture Zone; SFZ,
Spitsbergen Fracture Zone; SLTFZ, South Lena Trough Fracture Zone; YP, Yermak Plateau.
B. Free-air gravity, contour interval 10 mGal, compiled from Andersen and Knudsen (1998) and
Childers et al. (2001).
C. Magnetic anomalies, contour interval specified by the scale bar, from Verhoef et al. (1996).
57


-5˚

-1
-2

-1
-2
L

˚

-5˚


-1
-2

-1
L

-2
A AB 85˚ B 85˚

L
L
82 mGal
H ˚

FZ
L H 84˚ 300

3
H

5
L 50
H H

3
H 40
82 L 83˚
˚ 84H˚ 81
˚
A 30

3
H
3 20
L
82˚
10
1

L 80
˚
H

4
L 0
H
4 L 81˚
-10
4
NFZ

NB 83˚ 79 -20
˚
81
˚ 80˚ -30
2

NL

-40
TF

78
˚
Z

-50
79˚

-300
82˚
77˚
78˚
YP
80
SL
LV

H
˚
T

15˚

10˚
FZ

L
3


L 5˚

-5˚


-1
-2

-1
-2
85˚
81˚ C
2

1 82 nT
˚ 84˚ 1200
HD
79
˚ 200

H L 150
83˚
81 100
3 3 80˚ ˚
R

H
M

75
MD L
SF

50
82˚
Z

25
80
˚
78 0
HR

˚ H
2 81˚ -25

79˚ 79
-50
˚
H -75
H
80˚
2

-100
FZ

-150
L 78
˚
2

77˚ L 79˚ -200


H
RLBG
3-76 78˚
-1200
3

H
77˚
KR

78˚
100 km L
H
15˚

10˚
15˚

10˚

Legend
Epicentres
Plate boundary
0-100
Fracture zone Depth (km)
100-300
Lineament / alternate plate boundary
6.0 5.0 4.0 3.0
Synthetic flowline MS

Synthetic anomaly 5 (9.5 Ma) Focal mechanisms


Profile location
H L Bathymetric high; low Normal Reverse Strike-slip Composite
58

The 70 km long Molloy Ridge between the troughs of the Molloy and Spitsbergen
FZs rises by more than 3.6 km from the adjacent Molloy Deep. It has a clear central magnetic
anomaly and some symmetric lineations that are difficult to date (Vogt et al., 1981). Four
almost identical Harvard CMT focal mechanisms show normal faulting with oblique
components. The earthquake activity is high compared to the other plate boundary segments
in this province. To obtain a good match between morphology, magnetics and seismicity I
locate the plate boundary slightly to the northwest of the Molloy Ridge crest.
The Spitsbergen FZ is in a regional sense a mirror image of the Molloy FZ. Its
southern nodal basin, the Hayes Deep (Thiede et al., 1990), is about 1.1 km shallower than the
Molloy Deep, but has a comparable gravity low. The fracture zone bathymetry is well defined,
but changes into a 2.5 km deep saddle to the northwest where the plate boundary continues
into the Lena Trough. Therefore I have considered alternative plate boundary geometries,
noting that the Russian bathymetry (HDNO-VNIIOkeangeologia, 1999) indicates a
connection between the Molloy FZ and the Lena Trough parallel to the Spitsbergen FZ (Figure
6.4). This feature is, however, not evident in the IBCAO bathymetry (Jakobsson et al., 2000a).
The free-air gravity forms a linear low along the fracture zone, fitting the IBCAO bathymetry
best. Thus, the difference may be attributed to choice in contouring rather than inconsistent
data bases. The fracture zone trend is also sustained by the earthquake distribution, which is
linear and concentrated at the ridge-transform intersection, as for the Molloy FZ. The mode of
faulting is positively transform (Horsfield and Maton, 1970; Conant, 1972; Savostin and
Karasik, 1981). The best fit between data is achieved by locating the Spitsbergen FZ in
agreement with the IBCAO bathymetry (Jakobsson et al., 2000a) and Eldholm et al. (1990a).

RLBG 3-76

KR

Figure 6.3: Line interpretation of the multi-channel seismic profile RLBG 3-76, showing the
progradation of the western Svalbard margin over the accreting Knipovich Ridge (KR). The scale
shows the slopes if sediment velocities of 2.0 and 3.0 km/s are assumed. Profile location in
Figure 6.2. (After Sundvor et al., 1977.)
59

?
Figure 6.4: The Spitsbergen
Transform System plate
boundary location super-
imposed on the Russian
bathymetry (HDNO-VNII-
Okeangeologia, 1999). The
map indicates a possible
plate boundary path between
the Molloy Deep and the
southern Lena Trough
(dashed line). Two distinct
sub-basins, the southern of
which is termed the Lena
Valley, are also indicated.

The Lena Trough is here defined as the ~100 km wide and ~400 km long deep-water
province in the Fram Strait between the Greenland margin and the Yermak Plateau. It
comprises several high-relief basement peaks penetrating a variable sediment cover. The
ridge province separates the trough into two elongate sub-basins. These are most evident in
the Russian bathymetry (HDNO-VNIIOkeangeologia, 1999), where the southern basin is
denoted the Lena Valley (Figure 6.4). I interpret the basins to be underlain by oceanic crust
accreted along an oblique-spreading rift segment cut by a 24- and a 10-km-offset, dextral
transform fault. The transforms, denoted the South and North Lena Trough FZs, respectively,
connect ~4.0 km deep nodal basins and are associated with flowline-striking basement highs
60

of up to 1 km relief. Despite the small offsets, their bathymetry and gravity signatures
resemble those of the longer Molloy and Spitsbergen transforms. Farther north the connection
to the Gakkel Ridge is problematic due to sparse bathymetric and gravity coverage
(Jakobsson et al., 2000a; Childers et al., 2001). Nonetheless, I infer a curved extension of the
northern Lena Trough rift segment, similar to the Mohn-Knipovich ridge intersection
(Eldholm et al., 1990a). On the other hand, I cannot rule out another small-offset transform in
this area. Parallel magnetic anomalies, though of lower amplitude than on the Gakkel Ridge,
are present north of the fracture zone but are absent in the south (Vogt et al., 1979). However,
while the bathymetry indicates an axial valley near the eastern Lena Trough flank, the
magnetic anomalies favour a central location of the plate boundary. I choose the bathymetric
boundary but note that this may be a recent development which has reduced the offset across
the North Lena Trough FZ.
Sea-floor spreading becomes stepwise more oblique northwards in the Lena Trough.
The only available focal mechanism in the area confirms oblique, transform faulting (Savostin
and Karasik, 1981). However, the correlation between the plate boundary location and
seismicity is diffuse, with epicentres distributed over the full width of the Lena Trough.
Although the plate boundary location yields a reasonable fit to the other data sets, detailed
surveying may well yield a more segmented plate boundary agreeing better with the
seismicity.
At 83˚N the Lena Trough broadens into the Eurasia Basin and the trend of the ridge
province changes to that of the Gakkel Ridge. There is a continuous basin passage into the
Amundsen Basin, whereas the entrance to the Nansen Basin is barred by a 20-30 km wide,
elongate high rising more than 1.2 km above the basin. The high, termed the ‘Nansen FZ’ by
Heezen and Tharp (1975), forms a spine-shaped sill between the Yermak Plateau and the
Gakkel Ridge. Its composition is not clear, but its smooth flanks and rugged crest indicate
oceanic crust covered by sediments.

6.2 Gakkel Ridge


The 1800 km long Gakkel Ridge may have many small-offset transforms, 12 of which are
indicated in Figures 6.5-6.6, but is in a regional sense a continuous feature. Spreading is
generally orthogonal, but locally up to 20˚ oblique, at 0.26-0.66 cm/yr spreading rate. The
ridge divides the Eurasia Basin into the 350 km wide and 4.3 km deep Amundsen Basin and
the 400 km wide and 3.7 km deep Nansen Basin (Figure 1.1). The basin asymmetry is also
reflected by an abyssal plain with steep, local basement highs in the Amundsen Basin,
whereas the smooth sea floor in the Nansen Basin shoals gently to the east and towards the
continental margin. A characteristic kink in the ridge trend at ~60˚E marks a change in
Transform fault Spreading
Province Offset Max. age Reference
and fracture zone rate
km cm/yr m.y.

Molloy FZ 112, dextral 0.75 17 Johnson and Heezen (1967)


Spitsbergen Spitsbergen FZ 130, dextral 0.73 13 Johnson and Heezen (1967)
Transform
System South Lena Trough FZ 24, dextral 0.72 10 This study
North Lena Trough FZ 10, dextral 0.71 9 This study
FZ 6 16, dextral 0.64 55 Eldholm et al. (1990a)
FZ 8 13, dextral 0.58 33 This study
West FZ 11 20, dextral 0.55 33 This study
Gakkel Ridge FZ 12 5, sinistral 0.53 55 Eldholm et al. (1990a)
6.5-6.6), which divides the ridge province into two parts.

FZ 13 5, sinistral 0.52 18 Eldholm et al. (1990a)


FZ 15 8, dextral 0.43 33 Eldholm et al. (1990a)
East
FZ 16 16, dextral 0.42 10 This study

Table 6.1: Major transforms along the Arctic Plate Boundary. Spreading rate is calculated from the Eurasia-North America
anomaly 1-5 pole (Talwani and Eldholm, 1977). In the Eurasia Basin, maximum ages of transform faults are estimated
from identified magnetic anomalies offset by the fracture zones (Vogt et al., 1979). In the Spitsbergen Transform System,
where anomalies are more or less absent, ages are interpreted from bathymetry and gravity trends, and the assumption
of a constant spreading rate.
morphology eastwards. I associate the change with a transform region, i.e. FZs 11-13 (Figures
geophysical character and the transition from an accentuated to a smoother ridge
61
62

6.2.1 West Gakkel Ridge (Figure 6.5)


The axial province is wide, ~150 km, but becomes slightly narrower to the east.
Morphologically it can be divided into three compartments connected by the small-offset,
dextral FZs 6 and 8.
The ~200 km wide, massive ridge province between the northern Lena Trough and
FZ 8 corresponds in a plate reconstruction to the Yermak Plateau and the Morris-Jesup Rise
(e.g., Lawver et al., 1990). Its average depth is ~2.8 km, which is shallower than the ridge
segments east of FZ 8. Local peaks and troughs in the 0.8-4.0 km depth range dominate the
topography. Regionally described, the ridge province is a wedge-like feature which becomes
narrower eastwards and was delineated by Feden et al. (1979) as the “Yermak H-Zone”. The
wedge-shape is outlined by elongate valleys whose ~4.0 km depths are comparable to the rift
valley. The rift valley proper is not well defined though. The IBCAO bathymetry (Jakobsson et
al., 2000a) suggests that the plate boundary must either be centrally located in the ridge
province, thereby rising more than 1.5 km relative to the rift valley in the northern Lena
Trough, or turn around the northern ridge flank (Figure 6.2). On the other hand, the Russian
bathymetry (HDNO-VNIIOkeangeologia, 1999) favours a central location (Figure 6.4). This is
compatible with the potential field data, although I observe axial-like gravity lows over the
off-axis troughs. Additional support is gained from the relatively high earthquake activity
with normal and strike-slip focal mechanisms (Harvard CMT; Jemsek et al., 1986). The
concentration of seismicity in the central province and a ~5 km shift in the axial potential field
anomalies may reflect a young, dextral transform fault, FZ 5.
In a reconstructed position FZ 6 corresponds to the eastern flanks of the Yermak
Plateau and Morris-Jesup Rise, and even to offsets in the Svalbard margin morphology.
However, the bathymetric and potential field signature along the flowline trend in the Eurasia
Basin is weak. Its 16 km, dextral offset is also recorded by the -50 mGal free-air gravity
anomaly along the rift valley. The central magnetic anomaly decreases abruptly from >1000
nT to ~700 nT eastwards across FZ 6.
Between FZs 6 and 8 there is a gently S-shaped ridge segment characterised by high-
relief topography. The deep and continuous axial valley consists of about 12-15 km wide
(Jokat et al., 1995b) and 20-130 km long troughs, at some locations reaching 5.0 km depth. The
axial mountains commonly rise to 1.6-2.8 km, i.e., 1-2 km above the ocean floor and 2-3 km
above the axial valley. The axial troughs are separated by sills where the valley floor shoals by
more than 1.0 km. These sills are also imaged by the gravity anomaly. Basinwards they
continue as flowline-striking basement highs which may depict <5-km-offset fracture zones.
However, the earthquake activity along the segment is remarkably low.
63

85˚
86˚
˚
87
˚
90 ˚

85˚
80

86˚
˚
87
˚
90 80
˚

L
H
L

nT
60˚

100 150 200 1200


FZ 13

H
H

60˚
L

50˚

L
4
FZ 12
H

3
L
L

40˚

75
L

50
L
L

FZ 11 50˚

25
30˚
L

0
4 87
H
H
L

FZ 10

-25
H

˚
20
L

-50
FZ 9
H

˚
40˚ 86

-1200 -200 -150 -100 -75


L

H
H

FZ 8
H
H

L
H

10
85˚
4

3
4

H
H

4
L
B

L
L

84˚
30˚
FZ 7
Amundsen Basin

4
H
H


3
˚
C
L

82

˚
-20˚

-10
Nansen Basin

˚
87
L
L

85˚
86˚
˚
4

87
˚
90 80
˚
L

mGal
H
H

60˚
4

300
˚
20
L
H

50
L

50˚
˚
86

40
FZ 6
L

30
40˚
L

20
2
MJR

L
4

10
30˚
H

A
3
L

˚
10
85˚
˚
0

87
H

L
L

FZ 5
L
H
H

-10
L

˚
3 20
4

-20
H

˚
3

86
H

-30
L
L
H

4 ˚
-40
H

10
L

85˚
84˚
100 km

-50
3
L

4
-300

84˚
H

NLTFZ

˚
A

˚ 82
˚
-20˚

-10

82

-20˚

-1

Figure 6.5: West Gakkel Ridge. Location in Figure 1.1 and legend in Figure 6.2. MJR, Morris-Jesup
Rise.
A. Bathymetry (Jakobsson et al., 2000a) and seismicity.
B. Free-air gravity (Andersen and Knudsen, 1998; Childers et al., 2001).
C. Magnetic anomalies (Childers et al., 2001).
64

130˚

130˚
˚

˚
1 20
120
78˚
78˚

100 km
LS

nT

100 150 200 1200


80˚


11

75
82˚
80˚

50
25
84˚

0
10
3

-25
-50
˚
86˚ 90

-1200 -200 -150 -100 -75


11
asin

82˚
L

80˚
Nansen Basin
L

4
Amundsen B

˚
88
H

70˚
4
H

60˚
L
H

50˚
C

86˚
88˚
L
H

84˚

130˚
10
H
H

˚
L

120
H

78˚
FZ 16

mGal
H
4
H

300
H

FZ 15
H

50
4

80˚
L

˚
90
40
H

86˚
H

30
H


11

82˚
FZ 14
H

20
L

C
L
H

10

80˚
L

84˚

0
L

10
H
H

-10
L ˚

-20
88

˚
90
70˚ 86˚
L

-30
H

FZ 13 80˚
-40
H

˚
88

60˚
L

-50

4 70˚
H

FZ 12
-300
3

60˚
FZ 11
50˚ 50˚
A

86˚
88˚
H

86˚
88˚

Figure 6.6: East Gakkel Ridge. Location in Figure 1.1 and legend in Figure 6.2.
LS, Laptev Sea.
A. Bathymetry (Jakobsson et al., 2000a) and seismicity.
B. Free-air gravity (Andersen and Knudsen, 1998; Childers et al., 2001).
C. Magnetic anomalies (Verhoef et al., 1996).
65

The geophysical signature of FZ 8 is similar to FZ 6, with offset potential field


anomalies and well-defined, flowline-striking basement highs (Figure 5.3). However, the
earthquake activity is greater than on FZ 6. One focal mechanism (Jemsek et al., 1986) shows
normal faulting. Such events, however, may be ascribed both to solution inaccuracies and
anomalous faulting near the transform zone proper (Engeln et al., 1986; Wolfe et al., 1993).
East of FZ 8 the plate boundary turns slightly north, but the ridge morphology
resembles the segment to the west. On a regional scale it is not conclusive whether the axial
valley is continuous or broken into ~20 km long, orthogonally spreading segments.
Unfortunately, this question is not resolved by the epicentre distribution or the mixed focal
mechanisms (Harvard CMT; Savostin and Karasik, 1981). In fact, more than half of the events
are located outside the axial valley. I interpret two small fracture zones on this segment,
associated with both sills, flowline-striking topography and axial trend changes.
FZs 11-13 (Table 6.1), marking the transition to the East Gakkel Ridge, are all
associated with flowline-striking basement highs extending more than 150 km into the
Nansen and Amundsen Basins. The 20 km dextral offset of FZ 11 is the largest on the Gakkel
Ridge. The most pronounced bathymetric contrast in the Arctic Ocean occurs at FZ 12 where
the ridge rises to 391 m depth next to a 5000 m deep nodal basin (HDNO-VNIIOkeangeologia,
1999). The potential field data correlate well with the bathymetry, outlining both axial offsets
and the basinward extension of FZs. The focal mechanisms show large strike-slip
components. A characteristic feature is a 40 km left-hand seismicity offset agreeing with FZ 11
(Figures 4.7 and 6.5). The continuous axial valley, on the other hand, appears little affected by
the offset.

6.2.2 East Gakkel Ridge (Figure 6.6)


Sea-floor spreading is largely normal along the ~1100 km long plate boundary segment. It has
a minor trend change at 105-115˚E, a region in which I infer two small-offset transforms, FZs
15 and 16. The amount and quality of potential field anomalies are poorer than farther west,
as the province is not covered by the UNAP survey (Childers et al., 2001). Hence, the plate
boundary location is mainly based on bathymetry.
The ridge segment between FZ 13 and FZ 15 is gently curved. In the west, the
transition into the East Gakkel ridge is associated with lowering of ridge mounts, decrease of
bathymetric relief, change from axis-parallel ridges to discrete ridge mounts, dampening of
magnetic anomalies, and focusing of seismicity. The segment has been studied with
submarine-mounted side-scan sonars and gravimeters during the SCICEX programme
(Coakley and Cochran, 1998). In particular, the survey covers the location of a 209-event
earthquake swarm that occurred from January to August, 1999, near 85˚E. Recent axial
magmatic constructions including local lava flows suggest a relationship between volcanic
66

activity and the earthquake swarm (Müller and Jokat, 2000; Edwards et al., 2001). Moreover,
crustal accretion is indicated by 22 Harvard CMT solutions showing predominantly rift
mechanisms with variably reduced double-couple components. The background earthquake
activity appears low and clustered, however, comprising less than 10 teleseismic events a year
(Müller and Jokat, 2000).
The 15-20 km wide axial valley, at 4.2-4.4 km depth, is fairly continuous and has a
gentle valley floor. The Nansen Basin ridge flank is covered with sediments forming a 4.0 km
deep abyssal plain, which is penetrated by the 2.3-3.2 km deep axial peaks. Sediment
thickness within the rift zone varies from 0.1 to 1.3 km (Grachev and Karasik, 1974),
compatible with gravity modelling which indicates a crustal thickness of only 1-4 km
(Coakley and Cochran, 1998). The linear magnetic anomalies are well-developed but low-
amplitude along the entire plate boundary segment.
FZs 15 and 16 belong to a ~150 km long axial zone where three focal mechanisms
(Harvard CMT; Lazareva and Misharina, 1965; Franke et al., 2000) show strike-slip motion
contrasting with the composite modes of faulting on the adjacent ridge segments. Epicentres
and relatively high, ~1.8 km, ridge mounts indicate that the ridge axis is offset some 10 km to
the northeast between the transforms. There may be several other small-offset transforms, but
the elevated, flowline-striking topography only supports the existence of FZs 15 and 16.
The axial province becomes still narrower and smoother east of FZ 11, and the
sediments on the Nansen Basin ridge flank increase to as much as 4-6.5 km near the Laptev
Sea margin (Kristoffersen, 1990a; Kim and Verba, 1995). An anomalous, 5260 m deep basin at
81˚N is superimposed on the regional ridge bathymetry (Karasik and Pozdnyakova, 1979;
HDNO-VNIIOkeangeologia, 1999). Landward of the basin, the oceanic plate boundary can be
traced morphologically to about 60 km from the Laptev Sea shelf edge (Figure 6.7). High heat
flow and hydrothermal activity on the continental rise support this interpretation (Drachev et
al., 2001). Earthquakes are relatively abundant along the entire ridge segment and well
confined to the axial region. The focal mechanisms reveal fairly distinct zones of either normal
or strike-slip modes of faulting (Harvard CMT; Savostin and Karasik, 1981; Jemsek et al., 1986;
Sipkin et al., 1999; Franke et al., 2000), which may be attributed to a finer segmentation than
resolvable in this study.

