Prehistoric Seismic Basin Effects in The

Download as pdf or txt
Download as pdf or txt
You are on page 1of 38

The Ministry of National Infrastructures

Geological Survey of Israel

Prehistoric Seismic Basin Effects


in the Dead Sea Pull-apart

Z.B. Begin1, J.N. Louie2, S. Marco3, Z. Ben-Avraham3

1. Geological Survey of Israel


2. University of Nevada, Reno
3. Tel Aviv University

Report GSI/04/05 Jerusalem, June 2005


The Ministry of National Infrastructures
Geological Survey of Israel

Prehistoric Seismic Basin Effects


in the Dead Sea Pull-apart

Z.B. Begin1, J.N. Louie2, S. Marco3, Z. Ben-Avraham3

1. Geological Survey of Israel


2. University of Nevada, Reno
3. Tel Aviv University

Report GSI/04/05 Jerusalem, June 2005

1
Abstract

Site effect is the specific response to earthquakes that is characteristic of the attributes of
a site. The two- and three-dimensional shape of sedimentary basins may constitute an
important factor of site effects. In sediment-filled basins, in which a lens of soft
sediments overlies rocks with higher seismic velocities, two-dimensional resonance
patterns may prolong the duration of shaking, and induce a large amplification, much
larger than the one predicted from the corresponding one-dimensional analysis. The main
source of these phenomena is the development within the basins of surface waves,
including the vertically and elliptically polarized Rayleigh waves, and horizontally
polarized Love waves.

The Dead Sea Fault is an active left-lateral transform, posing a major seismic threat to the
population in its surroundings, and hence an earthquake-mitigating Building Code was
adopted in Israel. However, the present Code does not deal with basin effects. The
importance of these is here exemplified in the Dead Sea pull-apart, that generated a
destructive M6.2 earthquake in 1927. Within it, seismites dated to 60-16 kyBP represent
strong earthquakes, and their thickness is significantly different in three sites, 20 km
apart: one near Massada and two in the ‘Amiaz basin. The median thickness is 3, 6 and
12 cm, respectively, presumably reflecting differences in both the duration and
amplification of earthquakes.

These differences are plausibly explained by assuming that they are caused by basin
effects, due to the differences in both the shape of the basins in which the seismites are
located (the basin being deeper at the ‘Amiaz sites than at the Massada site), as well as
difference in location of the seismites, relative to the midline of the ‘Amiaz basin. Hence,
these seismites may constitute the longest and most detailed known record of prehistoric
site response.

2
Contents

1. Introduction ……………………………………………………….………………….. 1
2. The nature of basin effects …………………..………………………………………. 3
3. Geological setting …………………………………………………………………….. 8
4. The Lisan seismites ………………………………………………………………….. 8
4.1 Ground acceleration associated with the Lisan seismites
4.2 The significance of the thickness of the Lisan seismites
4.3 Thickness of the Lisan seismites and the possible role of site effects
5. Expected basin effects in the Massada and Ami'az areas …………………………… 16
5.1 One-dimensional calculation of site effects
5.2 The two-dimensional configurations of the Massada and Ami’az basins

6. Discussion …………………………...……………………………………………… 22
References ……………………………...………………………………………….… 25

3
Figures

Figure 1: Location map.


Figure 2. Simulation of basin effects in the Reno area.
Figure 3: Simulation of basin effects in the Los Angeles basin.
Figure 4: The effect of the Los Angeles basin on the duration of ground motion.
Figure 5: Basin effects in the Seattle basin.
Figure 6: A seismite in the Lisan Fm. within the Dead Sea pull-apart,
Figure 7: Thickness of Lisan seismites shown on the backdrop of lake water depth.
Figure 8: The relationship between the thickness of seismites in the Ein Gedi area and the
probability of their association with high earthquake intensity.

Figure 9: The relationship between the thickness of seismites in the PZ1 section and the
probability that they are overlain by gypsum.

Figure10: Location map.

Figure 11: Distribution of seismite thickness at the Massada and Ami’az plains.

Figure 12: Regression of the thickness of seismites between the PZ1 and PZ2 sites.

Figure 13: Thickness of graben fill in the Massada area.

Figure 14: Thickness of graben fill in the Ami’az plain.

Figure 15: Curves of depth versus shear-wave velocity for the Massada and PZ1 sites.

Figure 16: Response spectra of the Massada and PZ1 sites, based on one-dimensional
calculations.

Figure 17: A cross-section of the Dead Sea rift valley near the Massada site.

Figure 18: A cross-section of the Dead Sea rift valley near the Ami’az sites.

Figure 19: An east-west seismic profile across the Ami’az plain.

Figure 20: A cross-section of the Hula basin.

4
1. Introduction

The Dead Sea Fault (DSF) is an active fault zone forming the Arabian-Sinai plate boundary,
stretching from the spreading Red Sea to the Eastern Anatolian Fault (Fig. 1). All along the
transform it is accompanied by a conspicuous rift valley. Based on geological evidence it is
generally accepted that the DSF is a transform fault, along which the principal sense of
movement has been left lateral (Quennel, 1959; Freund, 1965; Freund et al., 1970; Garfunkel
et al., 1981; Girdler, 1990). The long-term left-lateral slip rate is at least 5 mm/year (105 km
in a maximum of 23 Ma) based on the geological evidence, and it is currently 3.3±0.4
mm/year based on GPS measurements (Wdowinski et al., 2004).

33oN
Fault
Sea

32oN
Dead

31oN

Israel-Sinai
plate

Arabian
30oN plate

0 km 50
35oE

Figure 1: Potentially active faults along the Dead Sea Fault (Bartov et al., 2002). Shaded
area denotes the Dead Sea rift valley. Arrows mark the sense of relative plate movement
along the transform.

1
The Dead Sea pull-apart is the deepest basin along the DSF, with a sedimentary Plio-
Pleistocene fill some 10 km thick, attesting to a rather rapid subsidence (Ben-Avraham,
1997). It is a tectonically active basin (Garfunkel et al., 1981), generating at present the most
intensive earthquake activity between the Gulf of Elat and the Lebanon Baqa'a, including a
destructive M 6.2 earthquake in 1927 (Shapira et al., 1993).

The Dead Sea Fault poses a major seismic threat to the population on both its sides, and an
assessment of seismic hazard in Israel is embedded in its Building Code (Israel Standard 413;
Shapira, 2002). Presently, the Code is not conservative in several aspects: It is based on a
10% probability that a certain peak ground acceleration at a site during the next 50 years will
be exceeded, but it assumes that the probability for the occurrence of strong earthquakes does
not increase with time; it includes a correction factor for soft sediments, but it does not treat
one-dimensional resonance effects in soft sediments overlying hard rocks (Zaslavsky et al.,
2002; Gvirtzman, 2004); and it does not consider basin effects.

