Seismic Tomography and Surface Deformation in by

Seismic Tomography and Surface Deformation in
Krfsuvik, SW Iceland
by
Jing Liu
B.S. Geophysics, China University of Geosciences (Wuhan), 2008
Submitted to the Department of Earth, Atmospheric, and Planetary Sciences
in Partial Fulfillment of the Requirements for the Degree
ACHUES
of
INS1W
MSSACHUSM
Master of Science in Geophysics
OF TECHNOLOGY
at the
Massachusetts Institute of Technology
MAR 0 3 2014
LIBRARIES
February 2014
C 2014 Massachusetts Institute of Technology. All rights reserved.
Signature of Author ...................
...............................................
Department of Earth, Atmospheric, and Planetary Sciences
November. 21, 2013
C ertified by
........
............................................
Michael C. Fehler
Senior Research Scientist-Earth Resources Laboratory
Thesis Supervisor
/7
J
Certified by.
Cecil & Ida Green Pro
&
.................................................
Bradford H. Hager
r, Director of Earth Resources Laboratory
Thesis Co-Supervisor
7
Accepted by
Y
Rob Van der Hilst
Schlumberger Professor
Head, Department of Earth, Atmospheric and Planetary Sciences
Seismic Tomography and Surface Deformation in
Krfsuvik, SW Iceland
by
Jing Liu
Submitted to the Department of Earth, Atmospheric, and Planetary Sciences
On November 21, 2013, in partial fulfillment of the
requirements for the Degree of Master of Science in Geophysics
Abstract
The Krfsuvik region of southwestern Iceland is a region of high potential for geothermal
energy that is currently experiencing seismic swarm activity and active surface
deformation. Understanding the subsurface structure of the area is of great scientific and
practical significance. Using permanent and temporary seismic stations deployed in the
region, we captured an earthquake swarm from Nov. 2010 to Feb. 2011 clustered around
the center of the Krfsuvik volcanic system. We studied the seismicity and Vp, Vs and
Vp/Vs ratio in this region by applying double difference tomography. Our tomography
result indicates a low velocity zone at a depth of about 6 kin, directly beneath the
earthquake swarm. At the same time, our relocation result delineates strike-slip and dipslip faults above and around this low velocity zone. Brittle-ductile transition is delineated
based on the distribution of the earthquakes in this area. In order to understand the
relation between the subsurface structure and the surface deformation, we modeled
surface deformation using the input parameters constrained from our tomography results.
We found that the main deformation is well captured by a pressure source yielding a
volume expansion of about 30x 106 m 3 at the depth of about 6 km, centered on the low
velocity zone detected in tomography. And the secondary deformation could be
explained by the normal and the right-lateral slip faults, whose patterns are delineated by
the earthquake relocations. The combination of the local stress caused by the expanding
source and regional stress that yields a combination of left-lateral shear and extension
might have triggered the earthquakes. Based on the low Vp, Vs and possibly high Vp/Vs
ratio at depth of ~6 km and its expanding property, the possibilities of supercritical water,
H2 0-rich partial melting with magma intrusion are discussed. The results of this thesis
provide new insights to understand the seismicity and surface deformation in volcanic
zones as well as provided important reference in exploration of new geothermal areas.
Thesis Supervisors:
Michael C. Fehler
.
Title: Senior Research Scientisph- EarthResouycpe Laboratory
Bradford H. Hager
Title: Cecil & Ida Green Professor, birector of Earth Resources Laboratory
2
Acknowledgements
I would like to thank the people for their help towards the completion of this thesis
during my two years of studying in the Earth Resource Laboratory at MIT.
First of all, I would like to thank my advisor, Dr. Michael Fehler for his support and
guidance throughout my study at MIT. I always feel lucky to meet and work with such
a reputable scientist as Dr. Fehler who offered me the chance to enter MIT as well as
the field of geophysics. I would also like to thank Prof. Brad Hager for his fruitful and
patient suggestions regarding my current as well as future work.
I would like to acknowledge the people in Earth Resource Laboratory (Micheal
Fehler, Bradford Hager, Maria Zuber, Robert van der Hilst, Alison Malcolm, Anna
Shaughnessy, Haijiang Zhang, Yingcai Zheng, Xingding Fang, Qin Cao, Hui Huang,
Xuefeng Shang, Chunquan Yu, Di Yang, Haoyue Wang, Lucas Willemsen, Alan
Richardson) for many helpful comments and exposure to different aspects of the
seismology and geodynamics problems in my work. I would also like to thank my
husband, Rong, for his support during difficult times.
This work was funded by US DOE, Iceland GEORG Program, Iceland & Swedish
National Science Foundations.
3
Contents
Chapter 1.................................................................................................................................
8
9
Introduction..............................................................................................................................
9
1.1 Plate Boundary of Iceland ..........................................................................................
10
1.2 Reykjanes Peninsula, SW Iceland ..................................................................................
1.3 Previous Study of Seismicity, Deformation and Structure in Krfsuvik........12
15
1.4 Surface Manifestations of Krysuvik Geotherm al Field .......................................
16
......................
...............................
al
Field
Geotherm
1.5 Our Study of the Krf'suvik
Chapter 2 ..................................................
.........................................................................
Seism ic Event Relocation and Tom ography.............................................................
2.1 Seism ic Data and Stations .............................................................................................
2.2 Theory of Event Relocation and Tom ography......................................................
2.3 Results of Event Relocation and Tom ography......................................................
2.3.1 Travel times used in this study..........................................................................................
2.3.2 Event Relocation Results ...........................................................................................................
2.3.3 Checkerboard Test .......................................................................................................................
2.3.4 Tom ography Results....................................................................................................................
.................................
Chapter 3 ....................................
.............................
......
19
19
19
21
22
22
23
26
31
36
Surface Deform ation Observation and M odeling...................................................
3.1 Geodetic Observation .....................................................................................................
3.2 Surface Deform ation Modeling...................................................................................
36
36
40
Chapter 4 ...............................................................................................................................
45
........ .................. ..................... 45
............................................
Discussion.....
45
4.1 Brittle-ductile Transition ................... ............................. .....................................
48
4.2 Possibilities of the low velocity zone.....................................................................
.............................
. ... .............................................
Chapter 5 . ........................
54
............... ...........................................................
54
................................................................................
56
Geotherm al M anifestation in Krysuvik.....................................................................
56
Conclusion ..........................................
Appendix A ..........................
Appendix B .....................................................................
.....................................................
GPS m easurem ents in Krysuvik ...................................................................................
References .........................
....
...................
4
... ............
.....
57
57
........................... 59
List of Figures
Figure 1-1: Iceland plate boundary. (Hjaltadottir, 2010). Fissure swarms are shown as
grey areas and central volcanoes enclosed with grey, thin circles. The black
dotted lines are the transform zones. HM: Hreppar micro-plate................. 10
Figure 1-2: Tectonic map of Reykjanes Peninsula and its location in SW Iceland
(inset). The blue lines are the plate boundary (solid-axial rifts, long dashed
line-propagating rift, dotted line-fracture zones). The hatched areas indicate
the high-temperature geothermal fields. Fissure swarms are shown as grey
areas in the inset, along the NE-SW direction. RR, Reykjanes Ridge; H,
Hengill volcano zone; WVZ, west Volcanic Zone; SISZ, South Iceland Seismic
12
Zone. M odified after Keiding et al., (2009) ................................................
Figure 1-3: Historical earthquakes in Krfsuvik from 1997 to 2006 (Keiding et al.,
2009). The two stars indicated the Mw5.9 earthquake in June 17, 2000 and
14
Mw5.0 earthquake in August 23, 2003 respectively. ...................................
Figure 1-4: Cumulative number of earthquakes in the Krfsuvik region from 2007 to
Sep. 2012 (continue with fig.3). The two diamonds indicate the two big swarm
15
events that happened in 2009 and 2011.......................................................
Figure 1-5: Geology map of the Krfsuvik area in our study. K: Lake Kleifarvatn; KV
18
(orange dot): Krysuvik Volcano. ................................................................
Figure 2-1: a. Seismic station distribution on the Reyjanes Peninsula from 2010 to
2011 (courtesy of Oli Gudmundsson). b. Seismic stations we used in this
study. The dashed black line indicates our study area. The orange dot is the
location of the Krfsuvik volcano. The small black dots are the earthquakes
20
from the SIL catalog. .................................................................................
Figure 2-2: a. Cumulative number of events from 1 't July 2010 to 3 1st December
2011 and the blue circle indicates the biggest jump that occurred on 2 7th
February 2011; b. Histogram of earthquake magnitudes from 1't July 2010 to
20
31st D ecem ber 2011. ...................................................................................
Figure 2-3: The residual histograms of Absolute P, S, Catalog differential P, S, and
24
C ross-correlation P, S. ...............................................................................
Figure 2-4: Earthquake events distribution in Krfsuvik area from Catalogue. .......... 25
5
Figure 2-5: Earthquake events distribution in Krfsuvik area after relocation. The
arrows on the top figure indicate the directions of N-S trending fault and NEtrending fault. ............................................................................................
25
Figure 2-6: Map view and cross sections presented in the following checkerboard test
and tom ography results..............................................................................
27
Figure 2-7: Checkerboard test of Vp for the map view in figure 2-6. Figures of the top
row are the input model; Figures of the middle row are the corresponding
recovered model using our acquisition geometry; Figures of the last row are
the corresponding resolvability distribution..............................................
28
Figure 2-8: Checkerboard test of Vp for the cross sections in figure 2-6. Figures have
the same meanings as in figure 2-8 .............................................................
28
Figure 2-9: Checkerboard test of Vs for the map view in figure 2-6. Figures have the
sam e m eanings as in figure 2-8..................................................................
29
Figure 2-10: Checkerboard test of Vs for the cross sections in figure 2-6. Figures have
the same meanings as in figure 2-8 .............................................................
29
Figure 2-11: Checkerboard test of Vp/Vs for the map view in figure 2-6. Figures have
the sam e m eanings as in figure 2-8 .............................................................
30
Figure 2-12: Checkerboard test of Vp/Vs for the cross sections in figure 2-6. Figures
have the same meanings as in figure 2-8.....................................................
30
Figure 2-13: Horizontal profiles of Vp, Vs and Vp/Vs ratio models at depths of 2,4,
and 6 km. The black solid line indicates the resolvability threshold of 0.7. The
black dashed lines are the other resolvability values...................................
34
Figure 2-14: Cross section profiles of Vp, Vs and Vp/Vs ratio models. The black solid
line indicates the resolvability threshold of 0.7. The black dashed lines are the
other resolvability values............................................................................
34
Figure 2-15: The isosurfaces of the Vp, Vs and Vp/Vs ratio. The regions where Vp <
5.75 km/s, Vs < 3.3 km/s and Vp/Vs > 1.89 are shown...............................