6.3 Laptev Sea (Figure 6.7)


The accreting Arctic plate boundary terminates at the Laptev Sea continental margin where it
changes into the continental Laptev Rift System. Onshore geology and seismic profiles from
the shallow Laptev shelf show a >500 km wide rift system in which sediments from the
Maastrichtian to present cover axis-parallel horsts and grabens and form a flat, <100 m deep
67

76˚
78˚
13

˚
80

A NSR
12 L


145
KU
˚
H
12 AR
0˚ New Siberian Islands

140
EL ˚
3
H BSNR
11

SH

MAGE86705
2

1 13

11

WLRB
ULR
ture

87722
rn Frac

SLRB
13
MAGE


Northe

100 km
˚ ˚
76 74
˚
72
76˚

76˚
78˚

78˚

13 13
˚

˚
80

80

0˚ 0˚

B
12

C
12

145 145
˚ ˚

12 12
0˚ 0˚

140 140
˚ ˚
11
11


13 13
5˚ 5˚
11
11


13 13
0˚ 0˚

˚ ˚ ˚
76 ˚ 76 74
˚

74
˚

72
72

mGal nT

-300 -50 -40 -30 -20 -10 0 10 20 30 40 50 300 -1200 -200 -150 -100 -75 -50 -25 0 25 50 75 100 150 200 1200

Figure 6.7: Laptev Sea. The transition from focused to defocused seismicity defines a 60 km wide COT
and a 150-km eastward offset in the axial region (grey box). Lineaments with some evidence of
strike-slip motion are shown by bold dashed lines, and the faults show the extent of major horsts
and grabens (Drachev et al., 1998). The proposed Laptev Sea microplate of Avetisov (1993) and
Franke et al. (2000) is shown by bold dashes. Location in Figure 1.1 and legend in Figure 6.2. AR,
Anisin Rift; BSNR, Bel’kov-Svyatoi Nos Rift; ELH, East Laptev Horst; KU, Kotel’nyi Uplift; NSR,
New Siberian Rift; SH, Stolbovoi Horst; SLRB, South Laptev Rift Basin; ULR, Ust’ Lena Rift.
A. Bathymetry (Jakobsson et al., 2000a) and seismicity.
B. Free-air gravity (Andersen and Knudsen, 1998).
C. Magnetic anomalies (Verhoef et al., 1996).
68

sea floor (Figure 6.8) (Roeser et al., 1995; Drachev et al., 1998). The 15-20 km thin continental
crust is cut by several deep, low-angle faults (Franke et al., 2000, in press). The sediment
thickness is >10 km in the western grabens and reaches 6.6 km farther east (Kim and Verba,
1995; Drachev et al., 1998). Both gravity and magnetic anomalies appear attenuated, but the
main horsts and grabens are clearly imaged by the free-air gravity field (Fujita et al., 1990a;
Drachev et al., 1998).
The earthquake distribution in Figure 6.7 and previous studies (Chapman and
Solomon, 1976; Fujita et al., 1990a,b; Franke et al., 2000) shows a small, right-hand seismicity
offset and an abrupt change from focused to defocused seismicity on the middle to upper
Laptev Sea continental slope at ~78˚N. The change occurs where the ridge axis intersects
lineaments linked to the transcurrent Northern Fracture (Drachev et al., 1998). Here, the focal
mechanisms (Harvard CMT; Jemsek et al., 1986; Cook, 1988; Sipkin et al., 1999; Franke et al.,
2000) show a transition from normal to strike-slip faulting. At the outer shelf there is a
relatively quiet zone which is replaced landwards by defocused seismicity. I relate these
changes to the transition from sea-floor spreading to crustal extension and thinning by
continental rifting. From the seismicity I conservatively estimate the width of the COT to less
than 60 km. Although defocused, the continental seismicity shows a clear N-S trend of
normal-faulting events with variable strike-slip components west of the New Siberian
Islands. This confirms the suggestion of Drachev et al. (1998) that the Bel’kov-Svyatoi Nos Rift
is currently the most active rift segment, and indicates that the plate boundary is offset some

Figure 6.8: Multi-channel seismic profiles MAGE 87722 and 86705 and associated gravity across the
Laptev Rift System. Note the character of the presently active Bel’kov-Svyatoi Nos Rift. Profile
locations in Figure 6.7. (After Drachev et al., 1998.)
A. Satellite-derived free-air gravity (Laxon and McAdoo, 1998).
B. Seismic interpretation. pK, basement; LU, lower seismic unit (Cretaceous-Lower Palaeocene);
MU, middle seismic unit (Eocene-Middle Miocene); UU, upper seismic unit (Upper Miocene-
Holocene).
69

150 km to the west at the COT. A more diffuse trend strikes NE-SW from the COT and shows
transform faulting with small dip-slip components. Also, I note that some events define a
weak N-S trend east of the New Siberian Islands (Figure 4.7).
71

7 DISCUSSION

The preceding chapter pointed at important aspects of the plate boundary morphology and
segmentation. Moreover, the new axis location constrains the structure, seismotectonics and
plate tectonic evolution of the Eurasia Basin and adjacent provinces. Below I attempt to
discuss this in a geodynamic and plate tectonic context. Particularly, I apply analogies with
similar plate boundary settings in the world oceans. The discussion includes the interplate
seismicity, which also constrains the regional structure of the Laptev Sea. Because of the
comprehensiveness of the Arctic Catalogue, some aspects of intraplate earthquakes are also
discussed. Finally, I propose a plate tectonic model for the evolution of the Fram Strait deep-
water passage, derived mainly from the present plate boundary geometry.

7.1 Seismicity of the Arctic Plate Boundary


The Gutenberg-Richter relationship, logN = a - bM (Equation 4.1) generally has high b-values
in MOR provinces (e.g., Francis and Porter, 1971). This observation has commonly been
related to low stress in the source region (Wyss, 1973) and implies that much of the
deformation associated with sea-floor spreading is conveyed aseismically. The common
mechanisms of faulting on ridge segments, i.e., dike intrusion and isostatic uplift of the
depressed axial valley, are associated with frequent but small earthquakes (e.g., Sykes, 1967).
The spreading of the MOR flanks is mainly caused by aseismic, thermal subsidence (e.g.,
Scholz, 1990), with earthquakes probably contributing only 10-20% to the required moment
release (Solomon et al., 1988). Stress build-up before rupture is larger on transforms, but only

Spitsbergen Transform System West Gakkel Ridge East Gakkel Ridge

With 1999 swarm


Without 1999 swarm
1000 1000 1000

logN = 7.7 - 1.2M


Number of events

logN = 6.7 - 1.0M


logN = 7.0 - 1.1M
100 100 100

10 10 10

1 1 1
4 5 6 4 5 6 4 5 6
Magnitude Magnitude Magnitude

Figure 7.1: Cumulative number of events as a function of magnitude for the Spitsbergen Transform
System and the Gakkel Ridge. Magnitude type is not specified because of many converted MS
values. The Spitsbergen Transform System appears to be in the highest state of stress.
72

the longest transforms show a good coupling between slip rates and earthquake stress release
(Brune, 1968; Kanamori and Stewart, 1976). The higher stress level at transform faults is
reflected by the majority of oceanic events being strike-slip, although most of the global MOR
system consists of rift segments (Frohlich, 2001).
From Figure 4.9d I calculated the Gutenberg-Richter relationship for all M S ≥ 4.0
earthquakes in each plate boundary province (Figure 7.1). The sparse events in the Laptev Rift
System constitute an incomplete sample set, but would presumably yield a low b-value
(Scholz, 1990). The three oceanic provinces have approximately log-linear distributions above
the MS 4.3 Arctic Catalogue completeness threshold. The Spitsbergen Transform System has
the lowest b-value, and hence appears to be in a slightly higher state of stress. I ascribe the
change to the larger proportion of transforms in the province, noting that conversion of the
magnitudes of Einarsson (1986) from mb to MS results in b-values of ~0.8 for transforms and
~1.5 for ridge segments of the Mid-Atlantic Ridge.
Figure 4.7 reveals a spatially clustered earthquake distribution. The epicentres tend
to be concentrated at ridge-transform intersections, in particular near the Molloy and
Spitsbergen FZs (Figure 6.2). Although the clustering may be induced by a short recording
period, few seismological stations and incorrect travel-time corrections, it may in fact be a
property of the Arctic seismicity. To evaluate the clustering, I applied the Arctic Catalogue
sorting criteria (Figure 4.4) and compiled a similar catalogue from the Mid-Atlantic Ridge
between Iceland and the Charlie-Gibbs FZ (Figure 7.2). A distinct clustering is observed also

36˚W 32˚W

58˚N 58˚N
e
idg
sR

56˚N 56˚N
jane
k
Rey

Figure 7.2: M S ≥ 3.0 earthquakes on the


Mid-Atlantic Ridge between the 54˚N 54˚N
Charlie-Gibbs FZ (CGFZ) and
Iceland, recorded by at least 12
stations between 1955 and 1999.
CGFZ
Spatial clustering appears both in the
Arctic Catalogue and in this better- 200 km
constrained catalogue. 52˚N 52˚N
36˚W 32˚W
73

in this much larger and better-constrained data set, in particular documented by the
concentration at the Charlie-Gibbs ridge-transform intersections and by an aseismic ridge
segment near 55˚N. Hence, spatial clustering appears to be neither a sampling-introduced
feature nor to be restricted to the Arctic seismicity (Kristoffersen et al., 1982; Einarsson, 1986;
Bergman and Solomon, 1990).
The earthquake concentration at ridge-transform intersections probably reflects a
relatively high stress build-up at the transition between extensional and transform motion.
Support for this is gained from the high inside corners, which are formed by upward flexure
of the lithosphere by a twisting moment exerted along the transform (Chen, 1989).
Microearthquake surveys at the Mid-Atlantic Ridge have revealed a high earthquake activity
with diffuse, triangular-like distributions cutting the inside corners, and a similar behaviour
may be observed on the Arctic plate boundary (Figure 7.3). Probably, this reflects internal
deformation of the corner along off-axis faults rather than a diffuse plate boundary at the
ridge-transform intersection (Rowlett, 1981).
Although the seismological and bathymetric expressions of the Arctic plate
boundary generally agree, two significant deviations occur. The epicentre distribution in the
Lena Trough is defocused over an area three to four times the width of the axial valley, and
several events are located north of the ridge west of FZ 11 (Figures 4.7 and 6.5). The latter
offset is not associated with systematic location errors. Rather, I note that slow-spreading
ridges, e.g., the Kolbeinsey Ridge north of Iceland, may respond to small perturbations in the
relative spreading direction by ridge propagation, asymmetric spreading and lateral ridge
2˚W

6˚E
4˚E
2˚E

79˚
30'
SF

N
Z

80˚
3
MF

H L 00'N
Z

79˚ 3
00'
N
R

H
M

4
L 79˚
30'N
8˚E

30 km 30 km

Figure 7.3: Seismicity at ridge-transform intersections.


A. Microearthquake survey at the eastern intersection of St. Paul’s Transform and the Mid-Atlantic
Ridge (3.2 cm/yr spreading rate). Epicentres are denoted by dots and ocean bottom
seismographs by squares. Depths greater than 4206 m are hatched, whereas stippled areas are
shallower than 3475 m. (After Francis et al., 1978.)
B. Arctic Catalogue epicentres at the Molloy Ridge (MR), spreading at 0.75 cm/yr. Bathymetry, in
km, from Jakobsson et al. (2000a). MFZ and SFZ, Molloy and Spitsbergen fracture zones,
respectively.
74

South Lena Trough FZ


North Lena Trough FZ
Spitsbergen FZ
Molloy Ridge

arm
Molloy FZ

9 sw

FZ 10
FZ 11
FZ 5

FZ 6

FZ 7
FZ 8
FZ 9
199
2000 Figure 7.4: Temporal distribution
of earthquakes along the
Arctic plate boundary.
1990 Swarms are common along
the entire plate boundary.
Earthquake activity is largest
in the Spitsbergen Transform
1980 System, where the distri-
bution may indicate earth-
quake episodicity.
1970

1960

0 500 1000 1500 2000 2500


km

migration (Appelgate, 1997). Thus, an initially continuous axial valley may develop into a
segmented geometry first reflected in the seismicity and later in the bathymetry and potential
field anomalies. Also, the offset near FZ 11 may indicate a recent episode of a long-term,
northward ridge migration (cf. Stein et al., 1977). If this is the case, the fact that the Nansen
Basin is wider than the Amundsen Basin may be accounted for.
In the Lena Trough, the crustal structure and properties are poorly known (Snow and
Hellebrand, 2001). The shallow events occurring in the proximity of a weakly constrained
COT off Greenland and the Yermak Plateau may have led to biased errors in the up to 45-
year-old epicentre locations. On the other hand, the defocused seismicity is a common feature
of both the Lena Trough and the exceptionally well-defined Knipovich Ridge valley (Eldholm
et al., 1990a). The off-axis seismicity along the Knipovich Ridge suggests stress release over a
wide region, and may be attributed to the following:
• The position of the spreading axis has probably shifted during the Neogene, because the
present Knipovich Ridge is asymmetrically located in the eastern Greenland Sea (Sundvor
and Eldholm, 1979);
• On the adjacent margin, the continental crust was downfaulted after the Eocene. Recent,
rapid loading by glacial fan construction has enhanced the local stress field (Byrkjeland et
al., 2000);
75

• Høgden (1999) observed a seismicity pattern consistent with incipient plate boundary
formation east of the Knipovich Ridge.
From the similar margin geology and oblique spreading I consider these factors relevant also
for the Lena Trough.
The earthquake activity varies during
the sampling period (Figure 7.4). Maximum 1e+25 1.0
earthquake activity occurs in the west,

Spreading rate (cm/yr)


particularly in the southern Spitsbergen

M0s (erg/yr)
Transform System, which is characterised by
large-offset transforms. From the b-values in 1e+24 0.5

Figure 7.1 I calculated the annual seismic


moment release ( M 0s ) to test the dependency on
spreading rate, which indicates the total
1e+23 0.0
moment of the spreading plates. However, there STS WGR EGR
Max. magnitude: 6.1 5.7 6.2
is no apparent correlation (Figure 7.5), possibly
because of a different number of large Figure 7.5: Seismic moment release rate
(boxes) and half-spreading rate on
earthquakes which contribute most to the midpoint (dots) for each oceanic
plate boundary province. MS of the
summed seismic moment. On the other hand,
largest recorded earthquake in each
the poor correlation may reflect different seismic province is indicated. EGR, East
Gakkel Ridge; STS, Spitsbergen
coupling and fault properties (Scholz, 1990). Transform System; WGR, West
Because seismic moment depends on fault area, Gakkel Ridge.

the different M 0s values may also reflect


differences in crustal thickness.
In addition to the spatial clustering discussed above, there is also a temporal
clustering of earthquakes (Figure 7.4). However, to avoid searching for trends and periodicity
in a random data set, I first tested Figure 7.4 for non-randomness (Wonnacott and Wonnacott,
1985). In a random sample of n observations the median will be intersected R times with
expectation and variance of
n
E(R) = --- + 1 (7.1)
2
and
n(n – 2) n – 1
var(R) = -------------------- ≈ ------------ , (7.2)
n(n – 1) 4
respectively. Assuming a normal distribution of R, the hypothesis that the earthquakes are
randomly distributed in time is only rejected at less than 85% level of significance. Bearing
this in mind, I observe that large events commonly recur at certain intervals (Figures 4.9c and
7.4). In fact, even events recorded before 1980 tend to coincide in time and space, suggesting
76

frequent earthquake swarms. The general swarm nature of MOR earthquakes has been
documented by Sykes (1970), among others, and swarms on the Arctic plate boundary by
Kristoffersen et al. (1982) and Müller and Jokat (2000).
All Arctic plate boundary segments deform seismically (Figure 7.4), without distinct
seismicity gaps, cf. Figure 7.2. The Spitsbergen Transform System may have a characteristic
~12-year cycle of interchanging high and moderate earthquake activity, the peak activity
occurring around 1973, 1985 and 1996. If so, it agrees with the periodicity of Mohn Ridge
earthquakes (Lindholm et al., 1990). The Gakkel Ridge does not exhibit any consistent
recurrence pattern, but the average recurrence period is between 5 and 10 years. An
uncommonly long interseismic, or quiet, period is found on the ridge segment between FZs 8
and 11, where no large earthquakes have occurred since 1987. The off-axis seismicity in this
area (Figures 4.7 and 6.5a) indicates that the ridge segment is an anomalous part of the plate
boundary.
The 1999 swarm east of FZ 13 (Müller and Jokat, 2000) is clearly imaged in Figure 7.4
as the single largest cluster of plate boundary earthquakes. The large number of events
occurring on an otherwise relatively quiet ridge segment implies that the region may be
abnormal in terms of stress build-up and crustal properties. However, the magnitude-
frequency distribution of the swarm has approximately the same b-value as the total Gakkel
Ridge seismicity (Figure 7.1). Accordingly, the region does not appear to be under anomalous
stress.