Basin effects are expressed by amplification of ground motions and the prolonging of their
duration (see below), and thus they may cause substantial damage even if the source
earthquake is distant or moderate. In contrast with one-dimensional site effects, basin effects
can not be recognized and anticipated by analyzing ambient noise, and at least moderate
earthquakes are needed to elucidate them. However, such earthquakes are infrequent along
the DSF, and hence, the study of prehistoric basin effects can enrich our understanding of
these effects. Also, since the 100 years of instrumental record of earthquakes near the DSF is
extended by the study of the historical seismic record (considered to be complete for M>6.5
in the last 2,000 years; Shapira, 2002), this record has to be critically viewed, as it may
include several inaccurate accounts tending towards exaggeration (Ambraseys and Karcz, 1992;
Avni et al., 2002; Karcz, 2004). In an attempt to extract magnitude values from ancient
damage descriptions, basin effects should also be considered.

The aim of this report is to analyze the possibility that basin effects played an important role
in the formation of seismites in the Dead Sea basin for a long period of about 45,000 years.

2
2. The nature of basin effects
In recent years it has become evident that the two- and three-dimensional shape of
sedimentary basins may constitute an important factor of site effects. It has been shown that
in sediment-filled basins, in which a lens of soft sediments overlies rocks with higher seismic
velocities, two-dimensional resonance patterns may prolong the duration of shaking, and
induce a large amplification, much larger than the one predicted from the corresponding one-
dimensional analysis. This was shown both in simulations (Bard and Bouchon, 1980, 1985;
Hill et al., 1990; Frankel and Stephenson, 2000; Olsen, 2000) and in field measurements
(Hanks, 1975; Liu and Heaton, 1984; Field, 1996; Joyner, 2000; Frankel et al., 2002; Pratt et
al., 2003; Pancha et al., 2004).

The main source of these phenomena is the development within the basins of surface waves,
including the vertically and elliptically polarized Rayleigh waves, and horizontally polarized
Love waves. Surface waves produce the largest displacements of shaking because their
amplitudes are tied to shear velocities. The lower the propagation velocity property of the
rocks, the larger the ground motions. In sedimentary basins the P velocity is often limited to a
minimum of 1.5 km/s, the P velocity of water-saturated sand or mud. Such materials can
have shear velocities an order of magnitude lower (down to 0.05 km/s), creating shear
displacements in surface waves that are much larger than P-wave displacements.

In the literature concerning basin effects one often finds that at the edge of sedimentary
basins seismic body waves are converted to surface waves. This notion needs some
elaboration, since this process is not the same type of wave conversion as that a body wave
undergoes at a discontinuity. Surface waves in sedimentary basins are basically multiple
reflections and multiply-reflected refractions. As body waves arrive at the edge of the basin
they are refracted, and then the body wave energy is trapped within the basin, between the
free surface and the basin interface, which acts as a strong reflector. Once trapped, they are
reflected back and forth, and out of the multitude of possible reflection paths some are
horizontal. Since this process takes some time until it becomes effective, the surface waves
may develop after a delay of several seconds. Hence, the observed late arrival of surface
waves in basins may be explained not only by their lower group velocity but also by the fact

3
that they take time to effectively develop out of the multitude of possible reflections. The fact
that the wave energy is trapped in the basin, and the low frequency that is typical of surface
waves, that is associated with lesser attenuation, contribute to the relative long duration of
strong ground motion. The increased duration of strong ground motion was shown both in
simulations (Bard and Bouchon, 1985; Frankel and Stephenson, 2000; Olsen, 2000) and in
actual measurements (Field, 1996; Lomnitz, 1996; Joyner, 2000; Frankel et al., 2002; Pratt et
al., 2003; Pancha et al., 2004).

Once this development has progressed and the process of constructive interference has
completed, surface waves develop whose energy propagates at their group velocity. Since the
group velocity is never exactly equal to the shear velocity in the medium, the wave has
effectively changed its velocity and its direction of propagation. This may be regarded as
"conversion" of the body waves to a new "phase", but two differences between the two
processes should be noted. (1) The development of surface waves in a basin is very
frequency-dependent, while a conversion (say from P to S waves) at a discontinuity is not,
and (2) while surface waves develop within a basin and only after the body waves enter into
it, the conversion of body waves happens instantaneously at the edge of the basin.

The damaging aspects of basin effects are thus three-fold: high amplification relative to rock
sites out of the basins; prolonged duration of strong motions (thereby increasing the number
of seismic cycles and thus increasing seismic loading), and the formation of large-
displacement surface waves. Here we study the possible influence of basin effects on
prehistoric seismites of the Lisan Fm. in the Dead Sea pull-apart.

4
Figure 2. Map of the Reno-area 3-dimensional velocity model (left) with basins represented,
and map view of 0.5-Hz synthetic wave propagation (right). Source: Pancha et al., 2004. Red,
yellow, and green colors represent horizontal seismic shaking while blue represents vertical
shaking. The Rayleigh wave from the earthquake epicenters (star) has passed through most of
the Reno basin (center right) in this view. The basins are trapping energy and amplifying
shaking. Concentration of basin amplification is observed near both basin centers and edges,
depending on basin geometries and earthquake location.

5
Figure 3: Simulation of basin effects in the Los Angeles basin (from Olsen, 2000). Closer to
the basin center ground motions are stronger and of longer duration (seconds). Numbers in
the basin denote seismic velocity, in km/s.

Figure 4: Simulation of basin effects in the Los Angeles basin (from Olsen, 2000). Colored
contours denote strong motion duration in 10 seconds interval, with 50 seconds (green) in the
middle of the basin. Bold black line denotes the coastline and other black lines denote main
roads.

6
A

Figure 5: Basin effects (from Pratt et al., 2003) in the Seattle basin in response to a remote
earthquake in Taiwan (Chi-Chi earthquake, 1999). A - seismograms in sites across the Seattle
basin; B - amplitude, relative to points out of the basin, for certain frequencies; C - seismic
velocities in the basin.

7
3. Geological setting
The study area is located in the western margins of the Dead Sea pull-apart. The structure of
this basin has been studied quite extensively (ten Brink and Ben-Avraham, 1989; Ben-
Avraham et al., 1990; Ben-Avraham et al., 1993; ten Brink et al., 1993; Ginzburg and Ben-
Avraham, 1997; Ben-Avraham, 1997; Aldersons et al., 2003). The study sites are located in
close proximity (1-2 km) to two active faults: the Jericho fault (Begin, 1975; Reches and
Hoexter, 1981; Gardosh et al., 1990; Niemi and Ben-Avraham, 1993, 1997; Shapira et al.,
1993; Lubberts and Ben-Avraham, 2002), and the western margin fault (Bartov and Sagy,
2004).