35
Figure 2-16: Resistivity at the depth of 5 km (Hersir et al., 2013). A low resistivity
body was observed at the similar location with the low velocity zone
discovered in this study...............................................................................
35
Figure 3-1: Displacement of GPS measurements in Krfsuvik area. The grey circles
indicate the uncertainties of each measurement (Michalczewska, 2012).
Different color arrows indicate measurements in different periods............. 37
Figure 3-2: The mean deformation in the InSAR Line of Sight direction from 2009 to
2011. The solid black lines indicate the volcanic systems (Michalczewska,
6
2012). The dashed black line reveals the secondary deformaton (dislocated
deformation of northwestern part of the main deformation)....................... 37
Figure 3-3: The algorithm of the cosine and sine rule on a sphere (Essentials of
38
G eophysics, M IT) .......................................................................................
Figure 3-4: Projection of the GPS measurements on the descending look angle of
39
TerraSA R-X ................................................................................................
Figure 3-5: Models and results simulated as the InSAR observation. Boxes al-3 are
map views of the deformation source geometry, and the red dot is the reference
point where deformation is taken to be 0. The numbers indicate the sources, 1
(black dot) for inflation source, 2 for normal fault and 3 for strike-slip fault.
The green lines indicate the strike directions of the faults and the red box
represents the fault plane for the normal fault. Plots bl-3 are the 3D views of
the corresponding map views. The cl-3 are the calculated displacement results
projected on the LOS of TerraSAR-X (descending). The silver boxes indicate
43
the corresponding area of the InSAR observation......................................
Figure 3-6: Seismic moment release starting from
1s
July 2010 to
3 1"
December
2011. The biggest jump occurred on 27th February 2011...........................
43
Figure 3-7: Comparison between the observed InSAR deformation and the simulated
deformation. The location of the expanding source mapped on the right picture
44
is (63.910N , 22.050W )...............................................................................
Figure 4-1: Longitude view of 3D plot with Vp (5.75 km/s) isosurface body and
earthquakes in our study area. The blue lines indicate the brittle-ductile
transition. Green, blue and red lines indicate BDT at 4, 5 and 6 km,
47
respectively. K : Lake Kleifarvatn. ..............................................................
Figure 4-2: Temperature profiles. Red, green and blue lines indicate temperature of
47
5500C at 4, 5 and 6 km , respectively. ........................................................
Figure 4-3: Rock melting studies (Grove et al., 2006). We focus on the dashed curve
of basalt melting with 1-2 wt% water. Other curved lines indicated different
kinds of rock in various studies. The red horizontal indicates the 0.2 GPa at ~6
51
km in our interested depth. .........................................................................
Figure 4-4: Relation between Vp, Vp/Vs ratio and volume of fraction of fluids
(Nakajima et al., 2001) in basaltic material (amphibolite). Solid and dashed
curves show the water inclusion and melt inclusion, respectively. Thin,
moderate, and thick curves denote the aspect ratio of 0.001, 0.01 and 0.1,
respectively. The black and gray stars indicate the possibly obtained Vp/Vs
ratio in this study. The horizontal dotted line indicates the Vp value (5.75 - 5.8
51
km/s) obtained in our tomography. ............................................................
7
List of Tables
Table 2-1: List of the RMS of residuals. ..............................................................
24
Table 2-2: Initial model (one-dimensional velocity model)...................................
26
8
Chapter 1
Introduction
1.1 Plate Boundary of Iceland
Iceland (fig. 1-1), between the Reykjanes ridge (RR) in the south and the Kolbeinsey
ridge (KR) in the north, is located at the intersection of the Mid-Atlantic Ridge and
the North-Atlantic mantle plume. Four rift zones constitute the plate boundary: The
Reykjanes Peninsula (RP), which is the on-land continuation of the RR, the western
volcanic zone (WVZ), which is a northward continuation of the RP, the eastern
volcanic zone (EVZ) and the northern volcanic zone (NVZ) (Einarsson, 1991). Most
seismicity and geothermal activity in Iceland is in some way related to the MidAtlantic plate boundary.
The northwest motion of the Mid-Atlantic Ridge relative to the mantle plume has
induced considerable offset of the active plate boundary from the Mid-Atlantic ridge.
The REVEL plate motion model indicates a relative North America (NA) - Eurasia
(EU) motion rate about 19.9 mm/yr in the direction about N102 0 E at 64 0 N, 20.5 0 W
(Sella et al., 2002). In response to this unstable situation, two transform zones have
developed: the south Iceland seismic zone (SISZ), connecting the RP to EVZ, the
Tjornes fracture zone (TFZ), which is composed of three lineaments: the Grimsey
lineament (GL), the Husavik-Flatey fault (HFF) and the less active Dalvik lineament
(DL), connecting the NVZ to the KR. Differential movements are taken up by these
fracture zones.
Because the fracture zones are very young and stress field that
9
induced them is changing with time, a complex series of faults have formed, which
are not oriented parallel to the spreading direction (Clifton and Einarsson, 2005;
Einarsson, 1976; Einarsson and Eiriksson, 1982).
KR TFZ
G.
66'
651
HM
64*
.
E VZ
km
Z
RRt--
0
-24'
-22*
-20'
-18'
-16'
50s
100-14.
Figure 1-1: Iceland plate boundary. (Hjaltadottir, 2010). Fissure swarms are shown as
grey areas and central volcanoes enclosed with grey, thin circles. The black dotted
lines are the transform zones. HM: Hreppar micro-plate.
1.2 Reykjanes Peninsula, SW Iceland
The Reykjanes Peninsula (RP) (fig.1-2) represents the northward prolongation of
the Reykjanes Ridge (RR). The plate boundary along the Peninsula is a zone of high
seismicity and recent volcanism, forming a transition between the Reykjanes Ridge
(RR) to the southwest, the Western Volcanic Zone (WVZ) to the northeast, and the
South Iceland Seismic Zone (SISZ) to the east. The zone of seismicity trends around
N80 0 E on the central of the Rekjanes Peninsula, but bends toward the south on the
western part of the peninsula before connecting to the Rekjanes Ridge. The obliquity
10
of the plate boundary indicates a combination of left-lateral shear and extension,
which was observed on the RP by a geodetic study conducted around 1968-1972
(Brander et al., 1976). Recently detailed studies by Keiding et al. (2008) show a leftlateral motion of ~18 mm/yr and opening of ~7 mm/yr below a locking depth of ~7
km.
The superposition and relative motion of the spreading plate boundary over the
mantle plume are manifested by the faults and volcanic systems in Iceland, typically
characterized by a central volcano and a fissure (dyke) swarm (Thordarson and
Larsen, 2007).
Thus, the main tectonic features on the Reykjanes peninsula are a
large number of NE-SW trending volcanic fissures, a series of NE-SW strking normal
faults and N-S striking right-lateral strike-slip faults crosscutting the normal faults
(Clifton and Kattenhorn, 2006). The fissures and faults are made up of right-stepping
en-echelon volcano-tectonic segments from west to east, the Reykjanes, Svartsengi,
Fagradalsfjall, Krfsuvik, Brennisteinsfjoll, Hvalhnukur and Hengill fissure swarms
(fig.1-2). In these fissure swarms, both earthquake swarms and main shock-aftershock
sequences are observed.
11
tI
'Ijai
T-~
RR
Bre
71
63,8
-22.8"
-22.6"
n
q~joII
rl
-22.4*
-22.2'
-22*
I km
-21.8
-21.6
Figure 1-2: Tectonic map of Reykjanes Peninsula and its location in SW Iceland
(inset). The blue lines are the plate boundary (solid-axial rifts, long dashed linepropagating rift, dotted line-fracture zones). The hatched areas indicate the hightemperature geothermal fields. Fissure swarms are shown as grey areas in the inset,
along the NE-SW direction. RR, Reykjanes Ridge; H, Hengill volcano zone; WVZ,
west Volcanic Zone; SISZ, South Iceland Seismic Zone. Modified after Keiding et al.,
(2009)
1.3 Previous Study of Seismicity, Deformation and
Structure in Krfsuvik
As one of the active volcanic systems on the Reykjanes Peninsula, the Krysuvik
volcanic system located in the center of the Peninsula consists of the Kr suvik
volcano (63.930N, 22.10 0 W) and NE-SW trending eruptive fissures and fractures with
a total length of about 55 km and width of about 13 km (Clifton and Kattenhorn,
2006; Thordarson and Larsen, 2007).
The fissure swarms are intersected by a series
of N-S trending faults (Clifton and Kattenhorn, 2006), which are known in a few
cases to be right-lateral strike slip faults (Einarsson, 1991).
The first seismic station was installed on Reykjanes Peninsula in 1991 by the
South Iceland Lowland (SIL) seismic network that operated by the Icelandic
Meteorological Office (IMO) and increased to seven by 1997. Since then, detailed
12
seismicity has been recorded.
In 1997-2000, a high background seismicity rate
(Clifton and Kattenhorn, 2006; Einarsson, 1991) was observed in Krfsuvik. During
June-July 1999, an intense swarm occurred during a period of long-lived swarms
(fig.1-3). On June 17, 2000, a big earthquake up to Mw5.9 occurred on a N-S striking
strike-slip fault that was believed to be triggered by an Mw6.5 earthquake in the
South Iceland Seismic zone. However, following the June 2000 events, there was a
sharp decrease in the seismicity rate in Krfsuvik (fig.1-3). On August 23, 2003, an
event with Mw5.0 occurred at a N-S striking strike-slip fault in the Kr suvik region.
This main shock was followed by more than 1000 aftershocks located at depths of ~1
to ~5 km. Part of the aftershocks were located at the same fault as the main shock,
while the other part was located at different NE-SW striking fault planes. Most
earthquakes in Kr suvik are centered at ~5 km deep, with a variation from surface to
~10 km (Keiding et al., 2009). The shallowest earthquakes are located in the central
Krf'suvik area, where considerable change in geothermal manifestation was observed
at the surface. The 2000-2006 seismicity around Krfsuvik was also accompanied by
corresponding strain changes.
The observed areal strain rate during 2000-2006
showed pronounced signals of areal expansion (Keiding et al., 2008) in Krfsuvik.
However, the simulated areal strain did not capture this areal expansion when the
stress change was only considered to be caused by spreading of the plate boundary.
During this period, a tomography study using five-months of local microearthquake
data recorded from April 10 to August 30 in 2005 suggested a strong decrease of Pand S- wave velocity and Vp/Vs ratio at the depth of 3-6 km broadly located beneath
Krfsuvik-Kleifarvatn geothermal area, which was interpreted as zones of increased
13
fluid temperature and pressure (Geoffroy and Dorbath, 2008).
However, instead of
independently inverting the Vp/Vs, they directly divide the Vp by Vs, which leads to
large uncertainties of the result of Vp/Vs.
8000
O
Krsuvfk
60(X)
4000
B
2000
U
1997 1998 1999 2000 2001 2002 2003 2004 2005 2006
Figure 1-3: Historical earthquakes in Krfsuvik from 1997 to 2006 (Keiding et al., 2009).