7.2 Intraplate Seismicity


Table 4.3 shows that plate boundary earthquakes dominate the Arctic Catalogue, but a
number of well-constrained events from the continental shelf areas are also present. On the
northeastern Svalbard margin some relatively large earthquakes overlap with pronounced
morphological features (Figure 7.6).
Six MS 3.6-4.7 earthquakes are located on the continental slope, four of these on the
Franz-Victoria Fan. The fan was rapidly deposited because of marginal downcutting and
mass transport in the Franz-Victoria Trough during the glacial period (Ostenso, 1974; Elverhøi
et al., 1995). Its thickness may exceed 3 km (Kristoffersen, 1990a). A high, >50 mGal, free-air
gravity anomaly embraces the bathymetric outline of the fan (Figure 7.6), suggesting it exerts
a large, uncompensated load on the marginal crust (Vogt et al., 1998). Hence, the most likely
earthquake trigger mechanism is the reactivation of break-up-related faults due to increased
vertical and flexural stress (Stein et al., 1989; Byrkjeland et al., 2000). Also, brittle fracturing
during sediment compaction and diagenesis may contribute (Bjørlykke, 1995). The focal
mechanism of event 1, which occurred simultaneously with event 2, shows a compressional
77

40˚
30˚
˚
20
˚
10
4 mGal
4 L MS
300
6.0
5.0 10 84˚
1
82 YERMAK 4.0
PLATEAU H -20
˚
3.0

4
3
-300
2 L

3 100 km
FZ 6

FRANZ- 83˚
5 VICTORIA
H FAN
81 4
˚ 6 1
3
3

2
2

1 82˚

FRANZ
80

-
˚

VICTO
8
L FRANZ
81˚
RIA TR
JOSEF
7 LAND
NORDAUST-
OUGH

LANDET
40˚
30˚

Figure 7.6: Earthquakes in northern Svalbard (Nordaustlandet) and adjacent marginal areas. UNAP
free-air gravity (Childers et al., 2001) is superimposed on IBCAO bathymetry (Jakobsson et al.,
2000a), with 0.25 km contours. The stippled area is erroneous, cf. Figure 3.4 (M. Jakobsson, pers.
comm., 2000). Note the possible prolongation of the FZ 6 onto the margin and Nordaustlandet.

mode of faulting (Savostin and Karasik, 1981). Events 5 and 6, although not overlapping a
high gravity anomaly, may also be explained by sediment loading.
Two events, 7 and 8, are located in Nordaustlandet, which is one of three major zones
of earthquake clustering in Svalbard (Mitchell et al., 1990). Bungum and Kristoffersen (1980)
noticed that the clustered seismicity of this kind depends on pre-existing zones of weakness,
regardless of the stress-generating mechanism. In fact, both events 7 and 8 and locally
detected events correlate well with mapped faults (Høgden, 1999). Moreover, the landward
extension of FZ 6 matches tectonic lineaments on the Svalbard coast and margin, and suggests
a relationship between the onshore active faults and the initial segmentation of the Gakkel
Ridge.
78

7.3 Crustal Accretion and Plate Boundary Segmentation


The regional trend of the Arctic plate
Nansen Basin Amundsen Basin
AXIS
boundary changes ~40˚ at the Lena Trough- 400
A UNAP
200
Gakkel Ridge transition. A similar, ~70˚

nT
0
change exists at the Mohn-Knipovich 80 -200
UNAP

mGal
intersection, where the present boundary 40
0
curves gently without significant offset
-40 0
IBCAO
(Figure 1.1). In a plate tectonic sense, both
2

km
these areas bound the opening of the
4
Greenland Sea. Sea-floor spreading in the
south commenced in the Early to Middle -80 -60 -40 -20 0 20 40 60 80

400
Miocene, and in the north in the Miocene
B UNAP
200

nT
(Kristoffersen, 1990b; Eldholm et al., 1994). 0
The continuous Mohn-Knipovich rift valley 80
UNAP
-200
mGal

40
is probably a young feature evolving from an
0
earlier transform region (Eldholm et al., -40 0
IBCAO
1990a). The curved axial geometry may be a
2

km
consequence of hot, ductile lithosphere.
4
However, the region of young crust in the
-80 -60 -40 -20 0 20 40 60 80
Lena Trough has both high heat flow
400
(Sundvor et al., 2000) and distinct, small- C GSC
200

nT
offset transforms (Figure 6.2). Assuming a 0
80 -200
similar evolution for the two areas, it may be Pogy D-1
mGal

40
inferred that the rift axis in the younger Lena 0

Trough region has not yet reached complete -40 0


IBCAO

continuity. 2
km

A smaller, but important trend 4

change occurs at ~60˚E, where the FZ 11-13


-80 -60 -40 -20 0 20 40 60 80
region divides the Gakkel Ridge into a high- km
relief western and a smoother eastern Figure 7.7: Magnetic, free-air gravity and
bathymetry profiles along anomaly 1-5
province (Figure 7.7). The change is also flowlines across the Gakkel Ridge. The
recorded by more subdued magnetic geophysical character changes from west
(A) to east (C). The values are projected
anomalies in the east. The larger sediment from gridded data sets (Table 3.1), except
the free-air-corrected Pogy D-1 profile
thickness in the eastern province, commonly
from Coakley and Cochran (1998). Profile
related to the proximity of the large Barents- locations in Figures 6.2, 6.5 and 6.6.

Kara and Laptev margin source areas


79

(Kristoffersen, 1990a), may have depressed the oceanic basement, thus smoothing the axial
relief. Assuming Airy isostasy and standard densities, I calculate that a 0.7-km increase in
sediment thickness may account for the average 0.2-km lowering of basement topography
near the axis. The increased basement depth may also contribute to the decreased magnetic
anomalies (Kovacs and Vogt, 1982).
The relief of the Amundsen Basin flank also decreases to the east, despite less
sediment cover (Kristoffersen, 1990a). The more elevated topography in the western ridge
province has been related to thick crust above a Yermak hot spot (Feden et al., 1979; Weigelt
and Jokat, 2001), whereas the lower topography in the east may reflect regionally denser or
thinner crust, as observed near FZ 15 (Coakley and Cochran, 1998). However, the velocity of
the >100-s-period Rayleigh and Love waves, which sample the upper mantle, does not
indicate lateral temperature variations beneath the Gakkel Ridge (Levshin et al., 2001). The
sparse rock samples and measurements of crustal thickness and heat flow preclude a detailed
comparison of the two ridge provinces. However, mantle heterogeneity along the Gakkel
Ridge is indicated by nearly undepleted mantle rocks near FZs 8 and 12, the FZ 12 sample
having the lowest Nd and Sr isotopic ratios (Mühe et al., 1993, 1997). The character of the
seismicity does not contrast the provinces significantly, although the epicentres are slightly
more focused and the seismic moment release rate appears higher in the east (Figure 7.5).
The ultra-slow spreading along the Gakkel Ridge lead me to consider whether the
two ridge provinces may reflect properties caused by the crustal accretion process. However,
linear magnetic anomalies, thin crust (Coakley and Cochran, 1998) and axial magmatic
constructions including local lava flows (Edwards et al., 2001) indicate a sea-floor spreading
framework. I therefore note that the topographic variation may reflect differences in the
production and emplacement of melt in the axial region. The production of melt in the upper
mantle depends mainly on temperature (Su et al., 1994), fluid content and pressure
conditions, i.e. a cold and dry mantle under high pressure will produce little, relatively
unfractionated melt and hence thin oceanic crust. In addition, lithospheric stresses (Shah and
Sempéré, 1998) and near-surface melt migration and magma plumbing (Grindlay et al., 1998)
may greatly modify the crustal accretion process and hence the topography.
The influence of spreading rate on the accretion process may be two-fold. First,
although crustal thickness appears fairly independent of spreading rate, slower spreading
tends to yield a higher topographic relief (Chen, 1992; Bown and White, 1994). This may be
explained by along-axis thickness variations of up to 3 km (Lin et al., 1990). Probably, the
variations arise from diapiric mantle upwelling, which may be more important beneath
slower-spreading ridges (Parmentier and Phipps Morgan, 1990). Thermal buoyancy effects
may also contribute by cooling and mixing depleted and undepleted mantle material. Thus,
80

Figure 7.8: Mantle flow model of along-axis segmentation at slow- to ultra-slow spreading ridges.
Persistent, diapiric mantle flow causes regions of thicker crust above the upwelling centres and
little melting and crustal accretion elsewhere. (After Bell and Buck, 1992.)

even small perturbations in mantle temperatures may cause relatively large thickness
variations (Su et al., 1994). Such perturbations are not necessarily reflected by the regional
surface-velocity structure (Levshin et al., 2001). Because the lower crust is too cold to undergo
rapid ductile flow, the diapiric mantle convection cells may become stationary with respect to
the axis and maintain the crustal thickness variations through time (Figure 7.8).
Second, at ultra-slow spreading rates, the axial upward flux of mantle material may
be low enough that conductive cooling reduces melting (Reid and Jackson, 1981; White et al.,
2001). The relatively cold axial region and low dynamics of the system may also impede the
circulation of hydrothermal fluids. As a consequence, the melt emplacement is highly
episodic and may be insufficient to fill the space created by the separating plates (Figure 7.9).
Hence, the oceanic crust may be block-faulted and thinned after its accretion, resulting in a
wide axial valley as in Figure 7.7c (Louden et al., 1996). This promotes the existence of a
critical spreading rate. The changes across the Melville FZ on the Southwest Indian Ridge,
spreading at ~0.8-cm/yr, may be ascribed to this concept (Cannat et al., 1999). Applying this
analogy, I infer that the ~0.5 cm/yr spreading rate near FZs 11-13 may represent a threshold
level for melt production. Thus the additional sediment loading effect in the eastern province
is superimposed on an axial morphology governed by the ultra-slow accretion process
(Figure 7.9).
The inheritance of pre-existing zones or lineaments of crustal weakness on the large-
scale plate boundary segmentation is demonstrated by the De Geer Zone, which has
governed the loci of two major shifts in the regional plate boundary azimuth (Eldholm et al.,
1987; Faleide et al., 1993). Pre-opening features within this region have influenced the location
of the fracture zones in the Spitsbergen Transform System, which correspond to offsets in the
Svalbard and Greenland margin morphology. In the Eurasia Basin a similar relationship is
indicated by the Laptev Sea COT; the basin-wide gravity signature of FZ 12; and by the FZ 6
81

West Gakkel Ridge East Gakkel Ridge


(>0.5 cm/yr) (<0.5 cm/yr)

AB NB AB NB

Figure 7.9: Combined sediment and accretion model for the two distinct Gakkel Ridge provinces. In the
eastern province the spreading rate is low enough that conductive cooling from above nearly
counterbalances the upward flow of hot mantle material. As a result, the melt supply is
insufficient to fill the space between the separating plates. The oceanic crust is stretched and
faulted, causing a low relief. In the western province the melt supply keeps pace with plate
spreading. Additionally, the thicker sediment cover in the east smooths the sea-floor topography
and depresses the oceanic crust. AB, Amundsen Basin; NB, Nansen Basin.

flowlines connecting the eastern flanks of the Yermak Plateau (Figure 7.6) and Morris-Jesup
Rise.
Most other small-offset fracture zones, of which more than those mapped in Figures
6.5 and 6.6 may exist, are probably recent features. I note that Martinez and Cochran (1988)
suggested that pre-opening structures have directed upper mantle convection cells in the
young Red Sea, thus governing also the small-scale axial segmentation. On the other hand,
even long-lived fracture zone clusters could be caused primarily by local thermal minima
between mantle convection cells (Bonatti, 1996). Closely spaced, small-offset fracture zones
may also have formed as cooling cracks during the ultra-slow crustal accretion (Schouten et
al., 1985; Sandwell, 1986) or in response to ephemeral changes in plate motion (Appelgate,
1997). Hence, the fracture zone spacing appears to be controlled by near-surface accretion
processes rather than the size and spacing of diapiric mantle convection cells (Grindlay et al.,
1998).

7.4 The Laptev Sea Continent-Ocean Transition


The Laptev Sea margin is in many respects analogous to the Senja margin in the SW Barents
Sea (Figure 7.10). Both margins were formed where the plate boundary intersects and
becomes offset by a regional tectonic lineament, the De Geer Zone between NE Greenland
and Svalbard, and the Northern Fracture in the Laptev Sea. The Laptev Sea earthquake
distribution changes from focused to defocused over a not more than 60 km wide area with
82

strike-slip focal mechanisms (Figure 6.7). This, and the presently active Bel’kov-Svyatoi Nos
Rift (Drachev et al., 1998), are consistent with a 150 km offset of the plate boundary where the
Gakkel Ridge terminates. By analogy with the Senja margin I infer that the offset reflects a
sheared margin segment. Moreover, I predict a structurally distinct COT even narrower than
indicated by the seismicity (cf. Breivik et al., 1999). Structural and seismic interpretation
shows an eastward migration of the ridge axis relative to the rift (Drachev et al., 1998).
Accordingly, the sheared nature of the margin may have existed since the opening of the
Eurasia Basin. The sharp COT was also indicated by Franke et al. (2000), who found
maximum focal depths and a decrease in the anelastic attenuation of shear waves south of
~78˚N (Figure 7.11). The change was ascribed to a drop in partial melting and interpreted as a
~100 km extension of the oceanic rift into the continental margin.
The width of the Eurasia Basin decreases only slightly to the east (Figure 3.9),
reflecting the varying distance to the rotation pole since opening, and linear magnetic
anomalies continue to the Laptev Sea continental slope (Karasik, 1974). This implies that: 1)
the entire Eurasia Basin opened almost simultaneously, splitting the narrow Lomonosov
Ridge continental block off Eurasia at ~55 Ma (e.g., Lawver et al., 1990); and 2) the location
and nature of the Laptev Sea COT has not changed during this period. Thus, the Laptev Sea
contrasts with other active ocean-continent transitions (Figure 7.12). In the Red Sea, for

Figure 7.10: The Norwegian-Greenland Sea,


reconstructed to chron 13n (~33 Ma). The
structural setting suggests that the
reconstructed, sheared Senja margin and
the present Laptev margin are similar.
Arrows indicate the change in relative
plate motion at chron 13n. Oceanic crust,
with identified magnetic anomalies, is
shaded. GFZ, Greenland Fracture Zone;
HFZ, Hornsund Fault Zone; SFZ, Senja
Fracture Zone; VVP, Vestbakken Volcanic
Province. (After Faleide et al., 1993.)
83

example, the spreading axis has propagated stepwise across closely spaced, pre-rift structures
since 4-5 Ma (Cochran and Martinez, 1988). Similarly, the 6-m.y.-old Woodlark back-arc basin
off Papua New Guinea is wedge-shaped and composed of progressively younger and
narrower oceanic segments towards the COT (Benes et al., 1997; Taylor et al., 1999). However,
magmatic underplating of continental crust is neither observed in the Woodlark Basin (Taylor
et al., 1999) nor at the Laptev Sea COT (Franke et al., 2000). The onset of sea-floor spreading in
the Gulf of California at 8-9 Ma may have been abrupt, but occurred much more obliquely
than in the other settings (Lyle and Ness, 1994). Therefore, the young and elongate Red Sea is
probably the nearest analogue to the early Eurasia Basin.
A main question is why the Eurasia Basin
has not propagated farther east during the past 55
m.y. One explanation may relate to anomalous
elastic properties of the Laptev Sea crust (Avetisov,
1993), and I note that several onshore foldbelts
continue offshore (Drachev et al., 1998). Thus, the
continental extension of the plate boundary
propagates through heterogeneous lithosphere. In
fact, the Laptev Sea may be considered a “locked”
zone of initially thickened crust that requires
increased extension prior to break-up, hence
delaying further opening (Figure 7.13a). Such
settings are described in the Atlantic, where the total
extension of neighbouring passive margin segments
and their conjugates may differ by more than 500 Figure 7.11: Seismological signature of
the Laptev Sea COT. Focal
km (Dunbar and Sawyer, 1996). The >10 km deep depths (middle) and anelastic
Laptev Sea basins indicate that the crust is greatly attenuation of shear waves
(bottom) are compared with
thinned and near break-up. margin bathymetry (top). (After
Franke et al., 2000.)
Another explanation is that stress may be
transferred a greater distance along the sheared margin and that some lithospheric extension
may take place east of the New Siberian Islands. There is, however, little earthquake and
geological evidence for this model (Figure 6.7). On the other hand, Levshin et al. (2001)
observed a low-velocity anomaly in the upper mantle beneath the Laptev and East Siberian
seas, not coincident with the earthquake activity (Figure 7.13b). This may reflect elevated
temperatures which may favour aseismic deformation over a much wider zone than the
Laptev Rift System proper.
84

A B

C
M
EX
IC
O

Figure 7.12: Ocean-continent transitions outside the Arctic plate


boundary region, plotted at equal scales (Mercator projection).
A. Red Sea, divided into three sections representing different stages in
the margin development. (After Martinez and Cochran, 1988.) PACIFIC
B. Woodlark Basin in Papua New Guinea, with past and present plate OCEAN
boundaries indicated. (After Benes et al., 1997.)
C. Gulf of California. (After Lyle and Ness, 1994.)
85

Avetisov (1993) proposed the formation of a Laptev Sea microplate if the present
tectonic setting is maintained. Franke et al. (2000) interpreted the epicentre distribution and
focal mechanisms to be consistent with this model, and in fact argued for its present existence.
My data show the same geometry (Figure 6.7). However, the Laptev Sea crust is extensively
faulted both on and off the inferred microplate boundaries (Drachev et al., 1998; Franke et al.,
in press). Hence, the observed seismicity trends may only reflect the present, dominant
tectonic activity within a rift system characterised by migrating rift and transfer zones
throughout the Cenozoic.

Figure 7.13: Possible explanations for the persistent


Laptev Sea COT during the Eurasia Basin
opening history.
A. “Locked zone” model of continental break-
up. The break-up, initiated from the left,
propagates rapidly to the end of a weak
trend in the pre-rifted nucleation zone, but
further propagation is delayed because of
stronger crust. Model dimensions 1000x1000
km. (After Dunbar and Sawyer, 1996.)
B. Elevated upper-mantle temperature east of
the Laptev Sea, interpreted from the
decreased group velocity of 150-s-period
Rayleigh waves relative to the reference
model. Note for comparison the negative
stand-out of the Iceland hot spot. (After
Levshin et al., 2001.)
86

7.5 Opening of Fram Strait


The plate boundary geometry within the Spitsbergen Transform System (Figure 6.2)
constrains the timing of the establishment of a deep-water passage through the Fram Strait.
The opening of this gateway is considered to have had major implications for the Cenozoic
ocean circulation and climate (Thiede and Myhre, 1996), and the various models for onset of
deep-water flow have been summarised by Eldholm et al. (1994). They pointed out that an
initiation coeval with the change in relative plate motion at chron 13n (~33 Ma) was incorrect
because the crust between NE Greenland and Svalbard had to undergo a 15-20-m.y. period of
extension between the change in rotation pole and onset of sea-floor spreading.
Correspondingly, a Middle to Late Miocene age was proposed. Post-chron 13n ages of 17-12
and 7.5-5 Ma have also been suggested from plate reconstructions and subsidence
considerations by Kristoffersen (1990b) and Lawver et al. (1990), respectively. Eiken and Hinz
(1993) inferred that the 3-10 m.y.-old contourites in the Fram Strait were deposited after the
formation of the passage.
I now predict that a deep-water passage was first achieved when a continuous
corridor of oceanic crust was established between the Greenland Sea and the Eurasia Basin,
i.e., when continent-continent translation along all transform faults (cf. Le Pichon and Hayes,
1971) within the Spitsbergen Transform System had ended. I address the question by:
1. Assuming that the present plate boundary segmentation has existed since opening; in fact,
the present fracture zones correspond to offsets in the Svalbard and Greenland margin
morphology, suggesting they have been active since the time of break-up;
2. Initially applying the 2000 m IBCAO isobath (Jakobsson et al., 2000) as an outer limit of
either continental crust or the Yermak Plateau volcanic constructions that were formed no
later than chron 13n;
3. Rotating Greenland relative to Eurasia by the Talwani and Eldholm (1977) anomaly 1-5
stage pole, assuming a constant angle of rotation during the 9.5-m.y. period; the validity of
the pole is supported by fracture zone trends along the entire Arctic plate boundary; and
4. Noting that most ridge flanks are symmetric with respect to the present axis.
An exception is the northwestern flank of the Molloy Ridge which is ~60 km wider than the
southeastern flank. I therefore follow Thiede et al. (1990) and infer a pre-chron 3n shift in axial
location between the Molloy and Spitsbergen fracture zones. Finally, I obtain a better fit to the
magnetic anomalies between the North Lena Trough FZ and the Gakkel Ridge by introducing
a recent, chron 1n, axial shift (Figure 7.14).
The plate reconstruction to 9.5 Ma (chron 5n) shows that the 2000 m contours are still
separated by about 50 km in the east-west direction (Figure 7.14) while overlap is achieved at
~30 Ma using the chron 5n-13n stage pole of Talwani and Eldholm (1977). The 9.5 Ma plate
87

boundary location relative to the 2000 m contour, modified by the Molloy Ridge and North
Lena Trough axial shifts, indicates an onset of sea-floor spreading near this time under my
assumptions. However, more space is available on either side of the Molloy Ridge, suggesting
an earlier onset, ~15 Ma, along this segment. If there is oceanic crust landward of the 2000 m
contour, the Lena Trough may also have started spreading earlier. However, observations of
oceanic basement and/or post-30 Ma volcanic constructions along the western flank of the
Yermak Plateau (Baturin et al., 1994; Jokat, 1998) make my assumptions reasonable. This
model yields an 18-23 m.y. period of extension prior to sea-floor spreading, similar to that
proposed by Ren et al. (in press) on the Vøring margin off central Norway.
The largely persistent fracture zone geometry, the two axial shifts, and the COT
location are the main uncertainties of my model. For example, large crustal areas basinward
of the 2000 m contour in the Lena Trough is not accounted for. This may imply an earlier
opening, probably combined with temporary axial shifts between the main fracture zones.
According to Figure 7.14, the final continent-continent separation across the
transforms was achieved between 7.5 Ma and 4.4 Ma. The latest separation, which occurred at
the Spitsbergen FZ, established the deep-water passage. A pre-9.5-Ma onset of sea-floor
spreading will make these dates older, but the plate geometry does not favour deep-water
conditions prior to the Late Miocene-Early Pliocene, i.e., 9.5-4.4 Ma. Thus, my reconstruction
model yields a younger passage than inferred previously. This implies that the far-field
sedimentary record needs to be examined for independent corroboration of a relatively
young, major deep-water gateway through the Fram Strait.
89