The lacustrine Lisan Formation, consisting of authigenic aragonite and gypsum layers
alternating with silt and sand detritus (Begin et al., 1974; Katz et al., 1977; Stein 2001), was
deposited within the Dead Sea rift valley during the period of 70,000-14,000 years B.P.
(Kaufman, 1971; Kaufman et al., 1992; Schramm et al., 2000; Haase-Schramm et al., 2004).
Within the Dead Sea basin, breccia beds in the LIsan Fm. and Holocene lacustrine
sediments, formed during the last 60,000 years, were interpreted as seismites (Fig. 6),
induced by M>5.5 earthquakes (Marco and Agnon, 1995; Marco, 1996; Marco et al., 1996;
Ken-Tor et al., 2001), by strong and remote earthquakes (Migowski et al., 2004) and by
nearby M>7 earthquakes (Begin et al., 2005).

4. The Lisan seismites


4.1 Ground acceleration associated with the Lisan seismites.
The mechanism of formation of the seismites is not clear. They were originally continuous
alternating laminae of aragonite and fine detritus, lying flat on the bottom of Lake Lisan.
These undisturbed laminae were later fluidized, brecciated, partly resuspended in the Lake
water, and then resettled (Marco et al., 1996). A mechanism for the formation of the Lisan
seismites was suggested, based on laboratory experiments (Hamiel, 1999). These
experiments showed that in order to develop waves in the boundary between mud and water,
the densities of which are similar, very large acceleration (>10g) is needed. On the other
hand, partial suspension of the mud layer into the overlying water layer was induced by

8
Figure 6: A 10 cm thick seismite in the Lisan Fm. within the Dead Sea pull-apart, showing
torn white aragonite laminae in silt-clay matrix. Undisturbed aragonite laminae top the
seismite. Scale shows centimeters.

waves that developed in the water surface at low accelerations. Hence it was suggested that
the seismites were formed due to high amplitude waves in shallow water. However, this
model can not explain the formation of seismites in depths greater than several tens of
meters.

On the other hand, the depth of water under which this brecciation took place can be
estimated through knowledge of Lake Lisan level curve (Bartov et al., 2002) and the
elevation of the sites, adjusted to sedimentation rate. It arises from the history of depth in the
Ami’az plain sites that relatively thick (10-15 cm) seismites in the Lisan Formation, were
formed even at water depths of about 100m (Fig. 7).

9
50 120

Hiatus

Water Depth at PZ1, m


Hiatus
40 80
Seismite thickness, cm

30 40

20 0

10

0
20 30 40 50 60
Age, k-years B.P

Figure 7: Thickness of Lisan seismites (bars) shown on the backdrop of the water depth in
which they were formed in Lake Lisan. Depth data are after Bartov et al., 2002. Age of
seismites and their thickness is from Begin et al., 2005.

For seismites in the Dead Sea basin 20 km north of Massada, with thickness 0.2 ≤ D ≤ 13 cm,
that were matched to historical earthquakes, the associated intensity was estimated as
VI<Imm<VIII (Migowski et al., 2004), which would translate to peak ground acceleration of
about 0.1-0.3g. However, this may be an underestimate as intensities were estimated from a
general magnitude-distance relationship without considering possible basin effects of the
Dead Sea rift.

4.2 The significance of the thickness of the Lisan seismites.


Accepting that the brecciated beds in the Lisan Formation were presumably formed by
liquefaction or fluidization following earthquakes, we assume that thicker seismites may
indicate stronger earthquake intensity. This seems reasonable in light of the correlation
between the Liquefaction Severity Index and earthquake magnitude (Youd and Perkins,
1987), and because stronger earthquakes cause more stress cycles at a site (Seed and Idriss,
1982), which in turn may bring about higher pore pressure within the sediment. This

10
relationship is difficult to demonstrate (Tuttle, 2001), but it is partially supported by data on
historic seismites in the Dead Sea basin (Migowski et al., 2004). A weak relationship
(p=0.11; note that the thickness range is much smaller than the range in the Lisan Formation)
is shown between the thickness of seismites and the probability of their being associated with
earthquake intensity >VII rather than VI-VII (Fig. 8).

2
1
is associated with intensity >VII
Probability that brecciated bed

0.8

0.6

0.4

0.2 p=0.11

11 4 2 2
0

0 5 10 15
Thickness of brecciated beds, cm

Figure 8: Logistic function, showing that thicker seismites in the Ein Gedi area within
the Dead Sea pull-apart have a higher probability to be associated with earthquake
intensity >VII (triangles) rather than with intensity VI-VII (squares). The graph is based
on the matching of seismites with historical earthquakes (Migowski et al., 2004). p
denotes the statistical significance of the fitted logistic function. Note that the thickness
range here is much smaller than the range in the Lisan Formation (Fig. 9). Since p > 0.05
we do not reject the null hypothesis that earthquake intensity is not related to the
thickness of the seismites. However, since p ~ 0.1 we view the data as indicating a weak
relationship between the two attributes.

For the ~30 seismites in the Lisan Formation at the Ami’az plain, dated to 60-16 kyBP, both
the thickness of seismites and the lithology of beds directly overlying them were used in
order to differentiate between stronger and weaker earthquakes represented in these sections.
It was found that thicker seismites have a higher probability of being overlain by laminated
gypsum and detritus (Fig. 9). This association was interpreted as arising from the mixing of
Lake Lisan water column due to high seiches induced by strong earthquakes, and served to

11
identify 11 M≥ 7 earthquakes that originated within the tectonically active Dead Sea pull-
apart (Begin et al., 2005). This model highlights the significance of seismite thickness as an
indicator of earthquake intensity.

2
or by laminated gypsum and detritus 1
Probability that brecciated bed
is overlain by a detritus bed

0.8

0.6

0.4

0.2

2 2 3
0

0 10 20 30 40 50
Thickness of brecciated beds (cm)

Figure 9: Logistic function, showing that thicker seismites in the PZ1 section have a higher
probability to be overlain by beds of laminated gypsum and detritus (circles) and by detritus
beds (triangles) rather than by beds of laminated aragonite and detritus (diamonds). Digits
above symbols denote number of beds with identical thickness. Full line is for both couplet
types (p = 0.02, where p denotes the p-value obtained in testing the null hypothesis that the
layering is random) and broken line is only for couplets of seismites overlain by laminated
gypsum couplets (p = 0.05). Since p ≤ 0.05 we reject the null hypothesis that the layering
pattern is not related to the thickness of the seismites.

