ArticlePDF Available

Finite-fault slip models for the 15 April 1979 (M-W 7.1) Montenegro earthquake and its strongest aftershock of 24 May 1979 (M-W 6.2)

Authors:

Abstract and Figures

We revisit the April 1979 Montenegro earthquake sequence to invert for finite-fault slip models for the mainshock of 15 April 1979 (Mw 7.1) and of the strongest aftershock of 24 May 1979 (Mw 6.2) using P, SH and SV waveforms, retrieved from IRIS data center. We also used body waveform modelling inversion to confirm the focal mechanism of the mainshock as a pure thrust mechanism and rule out the existence of considerable strike slip component in the motion. The mainshock occurred along a shallow (depth 7 km), low angle (14°) thrust fault, parallel to the coastline and dipping to the NE. Our preferred slip distribution model for the mainshock indicates that rupture initiated from SE and propagated towards NW, with a speed of 2.0 km/s. Moment was released in a main slip patch, confined in an area of L ∼ 50 km × W ∼ 23 km. The maximum slip (∼ 2.7 m) occurred ∼ 30 km to the NW of the hypocenter (location of rupture initiation). The average slip is 49 cm and the total moment release over the fault is 4.38e19 Nm. The slip model adequately fits the distribution of the Mw ≥ 4.3 aftershocks, as most of them are located in the regions of the fault plane that did not slip during the mainshock. The 24 May 1979 (Mw 6.2) strongest aftershock occurred ∼ 40 km NW of the mainshock. Our preferred slip model for this event showed a characteristic two-lobe pattern, where each lobe is ∼ 7.5 × 7.5 km2. Rupture initiated in the NW lobe, where the slip obtained its maximum value of 45 cm, very close to the hypocenter, and propagated towards the south-eastern lobe where it reached another maximum value — for this lobe — of 30 cm, approximately 10 km away from the hypocenter. To indirectly validate our slip models we produced synthetic PGV maps (Shake maps) and we compared our predictions with observations of ground shaking from strong motion records. All comparisons were made for rock soil conditions and in general our slip models adequately fit the observations especially at the closest stations where the shaking was considerably stronger. Through the search of the parameter space for our inversions we obtained an optimum location for the mainshock at 42.04°N and 19.21° E and we also observed that better fit to the observations was obtained when the fault was modeled as a blind thrust fault.
Content may be subject to copyright.
Finite-fault slip models for the 15 April 1979 (M
w
7.1) Montenegro
earthquake and its strongest aftershock of 24 May 1979 (M
w
6.2)
Christoforos Benetatos, Anastasia Kiratzi
Aristotle University of Thessaloniki, Department of Geophysics, 54124 Thessaloniki, Greece
Received 10 November 2005; received in revised form 9 March 2006; accepted 16 April 2006
Available online 27 June 2006
Abstract
We revisit the April 1979 Montenegro earthquake sequence to invert for finite-fault slip models for the mainshock of 15 April
1979 (M
w
7.1) and of the strongest aftershock of 24 May 1979 (M
w
6.2) using P, SH and SV waveforms, retrieved from IRIS data
center. We also used body waveform modelling inversion to confirm the focal mechanism of the mainshock as a pure thrust
mechanism and rule out the existence of considerable strike slip component in the motion. The mainshock occurred along a shallow
(depth 7 km), low angle (14°) thrust fault, parallel to the coastline and dipping to the NE. Our preferred slip distribution model for
the mainshock indicates that rupture initiated from SE and propagated towards NW, with a speed of 2.0 km/s. Moment was released
in a main slip patch, confined in an area of L50 km × W23 km. The maximum slip (2.7 m) occurred 30 km to the NW of
the hypocenter (location of rupture initiation). The average slip is 49 cm and the total moment release over the fault is 4.38e19 Nm.
The slip model adequately fits the distribution of the M
w
4.3 aftershocks, as most of them are located in the regions of the fault
plane that did not slip during the mainshock. The 24 May 1979 (M
w
6.2) strongest aftershock occurred 40 km NW of the
mainshock. Our preferred slip model for this event showed a characteristic two-lobe pattern, where each lobe is 7.5 × 7.5 km
2
.
Rupture initiated in the NW lobe, where the slip obtained its maximum value of 45 cm, very close to the hypocenter, and
propagated towards the south-eastern lobe where it reached another maximum value for this lobe of 30 cm, approximately
10 km away from the hypocenter. To indirectly validate our slip models we produced synthetic PGV maps (Shake maps) and we
compared our predictions with observations of ground shaking from strong motion records. All comparisons were made for rock
soil conditions and in general our slip models adequately fit the observations especially at the closest stations where the shaking
was considerably stronger. Through the search of the parameter space for our inversions we obtained an optimum location for the
mainshock at 42.04°N and 19.21° E and we also observed that better fit to the observations was obtained when the fault was
modeled as a blind thrust fault.
© 2006 Elsevier B.V. All rights reserved.
Keywords: Earthquake mechanics; Finite-fault; Montenegro; Earthquake source; Slip distribution
1. Introduction
On 15 April 1979 (GMT 06:19:48) a strong
earthquake (M
w
7.1) struck the coastal regions of
Serbia Montenegro and northern Albania. This event
caused extensive damage along 100 km of coastline
Tectonophysics 421 (2006) 129 143
www.elsevier.com/locate/tecto
Corresponding author. Tel.: +30 2310 998486; fax: +30 2310
998528.
E-mail address: kiratzi@geo.auth.gr (A. Kiratzi).
0040-1951/$ - see front matter © 2006 Elsevier B.V. All rights reserved.
doi:10.1016/j.tecto.2006.04.009
Fig. 1. a) Location map for the region of occurrence of the 1979 Montenegro sequence; Epicenters from Karakaisis et al. (1985) b) Distribution of
intensities and isoseismals for the 15 April, 1979 mainshock (Papazachos et al., 2001).
130 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
(Fig. 1a, b), killed 129 people, injured 1554 in former
Yugoslavia and Albania and left homeless more than
80,000 (Papazachos et al., 2001; Papazachos and
Papazachou, 2003). The cities that suffered more were
Bar, Budva, Kotor and Ulcinj in Yugoslavia; massive
landslides, significant damage in harbours, and ancient
monuments along the Montenegro coastline have been
also reported. There is a note in Papazachos and
Papazachou (2003), without any reference to their
original source, that a sea wave was reported in the
Adriatic Sea, where a ship sunk and sea side houses
were taken off by the sea wave along a line of 15 km
The aftershocks of the sequence formed two distinct
clusters, which operated simultaneously and in the
northern cluster the strongest aftershock occurred on
24 May 1979 (M
w
6.2).
The region of occurrence of the sequence is located
at the border of the Adria plate with the Eurasian
platform, where the folded alpine DinaridesAlba-
nides Hellenides mountain ranges exist, and conti-
nentalcontinental collision takes place. This area is
dominated by the IonianAdriatic faults that form a
zone of 30 km to 40 km in width, located at the outer
side of the orogene (Aliaj and Muco, 1980). Based on
GPS data, Oldow et al. (2002) proposed that the Adria
plate is fragmented into blocks from which the south-
eastern, that includes south-eastern Italy, northwestern
Albania and the western coasts of former Yugoslavia,
is moving northwards (3 to 10 mm/yr) colliding
with stable Eurasia.
