Open access peer-reviewed chapter

The Impact of Lake-Level Fluctuation on Earthquake Recurrence Interval over Historical and Prehistorical Timescales: The Case of the Dead Sea

Written By

Mariana Belferman and Amotz Agnon

Submitted: 27 June 2023 Reviewed: 03 October 2023 Published: 15 November 2023

DOI: 10.5772/intechopen.113357

From the Edited Volume

Science of Lakes - Multidisciplinary Approach

Edited by Ali A. Assani

Chapter metrics overview

45 Chapter Downloads

View Full Metrics

Abstract

We review the impact of large historical lake water-level changes on seismicity via the stress field of the shallow crust where devastating earthquakes nucleate. A novel backward earthquake simulation presented in this chapter can be used to investigate the geological record for the past ten millennia (presented in this study) and even more. The simulation is based on a theoretical model, which explains the variability in the recurrence interval of strong earthquakes. We suggest that the water-level changes in ancient lakes located in tectonic depressions along the Dead Sea transform could contribute to the observed differences. It is found that the increase in the water level moderates the seismic recurrence interval. Based on this empirical correlation together with mechanical considerations, an additional indication is established regarding the water-level reconstruction and location of earthquakes in the Dead Sea area. This indication is based on simulated earthquakes, by superimposing the effective normal stress change due to the reconstructed water-level change on the estimated tectonic shear stress accumulated since the preceding seismic event.

Keywords

  • earthquake recurrence interval
  • water-level change
  • lake
  • Dead Sea
  • poro-elasto-plasticity
  • strike-slip tectonic deformation
  • induced seismicity

1. Introduction

Earthquakes are one of the primary geological hazards, with historical and modern accounts serving as vivid testimony of the threats they pose. With catastrophic losses of human lives and health [1, 2], earthquakes may cause major economic losses and property damage [3, 4]. Earthquakes also play a significant role in sedimentary systems evolution, liquefying water-saturated clastic sediments by cyclic loading [5, 6, 7].

Despite the global, social, and scientific impact of earthquakes, in many cases, their triggering mechanisms remain poorly defined. Specifically, tectonic motion is the primary cause and consequence of earthquakes: For example., large earthquakes along transform faults are typically associated with slip rate and stress buildup and release (e.g., Dead Sea transform (DST)—[8, 9, 10, 11, 12, 13], San-Andreas fault—[14, 15] or North Anatolian fault—[16, 17, 18]).

However, earthquake triggers may have no direct relation to the local tectonic dynamics. Observations supported by theoretical calculations demonstrate that surface processes (e.g., glaciations, sedimentation, and water loading) are able to change significantly the stress field in the Earth’s crust [19, 20]. Field data from northern Europe and America suggest that melting of the ice sheets was connected to the paleo-seismologic evidence and to the development of late Pleistocene/early Holocene fault scarps [21, 22, 23, 24], whereas the seismicity in the tectonically stable Greenland and Antarctica may be solely due to the current ice unloading [25].

Modeling the slip-rate evolution of localized normal faults located below the glaciated Earth crust, Hetzel and Hampel [26], Hampel and Hetzel [27], and Hampel et al. [28] demonstrate that variations in the slip-rate in these faults depend on the changes in the differential stress in the crust, which are in turn connected to changes in the loading on the Earth surface. The glaciations induce a decrease in the differential stress in the crust and, thus, a decrease in the faulting rate. In contrast, during postglacial rebound and isostatic adjustment, the differential stress in the crust grows, resulting in increased rates of faulting, inducing in turn temporal clustering of the earthquakes.

Additionally, it has long been recognized that fluids can influence the occurrence of earthquakes. Injection of liquid into (or removal from) the subsurface is capable to induce seismicity [29, 30, 31, 32] even outside the region subjected to the direct fluid perturbations [33, 34, 35].

Moreover, seismic response in the field was observed to be connected to reservoir impoundments and their seasonal water level variations [36, 37, 38, 39, 40]. An increase in pore pressure may lead to two different kinds of seismic response: swarm-like rapid seismicity induced by immediate pore pressure buildup, and delayed seismicity associated with pore pressure diffusion from the reservoir to hypocentral depths [37, 41]. Delayed response, as observed in the field, is often associated with large magnitude earthquakes [37, 42]; it may extend significantly beyond the confines of the reservoir and may not show an immediate correlation with major changes in reservoir level.

Seismic activity change was associated with historical water-level changes in the Dead Sea lake [43, 44] and Salton Sea lake [45, 46, 47] at much longer timescales. The Salton Sea, a remnant of ancient Lake Cahuilla, lies on the southern San-Andreas Fault (SSAF) [48]. Luttrell et al. [47] conducted a study to examine how Coulomb stress perturbations influenced local faults in relation to changes in the level of Lake Cahuilla. Their objective was to assess whether the lake could potentially impact the timing of fault rupture. The researchers proposed that the loading resulting from the ancient Lake Cahuilla impeded fault failure along most of the SSAF, except for a potential small area within the lake boundary.

Subsequently, Brothers et al. [46] modeled changes in Coulomb stress due to the combined effects of lake loading caused by periodic flooding of Lake Cahuilla, movement of stepover faults, and increased pore pressure. They determined that these factors could elevate stress levels along the southern SSAF to the extent of triggering an earthquake. Over the past 1100 years, the periodic flooding of Lake Cahuilla has exhibited fluctuations of up to 100 meters. These variations have shown throughout this time period a significant correlation with the incidence of earthquakes along the southern, the least active segment, of the SAF.