In addition, in a recent study (Heifetz et al., 2005), based on considerations of the Kelvin-
Helmholtz Instability, the minimum ground acceleration needed to de-stabilize Lisan beds, as
the first stage in their brecciation, was shown to change with the square root of the bed
thickness, that is, bed thickness is proportional to the square of the minimum destabilizing
acceleration.

In summary, the thickness of the Lisan seisimites at a site seems to be related to local
earthquake intensity.

12
4.3 Thickness of the Lisan Lisan seismites and the possible role of site effects.
Three columnar sections of the Lisan Formation, each exhibiting some 30 seismites were
previously presented (Marco, 1996; Marco et al., 1996). One section is in the Massada plain
(M1, coordinates 18575/07965, Israel old grid; Fig 10) and two sections, 2 km apart, in the
Ami’az plain, some 25 km southwards: PZ2 (coordinates 18500/05800) and PZ1
(18450/05599).

35.5 0

33oN

32 0 32 0
Fault
Sea

32oN
Dead

M1
PZ
31oN

30oN

0 km 50
0 km 50
35oE
35.5 0

Figure 10: Location map (in area deleniated by box) , showing the M1 and PZ sites in the
context of active faults of the Dead Sea rift (Bartov et al., 2002). Grey area is the Dead
Sea, that was also the deepest basin in Lake Lisan.

The median thickness values for the M1, PZ2 and PZ1 sections are 3, 6, and 12 cm,
respectively (Fig. 11). The difference between the thickness distributions in the three
sections, tested through the Mann-Whitney test, is statistically significant (p= 0.018 for the
PZ2-PZ1 pair; p=0.0001 for the PZ2-M1 pair; p=0.000001 for the PZ1-M1 pair. p is the
probability that the difference between the distributions occurs randomly). The thickness
distributions describe the relationship between the sites in general, as is also depicted in the

13
regression of thickness values at PZ2 on PZ1 for those seismites that were matched between
the two sites (Fig. 12). However, it can be seen in Fig. 12 that in several cases DPZ1<DPZ2,
where D is the thickness of individual seismites.

It should be noted that there was a difference in the environments of deposition of the Lisan
Fm. at the Massada and Ami’az plain sites, with the former being about 50m deeper in Lake
Lisan. However, this difference does not explain the consistently thinner seismites at
Massada, since, as shown above, 10-15 cm thick seismites were formed at the Ami’az plain
even at a depth of about 100 m (Fig. 7).

100

M1
28
80 PZ2
of number of brecciaed beds

34
PZ1
Cumulative percent

29
60

40

20

0
0 10 20 30 40 50
Thickness of brecciated beds, cm

Figure 11: Cumulative percent of the number of seismites as a function of their thickness.
Source of data: Marco, 1996, Table 2.2. M1 is in the Massada plain, PZ1 and PZ2 are in the
Ami’az plain. Numbers denote the number of seismites in each site. The three distributions
are significantly different (see text).

14
50

e
lin
1
1:
40

Thickness of brecciated beds, cm


at PZ2 30

20

10
y= 0.49x+4.4
r=0.58
p=0.001
0
0 10 20 30 40 50
Thickness of brecciated beds, cm
at PZ1

Figure 12: Comparison of thickness of seismites matched between the PZ1 and PZ2 sites.
On average, the PZ1 beds are thicker than the PZ2 beds but for several seismites DPZ1<DPZ2.

A detailed correlation between seismites in the three sites has not been established yet,
although the PZ1 section is well dated (Haase-Schramm et al., 2004) and the M1 section is
partly dated in detail (Prasad et al., 2004). However, as the seismites were formed during the
same period in the three sections, it may be assumed that, in general, the same set of
earthquakes caused the formation of seismites in the two areas. There is no reason to assume
that earthquakes preferentially occurred south of the PZ1 site, rendering it closer to most
epicenters. Moreover, it was suggested (Begin et al., 2005) that 11 strong earthquakes, that
induced high seiches in Lake Lisan, occurred in the Lake deep basin, just north of the M1
site. Also, despite the mere 2 km distance between the PZ2 and PZ1 sections, they still
exhibit a significant difference in the thickness of their seismites. These considerations lead
us to assume that the cause of the difference in thickness of seismites between the three
sections may be a consistently different response to strong earthquakes in the three sites.
Since there is no conspicuous difference in the lithology of the seismites in the three sections,
we propose that the difference in seismite thickness may be explained by some site effects.
Applying the results of Heifetz et al. (2005; see above), and assuming similar lithology at

15
both sites, the 5-fold difference in the median thickness suggests, on average, a ~2-fold
higher ground acceleration at the Ami’az sites compared to the Massada site.

There is an additional, independent indication of the importance of the different local effects
between the Massada and Ami’az areas. In the Ami’az area, numerous young (<14 kYBP)
clastic dykes have been observed (Marco et al., 2002). It was suggested, on the basis of the
magnetic fabric in these dykes, that the clastics were fluidized and injected into fissures due
to strong earthquakes (Weinberger et al., 2003; Levi et al., submitted). Most of these dykes
appear near the center-line of the Ami’az plain. Similar clastic dykes are also present in the
Massada area, but they are fewer there (Weinberger, personal communication, 2004). This
may be interpreted as signifying higher earthquake intensities in the Ami’az sites relative to
the Massada area, and this seems to be in accord with the higher earthquake intensity that
may be assumed for the Ami’az sites compared to the Massada area, based on the difference
in seismite thickness.

5. Expected site effects in the Massada and Ami'az areas


5.1 One-dimensional calculation of site effect
The M1 site lies above a 1,300 m graben fill, which is probably underlain by ~200 m of rock
salt, with their interface dipping 20o eastwards (Fig 13). The rock salt is underlain by 200 m
chalk of the Mount Scopus Group and by hard carbonates of the Judea Group. The Ami’az
sites lie above soft sediments of the Lisan and Amora formations (Fig 14), 1,100 m thick at
PZ2 and 1,800 m thick at PZ1. These are underlain by some 2,000 m of rock salt, with their
interface dipping 20o southwards.

16
0875

0800
M1

0 km 2

1850 1900

Figure 13: Thickness (meters) of graben fill in the Massada area, underlain by rock salt. The
plain is bounded on the west by the Dead Sea pull-apart western margin fault. After: Israel
National Oil Company, 1992 (Ein Gedi). Coordinates are in the Old Israel Grid.