The Montenegro earthquake has been recorded by
stations at regional and teleseismic distances, of the
developing at that time GSN network, by a significant
number of short-period stations in the countries
surrounding the Adriatic Sea and by 29 SMA-1
accelerographs (Petrovski et al., 1979). The so far
published work concerns the characteristics of the
aftershock distribution (Kociaj and Sulstarova, 1980;
Hurtig et al., 1980; Console and Favali, 1981;
Karakaisis et al., 1985) the focal mechanism parameters
of the mainshock (Kanamori, 1979; Sulstarova, 1980;
Karakaisis et al., 1985; Baker et al., 1997) of the
foreshocks and aftershocks of the sequence (Kociaj and
Sulstarova, 1980; Karakaisis et al., 1985), as well as
stochastic strong ground motion simulations (Roume-
lioti and Kiratzi, 2002).
Here we study the rupture process of the two
strongest events of the sequence, of 15 April 1979,
Fig. 2. Focal mechanisms previously proposed for the mainshock and the solution obtained in this study using body wave waveform modelling.
Below each mechanism the strike/dip/rake is presented.
131C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
M
w
7.1 and of 24 May 1979, M
w
6.2 and we specifically
focus on the slip distribution onto the fault planes. We
use P and S waveforms from GSN stations, and the
finite-fault inversion technique of Antolik and Dreger
(2003). To better resolve our preferred slip models we
also invert for the focal mechanism of the mainshock as
there were diverse solutions available in the literature.
Furthermore we use our preferred finite-fault models to
calculate synthetic PGV (Peak Ground Velocity) maps
of the meizoseismal area, for both events, and compare
them with the maximum ground velocities that have
been recorded at the strong motion stations that are
available in the European Strong Motion Database
(Ambraseys et al., 2000).
2. Characteristics of the sequence
2.1. Focal mechanism of the mainshock (15 April 1979,
M
w
7.1)
Fig. 2 summarizes the previously published focal
mechanisms for the mainshock (see also Table 1).
The solutions vary from pure low angle thrust
faulting to reverse faulting combined with consider-
able strike slip motions. Most solutions indicate
faulting in a NWSEdirectionparalleltothe
coastline with the east dipping fault identified as
the fault plane (Karakaisis et al., 1985).
To better resolve the mechanism of the mainshock we
obtained digital waveforms from stations of the Global
Digital Seismograph Network (GDSN) from the Incor-
porated Research Institutions for Seismology (IRIS)
data center. We then used MT5 (Zwick et al., 1994) and
the procedures as described previously (e.g. Kiratzi and
Louvari, 2001; Benetatos et al., 2004, 2005 and in the
references therein) to invert for the focal mechanism
parameters (strike/dip/rake), centroid depth and seismic
moment, assuming a source represented as a point in
space and described in time by a source time function
consisting of overlapping isosceles triangles. In total we
use 7 P-waves and 7 SH-waves from stations in the
distance range from 30° to 90°. Prior to the inversion
waveforms have been filtered between 0.01 and 0.1 Hz
and convolved with a typical WWSSN 15100s long-
period instrument response. Green's functions have
been calculated using a half space of 6.5 km/s and
3.7 km/s for P- and S-waves, respectively and density
2.8 g/cm
3
. This model is adequate to explain the
simplicity of the Earth structure at the distance range we
use (Helmberger, 1974; Langston and Helmberger,
1975).
A single point source was adequate to obtain
satisfactory fit and our inversion yielded low angle
thrusting along a NWSE line (Fig. 3 and Table 1). Our
solution is very close to those of Baker et al. (1997) and
Boore et al. (1981). The source time function has a total
duration of 27 s, maybe large for an M
w
7.1 event, but
also close to the duration obtained by Baker et al. (1997).
Another indication for long source duration is the fact
that in the published Harvard CMT solution for this
event, the centroid time shift (the difference between the
centroid time and the origin time) which is interpreted as
half of the total duration, is 14.8 s, which implies a total
duration of 29 s. In the lower part of Fig. 3 we
Table 1
Source parameters for the 15 April 1979 (M
w
7.1) mainshock and for the 24 May 1979 (M
w
6.2) strongest aftershock previously published
Epicenter Depth
(km)
M
w
Nodal plane 1 Nodal Plane 2 P-axis T-axis Method
used
Ref
φ°N λ° E Strike° Dip° Rake° Strike° Dip° Rake° Az° Dip° Az° Dip°
15 April 1979
MS 06:19
––40 7.0 324 50 115 108 46 63 37 2 300 71 INV 1
332 55 1 241 89 145 292 23 191 25 INV 2
42.03 19.04 7.2 287 11 63 134 80 95 220 35 50 55 FMP 3
42.13 19.06 22 6.9/7.1 308 15 88 130 75 91 220 30 41 60 INV/FMP 4
––– 6.7 ML 319 85 42 225 48 174 84 24 191 32 FMP 5
41.93 18.99 1 7.1 312 44 45 186 61 124 252 10 146 59 FMP 6
42.09 19.21 12 6.9 316 14 90 136 76 90 226 31 46 59 INV 7
42.04 19.21 7
5)
7.1 300
(+10/8)
14
(+7/6)
88
(+12/5)
120 76 90 210 31 31 59 INV 7
24 May 1979
AF 17:23
––– 6.2 ML 37 41 154 147 73 52 265 19 17 48 FMP 4
42.15 18.71 7 6.3 299 50 31 188 67 136 247 10 146 47 FMP 5
42.26 18.75 6 6.2 320 35 90 140 55 90 232 10 52 80 INV 6
References: 1) Kanamori (1979),2)Sulstarova (1980),3)Boore et al. (1981),4)Console and Favalli (1981),5)Karakaisis et al. (1985),6)Baker et al.
(1997), 7) this study.
INV: Inversion, FMP: First motion polarities.
132 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
Fig. 3. a) Minimum misfit solution for the 15 April 1979 mainshock showing P- and SH-observed (solid) and synthetic (dashed) waveforms, which
are arranged azimuthally around the focal spheres. Station locations are plotted as capital letters on the focal spheres. Pand Taxes are also marked.
The second line on the header shows the strike, dip, rake, centroid depth (in km) and the scalar moment. This solution is similar to Baker et al. (1997).
b) Test of a solution when a strike-slip mechanism is assumed (arrows point out polarity mismatch).
133C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
investigate the fit to the waveforms if we assume
considerable strike-slip motion in the focal mechanism,
keeping all other parameters constrained. We clearly see
that the P-waves are not very sensitive, but in the SH
waveforms the fit deteriorates in most of the stations,
with polarity mismatch. Thus, for our further modelling
we accepted a thrust mechanism for the mainshock. For
the 24 May 1979, M
w
6.2 strongest aftershock we adopt
the focal mechanism of Baker et al. (1997) as listed in
Table 1.
