You are on page 1of 22

Journal of Structural Geology 83 (2016) 134e155

Contents lists available at ScienceDirect

Journal of Structural Geology


journal homepage: www.elsevier.com/locate/jsg

Partitioning of Holocene kinematics and interaction between the


Theistareykir Fissure Swarm and the Husavik-Flatey Fault, North
Iceland

 ttir b, F.A. Pasquare
 Mariotto c
A. Tibaldi a, *, F.L. Bonali a, P. Einarsson b, A.R.
Hjartardo
a
b
c

Department of Earth and Environmental Sciences, University of Milan-Bicocca, Milan, Italy


Institute of Earth Sciences, University of Iceland, Reykjavik, Iceland
Department of Theoretical and Applied Sciences, Insubria University, Varese, Italy

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 19 August 2015
Received in revised form
11 December 2015
Accepted 3 January 2016
Available online 7 January 2016

Our research is aimed at contributing to the general understanding of how transform-rift junctions work,
a topic that can be studied in exceptional detail in North Iceland, where the active transform HusavikFlatey Fault (HFF) connects with the Gudnnugja Fault (GF), the westernmost structure of the Theistareykir Fissure Swarm (TFS). We studied in the eld: i) offsets along the easternmost HFF, ii) the geometry and kinematics of 649 faults and 1208 tension fractures in the TFS, iii) the interactions among all
these structures. The HFF transtensional kinematics is compatible with the GF, which shows different
offsets north and south of the junction between these two faults. We suggest the possible prolongation of
the HFF beyond the junction, based on: i) the change in offsets and strikes of TFS normal faults, ii) the
chelon, NNW-SSE-striking normal faults and tension fractures with a slight right-lateral
presence of en-e
component, iii) the transition of some of the faults into tension fractures north of the prolongation of the
HFF, and iv) the decrease in the cumulated offset of all the faults north of the HFF prolongation. We
interpret these data as eld evidence of the rst stages of propagation of the HFF upward or across the
TFS: This has clear implications for dening the potential rupture length of the HFF and, hence, for
seismic hazard assessment.
2016 Elsevier Ltd. All rights reserved.

Keywords:
Theistareykir ssure swarm
Transform fault
Normal fault
Tension fracture
Iceland

1. Introduction
The emerged oceanic rift zone of Iceland provides fundamental
insights into the role of faulting vs magmatism. Remote tensile
stresses from crustal plate motions accumulate over decades before
being released during relatively short time periods. Deformation
takes place at the surface by generation of new fractures, incremental offset along pre-existing faults and tension fracture dilation,
and is usually accompanied by dyke intrusion (e.g. Bjornsson, 1985;
Gudmundsson, 1987a; Rubin and Pollard, 1988; Opheim and
Gudmundsson, 1989; Jonsson et al., 1997). Also at depth, extension may be accommodated by intrusions along vertical planes, as
already described in several classic studies (e.g. Walker, 1992, 1999;
Helgason and Zentilli, 1985; Gudmundsson, 1990, 1995). In Iceland,

* Corresponding author. Department of Earth and Environmental Sciences, University of Milan-Bicocca, P. della Scienza 4, 20126 Milan, Italy.
E-mail address: alessandro.tibaldi@unimib.it (A. Tibaldi).
http://dx.doi.org/10.1016/j.jsg.2016.01.003
0191-8141/ 2016 Elsevier Ltd. All rights reserved.

normal faulting at the surface is associated with the development


of scarps and tension fractures, whereas at a depth ranging from
few hundred metres to few kilometres, extension is accommodated
by shear faulting along planes dipping mostly at 65 e75 , as indicated by structural geology studies (Saemundsson, 1980;
m, 1991; Forslund and Gudmundsson,
Gudmundsson and Backstro
1991, 1992; Gudmundsson, 1992; Angelier et al., 1997;
 ttir et al., 2009, 2012, 2015; Hjartardo
ttir and Einarsson,
Hjartardo
2012) and focal mechanism analyses (Einarsson, 1991). Tension
fracture opening can also occur along tens-of-km-long structures in
the axial rift zones of Iceland, as related to magmatic pressure
induced by dyke intrusions (Gudmundsson, 1987a, 1987b, 1995;
Opheim and Gudmundsson, 1989; Sigmundsson et al., 2015). In
fact, at the tip of an advancing dyke, tensile stresses can increase
due to magma injection (Rubin and Pollard, 1988; Rubin, 1992) and
exceed the Coulomb failure threshold, triggering earthquakes in the
ttir and Einarsson, 1979; Einarsson and
brittle crust (Brandsdo
ttir, 1980; Feigl et al., 2000; Cattin et al., 2005; Wright
Brandsdo
et al., 2012; Sigmundsson et al., 2015).

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Despite the large number of works dealing with the detailed


description of fault and dyke swarms within onland rift zones,
usually transform faulterift junctions are described in oceanic environments. Geophysical studies of oceanic sea oor structures
cannot be conducted at the same level of detail as eld studies.
Therefore, more onland data about this type of junction are needed
to better constrain and understand their geometry, kinematics and
distribution of deformation. The latter issue is particularly
intriguing because rift zones can be subjected to deformation
induced by both regional tectonic stresses as well as magmatic
pressure linked to dyke intrusions. The way in which the geometry,
kinematics and deformation along a rift zone change in correspondence of a transform fault is something that deserves more
detailed investigations. In order to look into these topics, we
studied, by way of detailed eld surveys, the structural characteristics of the zone of interaction between an active transform fault
and an active rift zone in northern Iceland.
Here, the Theistareykir Fissure Swarm (TFS) (Fig. 1) is a good
example of an emerged oceanic rift affected by Holocene surface
activity as testied to by the presence of both faults and tension
ttir and Brandsdo
ttir, 2011; Hjartardo
ttir et al.,
fractures (Magnsdo
2015). The TFS stretches along a at area, therefore possible disturbances such as topographic effects on the geometry and kinematics of structures can be disregarded. Moreover, the studied area
is essentially composed of a horizontal sequence of basaltic lavas
dated at 14.5 and 2.4 ka BP west of the TFS and at 14.5 ka BP within
the TFS (Saemundsson et al., 2012); hence, also possible effects of

135

lithological heterogeneities at the surface can be disregarded, and


offsets along the different structures can be directly compared. The
TFS is NeS-elongated and most of its structures strike NeS, but
GPS-velocity data from the last decades show vectors (Jouanne
et al., 2006; Metzger et al., 2013) that are oblique with respect to
single faults and tension fractures as well as to the rift as a whole.
Also the horizontal GPS velocities measured all across Iceland
outside the rift zone and plate motion models indicate oblique
vectors relative to the TFS (Sella et al., 2002; Geirsson et al., 2006;

ttir et al., 2009).
Arnad
o
The western part of the TFS is connected to the major, active
transform Husavik-Flatey Fault (HFF) (Fig. 1); its total length
(comprising its offshore and onland sections) is around 100 km
(Metzger et al., 2013), whereas its onland section is 25 km long
(Gudmundsson, 2007; Garcia and Dhont, 2005; Einarsson, 2008).
Although, according to previous research the HFF ends at a junction
with the westernmost normal fault of the TFS, known as the Gudnnugja Fault (GF) (Fig. 2) (Gudmundsson et al., 1993;
 ttir and Brandsdo
ttir, 2011), other authors suggested a
Magnsdo
possible subsurface prolongation of the HFF towards the southeast, on the basis of earthquake and fracture distribution near the
 ttir et al., 2012). The HFF and
Kraa ssure swarm (KFS) (Hjartardo
the KFS have received much attention in the past (Saemundsson,
1974; Young et al., 1985; Bergerat et al., 1990, 2000;
Gudmundsson et al., 1993; Fjader et al., 1994; Angelier et al.,
2000; Garcia et al., 2002; Garcia and Dhont, 2005; Bergerat and
 ttir
 ttir,
Angelier,
2008;
Magnsdo
and
Brandsdo
2011;

Fig. 1. Tectonic setting of northeastern Iceland. The mid-Atlantic Ridge is here offset by the Hsavk-Flatey Fault and the Grmsey and Dalvk Lineaments. Yellow stars indicate
location of M > 6 historical earthquakes with indication of magnitude (M) and year (Stefansson et al., 2008). Orange stripes represent volcano-tectonic rift zones, also termed
 ttir et al., 2015). Faults and ssures (dark grey lines) are from Magnsdo
 ttir and Brandsdo
 ttir (2011) and Hjartardo
ttir et al. (2012; 2015). White
ssure swarms (after Hjartardo
triangles represent the main Quaternary central volcanoes. Box locates Fig. 2. Inset shows location of the area and volcano-tectonic rift zones of Iceland (Einarsson and Smundsson,
 ttir et al., 2015). DAL Dalvik Lineament; GRL Grimsey Lineament; HFF Husavik-Flatey Fault; TFS Theistareykir Fissure Swarm; KFS Kraa Fissure Swarm;
1987; Hjartardo
mar Fissure Swarm; KOR Kolbeinsey Ridge; KR Kraa central volcano (triangle); SB Skjalfandi Bay; FL Flatey island;
AFS Askja Fissure Swarm; FFS Fremrina
TH Theistareykir central volcano; HU Town of Hsavik. (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this
article.).

136

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

 Mariotto et al., 2015). Fault scarps and tension fractures identied by eld work
Fig. 2. Detailed geological map of the study area (based on Saemundsson et al., 2012 and Pasquare
and satellite image interpretation are also provided. Rose diagrams of strike are shown for the 649 fault scarps and the 1208 tension fractures mapped in the present work. The
dashed black lines separate the four sub-areas studied in the present work.

ttir et al., 2012; Maccaferri et al., 2013, and references


Hjartardo
therein) whereas the vertical offsets of faults within the TFS,
located between these structures, has not been studied in detail
until today.
The main goal of our work is to gain a better knowledge of the
structural evolution of an area that is subjected to high seismic and
volcanic hazards; particular emphasis is placed on assessing
whether the HFF is longer than previously supposed. This has clear
implication in terms of the evaluation of the potential rupture
length of the HFF and hence in terms of seismic hazard assessment.
Another major goal is to contribute to the general understanding of
how transform-rift junctions work, which has implications for the
understanding of other cases elsewhere. To accomplish these main
goals, we carried out eld studies to: i) characterize for the rst
time the long-term Holocene kinematics of all tension fractures and
faults along the main portion of the TFS; ii) quantify the offset

components along the eastern HFF and the TFS and iii) evaluate the
evidence of the possible prolongation of the HFF eastward across
the TFS as well as the interaction of these structures at the surface.