060

00
PZ2
Outcrops 10 Sedom
of the salt
Judea diapir
Group 1500
hard
carbonates PZ1

055

0 km 2
185

Figure 14: Thickness (meters) of graben fill in the Ami’az plain, underlain by rock salt. The
plain is bounded on the west by the Dead Sea pull-apart western margin fault and in the east
by the rising Sedom diapir. Note that the fill thinns north of PZ1, with the dip of the interface
being 25o. After: The Israel National Oil Company, 1992, Fig. 10. Coordinates are in the Old
Israel Grid.

17
0
Massada PZ1

Alluvium Alluvium
1
Depth, km

Rock salt
Mainly chalk

2 Mainly Rock salt


hard carbontaes
(Judea Group)
Sandstones
(Hazeva Fm.)
3
0 0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
Shear wave velocity, km/sec Shear wave velocity, km/sec

Figure 15: Depth versus shear-wave velocity for the subsurface at Massad a
and at the PZ1 site.

Massada
Amplification ratio

PZ1

Frequency, Hz

Figure 16: Response spectra of the Massada and PZ1 sites, based on a one-dimensional
calculation (data presented in Fig. 15, and with 5% damping), showing (vertical axis) the
amplification ratio. Note that the results indicate a higher amplification at Massada, in
contrast with the expected amplification based on interpretation of the thickness of the Lisan
Fm. seismites. Calculated by Z. Gvirtzman with the Strong Motion Simulation package
(Boore, 2000).

18
5.2 The two-dimensional configurations of the Massada and Ami’az basins
Consideration of the two- and three-dimensional configuration in the areas of Massada and
Ami’az reveals an important difference between them. This is depicted in Figures 17 and 18,
which compares the geological crossections of the valleys in which the PZ and M1 sections
are located. These crossections reflect present conditions and as for the Massada site no
significant changes have taken place since the formation of the Lisan Formation seismites.
One should consider, however, some change in the Ami’az area, because the Sedom Diapir
has been rising at a rate of 6-9 mm/year in the last 14,000 years and Mount Sedom was
somewhat lower than 280 m below sea level for the period 43-28 years BP (Weinberger et
al., accepted). This means that the closure of the valley sediments was probably not complete
at the time of formation of some of the Lisan seismites. However, compared to the vertical
dimension of ~2,000 m of the valley fill this difference does not seem to significantly change
the basic configuration.

The numerical calculations of Bard and Bouchon (1985) showed that the 2-D resonance
pattern depends on both the “shape ratio” of the valley fill (ratio of maximum sediment
thickness, h, to the valley half-width, l) and the location of sites within the valley, measured
by a dimensionless parameter x/l, where x is measured from the center of the valley. They
showed that the higher the h/l value the lower the velocity contrast between the fill and the
basement that is needed for the generation of 2-D resonance effects (see also Frankel and
Stephenson, 2000 for basin resonance with a low velocity contrast of 1.5). According to Bard
and Bouchon (1985) these effects decay from the valley mid-line (x/l=0) with increasing x/l,
as indeed shown by observations in the Seattle basin (Pratt et al., 2003), but not in
observations in southern California (Field, 1996) and Reno, Nevada (Pancha et al., 2004).
Joyner (2000) suggested that the amplitude of surface waves decays with increasing distance
of the site from the edge of the basin, thus opening the possibility that, depending on the
location of the earthquake source and the location of its reflecting point on the basin edge,
amplification may at times be greater nearer to the basin edge.

19
West M1 East
0
Alluvium
2
Depth, km

4 Lisan
diapir
6 Rock salt
Carbonates
and sandstones
8

10

0 5 10 15 20
Distance, km

Figure 17: An East-West crossection across the Dead Sea rift valley along coordinate 080.
Sources: Al-Zoubi and Ten Brink, 2001; Israel National Oil Company, 1992 (Ein Gedi
permit).

West PZ1 East


0
Alluvium Sedom Alluvium
2 diapir
Rock salt
Depth, km

6
Carbonates
8 and sandstones
10

0 5 10 15 20
Distance, km

Figure 18: Crossection across the Dead Sea rift valley along coordinate 055. Modified after
Al-Zoubi et al., 2002. Note the deep Ami’az valley west of the Sedom Diapir.

20
PZ1 PZ2
West projected East

Car
bon

top sa
lt
ates

1 km
Figure 19: An east-west seismic profile across the Ami’az plain, showing location of PZ1
and PZ2 sites in relation to the valley mid-line. Vertical scale approximates the horizontal
scale. Seismic profile is printed courtesy of M. Gardosh, the Israel Geophysical Institute.

In relation to these considerations it can be seen that for the M1 site (Fig. 17), h/l=0.4 and
x/l=0.8. For the PZ sites, h/l=1.0 (Fig 18), and relative to the valley mid-line, PZ1 lies at the
center with x/l =0.1 while PZ2 is closer to the valley edge with x/l =0.7 (Figure 14 and 19). It
is also noteworthy that in the Ami’az area there might exist a 3-D effect due to the southward
dip of the base of the fill (Fig. 14). Focusing of seismic waves (Gao et al., 1996) by the
western flank of the subsurface Lisan Diapir may affect the Massada area (Fig. 17).

It may be concluded from the above that stronger seismic resonance effects are expected in
the two Ami’az sites than in the Massada site, and that in the mid-valley PZ1 site stronger
resonance is expected than in the close-to-edge PZ2 site. This is in line with the average
distribution of seismite thickness D in the three sections, where DPZ1>DPZ2>DM1 (Fig. 11).
Noting, however, that in some cases DPZ1<DPZ2, it should be remembered that the response at
a site depends not only on its structure but also on the earthquake source location, which
influences the angle, azimuth, and type of incident waves (Field, 1996; Joyner 2000).

21
Discussion
The spatial distribution of the thickness of the Lisan seismites in the three sites at the
Massada and Ami’az areas is well in accord with expected basin effects in these two areas.
This leads us to conclude that the differences in thickness of the Lisan seismites there reflect
specific basin response to strong earthquakes, a response that persisted for some 45,000 years
for a variety of earthquake sources. Since site response in general has been defined as “the
unique behavior of a site, relative to other sites, that persists given all (or most) of the
potential sources of earthquake ground motion in the region” (Field, 1996), the Lisan
seismites may represent the longest and most complete known record of basin effects.

The persistence of the basin effects on the Lisan seismites highlights the possible
significance of such effects in present earthquake hazard assessments in Israel, where basins
with high h/l ratios do exist in populated areas (see Fig. 20 for an example). These act in
addition to channeling of seismic waves, as studied In the Dead Sea rift (Gottschamer et al.,
2002). On the other hand, taking basin effects into account may result in a decrease of
estimated historical earthquake magnitudes based on assessment of local intensity.

W E

Figure 20: West– east cross-section of the Hula pull-apart basin, showing h/l ratio of about 1
(after Rybakov et al., 2003). In such basins, large basin effects are expected.