2.2. Spatial distribution of the aftershock sequence
Fig. 4a summarises the temporal and spatial
distribution of the entire (April 1979January 1980)
aftershock sequence, with M
w
4.3 (data as relocated
Fig. 4. a) Distribution of aftershocks with M 4.3 of the period 15 April 197923 January 1980. Note that they form two distinct clusters clearly
connected with the two strongest events. b) Distribution of aftershocks with M4.3 until 23 May 1979, prior to occurrence of the strongest
aftershock, which shows that the northern cluster was already in operation. Epicenters as relocated in Karakaisis et al., 1985.
134 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
by Karakaisis et al., 1985). It is seen that the aftershocks
form two distinct clusters, one at the southern part of the
activated area, clearly connected with the mainshock
and the other at the northern part, clearly connected with
the strongest aftershock of 24 May 1979 (M
w
6.2). The
northern cluster was already in operation prior to the
occurrence of the strongest aftershock (Fig. 4b).
Actually the first M
w
4.3 event was recorded on the
northern cluster 12 min after the occurrence of the
mainshock (Karakaisis et al., 1985). The region between
the two clusters is free of aftershocks, at least with
M
w
> 4.3 that the data covers. This is an indication that
this part has already slipped during the mainshock and
was not capable to produce aftershocks at later stages, a
remark assumed in Karakaisis et al. (1985) and shown
later in this paper. The depth distribution of aftershocks
is constrained to the upper 1520 km (Console and
Favali, 1981).
3. Slip distribution obtained from body waveform
inversion
3.1. Data and method of analysis
We examined the slip models of the mainshock and
of the strongest aftershock using data from 10 GDSN
stations that were in operation in 1979 (Fig. 5). The data
set consists of 7 P and 10 S waveforms in the distance
range from 1218 to 9830 km. Initial records have been
deconvolved from the instrument response, integrated to
obtain ground displacement and bandpass filtered
Fig. 5. GSN locations of the stations (white triangles) whose waveforms were used in the inversions. The waveforms form GRFO and TATO (black
triangles) were not used in the inversion due to low signal to noise ratios.
Fig. 6. Variation of the variance reduction (%) obtained from the finite-
fault inversion for the mainshock slip model, with different rupture
speeds tested. The highest variance reduction achieved is for rupture
speed of 2.0 km/s.
135C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
between 0.01 and 0.15 Hz. We model 40 s of the P-
waves and 80 s of the S-waves (SV and SH
components). Green's functions were computed using
a frequency-wavenumber integration method (Saikia,
1994) and the Iasp91 velocity model (Kennett and
Engdahl, 1991) which has been proved adequate to
explain the timing and the frequency content of the
waveforms at the distances used here. Actually,
FKRPROG is a plane-layered frequency integration
code, which can be used also at teleseismic distances
when an earth flattening algorithm is used to flatten the
velocity structure, as applied here.
The inversion method is a linearized least-squares
scheme presented previously (e.g., Hartzell and Hea-
ton, 1983; Wald and Heaton, 1994; Dreger and
Kaverina, 2000; Kaverina et al., 2002; Antolik and
Dreger, 2003), that places several constrains on the
solution which include slip positivity and spatial-
derivative smoothing. The fault plane is divided into
nodes in both coordinate directions and the rupture is
constrained to propagate at a constant velocity from the
hypocenter. The slip velocity history of each subfault
can be a delta function or an isosceles triangle with a
constant duration (adopted here), and each node is
allowed to slip only once as the (circular) rupture front
passes. The rake angle is also forced to remain constant
over the fault plane. The total system of equations to be
solved is:
WdG
kD
gI
0
@1
AðmÞ¼
Wd
0
0
0
@1
A:ð1Þ
Fig. 7. a) Fault plane view of our preferred slip distribution model for the 15 April 1979 Montenegro (M
w
7.1) earthquake obtained using least-squares
inversion of digitally recorded waveforms. The white star denotes the initiation of rupture onto the fault plane. Rupture propagated from SE towards
NW. Moment is released in a main slip patch (L50 km × W23 km); the maximum slip (2.7 m) occurred 7 km to the NW of the locus of rupture
initiation. b) Surface projection of the slip model along with the aftershocks (M4.3) that followed its occurrence. There is a good anti-correlation
between the area that slipped during the mainshock and the regions that slipped during the aftershocks.
136 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
The matrix Gconsists of the subfault Green's
functions, Wweights and normalizes the data
according to the inverse amplitude, Dis a gradient
smoothing matrix for both coordinate directions, mis
the vector of scalar moment values for each subfault,
λand ηare empirically determined constants that
control the weighting of the constraints relative to the
data fit, and dis the data vector. The slip positivity
constraint requires m0, indicating slip along the
rake direction. The values of scalar moment for each
subfault are converted to equivalent slip values using
a rigidity of 3.5 × 10
11
dyn cm
2
.
Fig. 8. P, SH and SV observed (solid lines) and synthetic (dashed lines) waveforms produced using our preferred slip model for the 15 April 1979
event.
137C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
4. Slip models of the events
4.1. Slip model for the 15 April 1979 Montenegro
mainshock (M
w
7.1)
To perform the inversions we used a fault plane of
dimensions 90 km × 50 km, larger than what is expected
for an earthquake of magnitude M
w
7.1, in order to
allow the slip to move to its preferable position without
constraining it due to limited fault dimensions (Das and
Suhadolc, 1996). The orientation of the fault and the
hypocenter depth adopted are those derived from our
teleseismic inversion (see Table 1). The fault plane was
divided into subfaults of dimensions 2 km × 2 km along
strike and dip, resulting in 45 nodes along strike and
25 nodes along dip direction.
A considerable number of initial inversions were
devoted to test the parameter space in terms of different
locations for the initiation of the rupture on the fault
plane and of the position of the fault plane relative to
the hypocenter. We additionally tested different rupture
velocities ranging from 1.4 to 3.0 km/s with a step of
0.2 km/s and the highest total variance reduction and
hence the better fit to the waveforms was for a rupture
velocity of 2.0 km/s (Fig. 6), which indicates a rather
slow rupture process. Moreover, we tested the
Fig. 9. a) Fault plane view of our preferred slip distribution model for the 24 May 1979 strongest aftershock obtained as before. The white star denotes
the initiation of rupture onto the fault plane. Note the characteristic two-lobe pattern, (each lobe 7.5 × 7.5 km
2
). Rupture initiated in the NW lobe,
(maximum slip value 45 cm), and propagated towards the south-eastern lobe (maximum slip value 30 cm). b) Surface projection of the slip alongwith
the aftershocks (M 4.3) that followed its occurrence, for reason of comparison as in Fig. 7b.