2. Geologic-tectonic framework
The Northern Volcanic Zone (NVZ) is active since 8e9 Ma
(Saemundsson, 1974; Jancin et al., 1985; Young et al., 1985; Bergerat
and Angelier, 2008) and is composed of the following ve, about
NeS-striking rift zones: the TFS, the Kraa, Fremrin
amar, Askja, and
ll volcanic systems (Fig. 1) (Hjartardo
 ttir et al., 2015). Each
Kverkfjo
of these systems consists of 5e20 km-wide and 60-100 km-long
fracture swarms and a central volcano (Saemundsson, 1974). The
swarms are made of normal faults, eruptive ssures and tension
fractures that strike parallel to the rift. Most research efforts in this
area were dedicated to shedding light into the KFS (e.g., Angelier

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

 ttir
et al., 1997; Acocella et al., 2000; Dauteuil et al., 2001; Hjartardo
et al., 2012). The TFS is a 10 km-wide elongated area that is cut by
NeS-striking normal faults, eruptive ssures and characterised by
the Theistareykir central volcano (Opheim and Gudmundsson,
1989; Garcia and Dhont, 2005). The most relevant structure in
the TFS is the GF, a Holocene normal fault that represents the
western edge of this rift zone (Fig. 2).
The age of lava ows from the Theistareykjabunga lava shield
was constrained to about 14.5 ka BP in the studied area (Slater et al.,
2001; Stracke et al., 2003). The latest eruption took place about 2.4
ka BP and led to the emplacement of the Theistareykjahraun lava
ows, between the central shield and the HFF (Saemundsson et al.,
2012). Active vertical deformation at Theistareykir was assessed by
means of GPS-based methods (Metzger et al., 2011). The latest
major event affecting the area was the rifting episode that occurred
at the nearby Kraa volcanic system from 1975 to 1984 (Bjornsson,
1985; Buck et al., 2006). During that 9-years episode, several
volcano-tectonic events resulted in major displacement, marked by
horizontal (several meters) and vertical (a few meters) offsets
(Tryggvason, 1980, 1984, 1986). The extension was accompanied by
intense, propagating earthquake swarms with maximum magni ttir and Einarsson,
tude M < 5 (Einarsson, 1986, 1987; Brandsdo
 ttir, 1980; Wright et al., 2012).
1979; Einarsson and Brandsdo
The onland portion of the HFF runs for a length of 25 km
(Gudmundsson, 2007; Garcia and Dhont, 2005) through the
rnes peninsula (Fig. 1), as far as the western border of the HoTjo
chelon, mainly right-stepping,
locene rift zone; it is made of en-e
dextral strike-slip fault segments (Gudmundsson, 1993, 2007).
Near Husavik, the HFF separates Tertiary rocks to the north from
Upper Pleistocene rocks to the south (Saemundsson, 1974; Garcia
et al., 2002). The HFF, rst studied by Einarsson (1958) and mapped by Saemundsson (1974), is marked by a vertical offset at the
surface that amounts to 200 m at many sites (Gudmundsson, 1993).
Its overall vertical displacement may be as much as 1400 m
(Gudmundsson, 1993); this is in line with the 1100 m vertical
displacement off the coast of the Flatey fault documented by Thors
(1982). Saemundsson (1974) proposed a right-lateral displacement
ranging from 5 to 10 km, whereas, according to Young et al. (1985)
it may be as much as 20 km. Garcia and Dhont (2005) broke down
the HFF into three parallel fault lines and, as already observed by
Gudmundsson (1993), identied two major sag ponds along the
fault's trace, occupied by lakes, which may be pull-apart basins
generated by transtensive movements.
As mentioned above, the HFF joins with NeS-striking normal
faults of the TFS (Gudmundsson, 1993); the most meaningful of
these intersections occurs where the HFF joins the GF, making an
about 60 angle (Gudmundsson et al., 1993). Near the intersection,

Gudmundsson et al. (1993), Gudmundsson (2007) and Pasquare
Mariotto et al. (2015) described transpression and transtension
zones. The former are characterized by fragmented lava blocks
which form hills and irregular ridges, while the latter are marked
by pure tension fractures and small collapse structures. Previous
research in this area suggests that faults strike from NW to NNE
with a main strike around NeS, whereas tension fracture strike is
from NNW to NNE, with a main NeS strike (e.g. Saemundsson et al.,
2012).

3. Methodology
An in-depth eld work has been carried out in the area of Fig. 2
at 1:10,000 scale, to produce maps of the eastern HFF and the TFS.
Our study was aimed at studying three types of fractures: i) those
marked by exclusively vertical offsets, referred to as normal faults;

137

ii) those with dominant strike-slip offsets and a subordinate normal


throw, referred to as transtensional faults; and iii) those with
exclusively horizontal opening normal to the fracture strike (no
shear component parallel to the fracture plane), referred to as
tension fractures. In regard to the faults of the TFS, we chose to
focus on measuring vertical offsets because we wanted to assess
whether there might be differences in the amount of vertical throw
in two different areas, respectively located south of the hypothetical prolongation of the HFF into the TFS and north of it. Regarding
tension fractures, our goal was to quantify the motion vectors over a
time-window longer then GPS measurements available in the
literature.

3.1. Faults
Some normal faults are associated with a ssure at the base of
the fault scarp, representing faults with both vertical displacement
and dilation at the surface. At all faults we measured in the eld the
components of horizontal and vertical offsets by means of GPS and
tape. It is necessary to point out that along the HFF the quantication of the strike-slip offset component was possible only at a few
sites due to the lack of diffuse piercing points. A tape was used at
the fractures with offset 10 m, this being the most accurate
method of measurement with an error in the order of few cm. GPS
measurements were taken where vertical offsets are >10 m. All
fault traces were walked along their entire length between the two
tip points, apart from fault F8, which was not walked along its
entire length, outside the study area, due to logistical reasons.
However, we believe this does not affect the validity of our results,
since fault F8 was measured along a sufciently long segment. Due
to the frequent presence of quite continuous vertical scarps, up to
tens of meters high, which make it extremely difcult or impossible
to walk across faults, we employed the following methodology:
measurements were taken every 50e100 m as two persons, each
one carrying a GPS, walked along the rim of the footwall block and
along the surface of the hanging-wall block. At each point of
measurement, GPS instruments were left xed until the value
assumed stability, i.e. with the minimum error. Measurements
were conducted at a sufcient distance (usually in the order of
20e50 m) from the fault scarp to avoid local disturbances such as
block tilting or presence of talus deposits. The differences in GPS
altitude data were collected simultaneously at both sides of the
fault and were processed to show the variations in offset amounts
along the fault length. Errors are in the order of 2 m for GPS
measurements.
We distinguished between faults that can be traced continuously as individual faults, and fault systems that represent
alignments of grouped individual faults. The same distinction was
applied to individual tension fractures and tension fracture
systems. Fault and fracture systems may either contain single
fractures exactly aligned with each other, or they may show lateral
overlapping. We considered as belonging to the same fault or
fracture system the aligned structures with a distance between
their tips smaller than 200 m, or in the case of lateral overlapping,
the structures which are laterally spaced by less than 100 m. It is
worth mentioning that the above thresholds have been used just to
label the main faults. Throughout our eld mapping we were able
to make very detailed distinctions in regard to fracture architecture
and kinematics. Moreover, we assessed that some fractures, previously mapped as normal faults, are in fact tension fractures. We
hence mapped and measured in the eld the faults that compose
the TFS and the HFF (including their junction), determining their
strike, dip, dip-slip and strike-slip displacement components,

138

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

kinematics, block rotation and amount of tilting.


3.2. Tension fractures
In regard to the motion vectors across tension fractures, we
made measurements of the opening direction by means of a compass (corrected for the local magnetic declination) only at the sites
regarded as unaffected by possible local gravity and erosion effects.
Opening directions, in fact, were measured at all places where the
t between fracture rims could be reconstructed in the highest
possible detail and an absolute correspondence between irregularities on the two opposite rock walls could be determined. We
also paid attention to measuring only those tension fractures that
are far away from fault scarps in order to avoid possible local gravity
effects on the opening vectors. At least two measurements of the
line connecting the two opposite piercing points were taken, at
regular intervals along each tension fracture. As erosion and block
falls affected locally the tension fracture edges, and produced a
widening of the ssures, measurements of dilation amounts were
restricted to tension fractures showing a good t between walls.
After selecting the most reliable tension fractures, we measured in
the eld their strike, opening direction, and opening amount.
4. Results
4.1. General pattern of structures
We mapped a total of 649 individual fault scarps and 1208 individual tension fractures (Fig. 2). Fault scarp length distribution is
represented as a histogram with classes of 10 m in Fig. 3A, where it
is possible to notice that the length of fault scarps ranges from 10 m
up to 1340 m; we need to stress that also the shorter fault scarps
were considered as individual ones, because they are, indeed, evidence of fault activity in the area. The most represented classes are
in the range of 40e60 m. The best tting function of the length
distribution is given as a logelog graph in Fig. 3B. The distribution
in this graph ts a power function, with a correlation coefcient
R 0.81. The lowermost classes (below 30 m) are underrepresented due to the following: i) many short fractures (joints
and tension fractures) are hidden by soil and grass, and ii) the
smaller fractures are similar in size to structures of non-tectonic
origin such as cooling fractures in lavas, and hence were not systematically measured.
Tension fracture length distribution is represented as a histogram with classes of 10 m in Fig. 3C, with single tension fractures
ranging from a few meters in length up to 480 m. The most represented class is in the range of 30e40 m. The best tting function
of the logelog graph (Fig. 3D) ts again a power function, with a
correlation coefcient R 0.91.
Fig. 4A shows the azimuth distribution of fault strikes that range
from N310 to N60 , with a clear gap in the range between N60
and N130 . The most represented strike is in the range N00 -20
with a total of 273 faults, and a maximum value at N10 -20 . Fig. 4B
shows the azimuth distribution of tension fracture strikes; they
range from N310 to N70 with a clear gap between N71 and
N129 . The most represented fracture strike is in the range N00 20 , exactly similar to fault strikes, with a total of 343 tension
fractures. The most represented value is N00 -10 , thus slightly
rotated counter-clockwise with respect to the fault strike
maximum value.
The onshore segment of the HFF extends from the coastal town
of Husavik with a NWeSE strike down to the junction with the TFS
(Fig. 1). It is composed of several segments of different length,
striking N115 to N150 (mostly N125 -140 ). These faults are
characterized by different eld evidence and affect rocks with

different age from NW to SE; the north-western and central portions of the HFF affect rocks of pre-Holocene age, mostly attributed
to volcanic and sedimentary sequences of Miocene to Pleistocene
age (pre-Late Glacial Maximum) (Saemundsson, 1974; Garcia et al.,
2002). Fig. 5A shows the distribution of fault lengths vs the number
of individual faults along the central and north-western part of the
onland section of the HFF, suggesting that here the fault zone is
composed of a relatively limited number of individual segments,
each several hundred metres long. The shortest segment is 746 m
long, whereas the longest one is 5531 m long, with most segments
in the 2e4 km range. Most segments are parallel to each other,
locally showing a right-stepping arrangement.
Further to the south-east, the HFF affects younger volcanic rocks
that belong to the series of lava ows erupted from the Theistareykjabunga lava shield, dated about 14.5 ka BP in the studied area
(Slater et al., 2001; Stracke et al., 2003). It is worth noting that our
eld data indicate that the Theistareykjabunga lava shield is not a
single volcano with a main crater area; geological and stratigraphic
data indicate that there are several emission points (Fig. 2), most of
which are 10 m to a few tens of meters higher than the surrounding
lava eld. Most of the emission points are aligned about NeS to
NNE-SSW. Fig. 5B shows the distribution of fault lengths versus the
number of individual faults along this south-eastern part of the
HFF; the graph indicates that the fault zone here is composed of a
much greater number of individual segments and that segments
are generally shorter than in the central-northwestern part of the
HFF. In the south-eastern part of the HFF, the shortest segment is
10 m long, whereas the longest is 353 m long, whereas most segments range between 30 and 150 m. This sector of the HFF is
characterised by the presence of short normal fault swarms striking
N340 to N00 , linked by N125 -135 striking segments. From
aerial photos and previous eld surveys, it appears that the surface
trace of the HFF terminates against the GF, the westernmost fault of
the TFS, making an angle of about 60 (Gudmundsson et al., 1993;
 Mariotto et al., 2015).
Pasquare
The onland TFS has an overall length of 34 km. It is 4- to 6-kmwide in its southern sector, becomes 8-km-wide in the study area
south of the junction with the HFF, and eventually narrows to a
width of about 4 km northward, as far as the coastline. In the
southernmost sector of the TFS, faults and tension fractures strike
N00 to N15 , mostly N10 . In the study area, the structures strike
N350 to N10 up to the junction with the HFF. North of the junction, fractures strike N350 to N10 along the rst 12 km (mostly
N00 ), and further north they strike N05 -20 up to the coast.
Several individual fault scarps compose the architecture of the
TFS (Fig. 2). They are from parallel to sub-parallel to each other, and
spaced from 0.2 to 1 km. Most individual segments are aligned,
with limited overlapping zones ranging from a few meters to a
maximum of 100 m. The width of the overlapping zones is typically
from a few meters to a few tens of meters. The aligned individual
faults give rise to ten main fault zones. Most of these fault zones
gradually fade out and are substituted by tension fractures with the
same orientation as the fault. Other tension fractures are located
away from the faults, giving rise to ssure swarms. Although the
general features of the GF have been described in some previous
 Mariotto et al., 2015), a
works (Gudmundsson et al., 1993; Pasquare
detailed characterisation of the offsets along the TFS has never been
published.
Since one of the main goals of this work is to contribute to
understanding the possible prolongation of the HFF toward the
south-east, across the TFS, here we provide also an analysis of the
geometric characteristics of the various structures in relation with
the possible interaction with the HFF. In Fig. 6 we present the distribution of fault scarp strike and tension fracture strike (X axis) vs
minimum distance from the HFF and its prolongation (Y axis). This