Belferman et al. [44] took advantage of the high temporal resolution of the historical earthquake records for the Dead Sea Lake during the last two millennia. The relatively small water-level hikes of 15 m, characteristic for time intervals of centuries to millennia, were analyzed and shown to be able to exacerbate the seismicity pattern at the Dead Sea fault. Belferman et al. [43] demonstrated for the first time that plausible scenarios for lake-level history show striking correlations with the historical record of earthquake recurrence intervals. Additionally, the plausible correlation observed between these phenomena suggests a complementary indicator that could be used to constrain either the locations of historical earthquakes or lake-level fluctuations during periods of uncertainty in one of these variables.

More recently, Hill et al. [45] present a direct relationship between changes in Lake Cahuilla water level and earthquakes based on computed time-dependent Coulomb stress changes for the last 1100 years. The authors present coupled 3D finite element model that takes into account time-varying surface water loads, fault geometry, crustal poroelasticity, and viscoelastic relaxation in the ductile substrate. The simulation results indicate that the past six major earthquakes on the southern San-Andreas Fault probably occurred during high stands of Lake Cahuilla. It was shown that increases in lake-level result in positive Coulomb stress changes on most of the southern San-Andreas Fault, with stressing rates as high as two to three times the tectonic loading rate.

In this chapter, we summarize the results and methods to identify the correlation between the historic water-level reconstructions at the Dead Sea and seismicity patterns in the area over the past two millennia. We also offer an additional analytical model to cope with dependence on initial conditions, which can be used in the future for interpreting these phenomena for earlier extended periods.

Advertisement

2. The study area: Dead Sea

2.1 Tectonics

The Dead Sea Lake fills tectonic depressions along the Arabian and African plate boundaries. The Arabian Plate is rifting away from the African and Nubian plates along the active spreading axis of the Red Sea, with the Sinai-Levant Block lagging behind, sheared along the Dead Sea fault (Figure 1) [49, 50, 51].

Figure 1.

Tectonic plates in the Middle East. The Dead Sea fault is one of the Dead Sea transform (DST) fault system, which transfers the opening at the Red Sea to the east Anatolian fault. Jordan Valley (JV), Dead Sea (DS), Arava/Araba Valley, and Gulf of Aqaba (GoA) located along the DST.

The Dead Sea fault is the major fault system along the plate boundary, and one of the best studied continental transform faults in the world [8, 11, 12, 49, 52, 53, 54, 55]. The tectonic motion at the DST is characterized predominantly by a left-lateral strike-slip regime with a velocity of up to ∼5 mm/yr. along various segments [51, 56, 57, 58]. Motion is transferred from the opening Red Sea to the escaping Anatolian Block via sinistral shear motion along the Dead Sea transform (e.g., [49, 51, 59]). This sinistral strike-slip plate boundary comprises the basins of the Gulf of Aqaba/Elat, Araba/Arava, the Dead Sea, the Jordan, Hula, Beqa’a, Ghab, and Karasu, a distance of about 1000 km [60].

Many factors and processes operating on different scales may have a profound influence on the dynamics of the deformations in the Dead Sea region. One of them—the influence of the water bodies deep underneath the Jordan-Arava valley—is explored in depth in this study.

2.2 Seismicity

The Dead Sea fault system is a major source of prehistoric, historical, and contemporary earthquakes [8, 11, 61, 62, 63, 64, 65]. Tectonic motion on the DST has always been addressed as a primary cause of earthquakes in the region. Large earthquakes (M > 5.5) are typically associated with slip rate and stress buildup at the strike-slip faults of the transform [9, 10, 11, 13, 66, 67, 68, 69]. Yet, normal faults might release M > 6 earthquakes [64].

Due to paleo-seismological research, we have access to 220 kyr of earthquake records, obtained based on different dating methods (e.g., uranium-thorium dating, radiocarbon dating, infrared stimulated luminescence dating) of deformation or changes that occur in rock structures due to earthquake shaking (namely seismites) [65]. A promising environment for paleoseismological investigations is provided by caves because they serve well as repositories for geological information. Kagan et al. [70] constrained dates of 38 seismites from the Soreq and Har-Tuv caves, 40 km of the DST. This was based on interpretation and amalgamation of individual seismite ages for bracketing the chronology of the strongest (M7.5-8) earthquakes covering the time interval since 185 ka.

Kagan et al. [70] found in their study a correlation between the paleoseismic events that occurred during the past 75 kyr identified with speleothems with other independent paleoseismic records in the region. Some of them were recorded in Dead Sea cores [71] and correlate with damage at archeological sites [8, 72]. Others correlate with seismites in the Lisan Formation [73, 74] or in the pre-Lisan Formations [65, 75].

Earthquake imprints formed in laminated sediments of the Lisan Formation [76] such as intraclast breccias [7], represent an uninterrupted geological record of the earthquake in the Dead Sea region for the past 14 to 70 kyr. Porat et al. [77] inferred that injection of clastic dikes is one of the most impressive liquefaction features generated by strong, M ≥ 6.5 earthquakes. The authors constrained the liquefaction date to shortly after the deposition of the Lisan Formation (15 kyr) by using optically stimulated luminescence ages. The dikes ages vary between 15 and 10 ka, thus suggesting that the period following the draw-down of Lake Lisan period was accompanied by strong earthquakes.