138 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
sensitivity of our inversions in terms of the focal
mechanism parameters and their uncertainties, varying
the strike, dip and rake within their error limits. During
these tests some stations (e.g. ANTO) were more
sensitive to these changes and improved the fit, while
for other stations the fit was worse, so we decided it
was a good compromise to adopt the values from our
teleseismic inversion in the finite-fault inversion. From
these initial runs of the inversion we saw that our
inversions were stable and the position and magnitude
of slip remained also very stable. We observed that we
were getting a better fit to the waveforms and we were
predicting more accurately the level of strong ground
shaking, as it will be presented in the following
sections, when the epicenter of the mainshock was
taken at 42.04° N and 19.21° E (slightly modified from
Baker et al., 1997) and when we modelled the fault as a
blind thrust fault that did not reach the surface. When
we allowed the fault to break the surface we were
getting very high values of ground shaking on land,
which contradicted the observed intensities and the
observed accelerations.
Our preferred slip distribution solution for the 15
April 1979 mainshock is presented in Fig. 7a.
Rupture initiated from SE and propagated towards
NW. The main slip patch is confined in an area of
L50 km × W23 km that is in agreement with
empirical relations for reverse faults (Wells and
Coppersmith, 1994; Papazachos et al., 2004). The
main moment release occurred 30 km to the NW
of the hypocenter (location of rupture initiation) with
a maximum slip value of 2.7 m. The average slip is
49 cm and the total moment release over the fault is
4.38e19 Nm, in agreement with the results of our
moment tensor inversion. In Fig. 7b we present a
map view of the slip distribution of the mainshock
along with the aftershock epicenters (Karakaisis et
al., 1985) that occurred until 24 May 1979, day of
the occurrence of the largest aftershock (M
w
6.2).
The data set consists of 74 events with M
w
4.3.
There is a good anti-correlation between the areas
that have slipped during the mainshock and the
locations of the aftershocks. The slip model (Fig. 7a)
provides a good overall fit between the observed
Fig. 10. P and SH observed (solid lines) and synthetic (dashed lines) waveforms produced using our preferred slip model for the 24 May 1979 event.
139C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
Fig. 11. Synthetic Shake map, contours of Peak Ground Velocities (in cm/s) using forward modelling and our preferred finite-fault slip model (Fig. 7a)
for the mainshock. Black triangles denote strong motion station locations and to the right of each station code (see Table 2) the maximum observed
PGV (in cm/s ) is shown, for comparison. Note that our model adequately predicts the observed values especially those of the closer to the epicenter
(see text for further details).
Fig. 12. As in Fig. 11 but for the 24 May 1979 aftershock using the slip model of Fig. 8a (see Table 2 and text for more details).
140 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
and the synthetic P-, SH- and SV-waveforms (Fig.
8).
4.2. Slip model for the 24 May 1979 aftershock
(M
w
6.2)
We used a fault plane with strike 320° dip 35°
rake 90°(Baker et al., 1997) having dimensions
30 km × 20 km, with the same reasoning as before.
We divided the fault plane into nodes of 1 km× 1 km
along strike and dip, respectively. The location of the
aftershock was taken at 42.15° N, 18.708° E (Karakaisis
et al., 1985), placing it in the north western part of the
activated area. Initial runs concerned the sensitivity of
our inversion to the location of the fault relative to the
hypocenter, different values for rupture velocity (span-
ning the range from 1.2 km/s to 3.0 km/s; increment of
0.2 km/s ). These tests showed that the slip was very
stable and not sensitive to the position of the fault
relative to the hypocenter, while the optimum rupture
velocity was 2.6 km/s.
Our preferred slip model for the 24 May 1979
aftershock (Fig. 9a) shows a characteristic two-lobe
pattern which is also reflected in its surface projection
(Fig. 9b). Each lobe is 7.5 × 7.5 km
2
. Rupture initiated
in the NW lobe, where the slip obtained its maximum
value of 45 cm, very close to the hypocenter. Then
rupture propagated to the south-eastern lobe where it
reached another maximum value for this lobe of
30 cm, approximately 10 km away from the hypocenter.
In Fig. 10 the fit between the observed and synthetic
waveforms is presented. We were able to achieve a
satisfactory overall fit for P- and for the single S-
waveform used.
5. Predicted ground motions (Shake maps)
To calculate the predicted ground motions by our
finite-fault models we constructed a grid spacing
0.02° × 0.02° around the region of study. At each point
we calculated synthetic velocity records for rock site
conditions, following the method of Dreger and
Kaverina (2000), using our preferred finite-fault model
(moment weights distributed on the fault plane). Green's
functions were calculated, as previously, using the
discrete wavenumber code (FKRPROG; Saikia, 1994).
For this application we adopted a more refine velocity
model applicable to the area (Ambraseys et al., 2004).
Table 2
Information on the sites of the strong-motion stations that recorded the ground motions of the Montenegro mainshock and the largest aftershock
Station Lat
(°N)
Lon
(°E)
Description Site class 15 April 1979 mainshock 24 May 1979 aftershock
Distance
(km)
Observed
PGV cm/s
PGV (for Rock
class, B cm/s)
Distance
(km)
Observed
PGV cm/s
PGV (for Rock
class, B) cm/s
BAR 42.095 19.101 Bar,
Skupstina Opstine
Stiff soil 16 47 40.7 32 15.9 14.4
ULA 41.919 19.221 Ulcinj,
Hotel Albatros
Rock 21 21 21
ULO 41.911 19.249 Ulcinj,
Hotel Olimpic
Stiff soil 24 42.8 38.6 54 2.9 1.8
PETO 42.204 18.948 Petrovac,
Hotel Oliva
Stiff soil 25 31.3 27.4
TIS 42.43 19.261 Titograd,
Seismoloska Stanica
Rock 55 3.4 3.4
TIG 42.442 19.264 Titograd,
Geoloski Zavod
Rock 56 4.0 4.0
HRZ 42.457 18.531 Herceg Novi,
O.S.D. Pavicic
Rock 65 13.3 13.3 32 5.4 5.4
DUB 42.656 18.091 Dubrovnic
Pomorska Skola
Rock 105 3.3 3.3
BUD 42.284 18.831 BudvaPTT Alluvium 8 22.8 11.9
PET 42.204 18.949 Petrovac
Hotel Rivijera
Alluvium 16 13.8 7.3
KOTN 42.418 18.773 Kotor Nas Raki Rock 21 3.9 3.9
TIVA 42.425 17.708 TivatAerodrom Alluvium 22 7.8 3.1
KOTZ 42.436 18.769 KotorZovod za
Biologgiju Mora
Alluvium 23 7.9 3.4
(Distances refer to the epicentral distances that were used in the empirical attenuation relations).
141C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
Green's functions were band pass filtered between
0.01 Hz and 5 Hz in order to incorporate the higher
frequencies in the estimation of the near source ground
motions. In each grid point we calculated an average
velocity value (geometric mean), using the maximum
amplitudes of the two horizontal components.
The Shake maps for the mainshock and for the
aftershock are presented in Figs. 11 and 12, respectively.
To validate our results we used the observed PGV values
(also shown in Figs. 11 and 12) at the locations of the
available strong motion records (Table 2) for these two
events (Ambraseys et al., 2000). To have a common
reference all observed PGV's were converted to rock
site conditions (B class, NEHRP classification), using
empirical attenuation relations (Tromans and Bommer,
2002), and these are plotted on the maps. More
specifically, we used the empirical relations to calculate
the PGV values for rock, stiff soil and alluvium site
conditions and the difference in PGV's was subtracted
from the observed values of either stiff soil or alluvium
sites. In this way we unified the PGV values at all sites
as they were all situated on reported rock conditions.