L
84˚ 84˚ 84˚

3
Present 4.4 Ma 6.0 Ma

NL
L

T FZ
gh
L

Trou
H L
H
L YP 82˚ 82˚ 82˚

Lena
3

SL
L
BATHYMETRY

TF
H

Z
H
mGal H
300
50
40 L
30
20

SF
10

Z
0 80˚ 80˚ 80˚
-10
-20 HR
-30
-40
-50

M
FZ
-300
FREE-AIR GRAVITY
nT
1200
-5B
-5B

-5-
-5-

200 78˚ 78˚ 78˚


-

150
-

100
75

20˚

20˚

20˚


10˚

10˚

10˚
50
25
0 84˚ 84˚ 84˚
-3-

-25 7.5 Ma 9.5 Ma 30 Ma


-50
-75
-5-

-5-
-3A-

-3-
-3-

-100
-150
-200
-1200
MAGNETIC ANOMALY

Bathymetry 82˚ 82˚ 82˚


1000 m
2000 m
2500 m

Structures
Hornsund Fault Zone
80˚ 80˚ 80˚
Plate boundary
Extinct plate boundary

Isochrons
6.0 Ma
9.5 Ma
15.0 Ma 78˚ 78˚ 78˚
20˚

20˚

20˚


10˚

10˚

10˚
Figure 7.14: Plate tectonic reconstruction of the Fram Strait, correlated with present bathymetry and potential field Hornsund Fault Zone from Eldholm et al. (1987). Anomaly numbers implied by the reconstruction are indicated
anomalies (left). Persistent fracture zones and a constant Eurasia-Greenland anomaly 1-5 stage pole and angle in the magnetic overview. Oblique Mercator map projections. HR, Hovgård Ridge; MFZ, Molloy Fracture Zone;
(Talwani and Eldholm, 1977) are assumed. Bathymetry from Jakobsson et al. (2000a), free-air gravity from NLTFZ, North Lena Trough Fracture Zone; SFZ, Spitsbergen Fracture Zone; SLTFZ, South Lena Trough Fracture
Andersen and Knudsen (1998) and Childers et al. (2001), and magnetic anomalies from Verhoef et al. (1996). Zone; YP, Yermak Plateau.
91

8 SUMMARY AND CONCLUSIONS

In a plate tectonic framework, the high-latitude part of the Eurasia-North America plate
boundary links the Eurasia Basin to the Norwegian-Greenland Sea and the Atlantic (Figure
1.1). It comprises the Knipovich Ridge in the Greenland Sea and the Gakkel Ridge in the
Eurasia Basin, which are connected by the complex Spitsbergen Transform System.
Regionally, the initial plate geometry reflects two large Cretaceous-Palaeocene rift systems
offset by the mega-shear the De Geer Zone in the incipient Greenland Sea.
The Arctic plate boundary, extending from the northern Knipovich Ridge to
mainland Siberia, is special because 1) its proximity to the Eurasia-North America rotation
pole in northern Siberia causes ultra-slow spreading of 0.26-0.75 cm/yr, the lowest known in
the global MOR system; and 2) it encompasses an ocean-continent transition at the Laptev Sea
margin. Compared with other parts of the MOR system the Arctic plate boundary is poorly
studied. However, recently published bathymetry and potential field data and new and
improved earthquake records have increased the data base considerably. Therefore, the Arctic
plate boundary is re-examined with the objective of obtaining a better understanding of its
regional location, character, morphology and segmentation. The new data have been
integrated with older regional data sets and local bathymetry, potential field and seismic
profiles in a detailed interpretation of geophysical plate boundary indicators.
Where possible, the data have been processed for optimal display of the axial
province (Figures 3.9 and 4.7). In particular, earthquake epicentre data from 1955 to 1999 have
been processed in a tailored two-step procedure to: 1) assure that the best-constrained
epicentre report from each event is selected; and 2) filter the events while balancing the trade-
off between location accuracy and number of displayed events. This sorting shows that the
number of recording stations is probably the best diagnostic criterion for location accuracy.
The resulting Arctic Catalogue contains 2707 quality-evaluable events of which 1295 are well
enough constrained for the plate boundary analysis. It also includes 137 focal mechanisms
obtained from the literature.
Based on the structural framework and geophysical character I propose four Arctic
plate boundary provinces, the Spitsbergen Transform System; the West and East Gakkel
Ridge; and the Laptev Sea. The Spitsbergen Transform System can be subdivided into the
large-offset Molloy and Spitsbergen FZs, the interconnecting Molloy Ridge, and the Lena
Trough, where the small-offset South and North Lena Trough FZs have been inferred. Blocky
topography and the weak magnetic anomalies between the Molloy and North Lena Trough
FZs suggest young and migrating ridge segments but persistent fracture zones. The two
large-offset transforms are the most active segments of the Arctic plate boundary and account
for the relatively high state of stress in the province. The earthquakes are focused and have
92

almost uniform focal mechanisms. However, the Lena Trough earthquakes are more
scattered. By analogy with the Knipovich Ridge I ascribe this to off-axis features, i.e. old lines
of weakness, current migration of the axis and rapid sediment loading, or to incorrect travel-
time corrections in this poorly constrained region.
The linear, 1800-km-long Gakkel Ridge has a very deep axial valley nowhere offset
by more than 20 km. It bisects the Eurasia Basin into the poleward, abyssal Amundsen Basin
and the shallower and broader Nansen Basin. The FZ 11-13 transform region marks a change
in axial trend and divides the ridge into a western province of sharp and elongate crestal
blocks and an eastern province with smoother, discrete basement highs. The transition is also
associated with an eastward decrease of magnetic anomalies, increased sediment infill, and a
slight focusing of the low earthquake activity. This may be a combined effect of sediment
loading, stationary mantle diapirs, and the 0.5-cm/yr spreading rate as a threshold value for
the accretionary mode. The Gakkel Ridge may be migrating northwards, as indicated by the
seismicity offset near FZ 11.
At the Laptev Sea continental margin the >500 km wide Laptev Rift System replaces
the Gakkel Ridge. The continental seismicity is defocused but defines a clear trend in the
presently active Bel’kov-Svyatoi Nos Rift and a weaker trend to the southwest. Probably,
these trends reflect the current activity of a migrating rift rather than an incipient microplate.
Along with strike-slip focal mechanisms on the lower slope they indicate a not more than 60
km wide COT with a ~150-km-offset, dextral sheared margin segment along the Northern
Fracture. However, by analogy with the Senja margin I predict an even narrower COT.
Compared to other young ocean basins the Laptev Sea COT appears to have been a relatively
persistent feature. Probably, this is due to initially thickened crust forming a “locked” zone
which resists further opening, but in addition an upper-mantle heat anomaly may have
increased crustal ductility.
The new plate boundary location and the Arctic Catalogue sheds new light on the
structure, evolution and seismotectonics of the Eurasia Basin. For example, I show that the
Arctic interplate earthquakes are spatially clustered and occur in swarms like those of the
Mid-Atlantic Ridge. A 209-event swarm near 85˚E in 1999 (Müller and Jokat, 2000) is the
largest swarm recorded, but is not a special case in terms of stress. There may be a ~12-year
seismicity cycle in the Spitsbergen Transform System, whereas earthquake recurrence along
the Gakkel Ridge is variable.
FZ 6, FZ 12, the Spitsbergen Transform System fracture zones and the Northern
Fracture are traced basinwide and are related to pre-opening features that have governed the
evolution of the Greenland Sea and Eurasia Basin. The smaller transform segments are young
and probably result from the crustal accretion process.
93

With the assumption of persistent fracture zones and a minimum of axial shifts the
plate boundary location also constrains the evolution of the Fram Strait deep-water passage.
The formation probably occurred post-9.5 Ma, when sea-floor spreading commenced in the
Lena Trough, but prior to the continent-continent separation across the Spitsbergen FZ at 4.4
Ma (Figure 7.14). An 18-23 m.y. period of crustal extension is inferred before break-up.
The detailed verification, refinement and consequences of the new plate boundary
location depends on further studies of both the plate boundary province and the adjacent
ocean basins and margins. In this respect, some remaining questions are:
• How many fracture zones are there, and what are their basinward extent and tectonic
significance?
• How is the detailed crustal structure at the Laptev Sea continent-ocean transition?
• How old is the oceanic crust in the Fram Strait, where is the continent-ocean transition, and
is there any independent, far-field evidence of its opening?
• How has the Eurasia-Greenland-North America triple junction affected the plate
boundary, and what is the structure of the Yermak Plateau and Morris-Jesup Rise?
• How is the crustal accretion process at the ultra-slow-spreading Gakkel Ridge, and does
the significant morphology change represent a threshold in accretionary mode?
95

APPENDIX

A1 Abbreviations
The following abbreviations appear in the text and in the Arctic Catalogue (Appendix A3-A4):

AGP Arctic Gravity Project (Forsberg and LOREX Lomonosov Ridge Experiment, 1979
Kenyon, 1999) (Weber, 1979)
BCI Bureau Central International de Ma Million years before present
Sismologie, France mb Body-wave magnitude
BER Bergen, Norway MC Coda magnitude
BJI Beijing, China MD Duration (coda) magnitude
BRK Berkeley, USA
MJMA Japan Meteorological Agency
CMT Centroid-moment tensor focal
mechanism magnitude
COT Continent-ocean transition ML Local magnitude
DAG Danmarkshavn, Greenland MO Other magnitude
DNAG Decade of North American Geology, MOR Mid-oceanic ridge
1980-89 MOS Moscow, Russia
EID Experimental International Data MS Surface-wave magnitude
Center, USA MW Moment magnitude
EHB Engdahl et al. (1998) m.y. Million years
FUR Fürstenberg, Germany NAO Norwegian Seismic Array (NORSAR)
FZ Fracture zone NRL US Naval Research Laboratory
GEBCO General Bathymetric Chart of the NST Number of recording stations
Oceans (Johnson et al., 1979) OBM Ulan Bator, Mongolia
GMT Generic Mapping Tools (Wessel and OTT Ottawa, Canada
Smith, 1991) PAL Palisades, USA
GOL Golden, USA PAS Pasadena, USA
GPS Global Positioning System PDE Preliminary Determination of
GSC Geological Survey of Canada Earthquakes, US Geological Survey
HDNO Head Department of Navigation and PGC Pacific Geoscience Centre, Canada
Oceanography, Russia PMR Palmer, USA
HEL Helsinki, Finland RTI Ridge-transform intersection
HFS Hagfors, Sweden SCAMP Seafloor Characterization and
HMRG Hawaii Mapping Resource Group Mapping Pod
HRV Harvard, USA SCICEX Science Ice Exercise, 1993-99 (Coakley
IBCAO International Bathymetric Chart of the and Cochran, 1998)
Arctic Ocean (Jakobsson et al., 2000a) SYK Sykes (1965)
ISC International Seismological Centre, UiO University of Oslo
UK UNAP US-Norway Arctic Aerogeophysics
KMS Kort- og Matrikelstyrelsen, Denmark Program (Childers et al., 2001)
LAO Large Aperture Seismic Array, USA WWSSN World-Wide Standardised
LIP Large Igneous Province Seismograph Network
XXX Converted magnitude
96

A2 Making an Arctic Gravity Grid with GMT


The Arctic gravity grid (Figure 3.9b) was constructed from a gridded version of the UNAP
aerogravity data (Childers et al., 2001) and the KMS99 satellite and land gravity grid
(Andersen and Knudsen, 1998). The making of the combined grid is described briefly,
followed by a more detailed report on the display process. All steps apply the GMT 3.3.2
mapping and display software (Wessel and Smith, 1991). This consists of small routines that
may be combined to solve problems in a transparent and stepwise fashion.

Merging of grids.
1. Fourier-domain low-pass filtering of KMS99 data in order to remove high-frequent noise.
Linear tapering at 5 km wavelength, cutoff at 3 km (GMT program grdfft).
2. Resampling of both the UNAP and the KMS99 grid to 4x4’ cell spacing by means of a
simple averaging procedure (grdsample).
3. Merging of the grids yielding UNAP values inside the UNAP survey area, and KMS99
values elsewhere (grdmath).
4. One-to-one conversion of the grid to freepoint data (grd2xyz).
5. Averaging procedure blocking the data into 5x5’ cells and returning the mean value of each
cell, aimed at smoothing the edge effects between the data sets (blockmean).
6. Gridding to 4x4’ cell spacing by a radial simple weighted averaging procedure
(nearneighbor).

Display.
The resulting grid, grav_lonlat.grd, is registered with (longitude, latitude) coordinates
from 65˚ to 90˚N. In order to reduce the amount of data the data south of 70˚N are removed.
The program grdedit removes edge effects in the cropped grid.
grdcut grav_lonlat.grd -R-180/180/70/90 -Ggrav_crop_lonlat.grd
grdedit grav_crop_lonlat.grd -A
The resulting geographical grid is increasingly distorted towards the North Pole (Figure 3.2).
To moderate this, the geographical grid is reformatted into a cartesian grid of 0.0067 inch cell
spacing by the polar stereographic map projection (-JS). The reformatting involves
interpolation of grid values using a search radius of 0.05 inches, found by trial and error.
grdproject grav_crop_lonlat.grd -JS-20/90/9.5c -R-25/78/130/68r -
D0.0067 -S0.05 -Ggrav_crop_cartesian.grd
The cartesian grid is colour-coded by the program grdimage, which uses the gravity colour
table in Appendix A4, and illuminated from the upper left. The illumination file is produced
by the program grdgradient. Unfortunately, grdproject in GMT 3.3.2 has a bug that only
allows inches as a measure unit. As a consequence, the range of the cartesian grid must be
calculated in inches in order to set its correct linear projection (-JX). The trick is to use the
97

command mapproject corners.xy -JS-20/90/9.5c -R-25/78/130/68r, where


corners.xy is a file with the lower left and upper right map corner,
-25 78
130 68
and then convert the plot dimensions to inches.
grdgradient grav_crop_cartesian.grd -A315 -Gillum.grd -Nt1
grdimage grav_crop_cartesian.grd -R0/3.74016/0/9.60339 -JX9.5c/
24.3926c -P -Cgravity.cpt -Iillum.grd -K >! Figure3.9b.ps
The coastline is appended to the map by the program grdcontour, which traces the 0
contour from the IBCAO grid (Jakobsson et al., 2000a). This is a geographical grid, and
therefore the polar stereographic projection that corresponds to the linear projection must be
used. Also, a frame of geographical gridlines is drawn (-B).
grdcut ibcao_lonlat.grd -R-180/180/70/90 -Gibcao_crop_lonlat.grd
grdedit ibcao_crop_lonlat.grd -A
grdcontour ibcao_crop_lonlat.grd -R-25/78/130/68r -JS-20/90/9.5c -P -
Ba0g10f0/a5g5f1::WSEN -C100 -L-1/1 -Wc2/0 -K -O >> Figure3.9b.ps
Finally, a scale bar is drawn in the upper left map corner:
psbasemap -R0/2.1/0/9 -Jx1:1 -P -B0 -G255/255/255 -Y15.3926 -K -O >>
Figure3.9b.ps
psscale -D0.3/4.3/8/0.5 -B/:mGal: -Cgravity.cpt -I0.5 -L -P -O >>
Figure3.9b.ps
A3 Arctic Catalogue Focal Mechanisms
98