Migowski et al. [71] reconstructed an earthquake record using the dated Ein Gedi and Ze’elim cores seismites from the Dead Sea shores for the past 10 kyr. Later, for the last 4000 years, significantly more seismites were detected from the Ein Feshkha sedimentary sections in the Dead Sea area [64].

Recently, Liu et al. [65] reconstructed the longest archive of earthquakes from a 250 m core from the Deed Sea bottom, covering 220 kyr. Many of the seismites in the studies mentioned above, dated for the last 2 millennia, are matched by independent historical records [8, 68, 73, 78, 79, 80, 81, 82, 83, 84]. Using geological, archeological, biblical, historical, and seismological data, researchers infer part of the historical records for the Dead Sea area. Ben-Menahem [85] surmised that about 110 major earthquakes affected the area during the past 2500. Of these, 42 originated along the Dead Sea fault system itself, while 68 were imported from the Helenic-Cyprian arcs and the Anatolian-Elburz-Zagros fault systems.

Numerous publications list earthquakes that hit the Dead Sea and its surroundings during the last 2 millennia [8, 61, 62, 63, 79, 80]. Belferman et al. [44] filtered from the scores of listed events only the most destructive ones in the Dead Sea basin, typically causing local intensities of VII or higher in Jerusalem. For a minimal epicentral distance of 30 km, this would translate to a magnitude M > 5.7, according to the attenuation relations of Hough and Avni [86]. Considering site effects in the rocky terrain typical of the Jerusalem vicinity, where the near-surface shear velocity is relatively high, the modeled magnitudes exceed M6 (see Eq. (6) and Table 2: Beit HaKerem and Motza, which can be found in the reference [87]). In Belferman et al. [43], most earthquakes from the earlier list were amended. All earthquakes in Belferman et al. [43] were selected by simultaneously satisfying two criteria: (1) The acceptance regularizes the relation between recurrence intervals and lake level, and (2) they are corroborated by evidence from reassessed historical archives and/or seismites in the northern basin of the Dead Sea (Ein Feshkha and Ein Gedi sites).

2.3 Lake-level fluctuation

Several water bodies occupied the tectonic depression along the Dead Sea transform during the Neogene-Quaternary. The size and composition of these water bodies have changed through time, reflecting the changing climatic [88, 89, 90, 91, 92, 93, 94, 95] and physiographic [90, 96] conditions in the region. Layers of sediments accumulated in the tectonic depression reflecting several lake phases: the Neogene Sedom lagoon [97, 98, 99, 100, 101], the mid- to late-Pleistocene Lake Amora [93, 101, 102, 103], the last interglacial Lake Samra ∼130 kyr [104], the last glacial Lake Lisan 18 ∼ 70 kyr [101, 105, 106], and the Holocene to modern Dead Sea ∼15 kyr [92, 107].

The history of saline water bodies in the Dead Sea basin starts with the Sedom lagoon which has been related to the Neogene ingression of seawater into the Jordan-Arava valley. The reasons for the ingression and the exact age of the lagoon are not clear, and its disconnection from the open sea may be associated with various events that include eustatic changes in the Sea, the tectonic uplift of the Judea-Samaria Anticline, the formation of morphological sill in the Jezreel Valley, etc. [93].

Mid- to late-Pleistocene Lake Amora probably extended over a large part of the Dead Sea basin-Jordan Valley [93, 95]. Zak [101] suggested that the age of the Pleistocene sedimentary sequence at Mt. Sedom deposited by the Amora Formation lies between ∼1 Ma and ∼ 100 ka. The transition between Amora and Samra Formations corresponds to ∼130–140 ka, when the lake stood lower than ∼380 m below mean sea level (bmsl). The lithology and chronology of the Samra Formation suggest that the lake fluctuated mostly between ∼310 and ∼ 350 m bmsl [104], being significantly lower than the last glacial Lake Lisan (mostly ∼280 ± 40 m bmsl) [95, 96, 108, 109]) but higher than the Holocene Dead Sea (mostly ∼400 ± 20 m bmsl) [92, 107, 110, 111]. The lake-level curve for the past 150 kyr, presented in Figure 2, was digitized from Waldmann et al. [104].

Figure 2.

Water-level change in the last 150 kyr, for Samra, Lisan and the Dead Sea (DS) Lake. Derived from the work of Waldman et al. [104].

The most widespread sediments in the rift valley are those deposited at the bottom of Lake Lisan (Figure 2) that occupied the Dead Sea basin and adjacent basins between ∼70 and 18 ka., during the last glacial period in the northern latitudes [105, 112, 113, 114]. The age of the highest lake level of Lake Lisan ∼27–23 ka, [93, 95, 96, 108] or ∼ 26–24 ka [115, 116] corresponds to the time of a cold episode in the global climate, while the relatively low lake level of ∼55–30 ka. Refs. [95, 108, 115] corresponds to the warmer climatic conditions.

The curve of Holocene Dead Sea level fluctuations presented for past 10 kyr is based on the recognition and dating of shoreline deposits in the fan delta outcrops [88, 111, 117] or facies of drilled sediments [92, 107, 118, 119]. The Dead Sea fluctuated within the range of 370 to 430 mbsl during the past 10 kyr [92]. For the last centuries, a precise reconstruction of the Dead Sea level has been available [90].

During the past 4000 yr., the levels of the Dead Sea fluctuated within the range of 390 to 415 mbsl [111, 120]. Most of the time, the lake level was below 402 mbsl, and the Dead Sea was confined to its deep northern basin [111].