In general, the predicted PGV's for most of the
stations are in satisfactory agreement with the observa-
tions. For the mainshock the maximum ground shaking
(140 cm/s ) is predicted offshore, to the NW of the
epicenter and we do model adequately the high values
observed at the closest stations e.g. BAR, ULA and
PETO. We do not model the high value at ULO, which
is close to ULA where the recorded PGV is a factor of 2
smaller; besides ULO is not at the direction of the
rupture, so this high value could not be expected. In any
case, the mismatch can be attributed to model
insufficiency, misclassification of the site conditions of
the station, among other reasons. For the strongest
aftershock the maximum ground shaking (34 cm/s ) is
predicted in the sea and our slip model satisfactorily
predicts the PGV values at the closest stations (e.g BUD,
PET).
6. Conclusions
We revisited the April 1979 Montenegro earthquake
sequence, aiming to provide finite-fault slip models for
the mainshock of 15 April 1979 (M
w
7.1) and of the
strongest aftershock of 24 May 1979 (M
w
6.2) together
with other details of the rupture, using the available
waveforms from the IRIS network. We used body
waveform modeling inversion to recalculate the focal
mechanism of the mainshock and confirm its pure
thrust mechanism, (Boore et al., 1981; Baker et al.,
1997) and rule out the existence of considerable strike
slip component in the motion (Console and Favali,
1981; Karakaisis et al., 1985). Thus, the mainshock
occurred along a shallow (depth 7 km), low angle
(14°) thrust fault, parallel to the coastline and dipping
to the SE.
Our preferred slip distribution model for the main-
shock indicates that rupture initiated from SE and
propagated towards NW, with a speed of 2.0 km/s.
Moment was released in a main slip patch, confined in
an area of L50 km × W 23 km. The maximum slip
(2.7 m) occurred 7 km to the NW of the hypocenter
(location of rupture initiation). The average slip is 49 cm
and the total moment release over the fault is
4.38e19 Nm, in agreement with the results of our
moment tensor inversion. The slip model adequately fits
the distribution of the M
w
> 4.3 aftershocks, as most of
them are located in the regions of the fault plane that did
not slip during the mainshock.
The strongest aftershock of 24 May 1979 (M
w
6.2)
occurred 40 NW of the mainshock. Our preferred slip
model showed a characteristic two-lobe pattern. Each
lobe is 7.5 × 7.5 km
2
. Rupture initiated in the NW
lobe, where the slip obtained its maximum value of
45 cm, very close to the hypocenter. Then rupture
propagated to the south-eastern lobe where it reached
another maximum value for this lobe of 30 cm,
approximately 10 km away from the hypocenter.
We used the available strong motion records for the
two events to validate our slip models in an indirect
manner, by simply producing synthetic PGV (Shake
maps) and comparing our predictions with the observa-
tions. All comparisons were made for rock soil
conditions and in general our slip models adequately
fit the observations especially at the closest stations
where the shaking was considerably stronger.
At last, from the inversion parameter search, we
constrained the location of the mainshock epicenter. The
epicenter location which produced the best results, in
terms of matching the regions of ground shaking, was at
42.04° N and 19.21° E. In addition, in the same
parameter search we observed that we obtained the best
matching both in the synthetic waveforms and also to
the values of strong ground shaking when we model the
fault as a blind thrust fault that did not break the surface.
Acknowledgments
We acknowledge with thanks financial support from
the General Secretariat of Research and Technology
(GSRT), Ministry of Development of Greece. We are
grateful to Douglas Dreger, from the Seismological
Institute of Berkeley University, for his continuous
142 C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
support with the inversions. We thank the two
anonymous reviewers and the Editor for the constructive
comments and suggestions. Thanks are also due to the
IRIS consortium and all the affiliated networks for
providing the digital waveforms. Most of the figures
were produced using the GMT software (Wessel and
Smith, 1998).
References
Aliaj, S., Muco, B., 1980. The geological condition that generated the
earthquake of April 15, 1979. Proc. of the 17th Assembly of the
ESC, Budapest, 1980, pp. 7983.
Ambraseys, N., Smit, P., Berardi, R., Rinaldis, D., Cotton, F., Berge-
Thierry, C., 2000. European Strong - Motion Database, European
Council, Environment and Climate Research Programme (CD).
Ambraseys, N., Douglas, J., Sigbjörnsson, R., Berge-Thierry, C.,
Suhadolc, P., Costa, G., Smit, P.M., 2004. Dissemination of
European strong-motion data, vol. 2, using strong motion
datascape navigator (CD).
Antolik, M., Dreger, D., 2003. Rupture process of the 26 January 2001
M
w
7.6 Bhuj, India, earthquake from teleseismic broadband data.
Bull. Seismol. Soc. Am. 93 (3), 12351248.
Baker, C., Hatzfeld, D., Lyon-Caen, H., Papadimitriou, E., Rigo, A.,
1997. Earthquake mechanisms of the Adriatic Sea and western
Greece. Geophys. J. Int. 131, 559594.
Benetatos, Ch., Kiratzi, A., Papazachos, C., Karakaisis, G., 2004.
Focal mechanisms of shallow and intermediate depth earthquakes
along the Hellenic Trench. J. Geodyn. 37, 253296.
Benetatos, C., Kiratzi, A., Roumelioti, Z., Stavrakakis, G., Drakatos,
G., Latoussakis, I., 2005. The 14 August 2003 Lefkada Island
(Greece) earthquake: focal mechanisms of the mainshock and of
the aftershock sequence. J. Seismol. 9, 171190.
Boore, D.M., Sims, J.D., Kanamori, H., Harding, S., 1981. The
Montenegro, Yugoslavia, earthquake of April 15, 1979: source
orientation and strength. Phys. Earth Planet. Inter. 27,
133142.
Console, R., Favali, P., 1981. Study of the Montenegro earthquake
sequence (MarchJuly 1979). Bull. Seismol. Soc. Am. 71 (4),
12331248.
Das, S., Suhadolc, P., 1996. On the inverse problem for earthquake
rupture: the Haskell-type source model. J. Geophys. Res. 101 (B3),
57255738.
Dreger, D., Kaverina, A., 2000. Seismic remote sensing for the
earthquake source process and near-source strong shaking: a case
study of the October 16, 1999 Hector Mine earthquake. Geophys.
Res. Lett. 27, 19411944.
Hartzell, S.H., Heaton, T.H., 1983. Inversion of strong ground motion
and teleseismic waveform data for the fault rupture history of the
1979 Imperial Valley, California, earthquake. Bull. Seismol. Soc.
Am. 73, 15531583.
Helmberger, D.V., 1974. Generalized ray theory for shear dislocation.
Bull. Seismol. Soc. Am. 64, 4564.