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

139

Fig. 3. (A) and (C) Histograms showing the number of faults and tension fractures, respectively, as a function of their length (in metres), with classes of 10 m. (B) and (D) Same data
in a logelog diagram, showing the linear distributions revealing a power law.

analysis has been done by subdividing the study area into two
zones: the western one is related to the structures located south
and north of the outcropping surface trace of the HFF (Fig. 6A and
C), whereas the eastern zone encompasses the structures located
north and south of the theoretical prolongation of the HFF toward
the south-east, i.e. the zone where the HFF does not crop out
(Fig. 6B and D). In the eastern zone, the distance has been calculated
between each fault and a line that represents the ideal prolongation
of the HFF (e.g. Fig. 2).
The results for faults located in the western zone show that the
most complex pattern is along the HFF, where individual faults
strike from N00 to N50 and from N150 to N180 (Fig. 6A). At
increasing distance from the HFF, from about 0.5 km to 2.5 km,
faults strike N00 -10 and N160 -180 . At distances >2.5 km, faults
mostly strike N00 -40 . In the eastern zone (i.e. where the HFF does
not crop out), faults strike N00 to N22 and N125 to N180 , up to a
distance of about 0.5 km from the theoretical prolongation of the
HFF (Fig. 6B). At longer distances, fault strikes tend to cluster in the
N00 -40 and the N160 -180 range. It can also be noticed that the
N160 -180 range is quite homogeneous from a distance of 0.5 km
from the prolongation of the HFF to the whole distance of 6 km
from it. In other words, at distances greater than 0.5 km from the
prolongation of the HFF, both north and south of it, structures with
the same orientation as the HFF are missing in the TFS.

The results for the distribution of tension fracture strikes with


respect to the distance from the HFF in the western zone, show
once again that the most complex pattern is located along the HFF,
where tension fractures strike N00 -50 and N100 -180 (Fig. 6C).
With increasing distance, at about >0.4 km from the HFF, tension
fractures strike mostly from N00 to N30 and from N160 to
N180 . In the eastern zone (i.e. where the HFF does not crop out),
tension fractures strike N00 to N27 and N115 to N180 , up to a
distance of about 0.3 km from the theoretical prolongation of the
HFF (Fig. 6D). Further away, tension fractures mostly strike N00 40 and N150 -180 .
4.2. Husavik-Flatey Fault
4.2.1. Overall geometry
As already underlined, the HFF is composed of several fault
segments mostly striking N125 -140 . In the central and western
part, these segments are parallel to each other with local zones of
 Mariotto et al.,
overlap and a right-stepping geometry (Pasquare
2015). In the easternmost part, fault segments striking N125 135 are locally linked by fault swarms striking between N340 and
N00 (Fig. 2). The Holocene kinematics is difcult to quantify along
the central and western part of the HFF due to the lack of striated
fault planes. The offsets of landforms produced during the latest

140

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Fig. 4. Graphs of distribution of fault strikes (A) and tension fracture strike (B).

earthquake that struck the area in 1872 (Fig. 1) (Metzger et al., 2011;
Stucchi et al., 2013), have been obscured by erosion. This is mostly
because the central-western part of the HFF corresponds to a high
scarp with active and wide scree tongues, where a dense vegetation
cover is present (Fig. 7A). Nevertheless, eld observations allow us
to reconstruct the kinematics of the HFF here. Along the coast, there
is a natural section across the fault where it is possible to observe a
sequence of Pleistocene volcano-sedimentary, poorly cemented
deposits that lled an asymmetric depression south of the HFF
(Fig. 7B). These deposits, in turn, are faulted with a fault gouge
about 1 m thick and striking NWeSE. Older cemented deposits
cropping out north of the fault make up a block that is structurally
and topographically higher than the southern tectonic block. Along
skuldsvatn
the HFF, there are several depressions, such as the Ho
and Botnsvatn ones, partially lled with lakes, which have already
been interpreted as pull-apart structures (e.g. Fig. 7A)
(Gudmundsson, 1993; Garcia and Dhont, 2005). Since they are
located along overlapping fault segments with a right-stepping
arrangement, this indicates right-lateral kinematics. Fault scarps
are always present along the HFF and face southward, suggesting
the presence of a dip-slip offset component, compatible with
transtensional kinematics.
In the easternmost part of the HFF, offset deposits are volcanics

of Holocene age. They are dissected by NNW-SSE and NWeSE faults


(Fig. 2). The faults striking N340 to N00 show dominant dip-slip
(Fig. 8A) component and connect the NW-SE-striking fault
strands that have a right-stepping geometry. The NW-SE-striking
faults have a dominant right-lateral strike-slip component, as can
be seen, for example, in the picture of Fig. 7C. Several small pullapart depressions are located at right-stepping overlap zones and
push ridges occur at left-stepping overlap zones (Fig. 8B and C). A
dip-slip component is also present along the NW-striking fault
strands, showing a systematic down-sagging of the at lava eld
south of the HFF. The surface trace of the HFF terminates against the
GF through a gradual bending of the fault planes that from an
original NWeSE orientation, attain a NeS strike (Fig. 8D).

4.2.2. Offsets
Along the coastal outcrop of the HFF, at one location (Fig. 7B) we
observed that the Pre-Quaternary substrate crop out only NE of the
HFF, whereas on its SW block the substrate is covered by Quaternary deposits. In view of the about, we reckoned that a minimum
vertical displacement of 8e10 m must have occurred to account for
the observed setting.
The height of the HFF scarp from the Husavik area down the
Botnsvatn depression is from a few tens of meters up to 270 m.

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

141

Fig. 5. Graphs of fault length versus number of faults along the Husavik-Flatey Fault: (A) measurements along the western part where older rocks (pre-Quaternary) crop out, and (B)
measurements along the easternmost segment where late Quaternary rocks crop out.

Further to the south-east, the scarp is 100e140 m high from the


skuldsvatn area to the point where the Holocene lava deposits
Ho
crop out along the HFF trace (Fig. 2). Starting from this point the
scarp height abruptly drops down to a few tens of meters.
More in detail, the south-eastern section of the HFF is characterised by the two swarms of parallel NNW-SSE normal faults and
the NWeSE transcurrent faults (Fig. 2). The total cumulative offset
measured with the tape along the western swarm of NNW-SSE
faults is 22.5 m. These faults are marked by pure dip-slip kinematics, hence this value represents the net slip. The eastern swarm
of NNW-SSE faults has offsets totalling 20 m of net dip-slip.
The NWeSE transcurrent fault segments have a transtensional
right-lateral kinematics. The strike-slip component is difcult to
quantify. The dip-slip component, instead, has been systematically
measured and is shown in Fig. 9A. This offset component ranges
from a minimum of 5 m to a maximum of 17 m. The average value is
10.7 m. There is clear decrease of the cumulated total dip-slip
component from the central-western segment to the eastern
segment.
4.3. Theistareykir rift zone
In the following sections we describe in detail all the main faults
belonging to the TFS. We illustrate: i) the geometry of the fault

scarps; ii) the deformation style; and iii) vertical offsets along the
whole length of the studied structures. The reason for this approach
is to assess possible changes in the amount and style of deformation north and south of the hypothetical prolongation of the HFF.
4.3.1. Gudnnugja Fault
The GF is 5.74 km long and strikes between N10 and N20 . In
correspondence of the junction with the HFF, the GF bends and
attains a N160 strike. The GF scarp faces towards the west. The rst
evidence of the GF to the north is given by NNE-striking tension
fractures that gradually transition to a fault scarp with an initial
offset of 1e3 m along its rst 100 m of length. Further southward,
the offset gradually increases to 31 2 m and then it decreases to
15 2 m near the junction with the HFF (Fig. 9B). Along the stretch
of the GF north of the junction, the average offset value is 10.5 m.
Along this section, which is 1.5 km long, lavas are 14.5 ka old on
both sides of the fault. The lavas are offset along one single fault
scarp, without any rotation of the footwall block, or with very little
rotation (<10 ), considering an original horizontal topography
based on the morphology of the surrounding at area.
South of the junction with the HFF, the offset along the GF
rapidly goes up to 28 2 m and then reaches the maximum value of
33 2 m (Fig. 9B). The average offset is 19.3 m, with most values
between 15 and 22 m. Along this southern stretch, 14.5 ka old lavas

142

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Fig. 6. Upper graphs: distribution of fault scarp strike (X axis) versus distance from the HFF in meters (Y axis) for the zone of the outcropping HFF (A) and for the zone further east
(B). In this eastern zone, the distance has been calculated between each fault and a line that represents the hypothetical prolongation of the HFF towards the SE. Bottom graphs:
distribution of tension fracture strike (X axis) versus distance from the HFF in meters (Y axis) for the zone of the outcropping HFF (C) and for the zone further east corresponding to
its hypothetical prolongation (D).

crop out on the eastern block and 2.4 ka old lavas are observed on
the western block. Near the fault scarp, the hanging-wall block (i.e.
the western tectonic block) is marked by a monoclinal deformation style, represented by lavas dipping in the same direction as
the fault plane (i.e. towards the west) with dip angles in the
40 e65 range.
4.3.2. Fault F2
Fault F2 in Fig. 2 is 2.5 km long and is located 350 m north of the
GF. The fault is composed of two main aligned segments, 173 m
apart from each other. The southern segment extends northward in
the form of a tension fracture that overlaps the northern segment
with a left-stepping geometry. The step-over width (i.e. the
perpendicular distance between the two overlapping fault

segments) is 72e88 m. The scarps of the two segments face to the


west and there is no tilting of the hanging-wall block. From the
north, the rst main fault segment strikes N00 and the offset increases to 7 m, then decreases to 5 m toward the southern tip zone
(Fig. 9C). The offset of the second segment that strikes N350 to
N00 , rapidly increases to 6 m and then up to 8e9 m, with a marked
decrease toward the southern tip. The average offset of the entire F2
is 5 m.
4.3.3. Midveggur Fault (F3eF4)
The Midveggur Fault is composed of two segment: the northern
one (F3) departs a few tens of meters east of the northern section of
the GF (Fig. 2) and runs parallel to the GF with a NeS strike. The
second segment (F4) departs very close to the previous fault and

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

143

Fig. 7. Photos along the Husavik-Flatey Fault (HFF). (A) Western onshore termination of the fault zone running across the town of Husavik. A fault scarp up to 270 m high is evident.
(B) Outcrop of the HFF along the coast immediately north of Husavik; the two arrows indicate the fault zone. (C) Right-lateral offset of a lava tube along the easternmost part of the
HFF; the axis of the lava tube is coincident with the dashed line. (Photos by A. Tibaldi).

runs southward with a NeS strike. It is composed of seven closely


spaced and mostly aligned individual fault strands.
The total length of segment F3 is 4.71 km. The fault scarp faces to
the west and there is no tilting of the hanging wall block. The
offset along F3, north of the point of intersection with the theoretical prolongation of the HFF, is up to 4e5 m (Fig. 9D). Southward
of this point, the offset increases to 7 m, and is frequently around
5 m. The average offset north of the point of intersection with the
possible prolongation with the HFF is 2.3 m, whereas the average
offset is 3.8 m south of it.
F4 is 4.13 km long and it extends a little further south than the
area we studied in detail. The scarp faces to the west and there is a
local tilting of the hanging wall block with values < 10 . The offsets
along F4 are variable, reecting the segmentation of this fault
(Fig. 9E). From the northern tip, the offset gradually increases to
5 m, then decreases and increases again up to 18 2 m. Further
south the offset decreases to 5 m and then goes up to a range of
15e25 2 m, until it decreases toward the southern tip. The
average offset is 14.7 m.