The drop from the late nineteenth century high stand of 390 mbsl to the present (2023) elevation of 428 m bmsl (www.water.gov.il) resulted mainly due to artificial water diversion of runoff water for the basin, superimposed on the climatic trend [111]. The current rate of lake-level decline is significantly high. However, drops of similar magnitude are not unusual in the late Holocene record, which yields evidence of long periods of droughts in the region [111].

Advertisement

3. Poroelasticity and its role in induced seismicity

Water-level-induced seismicity was addressed in several studies of water reservoirs [41, 121, 122] and was also analyzed using the theory of linear poroelasticity [123, 124]. For the first time, a relationship of this kind was considered in the pull-apart basin of tectonic depressions (Dead Sea lake) for earthquake intervals of timescale 100–200 years in Belferman et al. [43, 44]. The increase in water level contributes to an immediate increase in vertical stress due to the weight of the water mass (the loading effect) and subsequently to the increase in horizontal stress due to elastic coupling controlled by Poisson’s ratio [44]. Resulting stress changes are then superimposed on the background tectonic stress field [44]. Further, generation or absence of post-impounding seismicity depends on several additional factors, presented in Figure 3, and discussed in Belferman et al. [44].

Figure 3.

Factors that have the potential to influence the seismic activity prompted by alterations in water levels. Please refer to the text for more information.

Based on the real-time observation of 30 cases of post-impounding seismicity it was indicated that the complex interaction of the water-induced stress with the state of pre-existing stress near the reservoir, together with geological and hydrologic conditions at the site, determines whether these stress changes are sufficient to generate the seismic activity [125]. It was indicated that in areas of high strain accumulation and high levels of natural seismicity, stress changes induced by the reservoir may be small compared to natural variations in contrast to regions of moderate strain accumulation (low to moderate natural seismicity). Thus, in the low to moderate natural seismicity regions, the post-impounding can usually be accompanied by earthquakes. In aseismic areas with slow strain accumulation, the reservoir-induced stresses may be insufficient to raise the stress level to a state of failure [125].

Seismic activity induced by surface water-level fluctuations is also affected by the faulting regime (Figure 3), determined in turn by the respective orientations of the three principal stresses [126]. In regions where the vertical compressive stress is not minimal (normal and strike-slip faulting), seismic activity is more sensitive to the change in the effective stress due to water-level change than in regions where it is (thrust faulting) [123, 125, 127, 128, 129].

In many cases (for instance, the Koyna, Aswan, and Oroville reservoirs), the hypocenter of earthquakes, which are associated with changes in water levels, does not occur directly under the reservoir. In these cases, the faults passing through the reservoir serve as a conduit for pore pressure diffusion [41, 121, 130]. This allows the influence of the reservoir to extend to greater distances and depths, where larger magnitude earthquakes take place.

In addition to the conditions mentioned above, based on observations that some reservoirs continue to be active, whereas others show seismic quiescence under similar water-level fluctuations [37], it was indicated in Belferman et al. [44] that seismicity can strongly depend on the hydromechanical properties of faults (Figure 3).

Advertisement

4. Methods

4.1 Modeling setup

We consider a vertical fault underneath the lake (or reservoir) bed, extending along the outplane axis y (Figure 4), embedded in 2D (plain strain) geometry of the upper crust. We assume that fault failure will result from a change in the effective stress normal to the fault [123, 124, 131, 132, 133, 134]. A step change in the water load, σw, (Figure 4a) is applied on the entire top boundary of the half-infinite structure modeled, representing a valley filled by the lake, where uniaxial strain conditions apply.

Figure 4.

This figure, derived from the work of Belferman et al. [44], illustrates the model used in this study. (a) A step-function water load, applied on a top boundary of the structure, representing a poro-elastic seabed; (b) schematic of the structure below the lake with a vertical fault oriented along the outplane axis y. the calculations are based on uniaxial strain conditions.

4.2 The impact of Lake-level fluctuation on pore pressure

As a result of the water load on the seabed during water-level (WL) change, considering the diffusion, the pore pressure (p) is a function of time (t) and sub-bottom depth (z). Belferman et al. [44] consider a step function load on the half-space, for which, following Roeloffs [123], the analytical solution is:

pzt=γps+1γpserfcz/4ctE1

where ps is the pore pressure on the top boundary of the half-space, γ is the loading efficiency, and c is the hydraulic diffusivity. For the post-diffusion stage erfcz4ct1 at t (where t is the diffusion time). For any earthquake hypocentral depth, z, the assumption of the post-diffusion stage is valid, if the diffusion time scale, indicating the travel time of excess pore pressure signal from seabed to hypocentral depth, tdiff=z2/c, is small compared to the earthquake’s recurrence interval (RI) (tdifftRI) [44]. For an average earthquake hypocentral depth underneath the DSF, z20km [135, 136, 137] and c=4m2/s justified for the faulted rock, the diffusion time scale indicating a travel time of excess pore pressure from lake bed to the hypocentral depth tdiff3yr. It is negligible compared to the RI of moderate-to-large earthquakes (M > 5.5) in this area, estimated as tRI=1600yr for Lake Lisan and as tRI=100yr for the Dead Sea, based on seismite analyses from the Peratzim Creek and Ein Gedi cores [7, 8, 71, 74].

The water-level variation in Lake Cahuilla has a different tendency compared to the Dead Sea. Subsequently, Hill et al. [45] consider the periodic loading on the poroelastic half-space for which the analytical solution following Roeloffs [123] is:

pz=γps+1γpsezω/2ceizω/2cE2

were ω is the angular frequency of water loading (ω=2πT, where T is period time). Similar to a step function load for the post-diffusion stage, for periodic loading, we obtain the long-term limit: ezω2ceizω2c=ezπTceizπTc1 at T.