The two stars indicated the Mw5.9 earthquake in June 17, 2000 and Mw5.0 earthquake in
August 23, 2003 respectively.
From May to October 2009, over 10,000 earthquakes were recorded by the
seismic stations deployed by the HYDRORIFT project (Kristjansdottir, 2013) around
Kr suvik (fig.1-4). The majority of the earthquakes were located at depths from 1.4 km
to 4 km and generally delineated N-S striking faults. The focal mechanism study
indicated that normal, strike-slip, and reverse faulting events took place on the same
apparent faults in the same swarms, which was suggested by the cause of non-optimally
oriented faults that could be explained by a component of volume increase. Using this
seismicity and the seismic stations deployed by the HYDRORIFT project, they detected a
low P-wave velocity anomaly located southwest of the Lake Kleifarvatn at the depth of 58 km (Kristjansdottir, 2013). Meanwhile, a low resistivity anomaly was observed in this
area by 3D magnetotelluric method at a shallower depth of ~5 km (Hersir et al., 2013).
14
During the period of the increase of the seismicity, uplift deformation was observed in
this area (Michalczewska et al., 2012)
1.6x10
4
> 1.41.2E
-s 0.8 -
E
2012
2009 2010 2011
year
Figure 1-4: Cumulative number of earthquakes in the Krfsuvik region from 2007 to Sep.
2012 (continue with fig.3). The two diamonds indicate the two big swarm events that
happened in 2009 and 2011.
'207
2008
1.4 Surface Manifestations of Krfsuvik Geothermal
Field
The Krysuvik high-temperature area is about 40 km 2 , located primarily at the
intersections of the fissures and the strike-slip faults (see map in Appendix A).
The
hyaloclastite unit is by far the most voluminous unit and hosts all geothermal
manifestations.
Welded lavas and spatter debris or scoria are also observed. Phreatic
craters are along zones of weakness as indicated by their linear distribution near faults. It
is widely believed that the numerous phreatic craters indicate the presence of some heat
source at depth, responsible for these explosive eruptions (Mawejje, 2007).
Most of the area are enclosed by the 15-500 C isotherm lines at the depth of 10-12
cm (Mawejje, 2007).
The area with isotherm (50 0C) was observed to be completely
altered to clays and, in most cases, it was active with fumaroles, mud pots, boiling
15
springs and sulphur (Appendix A).
Steam vents and fumaroles were observed to be
linearly distributed following fractures in some places. Recharge seems to be due to an
inflow of cold water from the northeast, probably from Lake Kleifarvatn. In addition to
steam, other gases are also produced. Gases like hydrogen sulfide could easily be
identified by the characteristic smell. Mud pots, like thick porridge boiling in a pot, were
only mapped at lower altitudes, especially at Seltnin (Southwest of Lake Kleifarvatn).
The clay material is mainly kaolinite, suggesting alteration by acid leaching. In these
mud pots, temperature as high as 950 C was recorded. Like mud pots, boiling springs
were also observed at lower altitudes, especially at Seltn'n. The hydrothermal
manifestations may be linked to an active heat source at depth.
1.5 Our Study of the Krfsuvik Geothermal Field
In our study, we focus on the Krfsuvik area enclosed in the dashed box in figure
1-5. The hyaloclastite units, surrounded by basaltic lava, are by far the most voluminous
units in the center of our study area. In the geologic map of figure 1-5, the rock types are
lava shield (> 7000 years old), prehistoric lavas (>2400 years old), hyaloclastite (older
and early Bruhnes), prehistoric lavas (>2400 years old), hyaloclastite (older and early
Bruhnes), compound lava on hyaloclastite and lavas of
8 th_9 th
century from left to right.
Only a few volcanic eruptions are reported to have occurred in historical times (the last
1100 years) in the vicinity of this area. Shallow level magma chambers and dense dyke
swarms are considered as the heat source for the geothermal system.
19 shallow exploration wells were drilled in the Krfsuvik area from 1945 to 1950.
Then another three additional wells were drilled in 1960 and an additional well was
16
drilled in 1995. During most of the 20 century, the local municipalities have been keen
on utilizing the geothermal fields in the Krfsuvik area for electricity production, large
scale domestic heating, and other purposes such as balneology, green houses, etc. So far,
however, no large-scale utilization has been realized.
In order to investigate whether there are some potential geothermal sources or risk
of magma intrusion around Krf'suvik, in 2010, Massachusetts Institute of Technology
(MIT), Reykjavik University (RU) and Uppsala University (UU) deployed 38 seismic
stations on the Reykjanes Peninsula and recorded the seismicity that occurred from 2010
to 2011 in the area around Krfsuvik. Based on these seismic data, we have studied the
seismicity in the region using double difference tomography to better relocate seismic
events and determine the velocity structure. The tomography reveals a low velocity zone
at the depth of about 6 km and the relocation of seismic events delineates the N-S
trending and NE-SW trending faults. Brittle-ductile transition and thermal profiles
are created based on the distribution of the earthquakes and the location of the low
velocity zone in this area. Interferometric Synthetic Aperture Radar (InSAR)
observation suggests 20-30 mm uplift as well as expansion of this area.
To better
understand the relations among seismicity, velocity structure and the observed
deformation, we use Coulomb 3 software (U.S. Geological Survey) to model the surface
deformation. Our result reveals that the first order deformation may be caused by the
inflation of magma sources at around 6 km, where the seismically determined low
velocity zone is located and the secondary deformation can be modeled using a NE-SW
trending normal fault and a N-S trending strike slip fault that are constrained from the
seismic event relocation. This result indicates that the seismicity in Kr suvik around
17
2010-2011 might be triggered by partial melting or supercritical water activities. Such
discovery provides new insights to understand the seismicity and surface deformation in
volcanic zones as well as provide important reference in exploration of new geothermal
areas.
w
R
iis E S OW5KI
Skiringar I Legends
HO
64-
I Potplochet.love#J
1 old, I
a,,,,
F orogdeq Neo
Sib63.95
- 2400
639
oe
1
0
OMM004or Oft 000/
(
& Vw
-
-
-22.4
ra~case-22we"Mw
- - b~
WoeWA
LOA
K
/
-
-
-22.2
-22
Longitude (degree)
-2-
-
-21.8
-2 6
fgns
G
L77
4004000
0
A'o.o
Figure 1-5
3
Geasoog
mapy ofteKfuikae
norstd.K
aeKliavt;K
pos
.404
do
S
Figure 1-5: Geology map of the Kry'suvi'k area in our study. K: Lake Kleifarvatn; KV
(orange dot): Krfsuvik Volcano.
Chapter 2 presents an overview and analysis of the seismic data, including the
theory we used, the checkerboard test and also the tomography results. Chapter 3 is a
discussion of the surface deformation observation and modeling. Finally, chapter 4
contains a discussion of the seismic and deformation results.
18
Chapter 2
Seismic Event Relocation and
Tomography
2.1 Seismic Data and Stations
The seismic data used in this study were recorded by the SIL network and
temporary seismic stations deployed by Massachusetts Institute of Technology (MIT),
Reykjavik University (RU) and Uppsala University (UU). In 2010, MIT, RU and UU
deployed 14, 8, and 16 stations, respectively, on the Rekjanes Peninsula (RP) to study the
geothermal areas (fig.2- 1 a). Each station has a three-component broadband seismometer
and a GPS synchronized clock. The seismic data are digitally recorded at a sample rate of
100 Hz. In our study, 23 seismic stations (5 from SIL, 5 from MIT, 5 from UU and 8
from RU) are used for the seismic event relocation and tomography in the Krfsuvik area
(fig.2-1b).
Figure 2-2 shows the cumulative number of earthquakes and distribution of
earthquake magnitudes from 1st July 2010 to 31st December 2011 of the SIL catalog.
One pronounced peak in activity at Krysuvik was observed in Feb. 2011 and one smaller
peak around Aug. 2011 (fig.2-2a). About 3853 seismic events recorded from Jul. 2010 to
Dec. 2011 by the SIL seismic network around Krf'suvik were selected for analysis. The
magnitude of most events is in the range of 0-2, which are called microearthquakes
(fig.2-2b).
19
/
YIMO station
V UU station
V MIT tation
V
U sltatN
b
64
ash
63.9
~ A
63.85-grv
svh
~
-22.6
-22.4
-22.2
-22
-21.8
-21.6
Longitude (degree)
Figure 2-1: a. Seismic station distribution on the Reyjanes Peninsula from 2010 to 2011
(courtesy of Oli Gudmundsson). b. Seismic stations we used in this study. The dashed
black line indicates our study area. The orange dot is the location of the Krfsuvik
volcano. The small black dots are the earthquakes from the SIL catalog.
4000
1000
a
800f
3000
6001
E 2000:
Cz
*400
N 1000
200f
100
200
300
days
400
500
0
1
2
Magnitude
3
4
Figure 2-2: a. Cumulative number of events from I't July 2010 to 3lst December 2011
and the blue circle indicates the biggest jump that occurred on 2 7 February 2011; b.
Histogram of earthquake magnitudes from 1 't July 2010 to 3 1 ' December 2011.
20
2.2 Theory of Event Relocation and Tomography
We applied the Double Difference Tomography (Zhang et al., 2009; Zhang and
Thurber, 2003) to simultaneously relocate the seismic events and construct travel time
tomography (Vp, Vs and Vp/Vs). Based on seismic ray theory, the arrival time T of a
body wave from an earthquake i to a seismic station k is written as a path integral,
k
sdI,
T, =t +f
S
- 1)s dl,
=
-T
(1)
(2)
where t' is the original time of the event i , s is the slowness vector and dl is the element
length of the integral path. A first order truncated Taylor series is used to linearize the
highly non-linear relationship in equation (1,2). For equation (2), S and P waves from the
same events are with the same origin times, thus the unknown origin times are removed
from this equation. The misfit between the observed (To ) and predicted (T cal) arrival
time rk can be expressed by the perturbations of the hypocenter parameters (
, dt') and
slowness parameters (ds):
"
-
ca
d
=
+ dt
ds d.
(3)
n=1 14
The double difference, the difference of residuals between the observed and theoretical
travel times in linked event pairs (ij)
at one station (k), drk , are used in the DD
earthquake location algorithm (Waldhauser and Ellsworth, 2000), which could be written
as:
dr
=
- =rk-k=(T-T=T-
21
obs~(Tk -Tlj)cal.
-
(4
The observed travel time difference
(T
-
Tk)
can be calculated by both the
waveform cross-correlation for similar waveforms and absolute catalog arrival times.