4.3.4. Faults F6eF7eF9


Faults F6, F7 and F9 strike dominantly N00 with a few bends
striking N350 and N10 . Their scarps face to the west and there is
no tilting of the hanging wall blocks. The most prominent bend is
located at the southern termination of F5 where the fault attains a
NWeSE strike parallel to the HFF (Fig. 2).
Fault F6 is 4 km long. It has offset values ranging mostly 2e6 m
with a peak at 10 m north of the possible junction with the prolongation of the HFF (Fig. 9G). From this junction, the offset increases up to 8 m and decreases to 6 m, then rising to 10 m and

nally gradually reaching zero offset at the southern tip. The


average offset is 2.7 m north of the junction and 4.5 m south of it.
Fault F9 is 2 km long (Fig. 2). From the north, F9 has offset values
abruptly reaching 5 m then decreasing more gradually to 1 m
(Fig. 9L). The average offset is 1.3 m.
Fault F7 is 3 km long (Fig. 2). From the north, F7 has offset values
abruptly reaching 10 m, then decreasing more gradually to 2 m, and
increasing again up to 8 m and then to 10 m (Fig. 9H). The average
offset is 1.3 m.
4.3.5. Togarahellir Fault (F5) and Klifveggur Fault (F8)
Togarahellir (F5) and Klifveggur (F8) faults are located south of
the theoretical prolongation of the HFF (Fig. 2). They are quite
rectilinear in plan view, parallel to each other with a N20 strike.
Their scarps face to the west and there is no tilting of their hanging
wall block.
Togarahellir Fault (F5) is 2 km long and has variable offset values
mostly between 4 and 9 m, with two peaks at 11 2 m in its central
part (Fig. 9F). The average offset is 6.9 m.
The studied part of the Klifveggur Fault (F8) is 2 km long; this
structure extends further south for another 1.87 km, but for logistical reasons it has not been possible to take measurements as far as
its southern tip. It has offset values ranging from 0.5 to 5 m,
gradually increasing southwards (Fig. 9I). The average offset is
2.6 m.
4.3.6. Fault F10
Fault F10 is 3 km long and strikes N355 -05 . The scarp faces to
the west and locally there is a slight tilting of the hanging wall
block. The fault trace is highly segmented but it shows very relevant

144

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Fig. 8. Photo along the eastern part of the Husavik-Flatey Fault (HFF). (A) Normal faults striking about NNW-SSE that link the right-stepping NW-SE-striking segments of the HFF.
(B) Example of a push ridge at left-stepping and a pull-apart depression at right-stepping zones of the HFF; as seen in the box, this arrangement is compatible with right-lateral
strike-slip kinematics. (C) A wide push ridge along the HFF; this structure is bounded by a reverse fault on the side where the person is standing. (D) Structures at the junction
between the HFF and the Gudnnugja Fault. Locations in Fig. 2. (Photos by A. Tibaldi).

offsets (Fig. 9M); from north to south offset values gradually increase up to 22.5 2 m and then decrease to 14 2, then going up
to 22 2 m and 22.5 2 m. The average offset is 11 m. The
observation of such a large offset, especially in comparison with the
limited fault length, suggests that this structure might be a prolongation of fault F8, as will be discussed later on.
4.4. Tension fractures
4.4.1. Strike
In Fig. 10 and related rose diagrams, we selected the 239 tension
fractures showing the most reliable evidence of their opening directions. The fractures mostly strike N00 -10 and N350 -360 in
decreasing order of frequency. Very subordinate sets of tension
fractures strike N330 -350 and N10 -30 . In the rift zone sub-area,
the tension fractures predominantly strike in a northerly direction,
i.e. between N350 and N10 . In the SW block, the tension fractures
strike mostly in the range N00 -10 , and very subordinately N340 350 and N10 -20 . North of the HFF, tension fractures strike
mostly N350 -10 , and subordinately N320 -330 . The major
change in the orientation of tension fractures occurs along the HFF
zone (Fig. 10). Here there is a higher dispersion of tension fracture
strikes than in the other studied zones. Most tension fractures have
a N70 -110 orientation, with a peak between N90 and N100 .
Very subordinate tension fractures strike in the range N30 -60 .
For the whole studied area, based on the results of our eld
structural surveys, we plotted tension fracture strike vs longitude
(easting) (Fig. 11A). On this graph it is possible to observe that the
tension fractures striking N00 -10 and N170 -180 are homogeneously widespread from west to east, and hence represent the

typical regional geometry of ssuring. As opposed to the above, the


tension fractures striking N10 -30 and those in the range N120 170 occur only in the central part of the area.
In Fig. 11B, we plotted tension fracture strike vs latitude
(northing) for the whole studied area. It is possible to observe that
in the upper part of the graph, the tension fractures show the
highest dispersion with values of strike ranging from N00 to N60
(mostly N00 -N30 ) and from N100 -180 (mostly N140 -180 ).
This area represents latitude values corresponding to the HFF zone.
Going further north, i.e. in the uppermost sector of the graph,
tension fracture strikes tend to cluster in the ranges N00 -25 and
N140 -180 . Especially south of the latitude corresponding to the
HFF, tension fracture strikes are much less dispersed and correspond to the ranges N00 -20 and N170 -180 .
4.4.2. Fracture opening directions
In regard to the opening direction of tension fractures (Fig. 10
and related rose diagrams), the most represented values are in
the range N90 -100 . Subordinately, the opening trends are the
N100 -110 , N80 -90 , N70 -80 , and N60 -70 , in decreasing order of frequency. By comparing the dominant opening directions
with the dominant tension fractures strike, the overall rose diagram shows that most opening occurs perpendicularly to tension
fracture strikes, but there is also an asymmetric distribution with
several values rotated clockwise with respect to the trend
perpendicular to fracture strikes. In other words, the two relative
distributions suggest that there is also a minor (10 on average)
right-lateral transtensional component along part of the NeS tension fractures. This right-lateral component of opening can be
recognized also by analysing the single sub-areas. Both in the SW

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

145

Fig. 9. (A) Graph of the vertical component of offset along the HFF measured in the eld. Offset values are provided from west to east. Graphs of offsets measured in the eld within
the TFS: GF (B), Fault F2 (C), Fault F3 (D), Fault F4 (E), Togarahellir Fault (F5) (F), Fault F6 (G), Fault F7 (H), Klifveggur Fault (F8) (I), Fault F9 (L), F10 (M).

block, along the Rift Zone, and north of the HFF, the dominant
opening directions are always rotated clockwise of about 10 with
respect to the dominant tension fracture strikes. Along the HFF,
opening directions are more widespread and complicated since

they reect the larger dispersion in fracture strikes.


For the whole studied area, we plotted also the opening directions of tension fractures vs easting (Fig. 11C). In the left part of
this graph (i.e. in the western part of the studied area), the opening

146

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Fig. 10. Map of the study area with tension fracture opening directions measured in the eld. Rose diagrams are provided for the 239 measurements. Rose diagrams show fracture
strikes (green) and opening directions (blue) in the four sub-areas (For interpretation of the references to colour in this gure legend, the reader is referred to the web version of this
article.).

directions of the westernmost tension fractures are in the range


N110 -140 . More to the east they are N105 -130 , and then N80 110 . This indicates a gradual anticlockwise rotation of opening
directions from west to east. Further east, in the central part, there
is a wider scattering of the data, reecting once again the values
measured also along the HFF. It is important to note that also in this
area, most values are in the range N80 -120 . In the right part of the
graph, the opening directions are much less dispersed and are
concentrated in the range N80 -100 , with some values ranging
N70 -125 . If we compare the data in the left hand part of the graph
with those in its right hand part, it is possible to highlight, also at a
more regional scale, the anticlockwise rotation of opening directions from west to east. From a statistical point of view, in the
whole studied area there is a predominance of opening directions
in the range N100 -110 and a peak at N112 .
In Fig. 11D we provided the opening directions of tension fractures with respect to latitude (northing). The opening directions in

the lower part of the graph represent the deformation far away
from the HFF. In the southern part of the studied area, opening
directions are in the range N65 -125 . In the central part of the
graph, corresponding to the HFF zone, opening directions are much
more dispersed, although most values are in the range N80 -110 .
In the uppermost part of the graph, opening directions tend to
cluster again in the range N65 -120 .
In order to better understand the possible variations in the geometry of the various structures with distance from the HFF, in
Fig. 12A we plotted the distribution of tension fracture strikes vs
distance from the HFF and its prolongation. It is possible to observe
that with distance from the HFF, the range of fracture strikes becomes smaller, with values mostly ranging N00 -20 and N170 180 at greater distances. At distances between 0.4 and 2.5 km,
tension fractures strike N00 -30 and N155 -180 . At distances
<0.4 km from the HFF, strikes have the largest dispersion ranging
N00 -50 and N115 -180 . This scattering can be observed also

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

147

Fig. 11. (A) Distribution of tension fracture strike versus easting. (B) Distribution of tension fracture strike versus northing. (C) Distribution of tension fracture opening directions
versus easting. (D) Distribution of tension fracture opening directions versus northing.

along the rift zone in correspondence of the theoretical prolongation of the HFF.
In Fig. 12B we plotted the distribution of opening directions of
tension fractures vs distance from the HFF and its prolongation. On
this graph it is possible to observe very clearly that the opening
directions tend to focus in the range N80 -120 with distance from
the HFF. At distances >2.5 km, opening values are N85 -120 . At
distances between 0.4 and 2.4 km, values are N65 -125 . At distances <0.4 km from the HFF, the opening trends have a large
dispersion ranging N25 -150 . A wide scattering, in the range
N25 -125 , can be observed also along the rift zone in correspondence of the theoretical prolongation of the HFF.
5. Discussion
5.1. General distribution and size of faults and tension fractures
Our analysis of the general fracture patterns along the HusavikFlatey Fault (HFF) zone and the Icelandic rift segment of the
Theistareykir Fissure Swarm (TFS) reveals a size-distribution of
fractures that ts well a power law (Fig. 3). A power law distribution has been recognized in different tectonic settings
(Gudmundsson, 1987a,b; Main et al., 1990; Scholz and Cowie, 1990;
Scholz et al., 1993; Davy, 1993; Hardacre and Cowie, 2003; Sonnette
et al., 2010) with faults covering a large range of dimensions. In a
very similar tectonic setting, represented by the Thingvellir Fissure
Swarm (Iceland), Sonnette et al. (2010) found that the fracture
length mostly goes from 10 m to about 300 m, with a total range of
values going from 3 m to a maximum of 1780 m. In our work we
distinguished between tension fractures and faults. Although both
show a power law distribution, the faults in our study area have
lengths with size distribution very similar to the one of the
Thingvellir Fissure Swarm, whereas the tension fractures resulted
to be shorter both in terms of the general distribution and the

maximum values. We believe that this difference is due to the fact


that faults tend to develop by increasing both offset and length, the
latter occurring by linkage and nucleation (Cladouhos and Marret,
1996; Manseld and Cartwright, 2001). In the study area, ssures
may evolve by transformation into faults, as suggested by some
tension fractures we found at the tips of normal faults. Therefore,
tension fractures here are shorter because they represent earlier
stages of evolution. The fractures of our area are also shorter than
those measured in the Reykjanes Peninsula in southwest Iceland,
which mostly have lengths from tens of meters to 3000 m (Clifton
and Kattenhorn, 2006). Also this observation testies to a less
developed fracture eld in our area.
The distribution of fracture azimuths seems quite dispersed
(Fig. 4), but in fact it is consistent with the presence of the two
interacting systems of the HFF and the TFS. The HFF is dominated by
NWeSE structures in its central and western onshore sectors,
whereas the eastern sector is dominated by NWeSE, NNWeSSE
and NeS structures. The sectors of the HFF are also characterised by
a difference in fracture length distribution: the central and western
sectors have a relatively smaller number of faults with average
greater lengths, whereas the eastern sector has more fault segments, which are shorter. We believe that both the different strike
and length distribution can be explained in terms of the different
inherited geological histories of the sectors. The central-western
sectors are characterised by a minor presence of NeS-striking old
normal faults both north and south of the HFF, as shown by the
strong decrease of these structures toward the west (e.g.
Saemundsson et al., 2012). More in detail, the analysis of the
stratigraphic succession north of the HFF reveals that basaltic lava
rocks of the same Miocene age are present both north of the
western sector of the HFF and north of the eastern sector. Moreover,
widespread lavas of Pleistocene age are also present north of the
central and eastern sectors of the HFF. Thus, by comparing rocks of
the same age, it is possible to observe that several old normal faults