22
Although the results of this study are indicative, it should be noted that they are preliminary,
as the variations in seismite thickness in the Lisan Formation are based on three geological
sections only and the actual basin effects have not been either measured or simulated. Only
few measurements of site response were made in the Dead Sea area (Zaslavsky et al., 2000),
and a part of the problem lies in the very low ambient noise in this area. It should be noted
that in previous attempts to use the spectral ratio of horizontal versus vertical components in
order to predict site effects in sedimentary basins, fundamental resonant frequency was well
predicted but amplification was not (Dravinski et al., 1996; Field 1996).

Hence we suggest that future work on this important prehistoric seismic record should
proceed along the following lines:

(1) Field work, with the aim of acquiring better knowledge of the seismites, in an attempt to
establish matching of seismites in the three sites and obtain more information on the lateral
changes in their thickness.

(2) Field measurements of site response to current earthquakes, by deploying broad-band


seismometers at the three sites and also in nearby rock sites for reference. Since the expected
frequency in the study area is about 0.1 Hz, and since strong earthquakes are rare in the Dead
Sea area, it is may be useful to take advantage of strong but more distant earthquakes, of
which some 20-30 per year are expected to be useful for this purpose (R. Hofstetter, personal
communication, 2004).

(3) Three-dimensional simulations of earthquakes, based on a detailed, updated 3-


dimensional velocity model of the Dead Sea basin. The simulations should employ different
earthquake scenarios, with the field observations, both seismite thickness and ground-motion
data, serving as constraints on their degree of realism. These may produce hypothetical pre-
historic strong earthquakes in the Dead Sea rift in the period 60-16 kyBP. Those hypotheses
that are found incompatible with the constraints will be rejected, thus advancing our
understanding of regional earthquakes and tectonics along this important plate boundary.

23
Acknowledgements
The concept underlying this report was shaped through fruitful discussions with Z.
Gvirtzman, who also helped with one-dimensional analysis of the site response. We
benefited from many helpful suggestions by R. Amit, G. Baer, V. Lyakhovsky, A. Salamon,
R. Weinberger, and Y. Zaslavski, that led to marked improvements of the manuscript. B.
Cohen helped with the publication of the report.

24
References

Aldersons, F., Z. Ben-Avraham, A. Hofstetter, E. Kissling and T. Al-Yazjeen, 2003. Lower-


crust strength under the Dead Sea basin from local earthquake data and rheological
modeling, Earth Planetary Science Letters 214, 129-142.

Ambraseys, N.N., and Karcz, Y., 1992. The earthquake of 1546 in the Holy Land. Terra
Nova 4: 254-262.

Al-Zoubi A, and ten Brink, U.S., 2001. Salt diapirs in the Dead Sea basin and their
relationship to Quaternary extensional tectonics. Marine and Petroleum Geology 18: 779-
797.

Al-Zoubi A, Shulman, H., and Ben-Avraham, Z., 2002. Seismic reflection profiles across the
southern Dead Sea basin. Tectonophysics 346: 61-69.

Avni, R., Bowman, D., Shapira, A., Nur, A., 2002. Erroneous interpretation of historical
documents related to the epicenter of the 1927 Jericho earthquake in the Holy Land.
Journal of Seismology 6: 469-476.

Bard, P-Y. and Bouchon, M.,1980. The seismic response of sediment-filled valleys. Part 2.
The case of incident P and SV waves. Bulletin of the Seismological
Society of America 70: 1921-1941.

Bard, P-Y. and Bouchon, M.,1985. The two-dimensional resonance of sediment-filled


valleys. Bulletin of the Seismological Society of America 75: 519-541.

Bartov, Y., and Sagy, A., 2004. Late Pleistocene extension and strike slip in the Dead Sea
Basin. Geological Magazine 141: 565-572.

Bartov. Y., Sneh, A., Fleischer, L., Arad, V., and Rosensaft, M., 2002. Potentially active
faults in Israel. Geological Survey of Israel Report GSI/29/2002, 8 pp.

Bartov, Y., Stein, M., Enzel, Y., Agnon, A., Reches, Z., 2002. Lake-levels and sequence
stratigraphy of Lake Lisan, the Late Pleistocene precursor of the Dead Sea. Quaternary
Research 57: 9-21.
Begin, Z.B., 1975. The geology of the Jericho sheet. Geological Survey of Israel Bulletin 67,
35 pp.

Begin, Z.B., Ehrlich, A., Nathan, Y., 1974. Lake Lisan, the Pleistocene precursor of the Dead
Sea. Geological Survey of Israel Bulletin 63, 30 pp.

Begin, Z.B., Steinberg, D.M., Ichinose, G.A and Marco, S., 2005. A 40,000 year
unchanging seismic regime in the Dead Sea rift. Geology 33: 257-260.

25
Ben-Avraham, Z., 1997. Geophysical framework of the Dead Sea, in: Niemi, T.M., Z.
Ben-Avraham and J. Gat, (eds.), The Dead Sea: The lake and its settings, Oxford
University Press, New York: 22-35,

Ben-Avraham, Z., U. ten Brink and J. Charrach, 1990. Transverse faults at the northern end
of the southern basin of the Dead Sea graben, Tectonophysics 180: 37-47.

Ben-Avraham, Z., T. M. Niemi, D. Neev, J.K. Hall and Y. Levy, 1993. Distribution of
Holocene sediments and neotectonics in the deep basin of the Dead Sea, Marine Geology
113: 219-231.

Ben-Menahem, A., 1991. Four thousand years of seismicity along the Dead Sea rift. Journal
of Geophysical Research 96B: 20,195-20,216.

Boore, D.M., 2000. SMSIM – FORTRAN programs for simulating ground motions from
earthquakes: version 2.0 – a revision of Open File Report 96-80-A, United States
Geological Survey, Open File Report 00-509.

ten Brink, U. and Z. Ben-Avraham, 1989. The anatomy of a pull-apart basin: seismic
reflection observations of the Dead Sea basin, Tectonics 8: 333-350.

ten Brink, U.S., Z. Ben-Avraham, R.E. Bell, M. Hassouneh, D.F. Coleman, G. Andreasen,
G. Tibor, and B. Coakley, 1993. Structure of the Dead Sea pull-apart basin from gravity
analyses, Journal of Geophysical Research 98: 21,877-21,894.

Dravinski, M., Ding, G., and Wen, K.-L., 1996. Analysis of spectral ratios for estimating
Ground motion in deep basins. Bulletin of the Seismological Society of America 86:
646-654.

Field, E.H., 1996. Spectral amplification in a sediment-filled valley exhibiting clear basin-
edge-induced waves. Bulletin of the Seismological Society of America 86: 991-1005.