Hurtig, E., Grosser, H., Neunhöfer, Vucinić, S., 1980. The 15.4.1979
earthquake in Yugoslavia. Aftershock studies, Proc. Of the 17th
Assembly of the ESC, Budapest, pp. 101104.
Kanamori, H., 1979. Seismological aspects of intra-continental
earthquakes. Proc. of the International Research Conference on
Intra-Continental Earthquakes, 1721 September 1979, Ohrid,
Yugoslavia, pp. 117127.
Karakaisis, G.F., Karakostas, B.G., Papadimitriou, E.E., Papazachos,
B.C., 1985. Properties of the 1979 Monte Negro (Southwest
Yugoslavia) Seismic Sequence. PAGEOPH 122, 2535.
Kaverina, A., Dreger, D., Price, E., 2002. The combined inversion of
seismic and geodetic data for the source process of the 16 October
1999 M
w
7.1 Hector Mine, California earthquake. Bull. Seismol.
Soc. Am. 92 (4), 12661280.
Kennett, B.L.N., Engdahl, E.R., 1991. Travel times for global
earthquake location and phase identification. Geophys. J. Int.
105, 429465.
Kiratzi, A., Louvari, E., 2001. Source parameters of the Izmit Bolu
1999 (Turkey) earthquake sequences from teleseismic data. Ann.
Geophys. (former Annali di Geofisica) 44, 3347.
Kociaj, S., Sulstarova, E., 1980. Some features of the foreshocks and
aftershocks of the earthquake of April 15, 1979. Proc. of the 17th
ESC Assembly, Budapest 1980, pp. 115119.
Langston, C., Helmberger, D.V., 1975. A procedure for modelling
shallow dislocation sources. Geophys. J. R. Astron. Soc. 42,
117130.
Oldow, J., Ferranti, L., Lewis, D., Campbell, J., D'Argenio, B.,
Catalano, R., Pappone, G., Carmignani, L., Conti, P., Aiken, C.,
2002. Active fragmentation of Adria, the North African promon-
tory, central Mediterranean orogen. Geology 30 (9), 779782.
Papazachos, B., Papazachou, C., 2003. The Earthquakes of Greece,
3rd edition. Ziti Publications, Thessaloniki, p. 286.
Papazachos, B., Savvaidis, A., Papazachos, C., Papaioannou, Ch.,
Kiratzi, A., Muco, B., Kociu, S., Sulstarova, E., 2001. Atlas of
isoseismal maps for shallow earthquakes in Albania and
surrounding area. Publ. Dep. Geophys. 10, 57.
Papazachos, B., Scordilis, E., Panagiotopoulos, D., Papazachos, C.,
Karakaisis, G., 2004. Global relations between seismic fault
parameters and earthquakes moment. Proc of the 10th International
Congress of Geological Society of Greece, 1417 April,
Thessaloniki, Greece, pp. 14821489.
Petrovski, J., Petrovski, D., Naumoski, N., Zelenovic, V., 1979. Strong
motion characteristics and acceleration distribution during Mon-
tenegro April 15, 1979 earthquake. Proc. of the International
Research Conference on Intra-Continental Earthquakes, 1721
September 1979, Ohrid, Yugoslavia, pp. 137155.
Roumelioti, Z., Kiratzi, A., 2002. Stochastic simulation of strong-
motion records from the 15 April 1979 (M 7.1) Montenegro
earthquake. Bull. Seismol. Soc. Am. 92 (3), 10951101.
Saikia, C.K., 1994. Modified frequency-wavenumber algorithm for
regional seismograms using Filon's quadrature; modelling of
Lg waves in eastern North America. Geophys. J. Int. 118,
142158.
Sulstarova, E., 1980. The focal mechanism of the April 15, 1979
earthquake sequence, aftershock studies. Proc. of the 17th
Assembly of the ESC, Budapest, pp. 161165.
Tromans, I.J., Bommer, J.J., 2002. The attenuation of strong-motion
peaks in Europe. Proceedings of the 12th European Conference on
Earthquake Engineering, London, Paper, vol. 394.
Wald, D.J., Heaton, T.H., 1994. Spatial and temporal distribution of
slip for the 1992 Landers, California, earthquake. Bull. Seismol.
Soc. Am. 84, 668691.
Wells, D.L., Coppersmith, K.J., 1994. New empirical relationships
among magnitude, rupture length, rupture width, rupture area, and
surface displacement. Bull. Seismol. Soc. Am. 84, 9741002.
Wessel, P., Smith, W.H.F., 1998. New improved version of the Generic
Mapping Tools released. EOS Trans. AGU 79, 579.
Zwick, P., McCaffrey, R., Abers, G., 1994. MT5 Program, IASPEI
Software Library, 4.
143C. Benetatos, A. Kiratzi / Tectonophysics 421 (2006) 129143
... The Southern Adriatic is one of the major seismogenic areas of the Central Mediterranean region, characterized by the strongest historically known and instrumentally recorded earthquakes in Croatia and Montenegro Benetatos and Kiratzi, 2006;Govorčin et al., 2020;Schmitz et al., 2020). Recently in this area, Faivre et al. (2021aFaivre et al. ( , 2021b reconstructed the RSL changes of the Elafiti islands in the Dubrovnik archipelago, which are attributed to the Holocene seismotectonic activity (Fig. 1). ...
Article
New high-resolution relative sea-level (RSL) proxy data obtained from Lithophyllum rims in the Adriatic allow us to distinguish major local, regional and global RSL driving processes during the past two millennia. RSL change on the Elafiti islands in the Dubrovnik archipelago (Southern Adriatic) has been significantly affected by local tectonic contributions, which vary spatially and increase southeastwards from Jakljan to Grebeni. Consequently, the RSL change on northwestward islands, Jakljan and ˇSipan, is still dominantly driven by linear regional glacioisostatic adjustment (GIA) processes estimated at ~0.34 mm/yr. However, GIA effects are occasionally cancelled out by local, non-linear coseismic uplifts of small magnitude and by variations in the global mean sea level (GMSL). Thus the rapid fall in GMSL of 0.26 mm/yr offset the GIA effects between approximately ~1000–1250 cal CE, resulting in temporal hiatuses in algal rim formation. After ~1800 cal CE, GIA rates were significantly amplified by GMSL rise, which exceed 0.9 mm/yr at Jakljan and ˇSipan and goes up to 1.4 mm/yr at Koloˇcep and Grebeni, confirming the acceleration of RSL rise over the past two hundred years. The new GMSL records from the northern and southern Adriatic, allow us to reconstruct the first highresolution GMSL curve for the Adriatic. We show that GMSL variability is in-phase with solar activity during the last two millennia, acting on cycles of ~350, 220 and 100-yrs. We also show that increased GMSL fall, which stopped the formation of algal rims around ~1000 and ~1600 cal CE, coincides with the global reduction in radiative forcing associated with the Oort and Maunder minima, with a drop in sea surface temperature (SST) and with increased salinity. Thus, our analyses revealed a consistent periodicity between the Adriatic GMSL signal, solar forcing, SST and salinity, with the most important cycle being the 350-year Great Solar Cycle.