148

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

Fig. 12. (A) Distribution of tension fracture strike (Y axis) and (B) of tension fracture opening direction (Y axis) versus distance from the HFF in meters (X axis). The different symbols
represent the different blocks introduced in Fig. 2.

are present essentially north of the eastern part of the HFF. We


interpret the presence also of the NNW-SSE and NeS structures
along the eastern HFF as the effect of the structural inheritance of
pre-existing NeS weakness zones. As already put forward by Mann
(2007) and Mann et al. (2007), the interaction between strike-slip
fault systems and inherited structures with a different orientation
in the basement, can induce the formation of strike-slip fault zones
with a complex geometry, characterised by the development of
several fault steps with releasing and restraining bends. Similar
conclusions have been reached also by way of analogue experiments (Autin et al., 2013).
As regards the different size distribution of fracture lengths
along the HFF, we stress that the eastern sector of the HFF affects an
area with younger lava ows of Holocene age. Part of these lavas,
such as the Skildingahraun unit, dated at about 14.5 ka BP by
Saemundsson et al. (2012), buried the previous HFF trace by owing northward. As a consequence, during the Holocene, the HFF was

forced to propagate upward through these lavas to reach the surface. As already outlined in several papers, a transcurrent fault that
has an about continuous trace in the basement may propagate
upward across the surface strata through shorter fault segments.
Only at a more advanced stage of rock rupture, the earlier fault
segments may link together giving rise to longer segments (e.g.
Martel et al., 1988; An and Sammis, 1996; Kim et al., 2003; Rovida
and Tibaldi, 2005), as already occurred along the central and
western part of the HFF, where older rocks crop out. This is also
consistent with the fact that our fracture lengths are generally
smaller than in other zones of Iceland, as underlined above.
In regard to the TFS, the general fault pattern is clearly dominated by structures striking N00 to N20 . However, a few important differences can be noted: the southern part of the studied rift
zone is dominated by faults striking in the range N10 -20 , whereas
in the northern part of the rift, faults striking N00 to N10 dominate. Moreover, our detailed eld surveys enabled us to recognize

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

149

that some structures previously mapped as faults, are in fact tension fractures, especially in the northern portion of the studied
area. As a consequence, the southern rift zone as dened by the
presence of normal faults, is mostly 6e8 km wide, whereas the
northern part is up to 4 km wide at most. We believe that such
differences should be interpreted, as will be made clear in the next
sections.

5.2. Fault offsets along the Husavik-Flatey Fault


Thors (1982) estimated a throw of 1100 m for the HFF offshore.
Onshore, Gudmundsson (1993) estimated a maximum throw of
1400 m. Saemundsson (1974) suggested a right-lateral component
of displacement of as much as 5e10 km, whereas Young et al.
(1985) suggested an amount of 20 km. If we consider a strike-slip
component of 15 5 km and a vertical component of 1400 m, we
obtain a gross estimation of their ratio 11 3.
The data we collected in the eld were mostly aimed at
measuring in high detail the vertical component of offset, due to
the difculties in individuating piercing points that may show the
transcurrent component. Only at a few sites it has been possible to
obtain complete measurements of the two components, such as in
the case shown in Fig. 7C.
The surface vertical offset along the central-western part of the
onshore HFF reaches 200 m at many sites (Gudmundsson, 1993).
We also obtained in the eld values from 10 m up to 270 m. These
values may reect particularly large vertical components of slip due
to the presence of some pull-apart basins, as those occupied by the
skuldsvatn lakes (Gudmundsson, 1993; Garcia
Botnsvatn and Ho
and Dhont, 2005) or other releasing bend zones. Anyway, we
wish to stress that all along the western, central and eastern part of
the HFF, a vertical component of motions is always present. We
therefore consider the onland part of the HFF as a transtensional
right-lateral fault, along which a dip-slip component is always
present and is not limited to releasing bends and pull-apart
depressions.
All these data suggest that the HFF has a very long history and
the high scarps in the zones where older rocks crop out, result from
cumulated incremental offsets. Also in the studied area, along the
eastern segment of the HFF, there is a 100-m-high scarp affecting
slightly tilted lavas of Pleistocene age, whereas further south-east
the scarp's height suddenly drops to a few tens of meters. To
summarize, along the easternmost trace of the HFF, the low values
of offset occur where the HFF cuts through Holocene lavas (dated at
14.5 to 2.4 ka BP), with displacements in the order of 17 m at
maximum. We suggest that also here the HFF should have a higher
scarp, which has been covered by the late PleistoceneeHolocene
lavas. The uppermost lavas of the 14.5 ka succession denitively
owed northward across the fault trace, whereas the lower parts of
this succession might have been deposited in onlap against the
previous fault scarp. Another fault scarp formed after the deposition of the 14.5 lavas, which in turn stopped or deviated the successive 2.4 ka old lavas. Field observations, in fact, indicate that the
most recent lava locally owed parallel to the fault scarp.
The dip-slip component systematically measured along the NWSE-striking fault segments of the eastern part of the HFF, indicates
values of 5e17 m. This is in quite good agreement with the dip-slip
values of 5e22.5 m measured along the swarms of NNW-SSE to
NeS faults that link the NWeSE segments. The greater values along
the NNW-SSE-striking faults are due to the different orientation
(Fig. 13): these faults link the NWeSE faults along zones of overstep
with a right-stepping geometry and thus the dip-slip motions along
the NNW-SSE faults are enhanced.

Fig. 13. Three-dimensional sketch of the variation in the dip-slip component of motions along the eastern Husavik-Flatey Fault in relation to the changes in strike
orientation.

5.3. Faulting along the Theistareykir Fissure Swarm


5.3.1. Deformation at the Gudnnugja Fault
As shown in Fig. 9B, the greatest offsets of the TFS are present
along the part of the GF located south of the junction with the HFF.
The southern segment of the GF has displacement values mostly in
the range 15e22 m, with a maximum of 33 2 m and an average
offset of 19.3 m. Along this southern segment, 14.5 ka old lavas crop
out on the eastern block and 2.4 ka old lavas occur on the western
block. We thus argue that the measured offsets may be regarded as
minimum because the original scarp cut into the 14.5-ka-old lavas
has been partially covered in onlap by the most recent lava ows.
Along the section north of the junction, most offset values are between 12 and 20 m, the maximum value is 31 2 m, and the
average value is 10.5 m. Along this section, the outcropping lavas
are 14.5 ka old on both sides of the fault, thus fully reecting the
post-14.5 ka displacement. These data indicate that the GF has
experienced different deformation amounts in the last 14.5 ka, both
for maximum and average offsets. Displacements are much higher
south of the junction with the HFF, especially taking into consideration that the scarp height of this southern part of the GF has
been decreased by the onlap of more recent lavas.
The GF shows also another difference at the junction with the
HFF: north of it, the hanging-wall block of the GF (i.e. the western
block) shows rotation along a horizontal axis in the order of a few
degrees, whereas south of the junction, the hanging-wall block is
characterised by a monoclinal geometry with lavas dipping towards
the west with dip angles up to 65 . The monoclinal geometry has
been explained in the literature by means of two main models. One
model considers that monoclines at the hanging-wall block of
normal faults may develop as tilting along the fault in a way similar
to drag folding, as suggested for the Thingvellir Fissure Swarm by
Sonnette et al. (2010). In a similar way, Grant and Kattenhorn
(2004) suggested that narrow monoclinal fold forms when there
is a vertical fracture attached to and above a normal fault with
upward fault propagation.
More recently, Trippanera et al. (2015) used analogue models to
assess the upper crustal deformation induced by dykes and
compared their ndings with eld observations along major rift
zones, including Icelandic ones. For example, south of the Kraa
caldera they documented, at a 2-km-long fault, the tilting of the

150

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

hanging wall block associated with contractional structures at its


base. They explained this deformational pattern by means of an
experimental conguration simulating the development of two
arcuate faults (delimitating the tilted hanging wall block) that
propagate upward from a medium-depth, rectangular intrusive
complex.
We need to stress that classical development of monoclinal folds
along normal faults is referred to sedimentary successions where
the rst stage is given by exural bending above the propagating
fault (Jackson et al., 2006). This is followed by interlayer slip,
beginning of upward propagation of the main fault, and development of secondary faults in the upper sector of the footwall block. A
further stage is given by the complete faulting across the succession
up to the surface with nal rotation of the blocks composing the
monocline limb. In our case we believe that the rigidity of lavas at
the shallow level prevents any folding and interlayer slip; the
monocline develops by initial rupture and cracking along the
developing hinge lines where stresses concentrate. Cracking is
accompanied by rigid rotation of blocks.
The second model postulates that the monocline results from
lavas owing from the footwall block towards the hanging-wall
block on the slope created by the fault (Holland et al., 2006;
Sonnette et al., 2010). We believe this model cannot be applied to
our study area because north of the Junction HFF-GF, the same
horizontal lavas occur on the footwall and hanging-wall blocks,
whereas south of the junction, the 2.4-ka-old lavas did not ow
above the fault scarp.
Therefore, we consider that the presence of the monocline south
of the junction HFF-GF can be explained in terms of the different
geological history. South of the junction, in fact, the GF scarp was
partially onlapped by the 2.4-ka old lava ows. As the fault moved
repeatedly, it propagated upwards while breaking through the new
lava ows. The movement of the normal fault was accommodated
rst by exure of the overlying lavas allowed by the formation of
fractures in the developing hinge zones. This was followed by
fracture growth by linkage so that the fault was able to cut through
the entire uppermost layers.
5.3.2. Deformation at the other normal faults
All the other faults studied in the TFS show scarps facing towards the west, like the GF, and do not show any presence of
monoclines. Only locally the lavas of the hanging-wall blocks are
very slightly tilted in the direction opposite to the fault dip. Since
these faults affect the same lavas dated at 14.5 ka BP, which have
the same horizontal attitude, the offsets and geometry of the faults
can be directly compared with each other.
The average offset value and the maximum offset value at each
fault do not show any gradual variation from the west side of the
rift to the east side; faults with the greater offset values, such as GF,
F4 and F10, alternate with faults with lower offset values. Regarding
in particular F10, this fault shows offset values that are too large
with respect to its total fault length (3 km), whereas similar offsets
measured at GF and F4 are more compatible with their total length
(5.74 km and 4.13 km respectively). If we take into consideration
that F8 prolongs southward outside the graph of Fig. 9I with offset
chelon geometry of F8 and
values in the order of 5e10 m, the en-e
F10 and the very high offsets measured at F10 suggest that these
two faults, in reality, are strands of the same structure. This interpretation is consistent with the values of step-over width of overlapping segments belonging to the same fault zone, as seen in other
zones of Iceland and elsewhere (Huggins et al., 1995; Willemse,
 Mariotto, 2015).
1997; Acocella et al., 2000; Tibaldi and Pasquare
It is interesting to observe the different degree of segmentation
of faults F3 and F4 with respect to their range of offset values
(compare Fig. 9DeE with Fig. 14). F3 has offset values mostly in the

Fig. 14. Example of two faults of the TFS that show different degrees of segmentation.
Faults F3 and F4 are shown in Fig. 2.

range of 3e5 m, a maximum value of 7 m, and an average offset of


3.2 m F4 has offset values mostly of 10e20 2 m, with peaks of
25 2 m, and an average offset of 14.7 m. The degree of segmentation is higher at F3 that is composed of 82 individual segments for
a total length of 5310 m, resulting in an average length of 65 m per
individual fault segment. F4 is composed of 40 individual segments
for a total length of 4124 m, resulting in an average length of 103 m
per individual fault segment. This example shows the proportional
link between fault offset and length of individual fault segments
and suggests that the faults of the TFS have different ages, although
they affect the same recent Holocene rocks. Several examples
elsewhere suggest that at the initial stage of formation of a normal
fault at the surface, the individual fracture segments are characterised by an underlapping geometry (Fig. 14a) (Acocella et al.,
2000). Successively, one or both tips propagate until they pass