Frankel, A.D., and Stephenson, W., 2000. Three-dimensional simulations of ground motions
in the Seattle region for earthquakes in the Seattle fault zone. Bulletin of the
Seismological Society of America 90: 1251-1267.

Frankel, A.D, Carver, D.L. and William, R.A., 2002. Nonlinear and linear site response and
basin effects in Seattle for the M 6.8 Nisqually, Washington, earthquake. Bulletin of the
Seismological Society of America 92: 2090-2109.

Freund, R., 1965. A model of the structural development of Israel and adjacent areas since
upper Cretaceous times. Geological Magazine 102: 189-205.

Freund, R., Garfunkel, Z., Zak, I., Goldberg, M., Weissbrod, T., and Derin, B., 1970. The
shear along the Dead Sea rift. Philosophical Transactions of the Royal Society of
London, A, 267: 107-130.

26
Frieslander, U., 1998. Velocity surveys. Israel Oil Company Ltd, Emunah-1, Appendix 8.

Gao, S., Liu, H., Davis, P.M., and Knopoff, L (1996). Localized amplification of seismic
waves and correlation with damage due to the Northridge earthquake. Bulletin of the
Seismological Society of America 85: S209-S230.

Gardosh, M., Reches, Z., and Garfunkel, Z., 1990. Holocene tectonic deformation along the
western margins of the Dead Sea. Tectonophysics 180: 123-137.

Garfunkel, Z., Zak, Y., and Freund, R., 1981. Active faulting in the Dead Sea rift:
Tectonophysics 80: 1-26.

Ginzburg A. and Z. Ben-Avraham, 1997. A seismic refraction study of the north basin of the
Dead Sea, Israel, Geophysical Research Letters 24, 2063-2066.

Girdler, R.W., 1990. The Dead Sea transform fault system. Tectonophysics 180: 1-13.

Gottschammer, E., F. Wenzel, H. Wust-Bloch and Z. Ben-Avraham, 2002. Earthquake


modeling in the Dead Sea Basin, Geophysical Research Letters 29, No. 12,
10.1029/2001GL013800.

Grosowicz, Y., 1979. Ein Gedi 3 Prospect. Tel Aviv, Oil Exploration (Investments), Report
78/31, 8 pp.

Gvirtzman, Z., 2004. Ground motion amplification in the Israeli Foothills: Empiric relations
between resonance measurements of ambient noise and geological structure. Geological
Survey of Israel Report GSI/17/04, 42 pp. (in Hebrew, with an English abstract).

Haase-Schramm, A., Goldstein, S.L., and Stein, M., 2004. U-Th dating of Lake Lisan
aragonite (late Pleistocene Dead Sea) and implications for glacial East Mediterranean
climate change: Geocimica et Cosmochimica Acta, v. 68, p. 985- 1005.

Hamiel, Y., 1999. Dynamic behavior under vibration: Laboratory experiments with
sediments, wave propagation in non-linear elastic medium and possible implications to
earthquakes. M.Sc. Thesis, The Hebrew University of Jerusalem, 61pp (in Hebrew).

Hanks, T.C., 1975. Strong ground motion of the San Fernando California earthquake: ground
displacements. Bulletin of the Seismological Society of America 65: 193-225.

Heifetz, E., Agnon, A., and Marco, S., 2005. Soft sediment deformation by Kelvin Helmholtz
Instability: A case from Dead Sea earthquakes. Earth and Planetary Science Letters (in
press).

27
Hill, J., Benz, H., Murphy, M., and Schuster, G., 1990. Propagation and resonance of SH
waves in the Salt Lake valley, Utah. Bulletin of the Seismological Society of America 80:
23-42.

Israel National Oil Company, 1992. Ami’az East-1 completion report.

Israel National Oil Company, 1998. Emunah-1, geological completion report

Joyner, W.B., 2000. Strong motion from surface waves in deep sedimentary basins. Bulletin
of the Seismological Society of America 90: S95-S112.

Karcz, I., 2004. Implications of some early Jewish sources for estimates of earthquake hazard
in the Holy Land. Annals of Geophysics 47: 759-792.

Kashai, E., 1980. Mezada 2, recommendation for drilling. Oil Exploration (Investment) Ltd,
Report 80/45, 22pp.

Katz, A. Kolodny, Y., Nissenbaum, A., 1977. The geochemical evolution of the Pleistocene
Lake Lisan-Dead Sea system. Geochimica et Cosmochimica Acta 41: 1609-1626.
Kaufman, A., 1971. U-series dating of Dead Sea basin carbonates. Geochimica et
Cosmochimica Acta 35: 1269-1281.

Kaufman, A., Yechieli, Y., and Gardosh, M., 1992. Reevaluation of the lake-sediment
chronology in the Dead Sea basin, Israel, based on new 230Th/U dates. Quaternary
Research 38: 292-304.

Ken-Tor, R., Agnon, A., Enzel, Y., Stein, M., Marco, S. and Negendank, J.F.W., 2001.
High-resolution geological record of historic earthquakes in the Dead Sea basin. Journal
of Geophysical Research 106, 2221-2234.

Larsen, B.D., Ben-Avraham, Z., and Shulman, H., 2002. Fault and salt tectonics in the
southern Dead Sea basin. Tectonophysics 346: 71-90.

Levi, T., Weinberger, R., Aifa, T., Eyal, Y., and Marco, S., Earthquake-induced clastic dikes
in the Dead Sea rift detected by anisotropy of magnetic susceptibility. (submitted).

Liu, H-L., and Heaton, T., 1984. Array analysis of the ground velocities and accelerations
from the 1971 San Fernando, California earthquake. Bulletin of the Seismological
Society of America 74: 1951-1968.

Lomnitz, C., 1996. The gravielastic equation and the emergence of gravity waves in large
earthquakes. Bulletin of the Seismological Society of America 86: 1120-1228.

Lubberts, R. K. and Z. Ben-Avraham, 2002. Tectonic evolution of the Qumran basin from
high-resolution 3.5-kHz seismic profiles and its implication for the evolution of the
northern Dead Sea basin., Tectonophysics 346: 91-113.

28
Marco, S., 1996. Paleomagnetism and paleoseismology in the Late Pleistocene Dead Sea
graben. Ph.D Thesis, The Hebrew University of Jerusalem, 95 pp.

Marco, S. and Agnon, A., 1965. Prehistoric earthquake deformations near Masada, Dead Sea
graben. Geology 23: 695-698.

Marco, S., Stein, M., Agnon, A. and Ron, H., 1996. Long-term earthquake clustering: a
50,000-year paleoseismic record in the Dead Sea graben. Journal of Geophysical
Research 101B: 6179-6191.