no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl
**********************************************************************************************************************
1 1959/03/01 00:31:19.7 74.77 8.14 5.4 PAL 5.4 XXX 5.4 XXX 33 069 70 157 167 68 022 025 30 120 00 3 Laz65
2 1959/10/05 18:27:45.5 83.53 114.09 5.7 PAL 5.7 XXX 5.7 XXX 20 111 84 -029 205 61 -172 161 15 065 25 1 Laz65
3 1960/12/03 20:20:58.9 76.61 131.17 5.1 PAL 5.1 XXX 5.1 XXX 28 167 72 347 18 4 Coo88
067 59 -035 176 61 -144 300 00 035 45 Laz65
4 1963/05/20 17:01:35.0 72.20 126.25 5.5 PDE 10 174 61 -011 270 80 -150 039 13 136 28 3 Coo88
156 60 -030 262 65 -146 028 03 120 41 Sav81
120 90 -155 030 65 000 252 17 348 17 Laz65
5 1964/07/21 09:56:17.6 72.09 130.14 5.4 EHB 12 175 43 -066 324 51 -111 068 03 177 73 1 Coo88
274 72 041 170 52 157 139 41 038 13 Koz84
6 1964/07/31 23:45:58.3 86.45 40.89 5.2 EHB 5.2 XXX 5.2 XXX 7 088 59 -135 331 53 -034 208 04 303 52 4 Sav81
7 1964/08/25 13:47:18.8 78.10 126.72 6.2 EHB 7.0 XXX 7.0 XXX 4 346 48 -089 164 42 -091 075 03 274 87 1 Jem86
339 34 -109 182 58 -077 263 12 127 74 Sav81
336 49 -113 188 46 -066 081 00 180 73 Syk67
8 1967/03/14 07:50:14.9 82.35 39.35 4.7 EHB 5.5 XXX 5.5 XXX 13 027 59 112 169 38 058 344 68 101 12 2 Sav81
9 1967/10/18 01:11:43.3 79.82 2.79 5.6 EHB 6.1 MOS 6.0 XXX 42 223 86 131 71 3 Hor70
10 1967/11/23 13:42:03.7 80.20 -0.64 5.7 EHB 5.9 MOS 6.1 XXX 16 220 84 128 74 3 Hor70
11 1968/01/03 07:37:53.1 72.19 1.44 5.1 EHB 4.3 XXX 5.0 XXX 6 030 60 -037 141 58 -145 086 01 354 46 4 Sav81
12 1968/04/07 05:16:23.7 81.50 -3.54 5.3 EHB 4.6 XXX 5.3 XXX 28 228 75 042 125 50 160 094 40 352 16 4 Sav81
13 1968/06/08 00:41:27.9 86.97 52.35 5.2 EHB 4.6 MOS 5.1 XXX 32 248 54 -108 096 40 -068 350 07 107 74 1 Sav81
14 1969/04/07 20:26:29.2 76.53 130.82 5.4 EHB 5.5 PDE 5.4 XXX 11 020 69 -027 120 65 -157 070 03 339 33 1 Coo88
10 313 49 -107 157 44 -072 055 02 155 77 Jem86
267 79 -054 011 38 -163 329 26 213 44 Koz84
170 49 -097 000 43 -083 265 05 358 85 Sav81
299 64 -105 151 30 -061 040 18 180 67 Cha76
323 54 -134 201 54 -046 082 00 172 56 Con72
15 1970/04/23 00:55:49.0 80.69 121.87 5.2 EHB 4.9 PDE 5.2 XXX 27 226 50 -023 332 72 -137 094 14 197 42 4 Sav81
16 1970/10/21 08:14:12.2 74.64 8.54 5.4 EHB 5.2 PDE 5.4 XXX 33 029 66 148 133 61 028 349 39 082 03 4 Sav81
17 1970/10/26 20:53:32.1 79.81 2.85 5.6 EHB 5.7 PDE 5.6 XXX 34 221 82 -154 127 64 -009 352 12 087 24 3 Sav81
197 81 138 76 Con72
18 1971/05/31 03:46:51.2 72.19 1.08 5.5 EHB 5.7 PDE 6.0 XXX 13 EHB 028 57 -038 141 59 -140 264 01 356 49 4 Sav81
052 64 222 54 Con72
19 1971/11/26 23:07:49.0 79.43 -17.99 5.1 EHB 4.4 XXX 4.4 XXX 18 190 52 -128 062 52 -052 126 00 036 61 4 Syk74
no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl
**********************************************************************************************************************
20 1972/11/19 20:10:51.8 80.41 -2.27 5.4 EHB 5.1 XXX 5.1 XXX 0 224 87 -176 134 86 -003 359 01 089 05 3 Sav81
21 1972/11/25 20:03:28.1 80.34 -1.91 5.6 EHB 5.1 XXX 5.8 XXX 20 243 76 -148 145 59 -016 011 12 108 32 4 Sav81
22 1973/11/09 13:42:42.3 86.05 31.35 5.4 EHB 4.9 XXX 4.9 XXX 1 045 63 -105 256 30 -062 146 17 285 68 1 Jem86
23 1973/12/15 23:31:42.8 74.14 147.05 4.9 EHB 4.8 XXX 4.8 XXX 33 167 64 134 281 50 035 127 50 227 08 4 Coo88
151 37 125 290 60 067 157 66 037 12 Koz84
264 82 -009 356 81 -172 310 01 220 12 Ave78
24 1975/02/26 04:48:53.1 84.96 98.68 5.3 EHB 5.6 PDE 6.0 XXX 2 313 41 -106 155 51 -076 235 05 120 78 1 Jem86
279 48 -135 156 58 -051 221 03 116 57 Sav81
25 1975/03/02 14:23:27.1 85.02 97.69 5.1 EHB 5.0 PDE 5.2 XXX 27 267 56 -142 154 60 -040 211 02 119 49 4 Sav81
26 1976/01/18 04:46:23.4 77.80 18.54 5.5 EHB 5.9 PDE 5.6 XXX 47 209 76 175 301 85 014 166 13 074 06 3 Sav81
4 108 85 008 018 82 175 333 09 243 02 Bun77
27 1976/09/16 03:26:53.6 84.29 0.95 5.4 EHB 5.5 PDE 5.4 XXX 2 030 51 -100 226 40 -078 128 06 248 81 1 Jem86
28 1976/11/12 14:47:24.8 72.37 -70.45 5.3 EHB 5.1 PDE 5.8 XXX 23 295 55 060 160 45 125 148 65 046 05 2 Ste79
054 73 026 316 65 161 277 30 183 05 Wet78
29 1977/04/23 14:49:08.3 75.22 134.56 5.0 EHB 4.2 PDE 4.8 EHB 10 180 45 -090 000 45 -090 090 00 200 90 4 HRV
30 1977/12/23 11:15:45.8 72.07 -0.15 4.6 EHB 4.7 PDE 5.2 EHB 15 031 41 -120 248 55 -067 322 08 211 69 4 HRV
31 1978/01/04 14:52:10.5 85.67 -24.69 4.9 EHB 4.6 ISC 5.2 EHB 36 222 86 014 130 76 176 087 13 356 07 3 Wet82
216 80 016 123 74 169 080 19 349 04 Gre82
15 201 66 015 105 76 155 061 27 154 07 4 HRV
32 1978/02/01 18:05:35.9 79.88 0.90 5.1 EHB 4.8 ISC 5.1 EHB 15 039 47 -011 137 82 -136 260 23 008 36 4 HRV
33 1978/02/05 16:07:12.4 78.36 -107.85 5.0 EHB 4.8 ISC 5.1 EHB 15 115 77 -008 206 82 -167 340 04 071 15 3 HRV
34 1979/03/13 19:21:32.5 74.75 8.70 4.9 EHB 4.5 ISC 5.0 EHB 15 042 20 -093 225 70 -089 314 25 137 65 4 HRV
35 1979/06/15 23:19:45.7 86.36 36.42 4.7 EHB 4.4 ISC 5.0 EHB 15 261 40 -087 077 50 -093 169 05 328 85 1 HRV
36 1980/02/01 17:30:29.2 73.05 122.54 5.4 EHB 5.3 ISC 5.3 EHB 22 274 71 -137 167 50 -025 036 13 138 43 4 Coo88
18 315 55 -078 114 36 -107 036 10 262 76 HRV
37 1980/03/19 01:48:57.0 83.53 114.79 4.7 EHB 4.9 ISC 5.0 EHB 15 280 75 178 010 88 015 236 12 144 09 3 HRV
38 1981/11/20 20:59:16.4 79.52 3.29 5.1 EHB 4.6 ISC 5.2 EHB 15 200 45 -090 020 45 -090 110 00 219 90 4 HRV
39 1982/06/11 11:41:51.0 85.62 87.43 4.9 EHB 4.9 ISC 5.2 EHB 15 306 45 -090 126 45 -090 216 00 325 90 1 HRV
40 1982/06/12 00:15:10.9 85.68 85.92 5.2 EHB 5.2 ISC 5.5 EHB 15 311 45 -090 131 45 -090 221 00 330 90 1 HRV
41 1983/01/15 06:43:59.3 73.13 5.78 5.3 EHB 4.8 ISC 5.2 EHB 15 222 45 -090 042 45 -090 132 00 241 90 4 HRV
99
no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
100

yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl


**********************************************************************************************************************
42 1983/06/10 02:13:27.6 75.49 122.59 5.6 EHB 5.4 ISC 5.6 EHB 22 EHB 115 68 -029 217 63 -155 167 03 075 36 4 Fra00
150 70 054 034 41 148 265 17 018 52 Coo88
155 84 -113 052 24 -014 265 35 042 46 Par88
144 72 054 031 40 151 258 19 014 50 Jem86
12 142 59 -118 008 40 -052 252 10 005 67 HRV
43 1983/12/05 20:54:24.6 73.80 8.84 5.2 EHB 4.4 ISC 4.9 EHB 15 180 45 -090 000 45 -090 090 00 200 90 4 HRV
44 1984/05/17 07:55:59.2 79.74 3.21 5.2 EHB 5.1 HRV 5.7 EHB 10 224 46 005 130 86 136 077 33 186 27 4 HRV
45 1984/08/12 15:28:06.6 74.15 135.01 5.0 EHB 20 EHB 282 77 038 182 53 164 149 36 047 15 4 Fra00
46 1985/07/25 17:45:07.1 83.98 -1.62 4.8 EHB 4.9 ISC 5.1 EHB 10 027 42 -111 234 52 -072 312 05 203 75 4 HRV
47 1985/07/25 22:03:36.1 85.23 101.22 4.7 EHB 4.7 ISC 4.9 EHB 10 335 20 -072 137 71 -096 231 26 036 63 1 HRV
48 1985/10/15 19:52:03.4 85.70 85.00 4.9 EHB 4.8 ISC 5.2 EHB 10 106 31 -106 304 60 -081 028 15 238 73 1 HRV
49 1985/10/15 22:29:54.0 85.67 84.57 4.8 EHB 4.5 ISC 5.0 EHB 10 325 26 -060 112 68 -103 213 22 359 65 1 HRV
50 1985/10/17 08:39:27.7 76.04 7.14 4.6 EHB 5.0 ISC 5.1 EHB 10 257 34 -112 104 59 -076 183 13 048 72 4 HRV
51 1985/12/31 06:57:19.0 73.25 6.93 4.8 EHB 4.9 ISC 5.4 EHB 10 223 10 -085 038 80 -091 129 35 307 55 1 HRV
52 1986/06/20 07:00:08.5 81.78 119.47 4.8 EHB 4.2 ISC 4.9 EHB 15 356 32 -073 156 60 -100 254 14 040 73 1 HRV
53 1986/10/08 00:09:24.6 80.30 -1.72 5.2 EHB 4.8 ISC 5.2 EHB 15 305 90 180 035 90 000 260 00 350 00 3 HRV
54 1987/02/22 01:22:33.0 78.83 125.91 5.3 EHB 5.0 ISC 5.3 EHB 15 000 48 -057 136 51 -122 247 02 342 66 4 HRV
55 1987/03/30 03:13:43.4 74.52 -130.60 5.6 EHB 4.7 ISC 5.2 EHB 15 235 24 134 008 73 073 255 59 112 26 1 HRV
56 1987/07/11 06:15:53.6 82.20 -17.85 5.6 EHB 5.1 ISC 5.5 EHB 15 027 50 -044 148 59 -131 266 05 003 56 4 Chu97
15 232 45 -068 022 49 -111 127 02 224 74 HRV
57 1987/08/03 07:37:43.6 86.88 62.86 4.9 EHB 4.5 ISC 5.0 EHB 15 148 33 -068 302 60 -104 042 14 180 72 4 HRV
58 1987/09/22 22:05:17.4 76.40 134.28 5.6 EHB 5.2 ISC 5.5 EHB 10 006 41 -069 159 52 -107 261 06 014 75 4 HRV
59 1987/11/25 17:28:02.7 73.63 118.69 5.1 EHB 4.0 EHB 8 EHB 136 84 -008 227 82 -174 182 01 091 10 3 Fra00
60 1987/12/13 21:05:05.6 74.38 -93.68 5.2 EHB 5.2 ISC 5.3 EHB 15 107 49 052 337 54 125 308 62 043 02 4 HRV
61 1988/01/01 14:36:12.6 74.62 130.92 5.1 EHB 4.8 ISC 5.1 EHB 15 175 29 -122 031 65 -073 108 19 329 65 1 HRV
62 1988/03/21 23:31:26.4 77.58 125.46 6.0 EHB 6.2 ISC 6.3 EHB 25 EHB 208 67 002 117 88 157 070 17 165 15 3 Fra00
15 178 34 -073 339 58 -101 076 12 217 75 HRV
63 1988/04/25 20:09:27.5 78.55 6.00 4.9 EHB 4.9 ISC 5.4 EHB 15 221 70 -001 311 89 -160 084 13 178 15 3 HRV
64 1989/08/05 06:55:53.3 76.11 134.51 5.3 EHB 5.3 ISC 5.3 EHB 14 EHB 350 50 -090 170 40 -090 080 05 260 85 1 Fra00
15 348 40 -093 172 50 -087 260 05 101 85 HRV
65 1989/09/17 12:01:35.0 79.10 2.62 4.8 EHB 5.0 EHB 15 214 45 -090 034 45 -090 124 00 233 90 4 HRV
66 1989/10/03 23:09:56.0 80.61 121.64 5.2 EHB 5.2 ISC 5.4 EHB 12 EHB 142 66 007 050 84 155 003 21 098 12 3 Fra00
15 134 56 -101 334 36 -074 232 10 010 76 HRV
67 1989/10/07 01:25:07.0 78.81 4.32 5.0 EHB 4.5 ISC 5.1 EHB 15 311 90 -180 041 90 000 086 00 176 00 3 HRV
no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl
**********************************************************************************************************************
68 1989/11/04 18:04:03.1 72.25 0.77 5.0 EHB 5.3 ISC 5.4 EHB 15 212 45 -090 032 45 -090 122 00 231 90 4 HRV
69 1989/11/04 18:17:14.4 72.24 0.78 5.2 EHB 5.2 ISC 5.3 EHB 15 223 45 -090 043 45 -090 133 00 242 90 4 HRV
70 1989/11/17 04:05:21.2 80.55 122.17 5.2 EHB 5.5 ISC 5.7 EHB 15 359 23 -046 133 74 -106 236 27 020 58 1 HRV
71 1990/03/13 00:33:01.2 73.30 134.83 5.6 EHB 5.2 ISC 5.3 EHB 19 EHB 055 51 -050 181 53 -128 298 01 030 60 4 Fra00
15 006 45 -090 186 45 -090 276 00 025 90 HRV
72 1990/05/27 21:49:35.5 74.21 8.82 5.6 EHB 5.6 ISC 5.7 EHB 15 195 45 -090 015 45 -090 105 00 214 90 4 HRV
73 1990/06/09 18:24:33.0 75.09 112.98 4.9 EHB 5.1 EHB 10 EHB 204 63 -012 300 80 -152 069 11 165 27 3 Fra00
74 1990/11/11 07:06:31.4 74.53 9.13 4.8 EHB 4.7 ISC 5.3 EHB 15 199 45 -090 019 45 -090 109 00 218 90 4 HRV
75 1991/03/01 01:57:05.3 72.14 126.87 5.2 EHB 5.0 EHB 15 EHB 068 64 165 165 76 027 029 28 294 08 3 Fra00
76 1991/06/11 07:16:35.8 84.44 108.13 5.6 EHB 5.2 ISC 5.4 EHB 22 EHB 224 66 016 127 76 155 084 28 177 06 3 Fra00
16 333 41 -050 105 60 -119 215 10 326 63 HRV
77 1991/06/18 23:01:37.8 82.15 118.92 5.0 EHB 4.3 EHB 16 EHB 152 59 -049 272 49 -137 214 05 116 55 4 Fra00
78 1991/07/02 21:24:06.7 72.89 12.29 5.5 EHB 4.6 ISC 5.2 EHB 17 181 42 107 339 50 076 189 78 079 04 4 HRV
79 1991/09/01 06:51:07.9 78.96 3.27 5.2 EHB 5.0 HRV 5.6 EHB 10 308 90 179 038 89 000 084 00 174 00 3 Fej96
35 309 90 -180 220 90 000 084 00 174 00 HRV
80 1991/10/17 17:46:59.3 86.97 62.44 5.2 EHB 4.7 ISC 5.2 EHB 15 072 43 -135 305 61 -057 013 10 265 59 4 HRV
81 1992/02/15 04:52:06.4 75.91 125.06 4.9 EHB 4.9 EHB 15 EHB 128 61 -028 233 66 -147 359 03 092 39 4 Fra00
82 1992/02/17 00:02:00.3 79.17 124.52 6.0 EHB 5.7 ISC 5.9 EHB 19 EHB 355 76 -069 117 25 -145 068 28 290 54 1 Fra00
15 355 54 -065 137 43 -120 067 06 322 69 HRV
83 1992/06/08 09:30:16.6 81.27 121.13 5.2 EHB 4.9 ISC 5.1 EHB 9 EHB 356 85 -070 099 21 -167 068 37 286 46 1 Fra00
15 001 51 -055 134 51 -125 067 00 336 64 HRV
84 1992/07/20 07:46:50.2 78.55 5.59 5.7 EHB 6.3 ISC 6.7 EHB 15 126 80 171 217 81 011 082 13 352 01 3 HRV
85 1992/08/11 04:03:47.6 80.18 -1.25 5.2 EHB 5.0 ISC 5.4 EHB 15 218 68 022 120 70 157 078 31 169 01 3 HRV
86 1992/09/09 13:08:55.7 76.20 7.20 5.8 EHB 5.8 ISC 6.1 EHB 15 220 23 -046 353 74 -106 097 27 241 58 1 HRV
87 1992/09/10 14:54:37.9 76.24 7.83 5.3 EHB 5.1 ISC 5.2 EHB 15 200 90 -180 290 90 000 335 00 065 00 3 HRV
88 1993/01/21 13:43:16.4 78.86 125.46 5.2 EHB 5.1 ISC 5.5 EHB 15 142 47 -125 007 53 -059 076 03 339 65 4 HRV
89 1993/02/12 10:52:07.2 79.17 124.41 5.0 EHB 4.5 ISC 5.0 EHB 21 EHB 356 64 -124 233 41 -041 110 13 220 57 4 Fra00
15 324 42 -130 193 59 -060 261 09 153 63 HRV
90 1993/02/23 11:56:30.1 86.92 55.39 4.7 EHB 4.7 ISC 5.1 EHB 16 056 46 -126 281 54 -059 351 05 251 65 4 HRV
91 1993/07/12 05:05:33.9 72.14 1.28 5.1 EHB 4.9 ISC 5.3 EHB 15 251 39 -034 009 69 -123 123 18 238 53 4 HRV
92 1993/08/03 06:15:58.4 85.26 91.37 4.9 EHB 4.4 EHB 15 EHB 323 16 -071 124 75 -095 218 30 026 60 1 Fra00
93 1993/08/10 19:36:22.7 83.04 -27.61 5.5 EHB 5.2 ISC 5.5 EHB 15 189 31 -013 290 83 -121 045 32 171 43 4 HRV
94 1993/09/23 20:04:04.3 78.50 7.28 4.8 EHB 4.7 ISC 5.1 EHB 15 186 33 -101 019 58 -083 104 12 310 76 1 HRV
101
no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
102

yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl


**********************************************************************************************************************
95 1993/10/05 21:28:07.7 77.69 126.27 5.0 EHB 4.7 ISC 5.2 EHB 18 EHB 191 63 -044 304 51 -146 250 07 152 49 4 Fra00
15 032 35 -065 183 59 -106 284 12 054 71 HRV
96 1994/01/26 12:07:18.2 79.50 4.00 5.2 EHB 4.9 ISC 5.2 EHB 15 193 45 -090 013 45 -090 103 00 212 90 4 HRV
97 1994/09/10 01:24:12.4 83.64 -2.27 5.0 EHB 4.8 ISC 5.2 EHB 19 206 29 -092 029 61 -089 117 16 301 74 1 HRV
98 1994/09/23 02:11:38.5 76.22 145.65 4.6 EHB 4.2 ISC 10 EHB 183 59 -106 032 35 -065 284 13 056 71 1 Fra00
99 1995/03/09 07:04:24.9 78.31 2.24 5.2 EHB 4.8 ISC 5.1 EHB 15 137 47 071 343 47 109 332 76 240 00 4 HRV
100 1995/08/03 01:16:43.3 80.30 -2.66 5.0 EHB 4.7 ISC 5.2 EHB 15 118 70 -177 027 88 -020 074 12 341 16 3 HRV
101 1995/10/04 09:17:32.1 75.99 7.25 5.2 EHB 4.9 ISC 5.3 EHB 15 220 32 -050 355 66 -112 101 18 231 62 4 HRV
102 1995/12/08 07:41:16.1 72.58 3.58 5.3 EHB 5.3 ISC 5.5 EHB 15 014 40 -141 252 66 -057 319 15 206 56 4 HRV
103 1996/05/11 04:38:37.0 80.57 -2.58 5.5 ISC 5.3 ISC 5.7 HRV 15 206 24 -067 001 68 -100 099 22 254 66 1 HRV
104 1996/06/22 16:47:14.0 75.78 134.63 5.7 ISC 5.7 ISC 6.2 OBM 10 ISC 339 60 -055 104 45 -135 045 09 300 59 4 Fra00
15 144 29 -112 349 63 -078 070 17 283 69 HRV
105 1996/08/20 00:11:00.0 77.87 7.63 5.2 ISC 4.9 ISC 5.2 HRV 10 184 62 084 017 29 101 080 72 278 17 2 Fej96
15 202 43 -075 002 48 -103 101 03 207 79 HRV
106 1997/04/19 15:26:33.5 78.45 125.82 5.8 PDE 5.0 PDE 5.5 PDE 14 311 03 -142 183 88 -088 271 43 095 47 1 Sip99
107 1998/03/21 16:33:11.0 79.89 1.86 6.0 PDE 6.2 PDE 6.2 HRV 21 126 88 179 216 89 002 081 02 351 01 3 HRV
108 1998/04/23 03:36:56.0 79.57 3.65 4.7 PDE 4.8 PDE 5.2 HRV 15 217 43 -063 001 53 -113 108 05 213 71 4 HRV
109 1998/10/18 22:09:19.2 75.61 86.28 5.3 PDE 4.7 PDE 5.0 HRV 15 125 37 -074 285 54 -102 024 09 156 77 1 HRV
110 1999/01/01 09:28:08.8 79.99 -111.36 5.3 PDE 4.3 PDE 5.1 HRV 15 183 36 134 314 65 063 183 61 062 16 4 HRV
111 1999/01/28 22:52:45.3 85.56 85.99 4.7 PDE 4.7 PDE 5.4 HRV 15 130 17 -087 306 73 -091 038 28 215 62 1 HRV
112 1999/02/01 04:52:40.8 85.64 86.10 5.2 PDE 4.7 PDE 5.2 HRV 15 325 41 -070 120 52 -106 221 06 334 76 4 HRV
113 1999/02/01 09:56:35.0 85.73 84.44 5.2 PDE 5.3 PDE 5.6 HRV 15 110 40 -106 311 52 -077 031 06 271 78 4 HRV
114 1999/02/01 11:56:00.8 85.57 87.14 5.2 PDE 5.6 PDE 5.7 PDE 15 301 42 -086 116 48 -093 208 03 347 86 1 HRV
115 1999/02/19 19:10:00.5 85.57 87.04 5.2 PDE 5.0 PDE 5.4 HRV 15 326 30 -055 107 66 -108 211 19 346 64 1 HRV
116 1999/02/22 08:02:11.2 86.28 73.39 5.3 PDE 4.9 PDE 5.2 HRV 15 110 43 -105 311 49 -076 031 03 285 79 4 HRV
117 1999/02/25 15:35:17.3 85.67 83.42 4.8 PDE 4.9 PDE 5.3 HRV 15 329 21 -071 128 70 -097 224 25 027 64 1 HRV
118 1999/03/01 17:46:46.3 85.69 86.03 5.0 PDE 5.0 PDE 5.5 HRV 15 324 20 -046 098 76 -104 200 29 350 57 1 HRV
119 1999/03/04 07:52:45.5 85.74 84.91 4.9 PDE 4.8 PDE 5.5 HRV 15 148 26 -103 343 65 -084 068 19 265 70 1 HRV
120 1999/03/06 07:21:38.2 85.67 84.87 4.6 PDE 4.8 PDE 5.1 HRV 15 350 38 -062 136 57 -110 240 10 000 70 4 HRV
121 1999/03/13 01:26:33.5 85.69 84.80 5.4 PDE 5.2 PDE 5.4 HRV 15 308 36 -076 112 56 -099 208 10 347 77 1 HRV
122 1999/03/19 08:00:34.8 85.71 85.39 4.6 PDE 4.7 PDE 5.2 HRV 15 124 16 -093 308 74 -089 036 29 218 61 1 HRV
123 1999/03/21 15:24:07.8 85.63 86.82 5.5 PDE 5.2 PDE 5.3 HRV 15 112 20 -111 314 72 -083 038 26 236 63 1 HRV
124 1999/03/28 08:28:08.5 83.95 -1.48 4.6 PDE 5.0 PDE 5.1 HRV 15 179 42 -141 059 65 -054 123 13 014 55 4 HRV
125 1999/03/28 21:33:44.1 85.64 86.26 5.1 PDE 5.2 PDE 5.8 HRV 15 155 22 -052 295 73 -104 036 27 185 60 1 HRV
no date time lat lon mb--ref Ms--ref Mw--ref d---ref plane1 plane2 t-axis p-axis c ref
yyyy:mm:dd hh:mm:ssss str dp slip str dp slip azi pl azi pl
**********************************************************************************************************************
126 1999/04/01 10:47:53.0 85.65 86.53 5.2 PDE 5.2 PDE 5.4 HRV 15 112 44 -099 304 47 -082 028 01 286 84 4 HRV
127 1999/04/13 02:09:22.3 73.22 6.65 5.0 PDE 4.7 PDE 5.1 HRV 15 063 31 -032 181 74 -117 292 24 059 53 4 HRV
128 1999/04/26 13:20:07.6 85.67 84.83 5.3 PDE 5.0 PDE 5.2 HRV 15 300 35 -083 112 56 -095 205 10 003 79 1 HRV
129 1999/05/18 20:20:16.1 85.63 86.15 5.2 PDE 5.4 PDE 5.5 HRV 15 098 46 -126 324 54 -059 033 05 293 65 4 HRV
130 1999/06/07 16:10:33.6 73.02 5.19 5.4 PDE 5.5 PDE 5.5 HRV 15 247 47 -038 006 63 -130 123 09 225 53 4 HRV
131 1999/06/07 16:35:46.7 73.08 5.45 5.3 PDE 5.4 PDE 5.5 HRV 15 246 39 -032 001 71 -125 117 18 232 52 4 HRV
132 1999/06/11 23:54:52.0 85.60 83.70 5.2 PDE 4.6 PDE 5.0 HRV 15 261 43 -127 126 57 -061 196 08 090 64 4 HRV
133 1999/06/18 19:47:25.2 85.68 85.77 5.4 PDE 4.9 PDE 5.0 HRV 15 283 45 -090 103 45 -090 193 00 302 90 1 HRV
134 1999/07/08 19:25:10.5 85.74 83.26 5.0 PDE 4.7 PDE 5.2 HRV 15 127 22 -132 351 74 -074 069 27 282 58 1 HRV
135 1999/08/02 06:40:58.3 85.69 84.13 4.9 PDE 4.3 PDE 4.9 HRV 15 355 45 -090 175 45 -090 265 00 014 90 4 HRV
136 1999/08/03 13:55:41.4 72.26 0.40 5.1 PDE 5.2 PDE 5.3 HRV 15 039 41 -097 229 49 -084 314 04 182 84 4 HRV
137 1999/08/08 13:45:44.2 85.84 82.15 5.0 PDE 4.7 PDE 5.1 HRV 15 299 46 -036 056 65 -130 173 11 278 52 4 HRV

Legend

no Event number References: Ave78, Avetisov (1978); Bun77, Bungum (1977);


date (yyyy:mm:dd), time Origin time Cha76, Chapman and Solomon (1976); Chu97, Chung and Gao
lat Latitude (1997); Con72, Conant (1972); Coo88, Cook (1988); Fej96,
lon Longitude Fejerskov et al. (1996); Fra00, Franke et al. (2000); Gre82,
mb--ref Body-wave magnitude with reporting agency Gregersen (1982); Hor70, Horsfield and Maton (1970); HRV,
Ms--ref Surface-wave magnitude with reporting agency Harvard CMT; Jem86, Jemsek et al. (1986); Koz84, Koz’min
Mw--ref Moment magnitude with reporting agency (1984); Laz65, Lazareva and Misharina (1965); Par88,
d---ref Depth (-km), from the focal mechanism solution unless noted Parfenov et al. (1988); Sav81, Savostin and Karasik (1981);
plane1 (slip str dp) Slip, strike and dip of nodal plane 1 Sip99, Sipkin et al. (1999); Ste79, Stein et al. (1979); Syk67,
plane2 (slip str dp) Slip, strike and dip of nodal plane 2 Sykes (1967); Syk74, Sykes and Sbar (1974); Wet78, Wetmiller
t-axis (azi pl) Azimuth and plunge of tensional axis and Horner (1978); Wet82, Wetmiller and Forsyth (1982).
p-axis (azi pl) Azimuth and plunge of pressure axis
c Category: 1=normal; 2=reverse; 3=strike-slip; 4=composite
ref Reference

Figure A.1 (overleaf): The highest-ranked focal mechanisms from the above table. Compressional areas are shaded. For centroid-moment tensor solutions the
shading may deviate from the best-fit nodal planes which assume a double couple of forces and are indicated by bold lines. Pressure axes are shown by
black dots and tensional axes by open circles.
103
104

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36

37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54

55 56 57 58 59 60 61 62 63 64 65 66 67 68 69 70 71 72

73 74 75 76 77 78 79 80 81 82 83 84 85 86 87 88 89 90

91 92 93 94 95 96 97 98 99 100 101 102 103 104 105 106 107 108

109 110 111 112 113 114 115 116 117 118 119 120 121 122 123 124 125 126

127 128 129 130 131 132 133 134 135 136 137
105

A4 Main Output Data (CD-ROM)


The enclosed CD-ROM contains the following files:
• README: Contents and format descriptions
• plate_boundary.xy: Plate boundary location (plot with GMT program psxy -M)
• arccat.merge: Arctic Catalogue epicentres (Table 4.1)
• arccat_filtered.merge: Plate boundary sorting of the Arctic Catalogue (Table 4.1)
• aki.gmt: Arctic Catalogue focal mechanisms in Aki and Richards convention, (plot with
psmeca -Sa)
• cmt.gmt: Focal mechanisms in CMT format (plot with psmeca -Sm)
• bathymetry.cpt: GMT colour table for bathymetry (Figure 3.9a)
• gravity.cpt: GMT colour table for gravity (Figure 3.9b)
• magnetics.cpt: GMT colour table for magnetics (Figure 3.9c)
107

REFERENCES

Andersen, O. B., and Knudsen, P., 1998, Global marine gravity field from the ERS-1 and Geosat
geodetic mission altimetry: Journal of Geophysical Research, v. 103, p. 8129-8138.
Andreassen, T. E. S., 1997, En geofysisk undersøkelse av Yermakplatået [Cand. Scient. thesis]: Oslo,
University of Oslo, 132 p.
Appelgate, B., 1997, Modes of axial reorganization on a slow-spreading ridge; the structural evolution
of Kolbeinsey Ridge since 10 Ma: Geology, v. 25, p. 431-434.
Austegard, A., and Sundvor, E., 1991, The Svalbard continental margin; crustal structure from analysis
of deep seismic profiles and gravity: University of Bergen, Seismo-Series no. 53, 31 p.
Avetisov, G. P., 1978, O mekhanizme ochaga odnogo Arkticheskogo zemletryaseniya, in Geofizicheskie
metody razvedki v Arktike: Leningrad, Nauchno-Issledovatel’skii Institut Geologii Arktiki, p.
145-148.
Avetisov, G. P., 1993, Some aspects of lithospheric dynamics of Laptev Sea: Physics of the Solid Earth, v.
29, p. 402-412.
Barazangi, M., and Dorman, J., 1970, Seismicity map of the Arctic compiled from ESSA, Coast and
Geodetic Survey, epicenter data January 1961 through September 1969: Bulletin of the
Seismological Society of America, v. 60, p. 1741-1743.
Baturin, D. G., 1987, Evolution of the northern Barents Sea in the area of junction with the Eurasian
ocean basin: Marine Geology, v. 27, p. 308-312.
Baturin, D., Fedukhina, T., Savostin, L., and Yunov, A., 1994, A geophysical survey of the Spitsbergen
margin and surrounding areas: Marine Geophysical Researches, v. 16, p. 463-484.
Bell, R. E., and Buck, W. R., 1992, Crustal control of ridge segmentation inferred from observations of
the Reykjanes Ridge: Nature, v. 357, p. 583-586.
Benes, V., Bocharova, N., Popov, E., Scott, S. D., and Zonenshain, L., 1997, Geophysical and morpho-
tectonic study of the transition between seafloor spreading and continental rifting, western
Woodlark Basin, Papua New Guinea: Marine Geology, v. 142, p. 85-98.
Bergman, E. A., and Solomon, S. C., 1990, Earthquake swarms on the Mid-Atlantic Ridge; products of
magmatism or extensional tectonics?: Journal of Geophysical Research, v. 95, p. 4943-4965.
Bjørlykke, K., 1995, Fracturing and brittle/ductile properties of sedimentary rocks in relation to
diagenesis and in situ stress, in Fejerskov, M., and Myrvang, A. M., eds., Proceedings of the
workshop; rock stresses in the North Sea, 13-14 February: Trondheim, Norwegian Institute of
Technology, p. 64-75.
Bonatti, E., 1996, Anomalous opening of the Equatorial Atlantic due to an equatorial thermal
minimum: Earth and Planetary Science Letters, v. 143, p. 147-160.
Bown, J. W., and White, R. S., 1994, Variation with spreading rate of oceanic crustal thickness and
geochemistry: Earth and Planetary Science Letters, v. 121, p. 435-449.
Breivik, A. J., Verhoef, J., and Faleide, J. I., 1999, Effect of thermal contrasts on gravity modelling at
passive margins; results from the western Barents Sea: Journal of Geophysical Research, v. 104, p.
15293-15311.
Brune, J. N., 1968, Seismic moment, seismicity, and rate of slip along major fault zones: Journal of
Geophysical Research, v. 73, p. 777-784.
Bullen, K. E., and Bolt, B. A., 1985, An introduction to the theory of seismology: Cambridge, Cambridge
University Press, 499 p.
Bungum, H., 1977, Two focal-mechanism solutions for earthquakes from Iceland and Svalbard:
Tectonophysics, v. 41, p. T15-T18.
Bungum, H., and Capon, J., 1974, Coda pattern and multipath propagation of surface waves at
NORSAR: Physics of the Earth and Planetary Interiors, v. 9, p. 111-127.
Bungum, H., and Kristoffersen, Y., 1980, A microearthquake survey of the Svalbard region; final report,
phase 1: Kjeller, Norway, NTNF/NORSAR, Technical Report no. 1/80, 28 p.
Byrkjeland, U., Bungum, H., and Eldholm, O., 2000, Seismotectonics of the Norwegian continental
margin: Journal of Geophysical Research, v. 105, p. 6221-6236.
Cande, S. C., and Kent, D. V., 1992, A new geomagnetic polarity time scale for the Late Cretaceous and
Cenozoic: Journal of Geophysical Research, v. 97, p. 13917-13951.
Cande, S. C., and Kent, D. V., 1995, Revised calibration of the geomagnetic polarity timescale for the
Late Cretaceous and Cenozoic: Journal of Geophysical Research, v. 100, p. 6093-6095.
108

Cannat, M., Rommevaux-Jestin, C., Sauter, D., Deplus, C., and Mendel, V., 1999, Formation of the axial
relief at the very slow spreading Southwest Indian Ridge (49˚ to 69˚E): Journal of Geophysical
Research, v. 104, p. 22825-22843.
Chan, W. W., and Mitchell, B. J., 1982, Synthetic seismogram and surface wave constraints on crustal
models of Spitsbergen: Tectonophysics, v. 89, p. 51-76.
Chapman, M. E., and Solomon, S. C., 1976, North American-Eurasian plate boundary in northeast Asia:
Journal of Geophysical Research, v. 81, p. 921-930.
Chen, Y., 1989, A mechanical model for the inside corner uplift at a ridge-transform intersection:
Journal of Geophysical Research, v. 94, p. 9275-9282.
Chen, Y. J., 1992, Oceanic crustal thickness versus spreading rate: Geophysical Research Letters, v. 19, p.
753-756.
Childers, V. A., McAdoo, D. C., Brozena, J. M., and Laxon, S. W., 2001, New gravity data in the Arctic
Ocean; comparison of airborne and ERS gravity: Journal of Geophysical Research, v. 106, p. 8871-
8886.
Chung, W.-Y., and Gao, H., 1997, The Greenland earthquake of 11 July 1987 and postglacial fault
reactivation along a passive margin: Bulletin of the Seismological Society of America, v. 87, p.
1058-1068.
Churkin, M., Jr., and Johnson, G. L., 1976, Plate 19, Arctic [map 1:16 000 000], in Choubert, G., Faure-
Muret, A., and Chanteux, P., eds., Geological world atlas: Paris, Unesco and Commission for the
Geological Map of the World.
Coakley, B. J., and Cochran, J. R., 1998, Gravity evidence of very thin crust at the Gakkel Ridge (Arctic
Ocean): Earth and Planetary Science Letters, v. 162, p. 81-95.
Cochran, J. R., and Martinez, F., 1988, Evidence from the northern Red Sea on the transition from
continental to oceanic rifting: Tectonophysics, v. 153. p. 25-53.
Conant, D. A., 1972, Six new focal mechanism solutions for the Arctic and a center of rotation for plate
movements [M. A. thesis]: New York, Columbia University, 18 p.
Cook, D. B., 1988, Seismology and tectonics of the North American plate in the Arctic; Northeast Siberia
and Alaska [Ph. D. thesis]: East Lansing, Michigan State University, 250 p.
Crane, K., Eldholm, O., Myhre, A. M., and Sundvor, E., 1982, Thermal implications for the evolution of
the Spitsbergen Transform Fault: Tectonophysics, v. 89, p. 1-32.
Crane, K., Sundvor, E., Foucher, J. P., Hobart, M., Myhre, A. M., and LeDouaran, S., 1988, Thermal
evolution of the western Svalbard margin: Marine Geophysical Researches, v. 9, p. 165-194.
Crane, K., Vogt, P. R., Sundvor, E., Shor, A., and Reed, T., IV, 1995, SeaMARC II investigations of the
northern Norwegian-Greenland Sea, in Crane, K., and Solheim, A., eds., Seafloor atlas of the
northern Norwegian-Greenland Sea: Norsk Polarinstitutt Meddelelser, v. 137, p. 32-143.
Dawes, P. R., and Peel, J. S., 1981, The northern margin of Greenland from Baffin Bay to the Greenland
Sea, in Nairn, A. E. M., Churkin, M., Jr., and Stehli, F. G., The Arctic Ocean: New York, Plenum
Press, The Ocean Basins and Margins, v. 5, p. 201-264.
Dehlinger, P., 1978, Marine gravity: Amsterdam, Elsevier, 322 p.
DeMets, C., Gordon, R. G., Argus, D. F., and Stein, C., 1990, Current plate motion: Geophysical Journal
International, v. 101, p. 425-478.
Desimon, A. I., and Karasik, A. M., 1979, Some features of bottom relief and sea-floor spreading on the
Knipovich Ridge, Arctic Ocean: Transactions (Doklady) of the U.S.S.R. Academy of Sciences,
Earth Science Sections, v. 247, p. 73-77.
Dibner, V. D., 1998, The geology of Franz Josef Land - an introduction: Norsk Polarinstitutt
Meddelelser, v. 151, p. 10-17.
Drachev, S. S., Savostin, L. A., Groshev, V. G., and Bruni, I. E., 1998, Structure and geology of the
continental shelf of the Laptev Sea, eastern Russian Arctic: Tectonophysics, v. 298, p. 357-393.
Drachev, S. S., Bauch, H., Kassens, H., Kaul, N., Chizhov, D., and Roudoy, A., 2001, The Laptev Sea; a
natural laboratory for addressing the processes of rupture of continental lithosphere and their
impact on natural environment: Journal of Conference Abstracts, v. 6, p. 756.
Dunbar, J. A., and Sawyer, D. S., 1996, Three-dimensional dynamical model of continental rift
propagation and margin plateau formation: Journal of Geophysical Research, v. 101, p. 27845-
27863.
Dziewonski, A.M., Chou, T.-A., and Woodhouse, J.H., 1981, Determination of earthquake source
parameters from waveform data for studies of global and regional seismicity: Journal of
Geophysical Research, v. 86, p. 2825-2852.
109