A pseudo-bending ray tracing algorithm (Um and Thurber, 1987), which finds the ray
paths and calculates the travel times, is applied to update the velocity models. In
order to calculate the relative, absolute event locations and the subsurface velocity
structures, the absolute arrival times, the waveform cross-correlation data, and the
catalogue differential arrival times are used in the inversion. For the Vp/Vs ratio in
equation (2), in order to avoid the problems that P and S have different raypaths and
Vs structure can not be inferred reliably from Vp and Vp/Vs values, we check P and S
raypaths and remove the corresponding S-P times from the Vp/Vs inversion if the
paths differ by more than a specified threshold. The complete inversion system is
solved by the LSQR algorithm for sparse linear equations and least-squares problems
(Paige and Saunders, 1982).
2.3 Results of Event Relocation and Tomography
2.3.1 Travel times used in this study
Due to the small magnitude of most events, P and S arrival time picking become the
most difficult but important part in our seismic study. Two rounds of arrival time picking
were performed. In the first round, we automatically picked the P arrivals using the ARAIC method (Sleeman and van Eck, 1999) and then applied it to identify the S arrivals in
a window after the P arrivals. In the second round, we manually corrected the P and S
automatic picks by visually inspecting the three-component recordings. Thus we ensured
the high quality of the final event catalog. For the P arrivals, the arrival times on vertical
components are selected for their clear onset. For the S arrivals, the arrival times on east
or north components, whichever are more visible, are selected. Therefore, shear wave
splitting is not taken into account. We used two kinds of differential travel times in our
22
calculation. One is catalog differential time, which is obtained by directly subtracting the
corresponding catalog travel time of the linked neighbor (Waldhauser, 2001). The other
one is cross-correlation differential time, which is obtained by cross-correlation between
two waves in our specified threshold. In summary, we have totally 37,459 and 35,493
absolute catalog P and S arrives, 628,764 and 511,162 pairs of catalog differential P and
S arrives, 7,485,607 and 9,559,917 pairs of cross-correlation differential P and S arrives,
32,925 absolute catalog S-P times, 451,370 pairs of catalog differential S-P times, and
2,280,581 pairs of cross-correlation differential S-P times.
2.3.2 Event Relocation Results
We analyze the residuals between the calculated and measured travel times to
evaluate the convergence of inversion. The root-mean-square (RMS) residuals after the
first, eleventh and twelfth round iterations for absolute P, S, catalog differential P, S and
cross-correlation P, S are listed in Table 2-1. From the RMS of the residuals, we could
find that the inversion converged till the 12th round iteration. Figure 2-3 shows the
histograms of the corresponding residuals after 12 iterations. The residuals of S spread
more widely than P, which might be contributed by the anisotropy in this area.
Figure 2-4 and 2-5 show the catalogue event distribution and the relocation result,
respectively. From the map view of the two results, we find that our relocation result
better delineates the N-S trending and NE-SW trending faults on the surface. In the SIL
catalog, the event depths are widely scattered in a range of 2-10 km. However, in our
relocation result, most earthquakes are concentrated at the depth of 1-6 km and are more
clustered. From the plot of the cross section Al-B1 in the relocation result, we find that
the NE-SW trending fault is deeper than the N-S trending faults. Meanwhile, the average
23
event depth after relocation is about 2 km shallower than the catalog events. Systematic
shifts in event depths might be caused by the difference in the velocity models. In our
calculation, the velocity model is updated.
Table 2-1: List of the RMS of residuals.
RMS (s)
Absolute
P
Absolute
S
Catalog
differential
Catalog
differential
Crosscorrelation
Crosscorrelation
P
S
P
S
1"t round
0.107
0.172
0.086
0.119
0.087
0.135
round
0.038
0.061
0.029
0.045
0.008
0.011
121h round
0.038
0.061
0.029
0.045
0.008
0.010
1 1 th
Absolute P
8000
Absolute S
4000
6000[
3000
~4000f*f
E
I
1000
2
6-X 104
-0.1
0
residuals (s)
01
.82y
02
residuals (s)
Catalog differential P
I
4.x
&
01 02
Catalog differential S
51
3.
4.
2
82
-01
0
residuals (a)
01
2
02
-01
0
01
residuals (s)
Cross-coalaion P
x
.A
Cross-correlation
5
02
S
61
4
3;
4
1
2
0
5
0
residuals (s)
0065
01
481
-005
0
00 5
0.1
residuals (s)
Figure 2-3: The residual histograms of Absolute P, S, Catalog differential P, S, and
Cross-correlation P, S.
24
63.9 6
C%4
Al
B1
I:
63.9
A.B1_
-22.19
-21.87
-22.03
*W*
*
*
o
*
t
0
2
E
Al
-11
4
o
-
.
-
65
10
5 km
1
-22.19
.
-22.03
Longftute (degree)
-21.87
Figure 2-4: Earthquake events distribution in Krysuvik area from Catalogue.
63.96
BA
.
As6-
A1 -- 63.91
1
-E
--
km
63.86
06
WI
A2
-22.19
-22.03
-21.87
Al
-0
(w) WdeQ
B1
2
-
41N-S trendihg
fault
8t
ted6
5km
1
-22.19
5
-
If-trending.
kmfault
-22.03
-21.87
Longitute (degree)
Figure 2-5: Earthquake events distribution in Krysuvik area after relocation. The arrows
on the top figure indicate the directions of N-S trending fault and NE-trending fault.
25
2.3.3 Checkerboard Test
The starting model for the inversion is a one-dimensional model (Table 2-2)
(Tryggvason et al., 2002). The X nodes, from west to east, are -50 km, -16 km, -8 kin, -4
km, -2km, 0 km, 2 km, 4 km, 8 km, 16 km, 50 km; The Y nodes, from south to north, are
-40 km, -6 km, -2 km, 0 km, 2 kn, 6 kin, 40 km; The Z nodes, in depth, are -100 km, 0
km, 2 km, 4 km, 6 km, 9 km, 12 km, 16 km, 21 km, 32 km, 200 km. The reference
location (0, 0) is (63.91 0 N, 22.030 W).
Table 2-2: Initial model (one-dimensional velocity model)
Depth (km)
-100
0
2
4
6
9
12
16
21
32
200
P velocity
3.4
3.6
5.6
6.4
6.6
6.7
6.8
7.0
7.1
7.4
8.0
1.78
1.78
1.78
1.78
1.78
1.78
1.78
1.78
1.78
1.78
1.78
(km/s)
Vp/Vs ratio
In order to quantitatively understand our 3D tomography results, we evaluate the
resolution of our tomography model using the acquisition geometry through a
checkerboard test. The synthetic velocity model includes a 3D checkerboard pattern of
positive and negative 5% perturbations in P velocity and negative and positive 5%
perturbations in S velocity relative to the 1D velocity model. Therefore the constructed
synthetic model of Vp/Vs is on the order of positive and negative 10% perturbations.
Synthetic travel times are computed in the perturbed model for the relocated hypocenters
and station distribution.
quantitatively,
In order to assess the recovered checkerboard models
we calculated the semblance between the exact
26
and recovered
checkerboard anomalies using a circular operator in the (x, y)-plane with a radius of 5 km
centered on each node (Zelt, 1998).
R=
Z'=(ti + nri)2
i=
(5)
2 EiM(t.2 + nrt2)'
where tj and ri are the true and recovered velocity anomalies at the ith node inside the
circle consisting of M nodes. R-values, the resolvability, are in the range of 0.1-1.0. The
results are considered resolvable when values are larger than 0.7. R-values are used to
evaluate the resolution for the following studies of the slices of 3D tomography results.
63.96
./
kmt=63.91
-63 .91
e)
-
-j
on3AA
/ ,4km/
-22.19
-22.03
-21.87
Longitute (degree)
Figure 2-6: Map view and cross sections presented in the following checkerboard test and
tomography results.
27
Deph-2 km
63.96
Deplhv4 km
63.96
63911
Depth-6 km
6396--
63.91
Ph5
83.91
83.86
-2219
-5
83.96
VP(%)
83.98
8391w
G3.91
63.91
83.86
-5
83.88
63.96
VP(%)
M96
83.96
03.91
e3.91
06
63.86
22.3
Longlute Idegre.)
-21.
83
Longiout. Idegre )
Longnute (dsgr..)
Figure 2-7: Checkerboard test of Vp for the map view in figure 2-6. Figures of the top
row are the input model; Figures of the middle row are the corresponding recovered
model using our acquisition geometry; Figures of the last row are the corresponding
resolvability distribution.
Lattudo=63.91 (dogr..)
Latktud0=63.93(dgr..)
10
Vp(%)
10
Vp(%)
0
10
1C
resolvabily
06
P
10
-22.19
-22.03
Longitute (dogre)
10
-21.87
-22.19
-22.03
Longitute (dsgrW)
-21.87
.5
0.
Figure 2-8: Checkerboard test of Vp for the cross sections in figure 2-6. Figures have the
same meanings as in figure 2-8.
28
Depth=4 km
Depth=2 km
Deoth-6km
Vs(%)
83.9f
5
En
83.91
I51
21 0
03.
-21.87
-2203
-2219
63.96
VS(%)
83.96
63.91
83.91
63.96 -22.19
63.86
M3
63.96
63.98
83.98
63.91
83.91
e3.91;
-2203
-21.87
-21.87
resolvabs~y
.7
0.5
-22.19
-2203
Longtu* (degree)
-21.87
-cr- I
-Cr-%A
-r. .O
'~~-2219
LoVgNute (degee)
-22.03
Long"ute (degree)
-21.87
04
Figure 2-9: Checkerboard test of Vs for the map view in figure 2-6. Figures have the
same meanings as in figure 2-8.
Latkude=63.91 (deg")
-22.19
-22.03
Lattude=63.93(degree)
Vs(%)
-21.87
Vs(%)
10
resovabily
0
0.7
S0.80.6
0.5
10
-22.19
-22.03
Longitute (degre.)
-21.87
10
-22.19
-22.03
Longitute (degree)
-21.87
04
Figure 2-10: Checkerboard test of Vs for the cross sections in figure 2-6. Figures have the
same meanings as in figure 2-8.
29
Depth.2 km
Depth=4 km
Depth-6 km
83.9f
10
5
- 3.91
0
.5
-22.
-2.3-21.07
11-10
39G
10
5
0
-5
8e
-21.87
-22.03
63.
m94
e3.96
e3.91
83.91
83.91
-2219
-22.03
Longotme (dgre)
-21.
-10
-22.19
-22.03
Longstut (dsgre.)
-21.87
-2219
Z03
..
-21.87
Longtikui fdsgrs.)
Figure 2-11: Checkerboard test of Vp/Vs for the map view in figure 2-6. Figures have the
same meanings as in figure 2-8.
Latltude=63.93(degree)
LatItude=63.91 (degree)
VpIVS(%)
10
101
10
-;L.1
.
-22.19
-22.03
-21.87
VpIVs(%)
5
Ic
10
10
resolvablity
0.8
0.7
0.6
0.5
-22.19
-22.03
Longfitu* (degree)
-21.87
10
-22.19
-22.03
Longitute (degre.)