Here, T is defined by the periodic change of water level. Hence, for a long period of time, the diffusion effect is negligible because the pore pressure at the depth reached the value at the top boundary of the half-space. This is expressed in Hill et al. [45] in the “memory” effect of pore pressure at depth, whereby subsequent lakes can contribute to higher pore pressure owing to the diffusive time lag of a previous lake superimposing on the next. This memory effect is pronounced only in one case where the time interval (T) between maximal lake stands is about 41 years (Lake F and Lake E in Figure 2, [45]).

Dead Sea lake WL change is distinctly different from the periodic trend that is characteristic of Lake Cahuilla. This may be due to the pulverization of fault rock on the southern San-Andreas Fault [138].

4.3 Analytical formulation

A poroelastic model with brittle yielding was formulated by Belferman et al. [44], utilizing the Mohr-Coulomb failure criterion. This model was used to calculate the changes in earthquake recurrence interval in response to the increase of water level at basins overlying strike-slip faults.

Within the poroelastic component of the model, produced by the WL change, horizontal stress changes normal to the strike-slip fault calculated under a uniaxial (vertical) strain condition (Eq. (10b) in [44]). This applies to a post-diffusion stage, i.e., when pore pressure at hypocentral depth equilibrates with the lake’s bed (Pf=σw). In the case discussed, it is appropriate to skip the diffusion stage since the diffusion time is much shorter than the interseismic period. In this case, the effective vertical stress change becomes zero, while the effective horizontal stress (Δσ) change becomes:

Δσ=12ν1νβ1σwE3

where β is Biot’s coefficient and ν is Poisson’s ratio, σw=ρgΔh, where ρ is density of water and g is the acceleration of gravity.

Further, under the general loading conditions, the critical value of shear stress (τc) responsible for yielding at the fault is connected to the effective stress normal to the fault (σn), according to the Mohr-Coulomb failure criterion (e.g., [139]):

τc=C+tanφσnE4

where σn=σnPf is effective normal stress affected by pore pressure (Pf), C is cohesion, and φ is the angle of internal friction at the fault.

In order to calculate the change of recurrence interval (∆t) induced by the water level change, we use the Coulomb failure envelope and Mohr circle [140] with the following initial characteristics:

σ0=σ1+σ32,τ0=0,R0=σ1σ32E5

where σ1 and σ3 is the minimal and maximal principal stress, which is horizontal for the strike-slip faults. In this case, if during the interseismic period, the water level increased by Δh, the horizontal stress change can be specified by Δσ=Δσxx=Δσyy, following the definitions of principal stresses in 2D [140]. This change manifested in the shift of the Mohr circle toward the origin by Δσ while for the pre-seismic stress state, the Mohr circle center is: σ0=σ0+Δσ (where Δσ is negative at the post-diffusion stage, see Figure 5b, [44]).

Figure 5.

The Dead Sea WL reconstructions for the last two millennia. The dashed curves are suggested by the literature sources. Turquoise anchor points follow Bookman et al. [111] used in WL interpretation, while one point (in dark blue) is shifted to left in error interval of ±45 yr. solid, black line water curve is a plausible scenario suggested by this study.

We assume that the far-field tectonic strain rate, ε˙xy, stays constant over any interseismic time period, Δt, [44] while the elastic strain, Δε, accumulated during this period can be presented by: Δε=ε˙xyΔt. Based on Hooke’s law and considering only horizontal shear strain, shear stress accumulation at the fault during Δt is:

Δτxy=2Gε˙xyΔt=CcosφtRIΔtE6

where tRI is a tectonic recurrence interval and C ≠ 0 at the healed fault (e.g., [141]). Hence, a permanent tectonic strike-slip stress rate can be defined as CcosφtRI [44]. The change of the shear stress during the interseismic period increases the radius of the Mohr circle while for the pre-seismic stress state, the radius is: R0=R0+Δτxy.

Byerlee’s law envelope [142] was used to define the strength of a seismogenic zone at the fault immediately after the earthquake. An initial stress state, σinit, is defined as a Mohr circle with a radius R0 centered at the point σ00 and restricted by the Byerlee’s law envelope. The larger hypocentral depth is associated with a larger σ0. Based on the results of Byerlee [142] and on additional laboratory experiments (for a review, see [143]), we adopt, for simplicity’s sake, the friction angle φ=0.54 and initial cohesion C=0.

The pre-seismic Coulomb failure envelope is defined by the nonzero cohesion coefficient, C0, specific for the healed fault zone (e.g., [141] and references therein) and friction angle φ=0.54. When the circle reaches the failure envelope, the rock fails at the fault oriented most favorably for sliding (in our case, it is the pre-defined strike-slip fault). The stress will then drop again to σinit.

After stress release, the time to the next earthquake, Δt, is calculated from the solution of the Mohr-Coulomb failure criterion for a strike-slip tectonic regime and a water level change, Δh, characteristic to the Dead Sea lake [44]:

ττ02+(σσ0+σ2=R0+CcosφtRIΔt2τ=C+tanφσE7

Solving this problem for single solution, we get Δt us a linear function of the water level change Δh:

t=C+tanφρgΔhtRICE8

4.4 Best fit random method of WL curve prediction

In the interpretation of water level over the past two millennia [92, 111, 144], there are several uncertainty. Specifically, the water-level dating method (Radiocarbon dating) could have an error of about ±45 yr., as estimated from the radiocarbon dating of shoreline deposits in fan delta outcrops [111]. The entire past bi-millennial Dead Sea level record is constrained by less than twenty “anchor points” (the data obtained by the dating collected from surveyed paleo-shorelines, [111]). However, historical water level records are quite precise elevation-wise, as they are obtained from different points around the lake [92, 111].