... Faulting structures in the region have been highly active, with notable events recorded in historical records. The largest of these being the 1979 M w 6.9 event occurring ∼100km north of the coastal city of Durrës (Benetatos & Kiratzi, 2006). In recent years, the occurrence of numerous moderate-large magnitude events (2016,2017,2018) Stations were deployed from approximately two weeks after the mainshock. ...
Preprint
Full-text available
Machine Learning (ML) methods have demonstrated exceptional performance in recent years when applied to the task of seismic event detection. With numerous ML techniques now available for detecting seismicity, applying these methods in practice can help further highlight their advantages over more traditional approaches. Constructing such workflows also enables benchmarking comparisons of the latest algorithms on practical data. We combine the latest methods in seismic event detection to analyse an 18-day period of aftershock seismicity for the $M_{w}$ 6.4 2019 Durr\"es earthquake in Albania. We test two phase association-based event detection methods, the EarthQuake Transformer (EQT; Mousavi et al., 2020) end-to-end seismic detection workflow, and the PhaseNet (Zhu & Beroza, 2019) picker with the Hyperbolic Event eXtractor (Woollam et al., 2020) associator. Both ML approaches are benchmarked against a data set compiled by two independently operating seismic experts who processed a subset of events of this 18-day period. In total, PhaseNet & HEX identifies 3,551 events, and EQT detects 1,110 events with the larger catalog (PhaseNet & HEX) achieving a magnitude of completeness of ~1. By relocating the derived catalogs with the same minimum 1D velocity model, we calculate statistics on the resulting hypocentral locations and phase picks. We find that the ML-methods yield results consistent with manual pickers, with bias that is no larger than that between different pickers. The achieved fit after relocation is comparable to that of the manual picks but the increased number of picks per event for the ML pickers, especially PhaseNet, yields smaller hypocentral errors. The number of associated events per hour increases for seismically quiet times of the day, and the smallest magnitude events are detected throughout these periods, which we interpret to be indicative of true event associations.
Article
The Southern Adriatic is recognized as one of the most seismically active regions of the Central Mediterranean. It hosted the strongest historically known and instrumentally recorded earthquakes in Croatia and Montenegro. We conducted a detailed study of algal rims and tidal notches along the Dubrovnik–Konavle coastal area, aimed to contribute in reconstruction of relative sea level (RSL) change, palaeoearthquakes and neotectonics during the last 4500 years. The RSL reconstruction based on high-resolution geochronology showed that the spatial vari ability of RSL change is largely controlled by local tectonics, leading generally to an alternation of periods of RSL rise and rapid falls caused by coseismic uplift events. During the interseismic periods RSL rose at rates from 1.03 to 0.7 mm/yr before AD, followed by a slowdown to ~0.43 mm/yr, and reaching ~1.3 mm/yr after 1850s. The only exception relates to a period of slow RSL drop probably related to the 3.2 ka cold event. Rapid RSL falls caused by coseismic uplift indicate the occurrence of several large-scale events during the 4th–6th centuries AD, the 800–1360 cal AD period, and in 1395, with hints of earlier events between 1395 cal BC and 565 cal AD. These instances correspond to either unknown or insufficiently documented earthquakes. Conversely, the earthquakes in 1520 and 1667 have been clearly identified. The most compelling evidence of coseismic uplift is associated with the 1667 Dubrovnik earthquake, with an estimated uplift of 40–60 ±15 cm documented along 40 km of coastline. Analyses of multiple sites with the same high-resolution marker show that Holocene tectonic formations vary along different fault segments with minimum uplift rates ranging from ~0.3 to ~0.9 mm/yr over ~1.5 ka. The main seismogenic source responsible for coseismic uplifts is attributed to the Dalmatian unit basal thrust and its NE-dipping splays at the Adria-Dinarides boundary.
Article
Full-text available
This work presents the application of timber-based retrofitting techniques to a case-study stone masonry church featuring a wooden roof from 18th century. From the static point of view, the original roof structure presented a number of undersized structural elements, and its members were poorly or not connected among each other and to the masonry, making the church vulnerable to seismic loads as well. Thus, the roof was retrofitted with wood-based techniques, including an overlay of plywood panels, against seismic actions. These affordable, rapid, easily realizable interventions enabled both the conservation and seismic retrofitting of the roof, providing an adequate load-carrying capacity for static loads, and an effective diaphragm action against seismic loads. The conducted numerical analyses showed that the realized interventions greatly improve the seismic behaviour of the building. Besides, when the additional energy dissipation provided by the plywood panels overlay is taken into account in the numerical model, the church would even potentially be able to fully withstand the expected seismic action of the site.
Article
Outer Albanides experienced a seismic sequence starting on 21 September 2019, with an Mw 5.6 earthquake, considered a foreshock, and culminated with the mainshock on 26 November 2019, followed by a paramount aftershock activity. We propose a model for the co-seismic slip distribution using InSAR, permanent, and campaign GNSS measurements. We tested two hypotheses: an earthquake on a thrust plane with the direction N160° and along with a back thrust. By varying the depth and dip angle for the first hypothesis and only the dip angle for the second hypothesis, we concluded the optimal solution is a blind thrust at a 15-km depth dipping eastward 40°, a maximum slip of 1.4 m, and an Mw 6.38. A GNSS time series obtained after 2020 shows two slow slip events (SSEs): the first one is 200 days after the mainshock up to 26 days, and the second one is 300 days after the mainshock up to 28 days. We tested three hypotheses: SSE along the basement thrust where the mainshock has been localized, SSE along the flat formed by the detachment layer of the cover, and SSE along these two faults. We concluded that SSE occurred along the detachment layer or along the two faults.
Article
Full-text available
Kosovo is located in a key position in the central-west part of the Balkans providing an opportunity to understand the far-field effects of distributed intracontinental deformation caused by the Aegean extension in the south and Adriatic compression in the west. It is also situated along the NE-SW trending Shkodër-Pejë transverse zone, where the Dinarides and Albanides-Hellenides orogenic belts are juxtaposed. While the instrumental seismicity of the country indicates the activity of this fault zone and many others, the active faults in the country were not discussed in detail in the current literature. In this study, we analysed both the geomorphic and structural features of major mountain front faults in western Kosovo (i.e., Pejë, Istog, Krojmië and Prizen faults) to reveal the relative assessment of their activities and kinematic characters. Geomorphic and morphometric analyses of all the studied four different mountain fronts indicated high activity and tectonic uplift rates of over 0.5 mm/a. On the other hand, according to the collected kinematic data from the observed fault planes, all the studied faults are of normal character representing a dominance of NW-SE-directed extension in western Kosovo, which is most probably caused by the rollback of subducting slab in the Hellenic trench.