-21.87
04
Figure 2-12: Checkerboard test of Vp/Vs for the cross sections in figure 2-6. Figures have
the same meanings as in figure 2-8.
30
Figure 2-6 is the map view and cross sections of our study area in this thesis and
it is used in the following checkerboard test and tomography results. Figure 27,8,9,10,11,12 show the map view and cross profiles of the checkerboard test results of
Vp, Vs and Vp/Vs. The recovered checkerboard and resolvability values show that the
Vp has resolution down to a depth of ~7 km on both the horizontal and vertical profiles.
The Vs models are resolved well at the depths of 2 and 4 km and lose resolution at the
depth of ~ 6km for the horizontal checkerboard test, but have resolution down to a depth
of ~6 km. While the Vp/Vs ratio almost loses resolution at ~6 km on both horizontal and
vertical profiles. From the analysis of the resolution test, we find that our results have
relatively better resolution of vertical profiles than horizontal profiles. This phenomenon
can be explained from the ray coverage. From figure 2-5, we find that the earthquakes are
concentrated within a depth range of 1-6 km. Thus the limitation of the depth of
earthquakes caused that at the depth of ~6 km, the cross rays in vertical directions are
more than that in horizontal directions.
2.3.4 Tomography Results
Figure 2-13 presents horizontal slices of the tomography results at various depths.
In general, the resolvability tests validated our results. Clearly higher Vp and Vs are
observed beneath Krf'suvik volcano (gray dot) and beneath Lake Kleifarvatn, but there is
a region of reduced Vp and Vs between them at the depth of 2 km. These are consistent
with the geology study in area. The rocks between the Krfsuvik volcano and Lake
Kleifarvatn are hyaloclastite, while around Krfisuvik volcano they are basaltic lavas. The
relatively higher Vp and Vs of the Lake Kleifarvatn might be related with its formation.
Lake Kleifarvatn is an in-land lake, no connections with other rivers or lakes. Thus it
31
might be a crater lake, which was formed when water filled the basin left by the collapse
of big magma chamber without sufficient support for its roof after eruption. Therefor a
remnant magma chamber might be the possible reason for the higher Vp and Vs. Lower
Vp/Vs ratio has been obtained in the lake and around the southwest part of the lake. After
the big earthquake in
17 th
June 2000, a lot of cracks and fissures under Lake Kleifarvatn
have been created and 12% of by volume of the lake drained down (Clifton et al., 2003).
While because of the high permeability in this region, circulation of water with steam
reduces the Vp/Vs ratio slightly. On the surface, hot springs are observed at the
corresponding area (Mawejje, 2007). At the depth of 6 km, a low Vp anomaly is detected
within the resolvability range, while low Vs and high Vp/Vs anomalies are lost
resolvability on the horizontal map view.
On the cross sections of our study area (fig.2-14), a low Vp velocity zone at a
depth of about 6 km is detected within the resolvability range. Most part of the low Vs
velocity zone is in the resolvability range. While the high Vp/Vs ratio zone is slightly out
of, but very close to the resolvability range. Figure 2-15 shows a 3D view of the low Vp
(-5.75 km/s), Vs (~3.3 km/s) and possible high Vp/Vs ratio (-1.89) anomaly isosurfaces
along with the locations of earthquakes occurred in the Krfsuvik area. The low Vp
anomaly is directly beneath the earthquake swarm. Most of the microearthquake events
occurred above or around this low P velocity zone. The low S velocity and high Vp/Vs
ratio isosurface bodies are slightly toward the west. The decreased resolvability of Vs and
Vp/Vs ratio at the depth of -6 km and the relatively worse RMS (fig.2-3) caused by
anisotropy and noise of the S onset might have caused the slight shift of the Vs and
Vp/Vs ratio isosurface bodies from the center of earthquake swarm.
32
This low Vp and Vs velocity zone was also reported by Geoffroy (2008), which
was interpreted as supercritical water because of the low Vp/Vs ratio he obtained by
dividing Vp by Vs. However, when directly dividing Vp by Vs, the uncertainties of
Vp/Vs ratio will increase because of both the inversion uncertainties of Vp and Vs. For
example, if the uncertainties of both Vp and Vs are +5%, the uncertainties in Vp/Vs ratio
will reach ~10%, which is crucial to interpret the change of Vp/Vs ratio.
tomography, we invert the Vp/Vs ratio directly.
In our
At the depth of ~6 kin, our Vp/Vs
inversion reveals a possible high Vp/Vs ratio body, but its exact extension is not well
resolved because the lack of deeper earthquakes. Meanwhile, the 3D Magneto Telluric
(MT) survey also obtained a low resistivity zone at the same location of our low Vp zone
(Hersir et al., 2013) (fig.2-16). This high conductivity and low velocity zone could be
caused by the filling of water in cracks (supercritical water) or partial melting. If it is
interpreted as partial melting, the rock matrix will change and Vs will decrease much
more than Vp. Thus we could expect a high Vp/Vs ratio, which is consistent with the
possible Vp/Vs anomaly detected in our tomography. The detailed discussion of the low
Vp and possible high Vp/Vs ratio around 6 km is addressed in Chapter 4.
33
Deeth.2 km
Vs(kmis)
e3-96
Depth.2 km
3.9W
S2
8
5
6
4
63.91
63.91
Vplyi
5
Depth.4 km
Dopth.4 km
vp(kwMs)
Deih-4 km
2
63.91
83.91
Vs(kwih
it
8
D*Dthu6 km
83.9
Ii
VD(kNVs)
iskm,
DepthuO km
Depth-6 km
83
2
(
63.91
4
63.91
1.8
683
63.86
Longoul*0(d69-1
LongouI
idegrw)
LongdM*
(dso")
Figure 2-13: Horizontal profiles of Vp, Vs and Vp/Vs ratio models at depths of 2,4, and 6
km. The black solid line indicates the resolvability threshold of 0.7. The black dashed
lines are the other resolvability values.
L&IM
63.93(deg")
Vp(kmws)
VsfkmIs)
LaMUdO=63.93(dsar..1
~
LatIude.63.93(degs,*)
2
S
VP/VS
10
.a:.udA4391(sgm)
ZI
10
LaUd*S3.91(dSgeej)
VP(kmf/s)
i
-22.19
-22.03
LongfItw (do"gr@I
-21.57
.6IV
-22.03
VS(km/3)
Lautlude=6391(dogt.)
0
10
9.0 -2
-21.07
Longkuw (dog
)
10
--.
IWV
-".%"
*Z1.O
Longitui. (degre)
Figure 2-14: Cross section profiles of Vp, Vs and Vp/Vs ratio models. The black solid
line indicates the resolvability threshold of 0.7. The black dashed lines are the other
resolvability values.
34
0
Isoratio: Vp/Vs=1.89
Isovelocity: Vs=3.3 km/s
Isovelocity: Vp=5.75 km/s
-
2i
2
2
-
4
-4~
4
53.__
~
YLk
L0a,9lI
0388-2203
22,19
(degrool
1.8
Langule (dP
)
6391
63.8
NdW-)
22
07
-21.7
22.03
19
Lat
Ider
.9
I
86
-22
03
19
2202
-21,87
2187
Figure 2-15: The isosurfaces of the Vp, Vs and Vp/Vs ratio. The regions where Vp < 5.75
km/s, Vs < 3.3 km/s and Vp/Vs > 1.89 are shown.
1000
7095
-
100
l}70
E
7090
30
7085
50
Initial model
7060
440
445
450
455
Figure 2-16: Resistivity at the depth of 5 km (Hersir et al., 2013). A low resistivity body
was observed at the similar location with the low velocity zone discovered in this study.
35
Chapter 3
Surface Deformation Observation
and Modeling
3.1 Geodetic Observation
Further support of the low velocity zone in the tomography result and the
relationship between the seismicity and the tectonics of the Krfsuvik region can be
obtained by evaluating the surface deformation observed in the region.
We found publicly-available GPS and InSAR analyses showing the ongoing
deformation in the region. The geodetic measurements between 2009 and 2011 show
significant uplift and time-variable expansion in the Krfsuvik area (fig.3-1,2).
The
Global Positioning System (GPS) data recorded by the permanent GPS stations (MOHA,
KRIV) and semi continuous GPS stations (LAMB, SELC), in the Krfsuvik area recorded
20-30mm uplift around Feb. 2011 (Appendix B). This area also exhibits expansion to
some extent around the southwest corner of Lake Kleifarvatn (Michalczewska, 2012).
Meanwhile, the InSAR measurements in this region also detected uplift and expansion
(Michalczewska, 2012).
Figure 3-2 shows the average unwrapped TerraSAR-X
interferograms spanning from 2009 to 2011. The main deformation is observed around
the southwest corner of Lake Kleifarvatn and a secondary deformation, a dislocated
deformation of northwestern part of the main deformation, is enclosed within the dashed
lines in fgiure 3-2. The deformation shows displacements up to 30 mm in the Line-OfSight (LOS) direction, although we do not know whether the orbit is ascending or
36
descending. The inclination of TerraSAR-X is 97.440. The ascending (Jul.18, 2010 -
Jun.14, 2011) and descending (Jul.24, 2010 - Jun.08, 2011) vectors of the look angles are
(-0.5664, -0.1259, 0.8145) and (0.6422,
-0.1276, 0.7559) (east, north, vertical)
respectively.
I- ~ J
'-
6~44000-
0
5
Figure 3-1: Displacement of GPS measurements in Krysuvik area. The grey circles
indicate the uncertainties of each measurement (Michalczewska, 2012). Different color
arrows indicate measurements in different periods.
64'
E2!
*21
17
11
-~11
S
-22',
Figure 3-2: The mean deformation in the InSAR Line of Sight direction from 2009 to
2011. The solid black lines indicate the volcanic systems (Michalczewska, 2012). The
dashed black line reveals the secondary deformaton (dislocated deformation of
northwestern part of the main deformation).
37
In order to determine the orbit of the InSAR measurements, we project the threecomponent measurements of the GPS measurements (measured from figures in Appendix
B) of four GPS stations using equation (6) to both the ascending and descending look
angles, respectively, of the InSAR in the Krfsuvik area.
dl,, = dh sin(G) + d, cos(G)
(6)
where G is the look angle and it could be obtained from the vectors of the look angle as
G = atan ((e
2
+ n 2 )/z) (e, n and z are the unit vectors in east, north and vertical
directions), dl, 5 is the displacement projected on the look angle of the satellite, dh and d,
are the horizontal and vertical displacement, respectively. d, can be obtained directly
from the GPS measurements, and df could be calculated using the east and north
measurements, de and dn.
dh = de cos(D) + dn sin(D)
(7)
where D is the cross-angle between the north and the LOS surface projection vector. We
use the cosine and sine rules on a sphere (fig.3-3) to find this cross-angle.