Consequently, the interpretations of the curves (Figure 5) are not identical and are not unambiguous, except for some limitations [43]. Accounting for these limitations, which include the “anchor points” (the data obtained by the dating collected from surveyed paleo-shorelines, [111]) of water level, the 10 million WL curves were generated [44] for the last bi-millennial interval, using a uniformly distributed random number generator. For the simulation, we are setting a ten-year time step.

The linear correlation between RI of widely recorded medium to large historical earthquakes (M > 5.5) available from the literature and WL interpolations was tested on the basis of an estimate of the value of the Pearson product-moment correlation coefficient, R. These statistics were used to assess the suitability of each randomly interpolated WL curve and for identifying out-of-correlation earthquakes.

4.5 Earthquake simulation

The analytical model presented in this section has been discretized with the time step of 1 year and applied to the sequence of WL change samples [44]. From the starting point (AD 33, see Table 1 which can be found in the Appendix of the reference [44]), the simulator moves forward with time, along the WL curve. For each step, the WL change, hi, and the amount of accumulated tectonic stress are calculated. After each stress release, the time to the next earthquake, t, is calculated from the solution of the Mohr–Coulomb failure criterion for a strike-slip tectonic regime. Eq. (9) takes the following discrete form:

τiτ02+(σσ0+12ν1νβ1ρgΔhi2=R0+CcosφtRIΔt2=τi=C+tanφσiE9

For each time step, the algorithm determines whether there is a single solution, or two, or nil. A case of no solutions means that the Mohr circle is yet to reach the failure envelope, as the accumulating tectonic stress and the WL increase are still insufficient. At this case, the Δt increase with the time steps. The system of Eq. (7) may have a single solution when the failure criterion is met at the end of some timestep or two solutions when it is met before the end of the timestep. A case of two solutions is rounded down to a case of a single solution if a time step (1 year) is small compared to the earthquake RI (about hundreds of years), where the Mohr circle reaches the failure envelope and the earthquake accrues. In this case, the calculated Δt saved and resets to zero. At the final stage, we get an array of RI.

4.6 The backward earthquake simulation approach

This formulation is based on the same basic assumptions that were proposed in Belferman et al. [43, 44], as well as set out in the “Analytical formulation of the direct model” section. Similar to the previous simulation, we assume that after an earthquake, the Mohr’s circle returns to the initial state of stress. But in contrast to the previous case, our starting point of reference is the last earthquake, as it is much better constrained. Considering that at the moment before the specific earthquake occurred, Mohr’s circle reaches the failure envelope, the set of Eq. (7) gets one solution. The horizontal effective stress change (Δσ) calculated from Eq. (3). While σw=ρgΔh, calculated for the WL change Δh at the moment of this specific earthquake.

The recurrence interval, Δt, calculated from Eq. (8), and based on this value, we calculate the time of the previous earthquake tp=tcΔt. Running this algorithm on the WL change during the period 2–10 kyr (based on the level history of [92]), combined with the WL change received from the best fit random method [44] including the anchor points [111] for the past two millennia, we get a new catalog of simulated earthquakes presented in Figure 6.

This formulation, like the previous one, is based on the solution of a system of equations (Eq. (7)) in the domain determined by positive definite values of Δt. This is only possible when the WL change (Δh) does not exceed the value Ctanφρg. When the WL change exceeds the defined value, the model jumps to the next point where the WL change is within the definition domain.

Advertisement

5. Results and discussion

Even though historical water-level records exhibit a significant degree of elevation accuracy, given that they are derived from diverse surveyed locations around the lake [92, 111], the dating of water levels might carry an approximate error of ±45 years and even more, as deduced from radiocarbon dating of sediment deposits in fan delta outcrops [111] or sediment cores [92]. The substantial differences in potential interpolations arise from the uncertainties in dating and resolution aspects of WL reconstructions. This was shown in Belferman et al. [43], using best fit random method WL curve generation. Ten million different curves were generated, constrained with anchor points established by field surveys and radiocarbon dating [111].

Dating historical earthquakes can exhibit a high level of precision, and the validation of accuracy occurs when various historical references align in agreement [61, 62, 63]. But in many cases, the epicenter of even a historical earthquake can be imprecise or not known.

This chapter extends our simulation to prehistoric time (10 ka), for which further challenges include identifying and dating shoreline and earthquake markers [71, 92]. The water-level curve is combined from the one indicated in Belferman et al. [43] (for 2 ka) with that provided by Migowski et al. [92] (for 2–10 ka). Our model for that level curve generates 50 earthquakes.

The recurrence intervals are simulated using the reverse earthquake simulation algorithm, while the starting point of reference is a simulated earthquake that occurred in 1907 CE. We start from this point because it is the last earthquake received from the earthquake simulation presented in Belferman et al. [43], which correlates with the documented earthquake from the field of 1927 CE. The recurrence intervals are depicted in Figure 6 as black points plotted against the water level (orange curve). The calculated years of simulated earthquakes are presented along the time axis by blue stars (Figure 6).