Dziewonski, A.M. and Woodhouse, J.H., 1983, An experiment in the systematic study of global
seismicity: centroid-moment tensor solutions for 201 moderate and large earthquakes of 1981:
Journal of Geophysical Research, v. 88, p. 3247-3271
Edwards, M. H., Kurras, G. J., Tolstoy, M., Bohnenstiehl, D. R., Coakley, B. J., and Cochran, J. R., 2001,
Evidence of recent volcanic activity on the ultraslow-spreading Gakkel Ridge: Nature, v. 409, p.
808-812.
Eidvin, T., Jansen, E., and Riis, F., 1993, Chronology of Tertiary fan deposits off the western Barents Sea;
implications for the uplift and erosion history of the Barents Shelf: Marine Geology, v. 112, p. 109-
131.
Eiken, O., and Hinz, K., 1993, Contourites in the Fram Strait: Sedimentary Geology, v. 82, p. 15-32.
Einarsson, P., 1986, Seismicity along the eastern margin of the North American Plate, in Vogt, P. R., and
Tucholke, B. E., eds., The western North Atlantic region: Boulder, Colorado, Geological Society
of America, The Geology of North America, v. M, p. 99-116.
Eldholm, O., and Sundvor, E., 1980, The continental margin of the Norwegian-Greenland Sea; recent
results and outstanding problems: Philosophical Transactions of the Royal Society of London, v.
A294, p. 77-86.
Eldholm, O., and Windisch, C. C., 1974, Sediment distribution in the Norwegian-Greenland Sea:
Geological Society of America Bulletin, v. 85, p. 1661-1676.
Eldholm, O., Faleide, J. I., and Myhre, A. M., 1987, Continent-ocean transition at the western Barents
Sea-Svalbard margin: Geology, v. 15, p. 1118-1122.
Eldholm, O., Karasik, A. M., and Reksnes, P. A., 1990a, The North American plate boundary, in Grantz,
A., Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean Region: Boulder, Colorado, Geological
Society of America, The Geology of North America, v. L, p. 171-182.
Eldholm, O., Skogseid, J., Sundvor, E., and Myhre, A. M., 1990b, The Norwegian-Greenland Sea, in
Grantz, A., Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean Region: Boulder, Colorado,
Geological Society of America, The Geology of North America, v. L, p. 351-364.
Eldholm, O., Myhre, A. M., and Thiede, J., 1994, Cenozoic tectono-magmatic events in the Northern
Atlantic: potential paleoenvironmental implications, in Boulter, M. C., and Fischer, H. C., eds.,
Cenozoic plants and climates of the Arctic: Berlin, Springer-Verlag, NATO ASI Series, v. I 27, p.
35-55.
Eldholm, O., Richter, W., and Tebenkov, A., in press, Volcanic Provinces around the Eurasian Basin;
interplay with tectonism: Polarforschung.
Elverhøi, A., Andersen, E. S., Dokken, T., Hebbeln, D., Spielhagen, R., Svendsen, J. I., Sørflaten, M.,
Rørnes, A., Hald, M., and Forsberg, C. F., 1995, The growth and decay of the Late Weichselian ice
sheet in western Svalbard and adjacent areas based on provenance studies of marine sediments:
Quarternary Research, v. 44, p. 303-316.
Embry, A. F., 1990, Geological and geophysical evidence in the support of the hypothesis of
anticlockwise rotation of northern Alaska: Marine Geology, v. 93, p. 317-329.
Engdahl, E. R., van der Hilst, R., and Buland, R., 1998, Global teleseismic earthquake relocation with
improved travel times and procedures for depth determination: Bulletin of the Seismological
Society of America, v. 88, p. 722-743.
Engeln, J. F., Wiens, D. A., and Stein, S., 1986, Mechanisms and depths of Mid-Atlantic Ridge transform
faults: Journal of Geophysical Research, v. 91, p. 548-578.
Engen, Ø., Eldholm, O., Bungum, H., and Faleide, J. I., 2001, Plategrensen i Arktis: Geonytt, Norsk
geologisk forenings landsmøte, Oslo, 8.-10. januar 2001; Program og sammendrag, p. 45.
Faleide, J. I., Vågnes, E., and Gudlaugsson, S. T., 1993, Late Mesozoic-Cenozoic evolution of the south-
western Barents Sea in a regional rift-shear tectonic setting: Marine and Petroleum Geology, v. 10,
p. 186-214.
Feden, R. H., Vogt, P. R., and Fleming, H. S., 1979, Magnetic and bathymetric evidence for the “Yermak
Hot Spot” northwest of Svalbard in the Arctic Basin: Earth and Planetary Science Letters, v. 44, p.
18-38.
Fejerskov, M., Lindholm, C. D., Bungum, H., Myrvang, A., Bratlie, R. K., and Larsen, B. T., 1996, Crustal
stresses in Norway and adjacent offshore regions, Final report for the IBS-DNM project, topic 1.3
“Regional Stress Field”: Trondheim, Norwegian Institute of Technology, 29 p.
Forsberg, R., and Kenyon, S., 1999, Arctic Gravity Project [background paper]: Arctic Gravity Project,
National Imagery and Mapping Agency, http://www.nima.mil/GandG/agp, 4 p.
Fowler, C. M. R., 1990, The solid earth; an introduction to global geophysics: Cambridge, Cambridge
University Press, 472 p.
110

Fox, P. J., and Gallo, D. G., 1984, A tectonic model for ridge-transform-ridge plate boundaries:
Tectonophysics, v. 104, p. 205-242.
Francis, T. J. G., and Porter, I. T., 1971, A statistical study of mid-Atlantic Ridge earthquakes:
Geophysical Journal of the Royal Astronomical Society, v. 24, p. 31-50.
Francis, T. J. G., Porter, I. T., and Lilwall, R. C., 1978, Microearthquakes near the eastern end of St. Paul’s
Fracture Zone: Geophysical Journal of the Royal Astronomical Society, v. 53, p. 201-217.
Franke, D., Kruger, F., and Klinge, K., 2000, Tectonics of the Laptev Sea - Moma ‘Rift’ Region:
investigation with seismologic broadband data: Journal of Seismology, v. 4, p. 99-116.
Franke, D., Hinz, K., Block, M., Drachev, S. S., Neben, S., Kos’ko, M. K., Reichert, C., and Roeser, H. A.,
in press, Tectonics of the Laptev Sea region in north-eastern Siberia.
Frohlich, C., 2001, Display and quantitative assessment of distributions of earthquake focal
mechanisms: Geophysical Journal International, v. 144, p. 300-308.
Fujita, K., Cambray, F. W., and Velbel, M. A., 1990a, Tectonics of the Laptev Sea and Moma rift systems,
northeastern USSR: Marine Geology, v. 93, p. 95-118.
Fujita, K., and Cook, D. B., 1990b, The Arctic continental margin of eastern Siberia, in Grantz, A.,
Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean Region: Boulder, Colorado, Geological
Society of America, The Geology of North America, v. L, p. 289-304.
Géli, L., and Blanc, F., 1998, Mapping the sedimentary basins of the Barents and Kara Seas using ERS-1
altimetry-geodetic mission: Marine Geophysical Researches, v. 20, p. 109-127.
Goethe, J. W., 1833, Faust, zweiter Teil: Stuttgart and Tübingen, J. G. Sottasch.
Grachev, A. F. and Karasik, A. M., 1974, Sea-floor spreading and tectonics of the Eurasia Basin;
Geotectonic implications to the prospects of mineral resources on the Arctic shelf: Leningrad,
Nauchno-Issledovatel’skiy Institut Geologii Arktiki, Trudy, p. 19-33 (in Russian).
Grachev, A. F., and Naryshkin, G. D., 1978, The main features of sea-bottom topography of the Eurasia
Basin, the Arctic Ocean: Proceedings of the Leningrad State University, v. 2, p. 94-102 (in
Russian).
Gradstein, F. M., and Ogg, J. G., 1996, A Phanerozoic time scale: Episodes, v. 19, p. 3-5.
Granser, H., 1987, Three-dimensional interpretation of gravity data from sedimentary basins using an
exponential density-depth function: Geophysical Prospecting, v. 25, p. 1030-1041.
Gregersen, S., 1982, Earthquakes in Greenland: Bulletin of the Seismological Survey of Denmark, v. 31,
p. 11-27.
Grikurov, G. E., 1999, New bathymetry map “Bottom Relief of the Arctic Ocean” [abs.]: St. Petersburg,
VNIIOkeangeologia.
Grindlay, N. R., Madsen, J. A., Rommevaux-Jestin, C., and Sclater, J., 1998, A different pattern of ridge
segmentation and mantle Bouguer gravity anomalies along the ultra-slow spreading Southwest
Indian Ridge (15˚30’E to 25˚E): Earth and Planetary Science Letters, v. 161, p. 243-253.
Gutenberg, B., and Richter, C. F., 1949, Seismicity of the earth and associated phenomena: Princeton,
Princeton University Press, 273 p.
Harland, W. B., 1969, Contribution of Spitsbergen to understanding of tectonic evolution of North
Atlantic region, in Kay, M., ed., North Atlantic geology and continental drift: American
Association of Petroleum Geologists Memoir, v. 12, p. 817-851.
Harland, W. B., 1997, The Geology of Svalbard: Bath, The Geological Society, Geological Society
Memoir no. 17, 521 p.
Harvard CMT, Harvard Seismology; Centroid-Moment Tensor Project: http://
www.seismology.harvard.edu/projects/CMT/
HDNO-VNIIOkeangeologia, 1999, Bottom relief of the Arctic Ocean [map 1 : 5 000 000]: St. Petersburg.
Heezen, B. C., and Tharp, M., 1975, Map of the Arctic region, 1:5 000 000: New York, American
Geographical Society.
Heidland, K., Hinze, H., Monk, J., Niederjasper, F., Schenke, H.-W., and Schöne, T., 1995, Multibeam
bathymetric data of the Molloy Deep, Hovgård Ridge and Vesterisbanken, in Crane, K., and
Solheim, A., eds., Seafloor atlas of the northern Norwegian-Greenland Sea: Norsk Polarinstitutt
Meddelelser, v. 137, p. 144-149.
Hellebrand, E. W., Snow, J. E., and Mühe, R., in press, Mantle melting on Gakkel Ridge (Arctic Ocean);
Abyssal peridotite spinel compositions: Chemical Geology.
Hicks, E. C., Bungum, H., and Lindholm, C. D., 2000, Stress inversion of earthquake focal mechanism
solutions from onshore and offshore Norway: Norsk Geologisk Tidsskrift, v. 80, p. 235-250.
Høgden, S., 1999, Seismotectonics and crustal structure of the Svalbard region [Cand. Scient. thesis]:
Oslo, University of Oslo, 142 p.
111

Horsfield, W. T., and Maton, P. I., 1970, Transform faulting along the De Geer Line: Nature, v. 226, p.
256-257.
Husebye, E. S., Gjøystdal, H., Bungum, H., and Eldholm, O., 1975, The seismicity of the Norwegian and
Greenland Seas and adjacent continental shelf areas: Tectonophysics, v. 26, p. 55-70.
Idriss, I. M., 1985, Evaluating seismic risk in engineering practice, in Proceedings of the 11th
International Conference on Soil Mechanics and Foundation Engineering, San Francisco, v. 1, p.
255-320.
Jackson, H. R., Reid, I., and Falconer, R. K. H., 1982, Crustal structure near the Arctic Mid-Ocean Ridge:
Journal of Geophysical Research, v. 87, p. 1773-1783.
Jackson, H. R., Johnson, G. L., Sundvor, E., and Myhre, A. M., 1984, The Yermak Plateau - formed at a
triple junction: Journal of Geophysical Research, v. 89, p. 3223-3232.
Jackson, H. R., Forsyth, D. A., Hall, J. K., and Overton, A., 1990, Seismic reflection and refraction, in
Grantz, A., Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean Region: Boulder, Colorado,
Geological Society of America, The Geology of North America, v. L, p. 153-170.
Jakobsson, M., 1999, First high-resolution chirp sonar profiles from the central Arctic Ocean reveal
erosion of Lomonosov Ridge sediments: Marine Geology, v. 158, p. 111-123.
Jakobsson, M., 2000, Mapping of the Arctic Ocean: Bathymetry and Pleistocene paleoceanography [Ph.
D. thesis]: Stockholm, Stockholm University, 94 p.
Jakobsson, M., Cherkis, N. Z., Woodward, J., Macnab, R., and Coakley, B., 2000a, New grid of Arctic
bathymetry aids scientists and mapmakers: Eos, Transactions, American Geophysical Union, v.
81, p. 89-96.
Jakobsson, M., Macnab, R., and Members of the Editorial Board, 2000b, International Bathymetric Chart
of the Arctic Ocean beta version; technical reference and user’s guide: IBCAO, http://
www.ngdc.noaa.gov/mgg/bathymetry/arctic/arctic.html, 14 p.
Jeffreys, H., and Bullen, K. E., 1940, Seismological tables: London, British Association for the
Advancement of Science.
Jemsek, J. P., Bergman, E. A., Nabelek, J. L., and Solomon, S. C., 1986, Focal depths and mechanisms of
large earthquakes on the Mid-Arctic Ridge system: Journal of Geophysical Research, v. 91, p.
13993-14005.
Johnson. G. L., and Heezen, B. C., 1967, Morphology and evolution of the Norwegian-Greenland Sea:
Deep-Sea Research and Oceanographic Abstracts, v. 14, p. 755-771.
Johnson, G. L., Monahan, D., Grønlie, G., and Sobczak, L., 1979, General Bathymetric Chart of the
Oceans (GEBCO) Sheet 5.17 [map 1 : 6 000 000]: Ottawa, Canadian Hydrographic Service.
Jokat, W., 1998, The sediment distribution below the North Greenland continental margin and the
adjacent Lena Trough: Polarforschung, v. 68, p. 71-82.
Jokat, W., Uenzelmann-Neben, G., Kristoffersen, Y., and Rasmussen, T., 1992, ARCTIC ‘91; Lomonosov
Ridge - a double-sided continental margin: Geology, v. 20, p. 887-890.
Jokat, W., Weigelt, E., Kristoffersen, Y., Rasmussen, T., and Schöne, T., 1995a, New insights into the
evolution of the Lomonosov Ridge and the Eurasian Basin: Geophysical Journal International, v.
122, p. 378-392.
Jokat, W., Weigelt, E., Kristoffersen, Y., Rasmussen, T., and Schöne, T., 1995b, New geophysical results
from the south-western Eurasian Basin (Morris Jesup Rise, Gakkel Ridge, Yermak Plateau) and
the Fram Strait: Geophysical Journal International, v. 123, p. 601-610.
Jones, M. T., Tabor, A. R., and Weatherall, P., 1994, GEBCO Digital Atlas [CD-ROM]: Birkenhead, British
Oceanographic Data Centre.
Kanamori, H., and Stewart, G. S., 1976, Mode of strain release along the Gibbs fracture zone, Mid-
Atlantic Ridge: Physics of the Earth and Planetary Interiors, v. 11, p. 312-332.
Karasik, A. M., 1974, Yevraziyskiy basseyn Severnogo Ledovitogo okeana s pozitsiy tektoniki plit, in
Lazurkin, V. M., Gramberg, I. S., Ravich, M. G., and Tkachenko, B. V., eds., Problemy geologii
polyarnykh oblastey Zemli, sbornik statey: Leningrad, NIRDA, p. 23-31.
Karasik, A. M., and Pozdnyakova, R. A., 1979, Basement depth versus its age in Eurasian Basin, Arctic
Ocean: Doklady Akademia Nauk SSSR, v. 248, p. 169-174.
Karlberg, T., 1995, En geofysisk undersøkelse av Hovgårdryggen [Cand. Scient. thesis]: Oslo,
University of Oslo, 104 p.
Kearey, P., and Brooks, M., 1991, An introduction to geophysical prospecting: Oxford, Blackwell
Science, 254 p.
Kennett, B. L. N., Engdahl, E. R., and Buland, R., 1995, Constraints on seismic velocities in the Earth
from traveltimes: Geophysical Journal International, v. 122, p. 108-124.
112

Kim, V. I., and Verba, V. V., 1995, The geological structure of the Laptev Shelf and adjacent parts of the
Eurasian Subbasin (in the context of planned drilling), in Kassens, H., Piepenburg, D., Thiede, J.,
Timokhov, L., Hubberten, H.-W., and Priamikov, S. M., eds., Russian-German cooperation: the
Laptev Sea system: Bremerhaven, Alfred Wegener Institute for Polar and Marine Research,
Reports on Polar Research, v. 176, p. 383-387.
Kovacs, L. C., and Vogt, P. R., 1982, Depth-to-magnetic source analysis of the Arctic Ocean region:
Tectonophysics, v. 89, p. 255-294.
Kovacs, L. C., Johnson, G. L., Srivastava, S. P., Taylor, P. T., and Vogt, P. R., 1990, Residual magnetic
anomaly chart of the Arctic Ocean region, in Grantz, A., Johnson, L., and Sweeney, J. F., eds., The
Arctic Ocean Region: Boulder, Colorado, Geological Society of America, The Geology of North
America, v. L, Plate 4.
Koz’min, B. M., 1984, Seismicheskie poyasa i mekhanismy ochagov ikh zemletryasenii: Moscow,
Nauka, 126 p.
Kristoffersen, Y., 1978, Sea-floor spreading and the early opening of the North Atlantic: Earth and
Planetary Science Letters, v. 38, p. 273-290.
Kristoffersen, Y., 1982, The Nansen Ridge, Arctic Ocean; some geophysical observations of the rift
valley at slow spreading rate: Tectonophysics, v. 89, p. 161-172.
Kristoffersen, Y., 1990a, Eurasia Basin, in Grantz, A., Johnson, L., and Sweeney, J. F., eds., The Arctic
Ocean Region: Boulder, Colorado, Geological Society of America, The Geology of North
America, v. L, p. 365-378.
Kristoffersen, Y., 1990b, On the tectonic evolution and paleoceanographic significance of the Fram
Strait gateway, in Bleil, U., and Thiede, J., eds., Geological history of the Polar Oceans: Arctic
versus Antarctic: Dordrecht, Kluwer Academic Publishers, NATO ASI Series, v. C 308, p. 63-76.
Kristoffersen, Y., and Husebye, E. S., 1985, Multi-channel seismic reflection measurements in the
Eurasian Basin, Arctic Ocean, from US ice station FRAM IV: Tectonophysics, v. 114, p. 103-115.
Kristoffersen, Y., and Talwani, M., 1977, The extinct triple junction south of Greenland and the Tertiary
motion of Greenland relative to North America: Geological Society of America Bulletin, v. 88, p.
1037-1049.
Kristoffersen, Y., Husebye, E. S., Bungum, H., and Gregersen, S., 1982, Seismic investigations of the
Nansen Ridge during the FRAM I experiment: Tectonophysics, v. 82, p. 57-68.
Laxon, S., and McAdoo, D., 1994, Arctic Ocean gravity derived from ERS-1 satellite altimetry: Science,
v. 265, p. 621-624.
Laxon, S., and McAdoo, D., 1998, Satellites provide new insights into polar geophysics: Eos,
Transactions, American Geophysical Union, v. 79, p. 69-74.
Lawver, L. A., Müller, R. D., Srivastava, S. P., and Roest, W., 1990, The opening of the Arctic Ocean, in
Bleil, U., and Thiede, J., eds., Geological history of the polar oceans; Arctic versus Antarctic:
Dordrecht, Kluwer Academic Publishers, NATO ASI Series, v. C 308, p. 29-62.
Lazareva, A. P., and Misharina, L. A., 1965, Stresses in earthquake foci in the Arctic seismic belt:
Izvestiya Academy of Sciences of the USSR, Physics of the Solid Earth, v. 1, p. 84-87.
Le Pichon, X., and Hayes, D. E., 1971, Marginal offsets, fracture zones, and the early opening of the
South Atlantic: Journal of Geophysical Research, v. 76, p. 6283-6293.
Le Pichon, X., Francheteau, J., and Bonnin, J., 1973, Plate tectonics: Amsterdam, Elsevier Scientific
Publishing Company, Developments in geotectonics v. 6, 300 p.
Levshin, A. L., Ritzwoller, M. H., Barmin, M. P., Villaseñor, A., and Padgett, C. A., 2001, New
constraints on the arctic crust and uppermost mantle; surface wave group velocities, Pn, and Sn:
Physics of the Earth and Planetary Interiors, v. 123, p. 185-204.
Lin, J., Purdy, G. M., Schouten, H., Sempéré, J. C., and Zervas, C., 1990, Evidence from gravity data for
focused magmatic accretion along the Mid-Atlantic Ridge: Nature, v. 344, p. 627-632.
Lindholm, C. D., and Bungum, H., 2000, Probabilistic seismic hazard; a review of the seismological
frame of reference with examples from Norway: Soil Mechanics and Earthquake Engineering, v.
20, p. 27-38.
Lindholm, C. D., Havskov, J., and Sellevoll, M. A., 1990, Periodicity in seismicity; examination of four
catalogs: Tectonophysics, v. 191, p. 155-164.
Louden, K. E., and Forsyth, D. W., 1976, Thermal conduction across fracture zones and the gravitational
edge effect: Journal of Geophysical Research, v. 81, p. 4869-4874.
Louden, K. E., Osler, J. C., Srivastava, S. P., and Keen, C. E., 1996, Formation of oceanic crust at slow
spreading rates; new constraints from an extinct spreading center in the Labrador Sea: Geology,
v. 24, p. 771-774.
113