Let a, b. and d be the angular distances between 3 points and A. B. and D the
angles as indicated in Figure
The cosine rule for a and A is:
cosa = cos bcosd + sinbsindcosA
and the sine rule:
sin a
sin A
sin b
sinB
sin d
sin!)
Figure 3-3: The algorithm of the cosine and sine rule on a sphere (Essentials of
Geophysics, MIT).
38
For our problem, compared with the satellite altitude, the Krfsuvik area is
considerably small, so we use the central latitude of this area, 63.920, as the average
angle we used in cosine and sine rules. d = 97.440 - 900, b = 900 - 63.92 0 (if the
latitude of the point we want to calculate is 63.920), A = 900, then we get D = 160
(fig.3-3). Therefore we use 160 as the cross angle between the north and the LOS surface
projection vector in the area we modeled in next section.
After projecting the GPS measurements onto both the ascending and descending
orbits of TerraSAR-X using equation (6), we find that the result presented in Figure 3-4
that is obtained using the descending orbit of TerraSAR-X is fairly consistent with the
InSAR observation in Figure 3-2. Therefore in the following simulation, we project our
simulated surface displacements to the descending look angle of TerraSAR-X.
I
63.96
/7
rnr n
. .
2()
63.94
~63.92
110
63.9
MSELC
4-J
63.88
*1'
/
r-3n86
/
0
WRIV
/__
-10
-22.15 -22.1 -22.05 -22 -21.95 -21.9
longitude (degree,
Figure 3-4: Projection of the GPS measurements on the descending look angle of
TerraSAR-X.
39
3.2 Surface Deformation Modeling
We used constraints from our seismic velocity model, which includes the NE-SW
and N-S faults and a low velocity zone, to model the observed deformation. We used the
Coulomb 3 software (U.S. Geological Survey) to set up models and calculate the surface
displacement in a half space using various sources of deformation.
We start our deformation modeling by assuming that there is an isotropic volume
source buried at a depth of d where the low velocity zone is located. While there is
generally a trade off between deformation source depth and the volume of the source for
generating a certain deformation, we remove this ambiguity by constraining the main
source of deformation within the low velocity zone identified by the seismic inversion at
the depth of 6 km. For our deformation modeling, we place an inflation source at the
location of x=l km, y=0 and z=6 km (fig.3-5al, bl) and we choose a reference point (6,
8) km on the surface (Michalczewska, 2012), where the deformation is set to be zero.
When we use a volume change of 30 x 106
iM3 ,
the surface displacement projected on the
LOS vector (fig.3-5cl) can be compared with the observed result plotted in figure 3-2.
The largest uplift and subsidence in our simulation are around 35 mm and -15 mm
respectively. While the largest uplift and subsidence obtained from InSAR are 27 mm
and -13 mm respectively, which are in approximate agreement with our model.
Comparison of Figure 3-5cl with Figure 3-2 shows that the deformation predicted
using a volume source centered at a depth of 6 km fits the observed pattern of InSAR
deformation rather well. However, detailed comparison shows that the northwestern part
of the observed deformation extends further to the northwest (dashed black line in fig.32) and this extension is not captured by only using the inflation source. To further
40
investigate this northwest extension of the deformation, we add a dip-slip fault at the
depth from 3 km to 6 km, dip angle of ~ 4 0 'and the length of - 10 km, as delineated
from our seismic event relocations (fig.2-5). Input parameters of the fault in our model
are set to be 400 of the dip angle, (-4, -4) km and (4, 4) km for the starting and ending
points projected onto the surface and 3 and 6 km of the starting and ending points of
depth (fig.3-5a2, b2). After testing the normal and thrust faults, we find that a normal
fault with slip value of 0.09 m can explain the dislocation of the northwest part from the
main area (fig.3-5c2). Using equation MO = MDA (y is rigidity, D is displacement and A
is the area of the fault), we obtainM0
Li
3
cm( )kidyne-cmi
=
1.5x 1024 dyne - cm, and M" = 5.4 (M, =
10.7 (Hanks and Kanamori,
1979)) from the above
input
nu
parameters.
However, compared with the deformation obtained from InSAR, the simulated
surface deformation of the northwestern part in the northeast direction is smaller than the
observed. To address this discrepancy, our next test is including the strike-slip faults we
observed in our relocation results. To simplify our model, we included one strike-slip
fault at depth from 0.5 to 3 km and length of -8 km delineated from our seismic
relocation results (fig.2-5). Therefore some input parameters of the fault in our model are
set to be (0, -4) km and (0.1, 4) km of the starting and ending points projected on the
surface and 0.5 and 3 km of the starting and ending points of depth (fig.3-5a3, b3). After
testing different parameters of this strike-slip fault, we find that a right lateral strike-slip
fault with slip value of 0.09 m can well explain the northeast dislocation of the
northwestern part from the main area. Using the input parameters, we obtain MO =
5.8x 1023 dyne - cm, and M, = 5.1.
After adding motion along both the normal and
41
strike-slip faults onto the inflation model, the modeled result matches well with both the
major as well as the secondary deformation in the InSAR observation (fig.3-5c3). The
largest values of the uplift and subsidence on both the simulated deformation are about 27
mm and -13 mm, respectively, which is in reasonable agreement with the observation.
The total moment release in our geodetic model is about 2.08x 10 2 4 dyne -cm
and
the
MO=
10
total
(
+10.7)
recorded
seismic
moment
release
estimated
using
equation
(Hanks and Kanamori, 1979) for the recorded moment magnitudes by
SIL is about 2.3x1022 dyne - cm (fig.3-6). This discrepancy might be caused by the
aseismicity in this region.
We successfully modeled the observed InSAR deformation by considering the
low velocity zone as an expanding source and incorporating the normal and right lateral
strike-slip faults (fig.3-7). This expanding source will cause local pressure change in this
region. Thus the formation of the earthquakes above and around this expanding source
might be related to this local pressure change caused by the expanding source.
An increased pressure of an expanding source could lead to a local stress field
above the source whose maximum stress is vertical and compressive and horizontal
stresses are extensional. We consider this stress field caused by the expanding source as
local stress. Such a stress field would favor the formation of normal faults.
The combination of the left-lateral and expansion tectonic stress field of SW
Iceland is considered as regional stress. In such stress field, right-lateral strike-slip faults
as well as normal faults can easily occur. According to the previous seismic and
geological observation, right-lateral strike-slip faults already existed in this area (Clifton
and Kattenhorn, 2006; Einarsson, 1991).
42
IIS
mm
c1
bi
al
10
30
.5
01
1 0
.5
10
0
-10
.20
10
Y (km)
-15
-10
0
X (Wm)
10
0
X (Smn)
*10
10
X(km)
15
c2
b2
a2
10
/
nm
15
/o
/
-
10
E-io.
*10
-is-
1 0
Y (km)
-10
0
X (km)
10
-10
to
X (kml)
10
X(km)
c3
b3
1a3
1!
10
2t
I
10
5.10
-20
0
.10
10
D1
10
N -15
-10
(m)
0
Y (km)
x10
0
X (SM)
-1(
-10
-10
0
10
0X(km)
Figure 3-5: Models and results simulated as the InSAR observation. Boxes al-3 are map
views of the deformation source geometry, and the red dot is the reference point where
deformation is taken to be 0. The numbers indicate the sources, 1 (black dot) for inflation
source, 2 for normal fault and 3 for strike-slip fault. The green lines indicate the strike
directions of the faults and the red box represents the fault plane for the normal fault.
Plots bl-3 are the 3D views of the corresponding map views. The cl-3 are the calculated
displacement results projected on the LOS of TerraSAR-X (descending). The silver boxes
indicate the corresponding area of the InSAR observation.
2.5x 1022
2
E.
1.5[
C:
1
0.5[
100
200
300
days
400
500
Figure 3-6: Seismic moment release starting from 1st July 2010 to 31st December
2011. The biggest jump occurred on 27th February 2011.
43
The NE-SW trending dip faults delineated by seismic relocation, interpreted as
normal faults in the surface deformation modeling, and the N-S trending strike-slip faults,
interpreted as right lateral strike-slip fault, are consistent with the combination of the
regional and local stress. Therefore, the formation of the NE-SW trending normal faults
and the N-S trending right lateral strike-slip faults could be explained as two possibilities:
1). The faults are pre-existing faults, and slip is triggered by a combination of the
regional and local stress field. 2). The combination of the regional and local stress field
generated these faults. A more detailed analysis of the change of strain and stress
(including both the plate spreading and local expanding) in Krfsuvik geothermal field is
needed to further confirm the cause of the seismicity.
mn 0
64.00
6400
20
0
10
0
635
to
10
-22.26
-22.05
-21.85
Longitute (degree)
Figure 3-7: Comparison between the observed InSAR deformation and the simulated
deformation. The location of the expanding source mapped on the right picture is
(63.91 0 N, 22.050W).
44
Chapter 4
Discussion
4.1 Brittle-ductile Transition
From the 3D view of the low Vp zone and distribution of the earthquake events, we
observe that all of the earthquakes have occurred right above or around this low velocity
zone (fig.4-1). The occurrence of the earthquakes indicated the local materials to be
brittle. The lack of earthquakes in this low velocity zone probably reveals relatively more
ductile materials at the depth of about 6 km. According to the distribution of earthquakes
we plot a brittle-ductile transition (BDT) boundary in our study area, above which 99%
of the earthquakes occurred (fig. 4-1). The depth of BDT boundaries varies in a range of
4-6 km from west to east. The transition from brittle to ductile deformation is most
strongly dependent on lithology and temperature. The depth of the BDT can be
considered as a first order isotherm (Foulger, 1995). On the surface of our study area,
50 0C temperature at the depth of ~12 cm has been measured and high temperature
recorded in well KR-2 in Krfsuvik geothermal field is 250 0C at the depth of -1.2 km
(Mawejje, 2007). The non-glassy basalts may deform in the brittle field up to 5500 C +
100 0 C and in ductile behavior between 8000C and 950 0 C and the state in between is
semi-brittle (Violay et al., 2012). Based on the above temperatures and the depths of
BDT and ductile behavior, we can estimate the geothermal profile in this area (fig.4-2).
The geothermal gradient is about 2000 C/km between 0 and 1 km. While at the depths
deeper than 1 km, based on the different BDT boundaries, three geothermal profiles are
45
determined. To the east area of the earthquake swarm, next to Lake Kleifarvatn, the
depth of the BDT is 6 km where the temperature is around 5500 C. Moving to the west,
further away from Lake Kleifarvatn, the depth of the BDT, where the temperature is
5500 C, becomes shallower, 5 to 4 km. Below BDT, the materials are semi-brittle or
ductile. Due to lack of water convection, the thermal gradient would be higher than that
from 1 km to BDT boundary (dashed line in fig.4-2). The closer to Lake Kleifarvatn, the
smaller of the thermal gradient is (fig.4-2). This observation implies stronger convection
near Lake Kleifarvatn, which can be explained by the high fracture density and high
permeability near Lake Kleifarvatn as discussed in previous literatures (Clifton et al.,
2003; Geoffroy and Dorbath, 2008). The rocks with higher permeability and stronger
convection usually have lower Vp/Vs raio, which is consistent with our tomography
result near Lake Kleifarvatn at the depth of 2-4 km (fig.2-13).