Figure 6.

Orange curve represents the best fit random water level [43] combined with Migowski et al. [92] vs. simulated and historic RIs, correspondingly. The blue dots mark the dates of the seismic events, while the black dots indicate the recurrence interval between these events.

Two types of spatial segments have been identified (gray bars in Figure 5). Light gray bar (∼3490–4020, 4370–4730, and 9380–10,000 years) indicates the time segments where the WL change exceeds the definition domain specified in our model. The dark gray bars indicate the time segments of earthquake sequence (4024–4065, 4724–5068, and 9141–9196) received for periods of height WL stand, with recurrence interval smaller than 35 yrs. The domain excludes the time segments where the WL change almost reaches the critical value of Ctanφρg (marked in dark gray in Figure 6).

Simulated earthquakes are presented vs. earthquakes from the literature in Figure 7. The blue points (Figure 7) indicate the simulated events supported by literature. Orange points (Figure 7) indicate simulated events, for which there is no evidence in the literature. A cluster of earthquakes was generated in these segments. In Figure 7, we left only the first and the last earthquake of the cluster.

Figure 7.

A comparison dates of historic vs. simulated earthquakes based on the suggested best-fit WL curve [43] combined with WL curve presented by Migowski et al. [71].

In the segments marked in light gray, our model did not perform the calculation. The earthquakes received in these segments were calculated in the previous step. Therefore, our model may not have generated earthquakes that occurred in these segments, such as the one indicated by Ben-Menahem [85] to 1560 B.C. or the seismite estimated by Migowski et al. [71] to have occurred around 7700 B.C.

Of the 50 simulated earthquakes during a period of 10,000 years, only 15 events were not confirmed in the literature (Figure 7). All these 15 events occurred before 3000 BCE. In other words, for all simulated earthquakes by our model within a span of 5000 years, there is evidence in the literature of a significant earthquake occurring in the Dead Sea area. Five earthquakes not confirmed by literature generated at 3068–3533 BCE. In Ben-Menachem’s 1991 study, the archeological excavations at Tleilat el-Rassul of Kenyon were referenced, indicating the possibility of destruction caused by earthquakes that occurred between 3300 and 3600 BCE. In addition, Migowski et al. [71] dated the possible earthquake to 3300–3367 BCE from Ein Gedi core. This evidence can satisfy the simulated events 339 3 BCE, 3533 BCE. Another prehistoric cluster (four events) of earthquakes was simulated in the period 7196–7410 BCE due to a relatively high water level. This period is not well represented by the archeological record due to scarcity of masonry structures.

The advantage of using the backward simulation method, presented for the first time in this chapter, is that it starts from the last earthquake along the Dead Sea fault, for which ample information is available. By contrast, the previous simulation [43] starts from the last earthquake of the studied period. Therefore, it is suitable for a period of 2000 years where earthquakes have historical evidence. The backward method enables us to simulate earthquakes that occurred in prehistoric times, including during the Lake Lisan period. For this period, the water level was significantly higher, and the data regarding both phenomena are subject to even greater uncertainty. Nevertheless, the paleoseismicity studies [8, 12] indicate that despite such high water levels, the seismic activity may have been subdued compared with activity for the Holocene Dead Sea.

However, it is worth noting that our model is limited by the maximum change in water level, and the water level during the Lisan period exceeds this limit. On this occasion, in future studies, it is worth integrating additional physical properties of Earth’s crust. Field data from Scandinavia and North America suggest that the melting of Pleistocene ice sheets was connected to paleoseismic activity [19, 21, 22, 23, 24, 145, 146]. Conversely, the low level of seismicity in Greenland and Antarctica may be due to the current ice load [25], which in turn induces changes in the differential stress in the Earth crust [26, 28].

Hampel et al. [20], based on numerical models that include one or several fault planes embedded in a rheologically layered lithosphere, indicate how loading and flexure of the lithosphere by ice or water decreases the slip rates of nearby faults, whereas unloading and rebound of the lithosphere accelerates the slip (see also [47]). Slip rate variations are caused by changes in the differential stress, which is the difference between the maximum (σ1) and minimum (σ3) principal stresses. When a fault is under a load, the differential stress decreases during loading and flexure of the lithosphere and increases during unloading and rebound. This leads to a decrease and subsequent increase in the fault slip rate.

This configuration needs to be considered when developing the earthquake simulation model for the Lisan period. Additional considerations for further studies are the mechanical condition of the recorder (namely the sediment bed) and the effect of water column height on the dynamics of seismite generation [147, 148].

Advertisement

6. Conclusions

The correlation model, developed in [44] and supported by the comparability of a simulated earthquake with a historical catalog for 2,000 years [43] imitates reaction of pore-elasto-plastic crust. This model is validated by the results obtained in this chapter for 10 millennia of geological records of earthquakes and shows a good correlation for the past five millennia.

Based on the findings presented in the Results-Discussion chapter, we indicate that flexure of the lithosphere, suggested by Hampel et al. [20], may not be significant over two or five millennia, but it should be considered when analyzing longer time periods as Lake Lisan.

Considering this in our future model, we expect to receive an explanation of the low seismicity pattern during the Lisan high stand and solve the model limitations where the WL change exceeds the definition domain, specified in our model.

Nevertheless, the new back-counting model presented in this work enables us to extend the relationship between WL changes and seismicity over a period span of ten millennia.