Ludwig, J. W., Nafe, J. E., and Drake, C. L., 1970, Seismic refraction, in Maxwell, A. E., ed., The Sea:
New York, John Wiley, p. 53-84.
Lyle, M., and Ness, G. E., 1991, The opening of the southern Gulf of California, in Dauphin, J. P., and
Simoneit, B. R. T., eds., The gulf and peninsular province of the Californias: Tulsa, American
Association of Petroleum Geologists, AAPG Memoir 47, p. 403-423.
Macdonald, K. C., 1982, Mid-ocean ridges; fine scale tectonics, volcanic and hydrothermal processes
within the plate boundary zone: Annual Review of Earth and Planetary Science, v. 10, p. 155-190.
Maher, H. D., Jr., 2001, Manifestations of Cretaceous High Arctic large igneous province in Svalbard:
Journal of Geology, v. 109, p. 91-104.
Marks, K. M., and Stock, J.-M., 1994, Variations in ridge morphology and depth-age relationships on
the Pacific-Antarctic Ridge: Journal of Geophysical Research, v. 99, p. 531-541.
Martinez, F., and Cochran, J. R., 1988, Structure and tectonics of the northern Red Sea; catching a
continental margin between rifting and drifting: Tectonophysics, v. 150, p. 1-32.
Mendi, C. D., Ruud, B. O., and Husebye, E. S., 1997, The North Sea Lg-blockage puzzle: Geophysical
Journal International, v. 130, p. 669-680.
Miller, A. D., Foulger, G. R., and Julian, B. R., 1998, Non-double-couple earthquakes; 2, Observations:
Reviews of Geophysics, v. 36, p. 551-568.
Mitchell, B. J., Bungum, H., Chan, W. W., and Mitchell, P. B., 1990, Seismicity and present-day tectonics
of the Svalbard region: Geophysical Journal International, v. 102, p. 139-149.
Mühe, R., Devey, C. W., and Bohrmann, H., 1993, Isotope and trace element geochemistry of MORB
from the Nansen-Gakkel Ridge at 86˚ north: Earth and Planetary Science Letters, v. 120, p. 103-
109.
Mühe, R., Bohrmann, H., Garbe-Schönberg, D., and Kassens, H., 1997, E-MORB glasses from the
Gakkel Ridge (Arctic Ocean) at 87˚N; evidence for the Earth’s most northerly volcanic activity:
Earth and Planetary Science Letters, v. 152, p. 1-9.
Müller, C., and Jokat, W., 2000, Seismic evidence for volcanic activity discovered in Central Arctic: Eos,
Transactions, v. 81, p. 265.
Myhre, A. M., and Eldholm, O., 1988, The western Svalbard margin (74-80˚N): Marine and Petroleum
Geology, v. 5, p. 134-156.
Nakano, H., 1923, Notes on the nature of forces which give rise to earthquake motions: Seismological
Bulletin of the Central Meteorological Observatory of Japan, v. 1, p. 92-120.
Nalivkin, D. V., 1973, Geology of the USSR, translated by Rast, N.: Edinburgh, Oliver and Boyd, 855 p.
Nettleton, L. L., 1976, Gravity and magnetics in oil prospecting: New York, McGraw-Hill, 464 p.
Nilsen, C. L., 1996, Sen-kenozoisk utvikling av Franz Victoria-renna, nordøstlige Barentshavet [Cand.
Scient. thesis]: Oslo, University of Oslo, 128 p.
Oldenburg, D. W., 1974, The inversion and interpretation of gravity anomalies: Geophysics, v. 39, p.
526-536.
Ostenso, N. A., 1974, Arctic Ocean margins, in Burk, C. A., and Drake, C. L., eds., The geology of
continental margins: New York, Springer-Verlag, p. 753-763.
Ostenso, N. A., and Wold, R. J., 1977, A seismic and gravity profile across the Arctic Ocean Basin:
Tectonophysics, v. 37, p. 1-24.
Parfenov, L. M., Koz’min, B. M., Grinenko, O. V., Imaev, V. S., and Imaeva, L. P., 1988, Geodynamics of
the Chersky seismic belt: Journal of Geodynamics, v. 54, p. 15-37.
Parker, R. L., 1972, The rapid calculation of potential anomalies: Geophysical Journal of the Royal
Astronomical Society, v. 31, p. 447-455.
Parmentier, E. M., and Phipps Morgan, J., 1990, Spreading rate dependence of three-dimensional
structure in oceanic spreading centres: Nature, v. 348, p. 325-328.
Perry, R. K., Fleming, H. S., Cherkis, N. Z., Feden, R. H., and Vogt, P. R., 1980, Bathymetry of the
Norwegian-Greenland and western Barents seas [map 1 : 2 333 230]: Geological Society of
America Map and Chart Series, MC-21.
Perry, R. K., Fleming, H. S., Weber, J. R., Kristoffersen, Y., Hall, J. K., Grantz, A., Johnson, G. L., Cherkis,
N. Z., and Larsen, B., 1985, Bathymetry of the Arctic Ocean [map 1 : 4 704 075]: Geological
Society of America Map and Chart Series, MC-56.
Pitman, W. C., III, and Talwani, M., 1972, Sea-floor spreading in the North Atlantic: Geological Society
of America Bulletin, v. 83, p. 619-646.
Press, F., and Ewing, W. M., 1952, Note on refracted waves in a layer: Geological Society of America
Bulletin, v. 63, p. 1356.
Rasmussen, E., and Fjeldskaar, W., 1996, Quantification of the Pliocene-Pleistocene erosion of the
Barents Sea from present-day bathymetry: Global and Planetary Change, v. 12, p. 119-133.
114

Raymo, M. E., Jansen, E., Blum, P., and Herbert, T. D., eds., 1999, Proceedings of the Ocean Drilling
Program, Scientific Results, v. 162: College Station, Texas (Ocean Drilling Program), 295 p.
Reid, I., and Jackson, H. R., 1981, Oceanic spreading rate and crustal thickness: Marine Geophysical
Researches, v. 5, p. 165-172.
Reksnes, P. A., and Vågnes, E., 1985, Evolution of the Greenland Sea and Eurasia Basin [Cand. Scient.
thesis]: Oslo, University of Oslo, 136 p.
Ren, S., Faleide, J. I., Eldholm, O., Skogseid, J., and Gradstein, F., in press, Late Cretaceous-Paleocene
tectonic development of the NW Vøring Basin: Marine and Petroleum Geology.
Ringdal, F., 1976, Maximum-likelihood estimation of earthquake magnitude: Bulletin of the
Seismological Society of America, v. 66, p. 789-802.
Roeser, H. A., Block, M., Hinz, K., and Reichert, C., 1995, Marine geophysical investigations in the
Laptev Sea and the western part of the East Siberian Sea, in Kassens, H., Piepenburg, D., Thiede,
J., Timokhov, L., Hubberten, H.-W., and Priamikov, S. M., eds., Russian-German cooperation: the
Laptev Sea system: Bremerhaven, Alfred Wegener Institute for Polar and Marine Research,
Reports on Polar Research, v. 176, p. 367-377.
Rowlett, H., 1981, Seismicity at intersections of spreading centers and transform faults: Journal of
Geophysical Research, v. 86, p. 3815-3820.
Rowley, D. B., and Lottes, A. L., 1988, Plate-kinematic reconstructions of the North Atlantic and Arctic:
Late Jurassic to present: Tectonophysics, v. 155, p. 73-120.
Sandwell, D. T., 1986, Thermal stress and the spacings of transform faults: Journal of Geophysical
Research, v. 91, p. 6405-6417.
Sandwell, D. T., and Smith, W. H. F., 1997, Marine gravity anomaly from Geosat and ERS 1 satellite
altimetry: Journal of Geophysical Research, v. 102, p. 10039-10054.
Savostin, L. A., and Karasik, A. M., 1981, Recent plate tectonics of the Arctic Basin and of northeastern
Asia: Tectonophysics, v. 74, p. 111-145.
Scholz, C. H., 1990, The mechanics of earthquakes and faulting: Cambridge, Cambridge University
Press, 439 p.
Schouten, H., Klitgord, K. D., and Whitehead, J. A., 1985, Segmentation of mid-ocean ridges: Nature, v.
317, p. 225-229.
Shah, A. K., and Sempéré, J.-C., 1998, Morphology and transition from an axial high to a rift valley at
the Southeast Indian Ridge and the relation to variations in mantle temperature: Journal of
Geophysical Research, v. 103, p. 5203-5223.
Sipkin, S. A., Bufe, C. G., and Zirbes, M. D., 1999, Moment tensor solutions estimated using optimal
filter theory; global seismicity, 1997: Physics of the Earth and Planetary Interiors, v. 114, p. 109-
117.
Smith, W. H. F., and Sandwell, D. T., 1994, Bathymetric prediction from dense satellite altimetry and
sparse shipboard bathymetry: Journal of Geophysical Research, v. 99, p. 21803-21824.
Snow, J., and Hellebrand, E., 2001, Magmatic and hydrothermal activity in Lena Trough, Arctic Ocean:
Eos, Transactions, American Geophysical Union, v. 82, p. 193-198.
Sobczak, L. W., Hearty, D. B., Forsberg, R., Kristoffersen, Y., Eldholm, O., and May, S. D., 1990, Gravity
from 64˚N to the North Pole, in Grantz, A., Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean
Region: Boulder, Colorado, Geological Society of America, The Geology of North America, v. L,
p. 101-118.
Solomon, S. C., Huang, P. Y., and Meinke, L., 1988, The seismic moment budget of slowly spreading
ridges: Nature, v. 334, p. 58-60.
Soper, N. J., and Higgins, A. K., 1991, Late Cretaceous - early Tertiary deformation, North Greenland, in
Trettin, H. P., ed., Geology of the Innuitian Orogen and Arctic Platform of Canada and
Greenland: Ottawa, Geological Survey of Canada, The geology of North America, v. E, p. 459-
465.
Srivastava, S. P., 1985, Evolution of the Eurasian Basin and its implication to the motion of Greenland
along the Nares Strait: Tectonophysics, v. 114, p. 29-53.
Stein, S., Melosh, H. J., and Minster, J. B., 1977, Ridge migration and asymmetric sea-floor spreading:
Earth and Planetary Science Letters, v. 36, p. 51-62.
Stein, S., Sleep, N. H., Geller, R. J., Wang, S.-C., and Kroeger, G. C., 1979, Earthquakes along the passive
margin of eastern Canada: Geophysical Research Letters, v. 6, p. 537-540.
Stein, S., Cloetingh, S., Sleep, N. H., and Wortel, R., 1989, Passive margin earthquakes, stresses and
rheology, in Gregersen, S., and Basham, P. W., eds., Earthquakes at North-Atlantic passive
margins; neotectonics and postglacial rebound: Dordrecht, D. Reidel, NATO ASI Series, v. C 266,
p. 231-259.
115

Su, W., Mutter, C. Z., Mutter, J. C., and Buck, W. R., 1994, Some theoretical predictions on the
relationships among spreading rate, mantle temperature, and crustal thickness: Journal of
Geophysical Research, v. 99, p. 3215-3227.
Sundvor, E., and Austegard, A., 1990, The evolution of the Svalbard margins: synthesis and new
results, in Bleil, U., and Thiede, J., eds., Geological history of the polar oceans; Arctic versus
Antarctic: Dordrecht, Kluwer Academic Publishers, NATO ASI Series, v. C 308, p. 77-94.
Sundvor, E., and Eldholm, O., 1979, The western and northern margin off Svalbard: Tectonophysics, v.
59, p. 239-250.
Sundvor, E., Eldholm, O., Gidskehaug, A., and Myhre, A. M., 1977, Marine geophysical survey on the
western and northern continental margin off Svalbard: University of Bergen Seismological
Observatory, Scientific Report no. 4, 36 p.
Sundvor, E., Gidskehaug, A., Myhre, A. M., and Eldholm, O., 1978, Marine geophysical survey on the
northern Svalbard margin: University of Bergen Seismological Observatory, Scientific Report no.
5, 46 p.
Sundvor, E., Myhre, A. M., Austegard, A., Haugland, K., and Gidskehaug, G., 1982, Marine geophysical
survey on the Yermak Plateau: University of Bergen Seismological Observatory, Scientific Report
no. 7, 29 p.
Sundvor, E., Eldholm, O., Gladczenko, T. P., and Planke, S., 2000, Norwegian-Greenland Sea thermal
field, in Nøttvedt, A., ed., Dynamics of the Norwegian margin: London, Geological Society
Special Publication, v. 167, p. 397-410.
Sykes, L. R., 1965, The seismicity of the Arctic: Bulletin of the Seismological Society of America, v. 55, p.
519-536.
Sykes, L. R., 1967, Mechanisms of earthquakes and nature of faulting on the mid-oceanic ridges:
Journal of Geophysical Research, v. 72, p. 2131-2153.
Sykes, L. R., 1970, Earthquake swarms and sea-floor spreading: Journal of Geophysical Research, v. 75,
p. 6598-6611.
Sykes, L. R., 1972, Mechanism of earthquakes and nature of faulting on the mid-oceanic ridges: Journal
of Geophysical Research, v. 72, p. 5-27.
Sykes, L. R., and Sbar, M. L., 1974, Focal mechanism solutions of intraplate earthquakes and stresses in
the lithosphere, in Kristjansson, L., ed., Geodynamics of Iceland and the North Atlantic area:
Boston, D. Reidel, p. 207-224.
Talwani, M., and Eldholm, O., 1977, Evolution of the Norwegian-Greenland Sea: Geological Society of
America Bulletin, v. 88, p. 969-999.
Taylor, B. Goodliffe, A. M., and Martinez, F., 1999, How continents break up; insights from Papua New
Guinea: Journal of Geophysical Research, v. 104, p. 7497-7512.
Telford, W. M., Geldart, L. P., and Sheriff, R. E., 1990, Applied geophysics: Cambridge, Cambridge
University Press, 770 p.
Thiede, J., 1988, Scientific cruise report of Arctic expedition ARK IV/3: Bremerhaven, Alfred Wegener
Institute for Polar and Marine Research, Reports on Polar Research v. 43, 237 p.
Thiede, J., and Myhre, A. M., 1996, Introduction to the North Atlantic-Arctic Gateways; plate tectonic-
paleoceanographic history and significance, in Thiede, J., Myhre, A. M., Firth, J. V., Johnson, G.
L., and Ruddiman, W. F., eds., Proceedings of the Ocean Drilling Program, Scientific Results, v.
151: College Station, Texas (Ocean Drilling Program), p. 3-23.
Thiede, J., Pfirman, S., Schenke, H.-W., and Reil, W., 1990, Bathymetry of Molloy Deep: Fram Strait
between Svalbard and Greenland: Marine Geophysical Researches, v. 12, p. 197-214.
Thiede, J., Myhre, A. M., Firth, J. V., Johnson, G. L., and Ruddiman, W. F., eds., 1996, Proceedings of the
Ocean Drilling Program, Scientific Results, v. 151: College Station, Texas, 685 p.
Vågnes, E., 1996, Cenozoic deposition in the Nansen Basin, a first-order estimate based on present-day
bathymetry: Global and Planetary Change, v. 12, p. 149-157.
Verhoef, J., Roest, W. J., Macnab, R., Arkani-Hamed, J., and members of the Project Team, 1996,
Magnetic anomalies of the Arctic and North Atlantic Oceans and adjacent land areas: Geological
Survey of Canada Open File 3125b.
Vogt, P. R., 1979, Amplitudes of oceanic magnetic anomalies and the chemistry of oceanic crust;
synthesis and review of magnetic telechemistry: Canadian Journal of Earth Sciences, v. 16, p.
2236-2262.
Vogt, P. R., 1986a, Geophysical and geochemical signatures and plate tectonics, in Hurdle, B. G., ed.,
The Nordic seas: New York, Springer-Verlag, p. 413-662.
116

Vogt, P. R., 1986b, The present plate boundary configuration, in Vogt, P. R., and Tucholke, B. E., eds.,
The Western North Atlantic Region: Boulder, Colorado, Geological Society of America, The
Geology of North America, v. M, p. 189-204.
Vogt, P. R., Taylor, P. T., Kovacs, L. C., and Johnson, G. L., 1979, Detailed aeromagnetic investigation of
the Arctic Basin: Journal of Geophysical Research, v. 84, p. 1071-1089.
Vogt, P. R., Perry, R. K., Feden, R. H., Fleming, H. S., and Cherkis, N. Z., 1981, The Greenland-
Norwegian Sea and Iceland environment; geology and geophysics, in Nairn, A. E. M., Churkin,
M., Jr., and Stehli, F. G., eds., The Arctic Ocean: New York, Plenum Press, The ocean basins and
margins, p. 493-598.
Vogt, P. R., Crane, K., and Sundvor, E., 1994, Deep Pleistocene iceberg ploughmarks on the Yermak
Plateau; sidescan and 3.5 kHz evidence for thick calving ice fronts and a possible marine ice
sheet in the Arctic Ocean: Geology, v. 22, p. 403-406.
Vogt, P. R., Jung, W.-Y., and Brozena, J., 1998; Arctic margin gravity highs: deeper meaning for sediment
depocenters?: Marine Geophysical Researches, v. 20, p. 459-477.
Vorren, T. O., Richardsen, G., Knutsen, S. M., and Henriksen, E., 1991, Cenozoic erosion and
sedimentation in the western Barents Sea: Marine and Petroleum Geology, v. 8, p. 317-340.
Weber, J. R., 1979, The Lomonosov Ridge Experiment; LOREX 79: Eos, Transactions, American
Geophysical Union, v. 60, p. 715-720.
Weber, J. R., and Sweeney, J. F., 1990, Ridges and basins in the central Arctic Ocean, in Grantz, A.,
Johnson, L., and Sweeney, J. F., eds., The Arctic Ocean Region: Boulder, Colorado, Geological
Society of America, The Geology of North America, v. L, p. 305-336.
Weigelt, E., 1998, The crustal structure and sedimentary cover of the Eurasian Basin, Arctic Ocean:
Results from seismic and gravity measurements: Bremerhaven, Alfred Wegener Institute for
Polar and Marine Research, Reports on Polar Research v. 261, 127 p.
Weigelt, E., and Jokat, W., 2001, Peculiarities of roughness and thickness of oceanic crust in the Eurasian
Basin, Arctic Ocean: Geophysical Journal International, v. 145, p. 505-516.
Wessel, P., and Smith, W. H. F., 1991, Free software helps map and display data: Eos, Transactions,
American Geophysical Union, v. 72, p. 441-446.
Wetmiller, R. J., and Horner, R. B., 1978, Canadian earthquakes - 1976: Seismological Service of Canada,
Seismological Series, no. 79, 75 p.
Wetmiller, R. J., and Forsyth, D. A., 1982, Review of seismicity and other geophysical data near Nares
Strait: Meddelelser om Grønland, Geoscience, v. 8, p. 261-274.
White, R. S., Minshull, T. A., Bickle, M. J., and Robinson, C. J., 2001, Melt generation at very slow-
spreading oceanic ridges; constraints from geochemical and geophysical data: Journal of
Petrology, v. 42, p. 1171-1196.
Wolfe, C. J., Bergman, E. A., and Solomon, S. C., 1993, Oceanic transform earthquakes with unusual
mechanisms or locations; relation to fault geometry and state of stress in the adjacent
lithosphere: Journal of Geophysical Research, v. 98, p. 16187-16211.
Wonnacott, T. H., and Wonnacott, R. J., 1985, Introductory statistics: New York, Wiley, 649 p.
Woodhouse, J. H., and Dziewonski, A. M., 1984, Mapping the upper mantle: three dimensional
modelling of Earth structure by inversion of seismic waveforms, Journal of Geophysical
Research, v. 89, p. 5953-5986.
Wyss, M., 1973, Towards a physical understanding of the earthquake frequency distribution:
Geophysical Journal of the Royal Astronomical Society, v. 31, p. 341-359.
Ziegler, W. H., Doery, R., and Scott, J., 1986, Tectonic habitat of Norwegian oil and gas, in Spencer, A.
M., ed., Habitat of hydrocarbons on the Norwegian continental shelf; proceedings of an
international conference: London, Graham & Trotman, p. 3-19.

You might also like