46
63.96
-K
63.91
. ..
4-0'.
km
63.86
0
..4
6
8
-22.19
-22.03
Longitute (degree)
-21.87
Figure 4-1: Longitude view of 3D plot with Vp (5.75 km/s) isosurface body and
earthquakes in our study area. The blue lines indicate the brittle-ductile transition. Green,
blue and red lines indicate BDT at 4, 5 and 6 km, respectively. K: Lake Kleifarvatn.
0
2
E
8
0
200
400
600
800
Temperature (degree Celsius)
Figure 4-2: Temperature profiles. Red, green and blue lines indicate temperature of
550 0 C at 4, 5 and 6 kin, respectively.
47
4.2 Possibilities of the low velocity zone
Seismic velocity varies depending on several physical conditions, such as
composition, saturation condition, temperature, and pressure. Composition might play
the fundamental role among the others. Meanwhile, changes of status of the rocks, such
as solid versus partial melting, would also influence the velocities. The seismic velocity
anomalies caused by pressure and temperature variations at the same depth are
considerably smaller compared with the effect of saturation of fluid or melting conditions
of inclusions in the rock matrix at the same depth (Nakajima et al., 2001). Through the
thermal expansion study (AT =
A,
a: thermal volume-expansion coefficient, AV, AT:
volume and temperature change, V: Volume), we find that the temperature change in our
low velocity zone is 45-100 0 C given the volume change of 30 x 106 m 3 from our
surface deformation, the volume of the low velocity zone of 20 x10 m 3 from our
seismic tomography result and a of 15 - 33x 10-6 /C
in most rocks. Then we obtain
the change of the Vp of 0.025-0.057 km/s using the relation of AVp/AT -0.57x
10-3 km/(s'C) (Christensen, 1979). This result is significantly smaller than ~1 km/s
obtained in seismic tomography. Therefore the change of the velocity might be caused by
the effect of saturation of fluid or partial melting in rocks. Here we propose 2 possibilities
for the low Vp in our study: 1) partial melting; 2) supercritical water.
1). Partial melting: Based on the low Vp, Vs and possible high Vp/Vs in our
tomography result, partial melting could be a reasonable explanation, considering the
H2 0-rich condition in our study area. For basaltic rock, partial melting usually happened
at temperature -1200
0 C.
However, when mixed with water, the partial melting
temperature would decrease (fig.4-3) (Grove et al., 2006; Liu et al., 1996). If the
48
temperature reaches -800'C at depth ~ 6 km (fig.4-2) and mixes with 1-2 wt% (volume,
3-6%) water, basalt melting could occur. Given the volume of low velocity anomaly
-20 km 3 obtained from the tomography, the volume of water required would be
0.6-1.2 km 3 .
After the big earthquake in 17th June 2000, 12% of the volume
(-0.1 km 3 ) of Lake Kleifarvatn disappeared (Clifton et al., 2003). The required 0.6 km 3
of water injection is in reasonable range of the water loss if considering the contribution
of Lake Kleifarvatn, ground water and precipitation. According to our possible Vp/Vs
ratio of -1.89 (black star in figure 4-4), the volume of the melt would be bigger than
10%. When basaltic rocks are subjected to secondary processes, such as magma
differentiation and remelting, rhyolite will be generated. Thus we would predict some
volume of silicic rocks exposed in Krf'suvik if rhyolite flowed out on the surface. This
kind of rhyolites generated by remelting of basalt has been observed in Iceland by several
geologists and geochemists (Gunnarsson et al., 1998; Jonasson, 1994). At the same time,
from figure 2-14, a tail of low Vp is observed in deeper (> 7 km) and located in the east
of the ~ 6km low velocity zone. If this low velocity tail represents the trace of magma
intrusion, then partial melting at the depth of ~ 6 km would be further surpported. If the
crustal magma were derived from some deeper magma chamber, such as in shallower
mantle (Thordarson and Larsen, 2007), there would be low Vp, Vs and high Vp/Vs ratio
zones gradually distributed in deep structure. However, due to the limited depth of
earthquakes in this study, our tomography has no resolution when the depth is deeper
than 8 km.
Magma intrusion is a popular explanation for the observed uplift and expansion
deformation by the geodetic measurements. Seismicity and deformation triggered by
49
magma intrusion has been observed in several volcanic systems in Iceland. In Krafla
volcano in north Iceland, the 1974 - 1975 seismic activity accompanied the Krafla rifting
when multiple eruptions and dike injections occurred. The 1994-1995 seismicity and
deformation at the Hengill triple junction, Iceland, is considered to have been triggered
by minor magma injection in a zone of horizontal shear stress (Sigmundsson et al., 1997).
While in Krfsuvik region, the continue recording of GPS and InSAR has been observing
periodic inflation and deflation on the surface (Keiding et al., 2008; Michalczewska et al.,
2012). The intrusion of the magma would cause the inflation of the surface and the
increased pressure would cause the change of the local stress that could trigger
earthquakes when combine with the regional left lateral and expansion stress field. This
continued earthquakes could be observed at the historical seismic recording (fig.1-3,4).
These earthquakes could increase the local permeability and the ground water will reach
as deep as about 6 km. When the water accumulates to be some certain amount, partial
melting at shallow depth (-6 km) occurs, which would further increase the local stress,
trigger big seismic swarms and cause significant surface uplift and expansion. The
phenomenon of big seismic swarms inserting in long-live earthquakes has been observed
and recorded (fig. 1,3). During the big earthquake swarm, the water of Lake Kleifarvatn
would be injected into underground (Michalczewska, 2012). Due to the circulation of this
cold water (fig. 4-2), partial melting might be cooling down and then surface deflation is
observed. Change of the stress state and precipitation of minerals might make fractures
become to seal and then inflation caused by magma intrusion happens, which indicates
the beginning of another circulation of inflation and deflation. We speculate that the
intrusion of the magma would respond for the long period seismicity and surface uplift
50
and expansion, while the partial melting at shallow depth may be the cause of the big
earthquake swarms and the circulation of the cool water is possible the reason for the
short period deflation.
4
I
/
P&Is
02
B
wet pwwiotmt
Sedhnent
3
I
thS study
N et al.
'94
02 L&W
0
72
4
Loet al.
,98
1
Basat
0.2
(~6kn )
F\
!!
600
-
1000
800
1200
1400
T OC
Figure 4-3: Rock melting studies (Grove et al., 2006). We focus on the dashed curve of
basalt melting with 1-2 wt% water. Other curved lines indicated different kinds of rock in
various studies. The red horizontal indicates the 0.2 GPa at ~6 km in our interested depth.
1.8O.
0,6.0-
6
-
Vp (km/i)
----- --- - -----
4F. Irk.
70
volume fraction of fluids (%)
1
Figure 4-4: Relation between Vp, Vp/Vs ratio and volume of fraction of fluids (Nakajima
et al., 2001) in basaltic material (amphibolite). Solid and dashed curves show the water
inclusion and melt inclusion, respectively. Thin, moderate, and thick curves denote the
aspect ratio of 0.001, 0.01 and 0.1, respectively. The black and gray stars indicate the
possibly obtained Vp/Vs ratio in this study. The horizontal dotted line indicates the Vp
value (5.75 - 5.8 km/s) obtained in our tomography.
51
2). Supercritical water: The substance that is above its critical temperature and
pressure is defined as a supercritical fluid. The supercritical fluids have a combination of
gas-like mass transfer and liquid-like solvating characteristics, which are considered as
highly compressible. The critical point of pure water occurs at 22.1 MPa and 374 0C from
the laboratory experiments that present the physical properties of supercritical fluids
(Palmer et al., 2004). From the thermal profile in figure 4-2, we could find that the fluids
or ground waters at the depth of about 6 km are probably in a supercritical state with high
pressures. However, Vs is more sensitive to rock matrix but less sensitive to the free gas
to water ratio. Therefore, we should expect low Vp/Vs ratios in a supercritical case,
which is inconsistent with our inversion results. Unfortunately, due to the lack of deep
earthquakes, the resolvability of the Vp/Vs ratio at 6 km depth is far from ideal. At the
same time, we have relatively well resolved Vp and Vs and consider the uncertainties are
in reasonable range, if we calculate the Vp/Vs ratio by directly dividing Vs by Vp from
the tomography result, then the Vp/Vs value at where the low Vp appears would be 1.74,
which is actually a relatively lower than the average value (1.8).
Therefore, the
possibility of supercritical water cannot be ruled out based on currently available data. If
we benchmark this Vp/Vs of 1.74 to figure 4-4, as shown by the gray star, we find that
the volume fraction of water would be about 1-2%. Given the volume of low velocity
anomaly -20 km 3 obtained from the tomography, the volume of water included would
be 0.2-0.4 km 3 . As mentioned above, after the big earthquake in
17 th
June 2000, 12%
of the volume (~0.1 kM 3 ) of Lake Kleifarvatn disappeared (Clifton et al., 2003). The
required 0.2 km 3 of water injection is on the same order of the water loss if considering
52
the contribution of Lake Kleifarvatn, ground water and precipitation.
Such supercritical fluid could possibly be responsible for the dilatation of this area
and increase the potential of generating or triggering faults as we discussed above. The
formation of the supercritical reservoir from water would increase the local pressure and
cause slightly surface uplift and expansion. When some deeper earthquakes occur, the
dramatic increase of permeability of deeper structure could lead to recharge of the
supercritical reservoir. The mixing of supercritical fluid with downward colder water
would lead to phase separation and formation of a high dynamic superheated vapor phase,
which would convey heat to the upper crust and cause dilatation and seismic events. The
releasing of vapor will make pore pressures of the reservoir suddenly go down and then
cause surface deflation. After sealing of fractures above the reservoir, another
supercritical reservoir begins to format. In this supercritical case, we speculate that the
formation of the supercritical reservoir are responsible for the long period seismicity and
surface uplift and expansion, while the phase change of the supercritical fluid may be the
cause of the big earthquake swarms and the afterward short period deflation.
The understanding of this low velocity zone is of great important for developing
potential geothermal heat source, mitigating risk of magma intrusion or understanding the
activity of the Mid-Atlantic Ridge.
Detailed analysis of the energy distribution and
attenuation of S waves at deeper part in our study area would provide more properties of
this low velocity zone.
Incorporation of deeper earthquakes would not only provide
better constraint of Vp/Vs ratio to distinguish whether this low velocity zone is partial
melting or supercritical water, but also provide velocity information in deeper structure,
which would help us to trace whether there is magma intrusion or not.