Article
The western Balkans occupy a region influenced by two major active tectonic processes: the collision between the Adriatic Region and the Dinarides in the west, and the extension of the Aegean Region and its surroundings as they move towards the Hellenic Trench. An understanding of the kinematics and dynamics of the western Balkans has significance for our understanding of continental tectonics in general, and is the object of this paper. The region is rich in observational data, with many well-studied earthquakes, good geodetic coverage by GNSS, and abundant exposure of active faulting and its associated geomorphology, especially within the Mesozoic carbonates that cover large sectors of the extensional areas. We first use such observations to establish the regional kinematic patterns, by which we mean a clarification of how active faulting achieves the motions observed in the deforming velocity field obtained from GNSS measurements. We then use geomorphological observations on the evolution of drainage systems to establish how kinematic and faulting patterns have changed and migrated during the Late Neogene-Quaternary. The kinematics, and its evolution, can then be used to infer characteristics of the dynamics, by which we mean the origin and effect of the forces that control the overall deformation. The principal influences are: (a) the distribution and evolution of gravitational potential energy (GPE) contrasts arising from crustal thickness variations and elevation, in particular the growth of topography by shortening in the Albanides-Hellenides mountain ranges and the high elevation of mainland Greece relative to the Mediterranean sea floor; (b) the ability of the boundaries of the region, along the Adriatic coast and in the Hellenic Trench, to support the forces arising from those GPE contrasts. The evolution in space and time indicates an interaction between the anisotropic strength fabric of the upper crust associated with faulting, and the more distributed and smoother patterns of flow that are likely to characterise the ductile deformation of the lower, aseismic part of the lithosphere — both of which influence the deformation on the scale of 100–200 km. The persistent argument about whether continental deformation is best described by a continuum or by rigid-block motions is largely a matter of scale and particular location: both are influential in establishing the patterns we see.
Article
Full-text available
We describe two 5–7 km long normal fault scarps (NFSs) occurring atop fault-related anticlines in the coastal ranges of the Dinarides fold-and-thrust belt in southern Montenegro, a region under predominant contraction. Both NFSs show well-exposed, 6–9 m high, striated, and locally polished fault surfaces, cutting uniformly northeastward-dipping limestone beds at high angles and documenting active faulting. Sharply delimited ribbons on free rock faces show different colors, varying karstification, and lichen growth and suggest stepwise footwall exhumation, which is typical of repeated normal faulting during earthquake events. Displacements, surface rupture lengths, and geometries of the outcropping fault planes imply paleoearthquakes with Mw≈6 ± 0.5 and slip rates of ∼ 0.5–1.5 mm yr−1 since the Last Glacial Maximum. This is well in line with (more reliable, higher-resolution) slip rates based on cosmogenic 36Cl data from the scarps for which modeling suggests 1.5 ± 0.1 mm yr−1 and 6–15 cm slip every 35–100 years during the last ∼ 6 kyr. The total throw on both NFSs – although poorly constrained – is estimated to ∼ 200 m and offsets the basal thrust of a regionally important tectonic unit. The NFSs are incipient extensional structures cutting (and postdating emplacement of) the uppermost Dinaric thrust stacks down to an unknown depth. To explain their existence in a region apparently under pure contraction, we consider two possibilities: (i) syn-convergent NFS development or – less likely – (ii) a hitherto undocumented propagation of extensional tectonics from the hinterland. Interestingly, the position of the extensional features documented here agrees with geodetic data, suggesting that our study area is located broadly at the transition from NE–SW-directed shortening in the northwest to NE–SW-directed extension to the southeast. While the contraction reflects ongoing Adria–Europe convergence taken up along the frontal portions of the Dinarides, the incipient extensional structures might be induced by rollback of the Hellenic slab in the southeast, whose effects on the upper plate appear to be migrating along-strike of the Hellenides towards the northwest. In that sense, the newly found NFSs possibly provide evidence for a kinematic change of a thrust belt segment over time. However, with a significantly higher probability, they can be regarded as second-order features accommodating geometrical changes in the underlying first-order thrust faults to which they are tied genetically.
Chapter
The December 29, 2020, Petrinja earthquake highlighted the significance of maintenance of buildings in seismically active regions. Many family houses in rural parts of central Croatia were damaged by the Petrinja earthquake. During the earthquake, the insufficient response of the buildings in the rural regions resulted in the majority of human injuries and fatalities. These deaths and catastrophic property losses in rural areas have underlined the importance of earthquake prevention and mitigation in these communities. A study of the seismic performance of structures in rural areas, as well as an evaluation of seismic risk and susceptibility, were conducted with the aim to improve the overall capability to prevent seismic disasters in rural areas. The structural characteristics, construction method, and application of seismic codes and building rules of two rural villages in the Osijek-Baranja County were researched, and they can be considered representative of the continental area of Croatia and surrounding countries. GIS technology is used to process and present information on the location, distribution, and other characteristics of structures.
Conference Paper
Full-text available
On the basis of seismotectonic anlysis we distinguished two fault zones of the crust: The Ionian-Adriatic fault zone and the zone of the Shkodra depressional faults. The earthquake of April 15, 1979 was generated from th first fault zone of the crust that is located on the outer side of the orogene of the Dinaride-Albanide-Helenide branch of the Mediterranean folded belt, at the border with platform. The seismogene block activated during the earthquake of April 15, 1979 was the underwater Ulqin-Budva bock with a length of 40 km.
Article
Full-text available
On the basis of the seismotectonic analysis we distinguish two deep fault zones of the crust: The Ionian-Adriatic fault zone and the zone of Shkodra depressional faults. The earthquake of April 15, 1979 was generated from the first deep fault zone of the crust that is located on the outer side of the orogene, of the Dinaride-Albanide-Hellenide branch of the folded alpine belt of the Mediterranean, at the border with the platform. The seismogene block activated during the earthquake was the under water Ulqin-Budva block with a length of 40 km.
Research
Full-text available
Macroseismic information for 57 events that occurred in Albania and the surrounding area between 1851 and 1990. All macroseismic intensities are in the MSK-64 scale and correspond to shallow (h<60km) medium up to large (7.1>M³4.5) earthquakes of this area (approximately 39.5oN-42.3oN, 19.2oE-21oE) for which an adequate number of reliable macroseismic information was available.
Article
The focal mechanism solution of the April 15, 1979 earthquake demonstrates that compressional forces are responsible for the seismic activity along the Adriatic coast. The nodal plane, identified with the plane of faulting strikes in the direction N 46°W and has a large dip. The mechanism corresponds to an almost vertical reverse faulting with dipslip motion, in which the western block overrides the eastern one. The focal mechanism solution has shown that this earthquake sequence was the result of the reactivation of the deep-faulted Ionian-Adriatic zone between the orogene and the stable foreland.
Article
Generalized ray expansions of the P, SH, and SV displacement potentials resulting from a point-source dislocation are evaluated at the surface of a layered half-space. The Cagniard-de Hoop technique is used to obtain the transient response. The results of this analysis are used to construct synthetic seismograms for a shear dislocation on a vertical fault plane. Comparisons of synthetic and observed seismograms for the Borrego Mountain earthquake (April 9, 1968) at teleseismic distances indicate an equivalent point-source depth of 9km with the far-field time function approximated by a step function with an exponential decay. This time function fits both the P and S wave forms. The apparent shift in corner frequency between the P and S waves for shallow events, as reported by some investigators, is explained by surface reflections. 26 refs.