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

each other and the two fracture segments become overlapping


(Fig. 14b) (Pollard and Aydin, 1984; Willemse, 1997; Peacock, 2002).
The stress conguration at overlapping tips produces a bending of
the propagating fractures that become hook-shaped on approaching each other, and nally can become linked (Fig. 14c) (Olson and
Pollard, 1989; Gudmundsson et al., 1993; Renshaw and Pollard,
1994). This results in the formation of more continuous and
sinuous structures, characterized by longer individual segments. At
the linkage stage, the offset becomes more homogenous along the
newly formed continuous fault and the offset decrease of the
overlapping zones tends to disappear. The average offset tends to
increase and also the maximum offset may increase in the central
part of the fault.
Based on the above-described evolutionary stages of segmented
faults, and on the fact that the GF has the largest offsets among the
TFS faults and an almost continuous trace, we can suggest that the
GF has the longer history of incremental motions. The other faults
located to the east are all characterised by high segmentation and
lower average and maximum offsets, and thus can be considered to
be less developed than the GF. The lower development can be due
to a younger age than the GF, or to a lower rate of fault activity. A
gradual trend towards the east cannot be recognized since, for
example, fault F7 shows characteristics suggesting that it is less
developed than fault F8 that is located further east.
5.4. Possible prolongation of the Husavik-Flatey Fault
The eld evidence based on surface fault rupture trace and
tension fractures, suggests that the HFF ends against the GF
 Mariotto et al., 2015). Very
(Gudmundsson et al., 1993; Pasquare
detailed eld investigations carried out across the possible prolongation of the HFF towards the south-east did not reveal any
outcropping fracture with the same NWeSE orientation. However,
establishing the real position and length of the HFF is of paramount
importance, since this fault is active and triggered several large
earthquakes in the past. While most earthquakes along the Icelandic rift zone have M < 4 (Einarsson and Bjornsson, 1979;
Einarsson, 1986, 2015; Stef
ansson et al., 2008), earthquakes with
much larger magnitudes occurred along transcurrent faults in Iceland. Four major earthquakes took place along the HFF in the past
200 yrs, all of which offshore (Fig. 1). In 1755, an M 7 earthquake
struck the Skjalfandi Bay and, in 1838, an M 6.5 event occurred near
the westernmost end of the HFF (Metzger et al., 2011). In 1872, two
M 6.5 earthquakes occurred, with epicentres near the island of
Flatey and the town of Husavk. Metzger et al. (2011), assuming that
seismic energy has been accumulating since these two latest
earthquakes, suggested that the present seismic potential of the
fault is equivalent to an Mw 6.8 0.1 event. Metzger et al. (2013),
based on GPS data and back-slip modelling, suggested that the HFF
nsson (2014) sugends at the TFS, and after that Metzger and Jo
gested an even larger moment magnitude (up to Mw 7.0) for a
potential earthquake, considering also an instantaneous rupture of
the entire HFF. Hence, a correct evaluation of the possible
maximum length of the HFF is very important.
ttir et al. (2012) suggested the possible continuation of
Hjartardo
the HFF east of the junction with the GF in the form of a buried fault.
They consider that the HFF can continue as far as the Kraa Fissure
System (KFS), based on three facts: i) an earthquake migration from
the KFS toward the HFF was observed during the Kraa rifting
episode; ii) the central graben of the KFS widens abruptly at the
HFF-KFS intersection, and iii) the maximum fracture density in the
KFS is accordingly found in this area.
The work we did along the HFF and in the TFS allows us to
provide a series of clues into this important issue. First of all, as
regards fault geometry in plan view, the GF strikes N10 -15 south

151

of the GF-HFF junction, and then bends to a N340 strike north of


the junction, and eventually, further north, it attains again a N10 15 orientation. Such bending exactly in correspondence of the HFF
cannot be a simple coincidence, also because the HFF attains an enchelon arrangement near the junction with individual segments
e
striking N330 -350 . If we look at the surface traces of the major
faults we can observe that also F8 and F5 bend and attain a local
N340 strike near the prolongation of the HFF. More in detail, the
plot of the strikes of individual fault segments with respect to the
distance from the outcropping HFF, reveals that a more complex
range of values is present at distances smaller than 500 m (Fig. 6A).
This complexity corresponds to the presence of R (Riedel) and R
 Mariotto et al.,
shears, and T fractures along the HFF (Pasquare
2015). By comparison with the plot of the strikes of individual
fault segments with respect to the distance from the prolongation
of the HFF (Fig. 6B), exactly the same pattern emerges, suggesting
the presence of a zone of complexity along this prolongation. The
identical situation is observed with the tension fractures, which
have more dispersed values of strike in correspondence and nearby
the outcropping HFF (Fig. 6C), at a maximum distance of 400 m, as
well as near the prolongation of the HFF (Fig. 6D).
Fig. 11D shows that also the tension fracture opening directions
become more dispersed in the central-northern part of the studied
area, i.e. in correspondence of the HFF zone. This can be appreciated
also on the map of the sites of measurement of opening directions
(Fig. 10): all along the HFF they trend ENE-WSW, whereas in the
remaining sectors most opening vectors trend around EeW to
WNW-ESE. If we look at the variations of opening directions with
longitude (Fig. 11C), it emerges that diverse orientations are
distributed across the whole studied area. The same point can be
made for the strike variation of tension fractures with latitude and
longitude (Fig. 11B and A). All these data mean that the complexity
in the orientations and opening directions of the tension fractures
is present both along the zone of outcropping of the HFF and along
its prolongation towards south-east. The same conclusion can be
drawn by looking at the graphs of variation of tension fracture
strike and opening direction with distance from the HFF (Fig. 12). In
particular, the fracture strike and opening direction of the tension
fractures measured along the TFS show an increase of dispersion at
distances <400 m from the prolongation of the HFF.
We argue that the change in the strike of the GF at the intersection with the HFF, points to the important role of the HFF that
interacts with the GF locally controlling its geometry. The HFF, that
shows complex geometries typical of strike-slip structures, induces
a re-orientation to the GF that becomes sub-parallel to the HFF near
the intersection. We discovered that identical variations in the
strike of the other faults of the TFS do occur along a strip corresponding to the possible prolongation of the HFF toward the southeast: in Fig. 15A we plotted only the segments of faults and fractures
of the TFS with a different orientation with respect to the dominant
NeS to NNE-SSW strike. It clearly emerges a strip of fracture segments that strike exactly parallel to the easternmost strands of the
HFF (N330 -340 ), where it bends towards the GF. These N330 340 -striking segments have an orientation with respect to the
prolongation of the HFF, which is consistent with the fractures that
develop during the earlier stages of formation of a transtensional
fault, as seen in analogue experiments by sand box in Agostini et al.
(2009) (Fig. 15B). In other analogue experiments with clay box, as
those by Clifton et al. (2000), the pattern of fracturing under
transtensional movements is similar although more complex and
with more fractures since these experiments represent a more
mature stage of deformation (Fig. 15C). The pattern of the HFF is
also strikingly similar to the fracture geometry at Billings, Montana
(Harding et al., 1985) (Fig. 15D), although it is specular, because the
Billings example is in a left-lateral transtensional zone. Also the

152

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

chelon fault segments of the TFS with a NNW-SSE


Fig. 15. (A) Zone of prolongation of the Husavik Flatey Fault (HFF) towards south-east as suggested by the presence of en-e
orientation. Note that this strip lies exactly on the prosecution of the NWeSE strand of the HFF that can be seen in the upper-left corner of the map. Diverging large arrows
indicate the average direction of extension measured in the eld. (B) Analogue experiment by sand box with development of fractures in the earlier stage of right-lateral transtensional movement along a NWeSE fault zone (modied from Agostini et al., 2009); note the similarity with our proposed prolongation of the HFF. (C) Analogue experiment by
means of a clay box with development of fractures under right-lateral transtensional movement along a NWeSE fault zone (modied from Clifton et al., 2000). (D) Pattern of
fractures at Billings, Montana (after Harding et al., 1985): they are similar but specular to the HFF because the Billings example is in a left-lateral transtensional zone.

kinematics of the faults of this strip representing the prolongation


of the HFF is consistent with the analogue examples and with the
Billings eld example, since all faults are characterised by dominant
normal dip-slip motions. We also stress that the opening vectors of
the tensions fractures of the TFS have a slight right-lateral
component that is also consistent with our interpretation.
Moreover, we highlight the northward decrease in the width of
the TFS (from 8 to 4 km) and the change in the displacement eld
north and south of the HFF; the GF shows a strong decrease of the
average and maximum offset north of the junction with the GF.
More to the east, the successive faults F3 and F6 of the TFS, also
show a clear decrease in the average offset and in the maximum
values of offset north of the HFF prolongation (Fig. 9D and G).
Further east, fault F10 transforms into tension fractures northward.
To better quantify these variations, in Fig. 16 we plotted the total
offset measured on all the fault segments of the TFS east of the GF,

and separating the measures of the fault segments located south


and north of the prolongation of the HFF. The large decrease of the
total displacement eld north of the HFF prolongation can be
clearly noticed.
Based on these data, we conclude that both the geometry of the
fractures and their displacement amounts along the TFS indicate
that the HFF extends further south-east than the junction with the
GF. Its surface trace has not been recognized up to today because it
is not represented by a well developed NWeSE through-going
chelon
fault; on the other hand, it is represented by short en-e
NNW-SSE normal faults that combine with the faults and tension
fractures of the TFS, with which the HFF interacts. The eastern HFF
is in an under-developed stage both west and east of the junction
with the GF, but west of it there are NWeSE fault strands that
connect the NNW-SSE faults yet. East of the junction with the GF,
the mostly hidden HFF is capable of interfering with the TFS by: 1)

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

153

Fig. 16. Plot of the total offset measured on all the fault segments of the TFS east of the GF, with distinction between the fault segments located south and north of the prolongation
of the HFF. The large total decrease in the displacement eld in the northern part of the TFS can be clearly observed.

decreasing the offsets of the TFS faults, 2) locally inuencing their


strike, 3) interrupting some of the faults, 4) inducing variations in
tension fracture strikes, and 5) in their opening directions.

6. Conclusions
In the eld we performed a detailed mapping of the eastern
sector of the NWeSE active transcurrent Husavik-Flatey Fault (HFF),
in North Iceland, as well as of the NeS Theistareykir Fissure Swarm
(TFS). We systematically measured the components of Holocene
fault offsets and the opening directions of tension fractures. Previous studies suggested that the surface trace of the HFF ends in
correspondence of the junction with the westernmost fault of the
TFS, known as Gudnnugja Fault (GF).
We mapped 649 fault scarps and 1208 tension fractures from
10 m to kilometric scales. The central and western onshore sectors
of the HFF are made of dominant NW-SE-striking, major fault
segments with some bending and pull-apart basins. The eastern
sector of the HFF, west of the junction with the GF, is composed of
short NW-SE-striking fault segments linked by NNW-SSE normal
faults.
The GF shows larger offsets south of the junction with the HFF,
suggesting that part of its extension is accommodated by transferring the slip to the HFF. Moreover, the junction is characterised
by the intersection between NNW-SSE and NeS faults. A series of
similar features have been recognized further south-east along a
NWeSE corridor, providing clues into the interpretation of the
prolongation of the HFF across the TFS:
- All the normal faults of the TFS have larger average and
maximum offsets south of the proposed prolongation of the
HFF;
- the TFS width passes from 8 to 4 km north of the HFF
prolongation;
- some faults transition to tension fractures northward of this
prolongation;
- a change in strike of both faults and tension fractures occurs
along the NWeSE corridor;

- opening directions of tension fractures become more complex


along the corridor;
chelon NNW-SSE-striking normal faults and tension s- en-e
sures with a slight right-lateral component of opening are present only along the corridor.
We argue that all these are evidence of the rst stage of
development of the HFF across the TFS and of its interaction with
TFS structures. Another possibility is that an older segment of the
HFF is lying below the Holocene lavas of the TFS and has been
poorly active during this period, with the consequence that it has
not propagated upward yet. The eastern HFF is in an underdeveloped stage both west and east of the junction with the GF,
but west of it the deformation was large enough to induce faulting
at the surface along NWeSE strands that connect NNW-SSE normal
faults. East of the junction with the GF, the HFF is still mostly hidden
under the late Quaternary lava ows, but it is capable of creating a
pattern of interference with the TFS represented by a system of enchelon NNW-SSE fault segments and tension fractures.
e
The different geometry, deformation amount and kinematics of
the HFF, from north-west to south-east, can be explained in terms
of the different age of the involved rocks along the various sectors
of the HFF, which become younger towards the rift axis. As rocks
rejuvenate towards the south-east, the HFF passes from a mature
stage to the west to a very incipient stage to the east; moreover, the
HFF has to interact with the wide fault zone of the rift through a
complex fracture pattern. This can be considered as one of the best
examples of propagation of a transform fault tip towards the axis of
a volcanic ridge.