53
Chapter 5
Conclusion
The Krfsuvik geothermal field is considered as one of the most active geothermal
fields in SW Iceland. Large earthquake swarms have been recorded from Nov. 2010 Feb. 2011 in the Krfsuvik area.
Using the seismicity recorded by 23 seismic stations
deployed by SIL, UU, RU and MIT, we delineated shallow N-S trending strike-slip faults
and a relatively deeper NE-trending dip fault through earthquake relocations. We also
detected a low velocity zone around 6 km under the earthquake swarm, which is in the
similar location with a low resistivity zone detected by previous 3D MagnetoTelluric
(MT) survey(Hersir et al., 2011). Resolvability tests were performed to validate the
tomography results. Based on the distribution of the earthquakes, we delineated the
brittle-ductile boundary in Krysuvik geothermal field.
From the InSAR data, we observed the expansion and uplift from 2009 to 2012 in
this area. Such deformation might indicate more complex geodynamic behavior than
simple plate-motion strain accumulation. Using our seismic model as a constraint, we
considered both point source volume change and faults as possible sources of the
deformation. A point source at the depth of 6 kin, supported by the seismic tomography,
well modeled the main deformation, which further supported the low velocity zone. The
NE-SW trending dip fault, interpreted as a normal fault, and the N-S trending fault,
interpreted as right-lateral strike slip fault, well simulated the secondary order
deformation.
54
The potential geological explanations of supercritical water, H2 0-rich partial
melting with magma intrusion are discussed for the detected low velocity zone.
Further studies with deeper earthquakes are needed to obtain high fidelity Vp/Vs ratios
for better understanding the cause of the low velocity zone. The results of this thesis
provided new insights to understand the seismicity and surface deformation in volcanic
zones as well as provided important information in exploration of new geothermal areas.
55
.
.
...........
...
.
..
.....
.
..
.....
Appendix A
Geothermal Manifestation in
Krysuvik
Geology mapping area
Iy
Hot Springs, Mud Pots
f
f
I
I,
[
Mud
Pot
(Mawejje,
2007)
Sulphur needles
56
Appendix B
GPS measurements in Krysuvik
LI4dA)a*,nt
21
MR1
,
kb
t
1
44)
20
24)
z
1'
r
1)
-
2))
-4"
7(917
NON
2409
ZO 10
1
i
:M2
-
I --
2~' .2RIlop
2013
40
410
:(Krziffi
_I
N
211
A-111 awl
0)
-21,
40
201i7
2(11111
201$)
24410
24)11
40
2*112
iL I
'41)
I
2)MI
:IwM$
41)
2w'
240
2
1 11
60
2017
20WK
204
2010
2011
2112
iI,
201
(http://strokkur.raunvis.hi.is/-sigrun/KRIV.html).
57
i
~4kt( 1)'~ N
14~4n~
~~nddr'
410
~
-.
.2
.
.~ .4 .
. .
.
40
20-17
ILt
2"11
2012h 'II
-40
40)
ALI
E
E
4)
~~ ~ ~2010
2011
2huI2
2x7
~
2~'
(http://strokkur.raunvis.hi.is/~sigrun/KRIV.html).
58
2o
zmi
yo:
o
References
Brander, J.L., Mason, R.G., Calvert, R.W., 1976. Precise Distance Measurements in
Iceland. Tectonophysics 31, 193-206.
Christensen, N.I., 1979. Compressional Wave Velocities in Rocks at HighTemperatures and Pressures, Critical Thermal-Gradients, and Crustal Low-Velocity
Zones. Journal of Geophysical Research 84, 6849-6857.
Clifton, A., Einarsson, P., 2005. Styles of surface rupture accompanying the June 17
and 21, 2000 earthquakes in the South Iceland Seismic Zone. Tectonophysics 396,
141-159.
Clifton, A.E., Kattenhorn, S.A., 2006. Structural architecture of a highly oblique
divergent plate boundary segment. Tectonophysics 419, 27-40.
Clifton, A.E., Pagli, C., Jonsdottir, J.F., Eythorsdottir, K., Vogfjoro, K., 2003. Surface
effects of triggered fault slip on Reykjanes Peninsula, SW Iceland. Tectonophysics
369, 145-154.
Einarsson, P., 1976. Relative location of earthquakes within the Tjornes Fracture
Zone. Soc. Sci. Isl., 45-60.
Einarsson, P., 1991. Earthquakes and Present-Day Tectonism in Iceland.
Tectonophysics 189, 261-279.
Einarsson, P., Eiriksson, J., 1982. Earthquake fractures in the districts Land and
Rangavellir in the South Iceland Seismic Zone. Jokull 32, 113-120.
Foulger, G.R., 1995. The Hengill Geothermal Area, Iceland - Variation of
Temperature-Gradients Deduced from the Maximum Depth of Seismogenesis. J
Volcanol Geoth Res 65, 119-133.
Geoffroy, L., Dorbath, C., 2008. Deep downward fluid percolation driven by localized
crust dilatation in Iceland. Geophys Res Lett 35.
Grove, T.L., Chatterjee, N., Parman, S.W., Medard, E., 2006. The influence of H20
on mantle wedge melting. Earth Planet Sc Lett 249, 74-89.
Gunnarsson, B., Marsh, B.D., Taylor, H.P., 1998. Generation of Icelandic rhyolites:
silicic lavas from the Torfajokull central volcano. J Volcanol Geoth Res 83, 1-45.
59
Hanks, T.C., Kanamori, H., 1979. A moment Magnitude Scale. Journal of Geophysical
Research 84, 2348-2350.
Hersir, G.P., Arnason, K., Vilhjalmsson, A.M., 2013. 3D inversion of magnetotelluric
(MT) resistivity data from Krysuvik high temperature geothermal area in SW
Iceland., In Proceedings of the thirty-eight Workshop on Geothermal Reservoir
Engineering.
Hjaltadottir, S., 2010. Use of relatively located micro-earthquakes to map fault
patterns and estimate the thickness of the brittle crust in Southwest Iceland, Icelandic
Meteorological Office.
Jonasson, K., 1994. Rhyolite volcanism in the Krafla central volcano, north-east
Iceland. Bull Volcanol 56, 516-528.
Keiding, M., Arnadottir, T., Sturkell, E., Geirsson, H., Lund, B., 2008. Strain
accumulation along an oblique plate boundary: the Reykjanes Peninsula, southwest
Iceland. Geophys J Int 172, 861-872.
Keiding, M., Lund, B., Arnadottir, T., 2009. Earthquakes, stress, and strain along an
obliquely divergent plate boundary: Reykjanes Peninsula, southwest Iceland. J
Geophys Res-Sol Ea 114.
Kristjansdottir, S., 2013. Microseismicity in the Krysuvik Geothermal Field, SW
Iceland, from May to October 2009, Faculty of Earth Sciences. University of Iceland.
Liu, J., Bohlen, S.R., Ernst, W.G., 1996. Stability of hydrous phases in subducting
oceanic crust. Earth Planet Sc Lett 143, 161-171.
Mawejje, P., 2007. Geothermal Exploration and Geological Mapping at Seltun in
Krysuvik Geothermal Field, Reykjanes Peninsula, SW-Iceland, Geothermal Training
Programme. United Nations University, Reykjavik, pp. 257-276.
Michalczewska, K., 2012. Crustal deformation in the Krfsuvik area, in: System,
D.R.o.G. (Ed.).
Michalczewska, K., Hreinsdottir, S., Arnadottir, T., Hjaltadottir, S., Agustsdottir, T.,
Guomundsson, M.T., Geirsson, H., Sigmundsson, F., Guomundsson, G., 2012.
Inflation and deflation episodes in the Krysuvik volcanic system, AGU fall meeting,
San Francisco, USA.
Nakajima, J., Matsuzawa, T., Hasegawa, A., Zhao, D.P., 2001. Three-dimensional
structure of V-p, V-s, and V-p/V-s beneath northeastern Japan: Implications for arc
magmatism and fluids. J Geophys Res-Sol Ea 106, 21843-21857.
60
Paige, C.C., Saunders, M.A., 1982. Algorithm-583 - Lsqr - Sparse Linear-Equations
and Least-Squares Problems. Acm T Math Software 8, 195-209.
Palmer, D.A., Fernanadez-Prini, R., H., A.H.(Eds.), 2004. Aqueous systems at
Elevated temperatures and pressures; physcial chemistry in water, steam and
hydrothermal solutions. Elsevier Ltd., Oxford, UK, 745.
Sella, G.F., Dixon, T.H., Mao, A.L., 2002. REVEL: A model for Recent plate
velocities from space geodesy. J Geophys Res-Sol Ea 107.
Sigmundsson, F., Einarsson, P., Rognvaldsson, S.T., Foulger, G.R., Hodgkinson, K.R.,
Thorbergsson, G., 1997. The 1994-1995 seismicity and deformation at the Hengill
triple junction, Iceland: Triggering of earthquakes by minor magma injection in a
zone of horizontal shear stress. J Geophys Res-Sol Ea 102, 15151-15161.
Sleeman, R., van Eck, T., 1999. Robust automatic P-phase picking: an on-line
implementation in the analysis of broadband seismogram recordings. Phys Earth
Planet In 113, 265-275.
Thordarson, T., Larsen, G., 2007. Volcanism in Iceland in historical time: Volcano
types, eruption styles and eruptive history. J Geodyn 43, 118-152.
Tryggvason, A., Rognvaldsson, S.T., Flovenz, O.G., 2002. Three-dimensional
imaging of the P- and S-wave velocity structure and earthquake locations beneath
Southwest Iceland. Geophys J Int 151, 848-866.
Um, J., Thurber, C., 1987. A Fast Algorithm for Two-Point Seismic Ray Tracing. B
Seismol Soc Am 77, 972-986.
Violay, M., Gibert, B., Mainprice, D., Evans, B., Dautria, J.M., Azais, P., Pezard, P.,
2012. An experimental study of the brittle-ductile transition of basalt at oceanic crust
pressure and temperature conditions. J Geophys Res-Sol Ea 117.
Waldhauser, F., 2001. HypoDD -- A program to compute double-difference
hypocenter locations, Open-File. U. S. Geological Survey.
Waldhauser, F., Ellsworth, W.L., 2000. A double-difference earthquake location
algorithm: Method and application to the northern Hayward fault, California. B
Seismol Soc Am 90, 1353-1368.
Zelt, C.A., 1998. Lateral velocity resolution from three-dimensional seismic
refraction data. Geophys J Int 135, 1101-1112.
Zhang, H.J., Thurber, C., Bedrosian, P., 2009. Joint inversion for Vp, Vs, and Vp/Vs
at SAFOD, Parkfield, California. Geochem Geophy Geosy 10.
61
Zhang, H.J., Thurber, C.H., 2003. Double-difference tomography: The method and its
application to the Hayward Fault, California. B Seismol Soc Am 93, 1875-1889.
62