Marco, S., Weinberger, R., and Agnon, A., 2002. Radial clastic dykes formed by a salt diapir
in the
Dead Sea rift, Israel. Terra Nova 14: 288-294.

Migowski, C., Agnon, A., Bookman (Ken-Tor), R., Negendank J.F.W. and Stein, M., 2004.
Recurrence pattern of Holocene earthquakes along the Dead Sea Transform revealed by
varve-counting and radiocarbon dating of lacustrine sediments. Earth and Planetary
Science Letters. 222: 301-314.

Niemi, T.M. and Z. Ben-Avraham, 1994. Evidence of Jericho earthquakes from slumped
sediments in the Jordan River delta, Geology 22: 395-398.

Niemi, T.M. and Ben-Avraham, Z., 1997. Active tectonics in the Dead Sea basin, in: Niemi,
T.M., Ben-Avraham, Z. and Gat, J., (eds.), The Dead Sea: The lake and its settings,
Oxford University Press, New York, p. 73-81.

Olsen, K.B., 2000. Site amplification in the Los Angeles basin from three-dimensional
modeling of ground motion Bulletin of the Seismological Society of America 90: S77-
S94.
Pancha, A., J. G. Anderson, J. N. Louie, A. Anooshehpoor, and G. Biasi, 2004. Data and
simulation of ground motion for Reno, Nevada: presented at 13th World Conference on
Earthquake Engineering, Vancouver, B.C., Aug. 1-6, paper no. 3452.

Prasad, S., Vos, H., Negendank, J.F.W., Waldmann, N., Goldstein, S., and Stein, M., 2004.
Evidence from Lake Lisan of solar influence on decadal –to centennial-scale climate
variability during Marine Oxygen isotope Stage 2. Geology 32: 581-584.

Pratt, T.L., Brocher, T.M., Weaver, C.S., Creager, K.C., Snelson, C.M., Crossson, R.S.,
Miller, K.C., and Trehu, A.M., 2003. Amplification of seismic waves by the Seattle
basin, Washington State. Bulletin of the Seismological Society of America 93: 533-545.

Quennel, A.M., 1959, Tectonics of the Dead Sea rift. 20th International Geological Congress,
Mexico: 385-405.

29
Reches, Z., and Hoexter, D.F., 1981. Holocene seismic and tectonic activity in the Dead Sea
area: Tectonophysics, v. 80, p. 235-254.

Rybakov, M., Fleischer, L., and ten Brink, U., 2003. The Hulla Valley subsurface structure
inferred from gravity data. Israel Journal of Earth Sciences 52: 113-122.

Salamon, A., Hofstetter, A., Garfunkel, Z. and Ron, H., 2003. Seismotectonics of the Sinai
subplate – the eastern Mediterranean region. Geophysical Journal International 155: 149-
173.
Schramm, A., Stein, M., Goldstein, S.L., 2000. Calibration of the 14C time scale to >40 ka by
234U – 230Th dating of Lake Lisan sediments (last glacial Dead Sea). Earth and Planetary
Science Letters 175: 27-40.

Seed, H. and Idriss, I.M., 1982. Ground Motions and Soil Liquefaction During Earthquakes:
Earthquake Engineering Research Institute, Oakland, CA, 134 pp.
Shapira., A., Avni, R., and Nur, A., 1993. New estimate of the Jericho earthquake epicenter
of July 11, 1927. Israel Journal of Earth Sciences 42: 93-96.
Shulman, H., 1992. Seismic review. In: The Israel National Oil Company, Ltd., Ami’az East-
1 completion report: 19-24.

Shulman, H., Kashai, E., Salhov, S., Gardosh, M., and From, Y., 1992. Ein Gedi-3,
Recommendation for drilling. The Israel National Oil Company, Ltd., 34 pp.

Stein, M., 2001. The sedimentary and geochemical record of Neogene-Quaternary water
bodies in the Dead Sea basin – inferences for the regional paleoclimatic history. Journal
of Paleolimnology 26: 271-282.

Tuttle, M.P., 2001. The use of liquefaction features in paleoseismology: lessons learned in
the New Madrid seismic zone, central United States. Journal of Seismology 5: 361-380.

Wdowinski, S., Bock, Y., Baer, G., Prawirodirdjo, L., Bechor, N., Naman, S., Knafo, R.,
Forrai, Y., and Melzer, Y., 2004. GPS measurements of current crustal movements along
the Dead Sea fault: Journal of Geophysical Research 109, BO5403: 1-16.

Weinberger, R., Begin, Z.B., Waldmann, N., Gardosh, M., Baer, G., Frumkin, A.,
Wdowinski, S., (accepted). Quaternary rise of the Sedom Diapir, Dead Sea basin. In:
Enzel, Y., Agnon, A. and Stein, M., New Frontiers in Dead Sea Paleoenvironmental
Research. Geologiocal Society of America Special Publication.

Weinberger, R., Porat, N., Levy, Z., and Marco, S., 2003. Clastic dykes in the Lisan
Formation: new indications for Holocene earthquakes and salt tectonics. Geological
Survey of Israel Report GSI/19/2003, 10 pp (in Hebrew).

30
Youd, T. L. and Perkins, D.M., 1987. Mapping of liquefaction severity index. Journal of
Geotechnical Engineering 113: 1374-1392.

Zaslavsky, Y, Shapira, A., Gorstein, M., Kalmanovich, M., Perelman, N., Giller, V., Livshits,
D., Giller, D., Dan, I., Akseinko, T., and Ataev, G., 2002. Generalization of site effects
for earthquake scenario application, the Coastal Plain area, Final Report, The
Geophysical Institute of Israel, Report 595/274/02.

Zaslavsky, Y., Shapira, A., and Arzi, A.A., 2000. Amplification effects from earthquakes and
ambient noise in the Dead Sea rift (Israel). Soil Dynamics and Earthquake Engineering
20: 187-207.

31
- . " "
. -
- ,
. - ,
, ,
.

,
. , .
(M 6.2) ,
, 16-60 , .1927
. , " 20
, , , , " 12 ,6 ,3
.

,
, ( )
.
. -

32
‫סס‬ ‫‪-‬‬ ‫פ ‪ -‬ס‬

‫‪3‬‬
‫‪- . ,3‬‬ ‫‪. ,2‬‬ ‫‪. .' ,1‬‬ ‫‪. .‬‬

‫‪.1‬‬
‫‪,‬‬ ‫‪.2‬‬
‫‪.3‬‬

‫" ‪,‬‬ ‫‪,‬‬ ‫‪GSI /04/05‬‬


‫‪2005‬‬

You might also like