Acknowledgements
We thank the Iceland Meteorological Ofce for the seismic data
of the studied area. This work is a contribution to the International
Lithosphere Program - Task Force II. We are grateful to two anonymous reviewers for their precious comments and suggestions,
which enabled us to greatly improve the overall quality of the
paper.

154

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155

References
Acocella, V., Gudmundsson, A., Funiciello, R., 2000. Interaction and linkage of
extension fractures and normal faults: examples from the rift zone of Iceland.
J. Struct. Geol. 22 (9), 1233e1246.
Agostini, A., Corti, G., Zeoli, A., Mulugeta, G., 2009. Evolution, pattern, and partitioning of deformation during oblique continental rifting: Inferences from
lithospheric-scale centrifuge models. Geochem. Geophy. Geosy. 10 (11).
An, L.-J., Sammis, C.G., 1996. Development of strike-slip faults. Shear experiments in
granular materials and clay using a new technique. J. Struct. Geol. 18 (8),
1061e1077.
Angelier, J., Bergerat, F., Dauteuil, O., Villemin, T., 1997. Effective tension-shear relationships in tension ssure swarms, axial rift zone of northeastern Iceland.
J. Struct. Geol. 19 (5), 673e685.
Angelier, J., Bergerat, F., Homberg, C., 2000. Variable coupling explains complex
tectonic regimes near oceanic transform fault: Flateyjarskagi, Iceland. Terra
Nova 12, 97e101.

 ttir, T., Lund, B., Jiang, W., Geirsson, H., Bjo
rnsson, H., Einarsson, P.,
Arnad
o
Sigurdsson, T., 2009. Glacial rebound and plate spreading: results from the rst
countrywide GPS observations in Iceland. Geophys. J. Int. 177, 691e716.
Autin, J., Bellahsen, N., Leroy, S., Husson, L., Beslier, M.O., d'Acremont, E., 2013. The
role of structural inheritance in oblique rifting: insights from analogue models
and application to the Gulf of Aden. Tectonophysics 607, 51e64.
Bergerat, F., Angelier, J., 2008. Immature and mature transform zones near a hot
rnes Fracture Zone (Iceland).
spot: the South Iceland Seismic Zone and the Tjo
Tectonophysics 447, 142e154.
Bergerat, F., Angelier, J., Villemin, T., 1990. Fault systems and stress patterns on
emerged oceanic ridges: a case study in Iceland. Tectonophysics 179, 183e197.
Bergerat, F., Angelier, J., Homberg, C., 2000. Tectonic analysis of the Husavk-Flatey
Fault (northern Iceland) and mechanisms of an oceanic transform zone, the
Tjornes Fracture Zone. Tectonics 19, 1161e1177.
Bjornsson, A., 1985. Dynamics of crustal rifting in Iceland. J. Geophys. Res. 90,
151e162.
 ttir, B., Einarsson, P., 1979. Seismic activity associated with the September
Brandsdo
1977 deation of the Kraa central volcano in NE-Iceland. J. Volcanol. Geoth.
Res. 6, 197e212.
ttir, B., 2006. Tectonic stress and magma chamber
Buck, W.R., Einarsson, P., Brandsdo
size as controls on dike propagation: constraints from the 1975e1984 Kraa
rifting episode. J. Geophys. Res. 111 (B12), B12404.
Cattin, R., Doubre, C., De Chabalier, J.-B., King, G., Vigny, C., Avouac, J.-P., Ruegg, J.-C.,
2005. Numerical modelling of Quaternary deformation and post-seismic
displacement in the AsaleGhoubbet rift (Djibouti, Africa). Earth Planet. Sci.
Lett. 239, 352e367.
Cladouhos, T.T., Marret, R., 1996. Are fault growth and linkage models consistent
with power-law distributions of fault lengths? J. Struct. Geol. 18, 281e293.
Clifton, A.E., Kattenhorn, S.A., 2006. Structural architecture of a highly oblique
divergent plate boundary segment. Tectonophysics 419 (1), 27e40.
Clifton, A.E., Schlische, R.W., Withjack, M.O., Ackermann, R.V., 2000. Inuence of rift
obliquity on fault-population systematics: results of experimental clay models.
J. Struct. Geol. 22, 1491e1509.
Dauteuil, O., Angelier, J., Bergerat, F., Verrier, S., Villemin, T., 2001. Deformation
partitioning inside a ssure swarm of the northern Icelandic rift. J. Struct. Geol.
23, 1359e1372.
Davy, P., 1993. On the frequency-length distribution of the San Andreas fault system.
J. Geophys. Res. 98, 12141e12151.
rnes-Ba
rardalur in
Einarsson, T., 1958. A survey of the geology of the area Tjo
northern Iceland including paleomagnetic studies. Soc. Sci. Isl. 32, 1e79.
Einarsson, P., 1986. Seismicity along the eastern margin of the North American
Plate. In: Vogt, P.R., Tucholke, B.E. (Eds.), The Geology of North America, The
Western North Atlantic Region, M. Geological Society of America, pp. 99e116.
Einarsson, P., 1987. Compilation of earthquake fault plane solutions in the North
Atlantic and Arctic Oceans. In: Kasahara, K. (Ed.), Recent Plate Movements and
Deformation, Geodynamics Series, 20. Am. Geophys. Union, pp. 47e62.
Einarsson, P., 1991. Earthquakes and present-day tectonism in Iceland. Tectonophysics 189, 261e279.
kull 58,
Einarsson, P., 2008. Plate boundaries, rifts and transforms in Iceland. Jo
35e58.
Einarsson, P., 2015. Mechanisms of Earthquakes in Iceland. Encyclopedia of earthquake Engineering. In: Beer, M., Patelli, E., Kougioumtzoglou, I.A., Au, S.-K.
(Eds.). Springer Verlag, Berlin, Heidelberg. http://dx.doi.org/10.1007/978-3-64236197-5_298-1.
kull 29, 37e43.
Einarsson, P., Bjornsson, S., 1979. Earthquakes in Iceland. Jo
ttir, B., 1980. Seismological evidence for lateral magma
Einarsson, P., Brandsdo
intrusion during the July 1978 deation of the Kraa volcano in NE-Iceland.
J. Volcanol. Geoth. Res. 47, 160e165.
Einarsson, P., Smundsson, K., 1987. Earthquake epicenters 1982-1985 and volcanic
ur, Reykjavk (Eds.), I
systems in Iceland (map). In: Sigfsson, .I., Menningarsjo
rn Sigurgeirsson.
Hlutarins Eli, Festschrift for orbjo
Feigl, K., Gasperi, J., Sigmundsson, F., Rigo, A., 2000. Crustal deformation near
Hengill volcano Iceland 1993e1998: coupling between magmatic activity and
faulting inferred from elastic modelling of satellite radar interferometry.
J. Geophys. Res. 105, 25,655e25,670.
Fjader, K., Gudmundsson, A., Forslund, T., 1994. Dikes, minor faults and mineral
veins associated with a transform fault in North Iceland. J. Struct. Geol. 16,

109e119.
Forslund, T., Gudmundsson, A., 1991. Crustal spreading due to dikes and faults in
Southwest Iceland. J. Struct. Geol. 13, 443e457.
Forslund, T., Gudmundsson, A., 1992. Structure of tertiary and Pleistocene normal
faults in Iceland. Tectonics 11, 5748.
Garcia, S., Dhont, D., 2005. Structural analysis of the HsavkeFlatey transform Fault
and its relationships with the rift system in Northern Iceland. Geodin. Acta 18,
31e41.
Garcia, S., Angelier, J., Bergerat, F., Homberg, C., 2002. Tectonic analysis of an oceanic
transform fault zone revealed by faulteslip data and earthquake focal mechanisms: the HusavikeFlatey Fault, Iceland. Tectonophysics 344, 157e174.

ttir, T., Vo
lksen, C., Jiang, W., Sturkell, E., Villemin, T.,
Geirsson, H., Arnad
o
Einarsson, P., Sigmundsson, F., Stef
ansson, R., 2006. Current plate movements
across the Mid-Atlantic Ridge determined from 5 years of continuous GPS
measurements in Iceland. J. Geophys. Res. 111, B9.
Grant, J.V., Kattenhorn, S.A., 2004. Evolution of vertical faults at an tension plate
boundary, southwest Iceland. J. Struct. Geol. 26 (3), 537e557.
Gudmundsson, A., 1987a. Geometry, formation and development of tectonic fractures on the Reykjanes Peninsula, Southwest Iceland. Tectonophysics 139,
295e308.
Gudmundsson, A., 1987b. Tectonics of the thingvellir ssure swarm, SW Iceland.
J. Struct. Geol. 9, 61e69.
Gudmundsson, A., 1990. Emplacement of dikes, sills and crustal magma chambers
at divergent plate boundaries. Tectonophysics 176 (3), 257e275.
Gudmundsson, A., 1992. Formation and growth of normal faults at the divergent
plate boundary in Iceland. Terra Nova 4, 464e471.
Gudmundsson, A., 1993. On the structure and formation of fracture zones. Terra
Nova 5 (3), 215e224.
Gudmundsson, A., 1995. Infrastructure and mechanics of volcanic systems in Iceland. J. Volcanol. Geoth. Res. 67, le22.
Gudmundsson, 2007. Infrastructure and evolution of ocean-ridge discontinuities in
Iceland. J. Geodyn. 43, 6e29.
m, K., 1991. Structure and development of the Sveinagja
Gudmundsson, A., Backstro
graben, Northeast Iceland. Tectonophysics 200, 111e125.
lfsson, S., Jo
 nsson, M.Th, 1993. Structural analysis of a
Gudmundsson, A., Brynjo
transform fault-rift zone junction in North Iceland. Tectonophysics 220,
205e221.
Hardacre, K.M., Cowie, P.A., 2003. Variability in fault size scaling due to rock
strength heterogeneity: a nite element investigation. J. Struct. Geol. 25,
1735e1750.
Harding, T.P., Vierbuchen, R.C., Christie-Blick, N., 1985. Structural styles, plate tectonic settings, and hydrocarbon potential of divergent (transtensional) wrench
faults. In: Biddle, K.T., Christie-Blick, N. (Eds.), Strike-slip Deformation, Basin
Formation, and Sedimentation. Soc. Econom. Pa, 37, pp. 51e77.
Helgason, J., Zentilli, M., 1985. Field characteristics of laterally emplaced dikes:
rdur, eastern Iceland.
Anatomy of an exhumed Miocene dike swarm in Reydarfjo
Tectonophysics 115 (3), 247e274.

ttir, A.R.,
ll ssure swarm and the eastern
Hjartardo
Einarsson, P., 2012. The Kverkfjo
boundary of the Northern Volcanic Rift Zone,. Icel. Bull. Volcanol. 74, 143e162.
http://dx.doi.org/10.1007/s00445-011-0496-6.

ttir, A.R.,
Hjartardo
Einarsson, P., Sigursson, H., 2009. The ssure swarm of the
Askja volcanic system along the divergent plate boundary of N Iceland. Bull.
Volcanol. 71, 961e975. http://dx.doi.org/10.1007/s00445-009-0282-x.

ttir, A.R.,
Hjartardo
Einarsson, P., Bramham, E., Wright, T.J., 2012. The Kraa ssure
swarm, Iceland, and its formation by rifting events. Bull. Volcanol. 74,
2139e2153.

ttir, A.R.,
 ttir, S., Bjo
rnsdo
ttir, ., Brandsdo
ttir, B.,
Hjartardo
Einarsson, P., Magnsdo
2015. Fracture systems of the Northern Volcanic Rift Zone, Iceland e an onshore
part of the Mid-Atlantic plate boundary. In: Wright, T.J., Ayele, A., Ferguson, D.J.,
Kidane, T., Vye-Brown, C. (Eds.), Magmatic Rifting and Active Volcanism. Geol.
Soc. Lond. Spec. http://dx.doi.org/10.1144/SP420.1. Pub. 420.
Holland, M., Urai, J.L., Martel, S., 2006. The internal structure of fault zones in
basaltic sequences. Earth Planet. Sci. Lett. 248 (1), 301e315.
Huggins, P., Watterson, J., Walsh, J.J., Childs, C., 1995. Relay zone geometry and
displacement transfer between normal faults recorded in coal-mine plans.
J. Struct. Geol. 17, 1741e1755.
Jackson, C.A.L., Gawthorpe, R.L., Sharp, I.R., 2006. Style and sequence of deformation
during tension fault-propagation folding: examples from the Hammam Faraun
and El-Qaa fault blocks, Suez Rift. Egypt. J. Struct. Geol. 28 (3), 519e535.
Jancin, M., Young, K.D., Voight, B., Aronson, J.L., Saemundsson, K., 1985. Stratigraphy
and K/AR ages across the west ank of the northeast Iceland Axial Rift Zone, in
relation to the 7 MA volcano-tectonic reorganization of Iceland. J. Geophys. Res.
Solid Earth 90 (B12), 9961e9985.
Jonsson, S., Einarsson, P., Sigmundsson, F., 1997. Extension across a divergent plate
boundary, the eastern Volcanic Rift Zone, south Iceland, 1967e1994, observed
with GPS and electronic distance measurements. J. Geophys. Res. 102,
11,913e11,929.
Jouanne, F., Villemin, T., Berger, A., Henroit, O., 2006. Rift-transform junction in
North Iceland: rigid blocks and narrow accommodation zones revealed by GPS
1997-1999-2002. Geophys. J. Int. 167, 1439e1446.
Kim, Y.-S., Peacock, D.C.P., Sanderson, D.J., 2003. Mesoscale strike-slip faults and
damage zones at Marsalform, Gozo Island, Malta. J. Struct. Geol. 25, 793e812.
nsson, S., 2013. The stress shadow induced
Maccaferri, F., Rivalta, E., Passarelli, L., Jo
by the 1975e1984 Kraa rifting episode. J. Geophys. Res. Solid Earth 118 (3),
1109e1121.

A. Tibaldi et al. / Journal of Structural Geology 83 (2016) 134e155


ttir, S., Brandsdo
ttir, B., 2011. Tectonics of the eistareykir ssure swarm.
Magnsdo
kull 61, 65e79.
Jo
Main, I.G., Meredith, P.G., Sammonds, P.R., Jones, C., 1990. Inuence of fractal aw
distributions on rock deformation in the brittle eld. In: Knipe, R.J., Rutter, E.H.
(Eds.), Deformation Mechanisms, Rheology and Tectonics. Geol. Soc. London
Spec. Publ. 54, pp. 81e96.
Mann, P., 2007. Global catalogue, classication and tectonic origins of restraining
and releasing bends on active and ancient strike-slip fault systems. In:
Cunningham, W.D., Mann, P. (Eds.), Geol. Soc, 290. Spec. Publ., London,
pp. 13e142.
Mann, P., Demets, C., Wiggins-Grandison, M., 2007. Toward a better understanding
of the Late Neogene strike-slip restraining bend in Jamaica: geodetic, geological,
and seismic constraints. In: Cunningham, W.D., Mann, P. (Eds.), Tectonics of
Strike-slip Restraining and Releasing Bends, Geol. Soc, 290. Spec. Publ., London,
pp. 239e253.
Manseld, C., Cartwright, J., 2001. Fault growth by linkage: observations and implications from analogue models. J. Struct. Geol. 23, 745e763.
Martel, S.J., Pollard, D.D., Segall, P., 1988. Development of simple fault zones in
granitic rock, Mount Abbot quadrangle, Sierra Nevada, California. Geol. Soc. Am.
Bull. 100, 1451e1465.
nsson, S., 2014. Plate boundary deformation in North Iceland during
Metzger, S., Jo
1992e2009 revealed by InSAR time-series analysis and GPS. Tectonophysics
634, 127e138.
 nsson, S., Geirsson, H., 2011. Locking depth and slip-rate of the
Metzger, S., Jo
Husavk Flatey fault, North Iceland, derived from continuous GPS data
2006e2010. Geophys. J. Int. 187, 564e576.
Metzger, S., Jonsson, S., Danielsen, G., Hreinsdottir, S., Jouanne, F., Giardini, D.,
rnes Fracture Zone, North IceVillemin, T., 2013. Present kinematics of the Tjo
land, from campaign and continuous GPS measurements. Geophys. J. Int. 192,
441e455.
Olson, J., Pollard, D.D., 1989. Inferring paleostresses from natural fracture patterns:
A new method. Geology 17 (4), 345e348.
Opheim, J., Gudmundsson, A., 1989. Formation and geometry of fractures, and
related volcanism, of the Kraa ssure swarm, northeast Iceland. Bull. Geol. Soc.
Am. 101, 1608e1622.
 Mariotto, F., Bonali, F.L., Tibaldi, A., Rust, D., Oppizzi, P., Cavallo, A., 2015.
Pasquare
Holocene displacement eld at an emerged oceanic transform-ridge junction:
the Husavik-Flatey Fault - Gudnnugja Fault system, North Iceland. J. Struct.
Geol. 75, 118e134.
Peacock, D.C.P., 2002. Propagation, interaction and linkage in normal fault systems.
Earth Sci. Rev. 58 (1), 121e142.
Pollard, D.D., Aydin, A., 1984. Propagation and linkage of oceanic ridge segments.
J. Geophys. Res. 89, 10,017e10,028.
Renshaw, C.E., Pollard, D.D., 1994. Are large differential stresses required for straight
fracture propagation paths? J. Struct. Geol. 16, 817e822.
Rovida, A., Tibaldi, A., 2005. Propagation of strike-slip faults across Holocene
volcano-sedimentary deposits, Pasto, Colombia. J. Struct. Geol. 27 (10),
1838e1855.
Rubin, A., 1992. Dike-induced faulting and graben subsidence in volcanic rift zones.
J. Geophys. Res. 97, 1839e1858.
Rubin, A.M., Pollard, D.D., 1988. Dike induced faulting in rift zones in Iceland and
Afar. Geology 16, 413e417.
Saemundsson, K., 1974. Evolution of the axial rifting zone in northern Iceland and
the Tjornes fracture zone. Geol. Soc. Am. Bull. 85, 495e504.
Saemundsson, K., 1980. Outline of the geology of Iceland. Jokull 29, 7e28.
Saemundsson, K., Hjartarson, A., Kaldal, I., Sigurgeirsson, M.A., Kristinsson, S.G.,
Vikingsson, S., 2012. Geological map of the northern volcanic zone, Iceland.
Northern Part 1: 100.000. Iceland GeoSurvey and Landsvirkjun, Reykjavik.

155

Scholz, C.H., Cowie, P.A., 1990. Determination of total strain from faulting using slip
measurements. Nature 346, 837e839.
Scholz, C.H., Dawers, N.H., Yu, J.-Z., Anders, M.H., Cowie, P.A., 1993. Fault growth and
fault scaling laws: preliminary results. J. Geophys. Res. 98, 21,951e21,961.
Sella, G.F., Dixon, T.H., Mao, A., 2002. REVEL: a model for Recent plate velocities
from space geodesy,. J. Geophys. Res. 107 (B4), 2081. http://dx.doi.org/10.1029/
2000JB000033.

ttir, S., Vogfjo
rd, K.S., Ofeigsson,
Sigmundsson, F., Hooper, A., Hreinsdo
B.G., Rafn
Heimisson, E., Dumont, S., Parks, M., Spaans, K., Gudmundsson, G.B., Drouin, V.,

ttir, T., Jo
 nsdo
 ttir, K., Gudmundsson, M.T., Ho
gnado
ttir, T., Mara
Arnad
o
ttir, H., Hensch, M., Einarsson, P., Magnsson, E., Samsonov, S.,
Fridriksdo

ttir, B., White, R.S., Agstsd
ttir, T., Greeneld, T., Green, R.G., Rut
Brandsdo
o
 Pedersen, R., Bennett, R.A., Geirsson, H., La Femina, P.C.,
ttir, A.,
Hjartardo
rnsson, H., P
llhoff, M., Braiden, A.K.,
Bjo
alsson, F., Sturkell, E., Bean, C.J., Mo
Eibl, E.P.S., 2015. Segmented lateral dyke growth in a rifting event at
rarbunga volcanic system. Icel. Nat. 517, 191e195.
Ba
nvold, K., Shimizu, N., 2001. Melt generation and
Slater, L., McKenzie, D., Gro
movement under Theistareykir, NE Iceland. J. Petrol 42, 321e354.
Sonnette, L., Angelier, J., Villemin, T., Bergerat, F., 2010. Faulting and ssuring in
active oceanic rift: surface expression, distribution and tectonicevolcanic
interaction in the Thingvellir Fissure Swarm. Icel. J. Struct. Geol. 32, 407e422.
nsson, R., Gumundsson, G.B., Halldo
rsson, P., 2008. Tjo
rnes fracture zone.
Stefa
New and old seismic evidences for the link between the North Iceland rift zone
and the Mid-Atlantic ridge. Tectonophysics 447 (1e4), 117e126. http://
dx.doi.org/10.1016/j.tecto.2006.09.019.
de, F.,
Stracke, A., Zindler, A., Salters, V.J.M., McKenzie, D., Blichert-Toft, J., Albare
nvold, K., 2003. Theistareykir revisited. Geochem. Geophys. Geosyst 4 (2),
Gro
8507.
Stucchi, M., Rovida, A., Capera, A.G., Alexandre, P., Camelbeeck, T., Demircioglu, M.B.,
Giardini, D., 2013. The SHARE European earthquake catalogue (SHEEC)
1000e1899. J. Seismol. 17 (2), 523e544.
Thors, K., 1982. Shallow seismic stratigraphy and structure of the southernmost part
kull 32, 107e112.
of the Tjornes Fracture Zone. Jo
 Mariotto, F., 2015. Structural Geology of Active Tectonic Areas
Tibaldi, A., Pasquare
and Volcanic Regions. Lulu Press, Raleigh, US, p. 203.
Trippanera, D., Ruch, J., Acocella, V., Rivalta, E., 2015. Experiments of dike-induced
deformation: insights on the long-term evolution of divergent plate boundaries. J. Geophys. Res. Solid Earth 120. http://dx.doi.org/10.1002/2014JB011850.
Tryggvason, E., 1980. Subsidence events in the Kraa area, North-Iceland, 19751979. J. Geophysics-Zeitschrift fur Geophys 47, 141e153.
Tryggvason, E., 1984. Widening of the Kraa ssure swarm during the 1975-1981
volcano-tectonic episode. Bull. Volcanol. 47, 47e69.
Tryggvason, E., 1986. Multiple magma reservoirs in a rift zone volcano: ground
deformation and magma transport during the September 1984 eruption of
Kraa, Iceland. J. Volcanol. Geotherm. Res. 28, 1e44.
Walker, G.P., 1992. Coherent intrusion complexes in large basaltic volcanoesda
new structural model. J. Volcanol. Geoth. Res. 50 (1), 41e54.
Walker, G.P., 1999. Volcanic rift zones and their intrusion swarms. J. Volcanol. Geoth.
Res. 94 (1), 21e34.
Willemse, E.J., 1997. Segmented normal faults: correspondence between three
dimensional mechanical models and eld data. J. Geophys. Res. 102, 675e692.
Wright, T.J., Sigmundsson, F., Ayele, A., Belachew, M., Brandsdottir, B., Calais, E.,
Ebinger, C., Einarsson, P., Hamling, I., Keir, D., Lewi, E., Pagli, C., Pedersen, R.,
2012. Geophysical constraints on the dynamics of spreading centres from rifting
episodes on land. Nature Geoscience. http://dx.doi.org/10.1038/NGEO1428.
Young, K.D., Jancin, B., Orkan, N.I., 1985. Transform deformation of tertiary rocks
along the Tjornes fracture zone, North Central Iceland. J. Geophys. Res. 90,
9986e10010.

You might also like