Advertisement

Acknowledgments

This project was supported by grants from the Ministry of Energy (grant no. 213-17-002) and the German–Israeli Foundation for Scientific Research and Development (GIF; grant no. I-1280-301.8). We would like to express our gratitude to Dr. Regina Katsman and Prof. Zvi Ben-Avraham for their valuable contributions in initiating this chapter.

References

  1. 1. Daito H, Suzuki M, Shiihara J, Kilgore PE, Ohtomo H, Morimoto K, et al. Impact of the Tohoku earthquake and tsunami on pneumonia hospitalisations and mortality among adults in northern Miyagi, Japan: A multicentre observational study. Thorax. 2013;68:544-550
  2. 2. Xu Y, Huang J, Zhou J, Zeng Y. Patterns of abdominal injury in 37 387 disaster patients from the Wenchuan earthquake. Emergency Medicine Journal. 2013;30:538-542
  3. 3. Mileti D. Disasters by Design: A Reassessment of Natural Hazards in the United States. Washington, DC: Joseph Henry Press; 1999. DOI: 10.17226/5782
  4. 4. Noriega GR, Ludwig LG. Social vulnerability assessment for mitigation of local earthquake risk in Los Angeles County. Natural Hazards. 2012;64:1341-1355. DOI: 10.1007/s11069-012-0301-7
  5. 5. Allen JRL. Sedimentary Structures: Their Character and Physical Basis. Vol. 30. Amsterdam, The Netherlands: Elsevier; 1982. (Developments in Sedimentology)
  6. 6. Youd T, Perkins D. Mapping of liquefaction severity index. Journal of Geotechnical Engineering. 1987;113:1374-1392
  7. 7. Agnon A, Migowski C, Marco S. Intraclast breccias in laminated sequences reviewed: Recorders of paleo-earthquakes. In: Enzel Y, Agnon A, Stein M, editors. New Frontiers in Dead Sea Paleoenvironmental Research. Vol. 401. Colorado, US: Geological Society of America; 2006. p. 195. DOI: 10.1130/2006.2401(13)
  8. 8. Agnon A. Pre-instrumental earthquakes along the Dead Sea rift. In: Garfunkel Z, Ben-Avraham Z, Kagan E, editors. Dead Sea Transform Fault System: Reviews. Dordrecht: Springer; 2014. pp. 207-261. DOI: 10.1007/978-94-017-8872-48
  9. 9. Ben-Menahem A, Nur A, Vered M. Tectonics, seismicity and structure of the Afro-Eurasian junction—The breaking of an incoherent plate. Physics of the Earth and Planetary Interiors. 1976;12(1):1-50. DOI: 10.1016/0031-9201(76)90005-4
  10. 10. Ferry M, Meghraoui M, Karaki NA, Al-Taj M, Amoush H, Al-Dhaisat S, et al. A 48-kyr-long slip rate history for the Jordan Valley segment of the Dead Sea fault. Earth and Planetary Science Letters. 2007;260(3):394-406. DOI: 10.1016/j.epsl.2007.05.049
  11. 11. Garfunkel Z. Internal structure of the Dead Sea leaky transform (rift) in relation to plate kinematics. Tectonophysics. 1981;80(1–4):81-108. DOI: 10.1016/0040-1951(81)90143-8
  12. 12. Marco S, Klinger Y. Review of on-fault Palaeoseismic studies along the Dead Sea fault. In: Garfunkel Z, Ben-Avraham Z, Kagan E, editors. Dead Sea Transform Fault System: Reviews. Modern Approaches in Solid Earth Sciences. Vol. 6. Dordrecht: Springer; 2014. DOI: 10.1007/978-94-017-8872-4_7
  13. 13. Niemi T, Zhang H, Atallah M, Harrison JB. Late Pleistocene and Holocene slip rate of the northern Wadi araba fault, Dead Sea transform, Jordan. Journal of Seismology. 2001;5:449-474. DOI: 10.1023/A:1011487912054
  14. 14. Schwartz DP, Coppersmith KJ. Fault behavior and characteristic earthquakes: Examples from the Wasatch and San Andreas fault zones. Journal of Geophysical Research: Solid Earth. 1984;89(B7):5681-5698. DOI: 10.1029/JB089iB07p05681
  15. 15. Sieh KE. Slip along the San Andreas fault associated with the great 1857 earthquake. Bulletin of the Seismological Society of America. 1978;68(5):1421-1448. Available from: https://authors.library.caltech.edu/49783/1/1421.full.pdf
  16. 16. Barka AA. The north Anatolian fault zone. In: Annales Tectonicae. Vol. 6. Bristol, England: Bristol University Press; 1992. pp. 164-195. Available from: ftp://ftp.ecn.purdue.edu/Barka.PDF
  17. 17. Stein RS, Barka AA, Dieterich JH. Progressive failure on the North Anatolian fault since 1939 by earthquake stress triggering. Geophysical Journal International. 1997;128:594-604
  18. 18. Toksöz MN, Shakal AF, Michael AJ. Space-time migration of earthquakes along the north Anatolian fault zone and seismic gaps. Pure and Applied Geophysics. 1979;117:1258-1270. DOI: 10.1007/BF00876218
  19. 19. Muir-Wood R. Deglaciation seismotectonics: A principal influence on intraplate seismogenesis at high latitudes. Quaternary Science Reviews. 2000;19:1399-1411. DOI: 10.1016/S0277-3791(00)00069-X
  20. 20. Hampel A, Lüke J, Krause T, Hetzel R. Finite-element modelling of glacial isostatic adjustment (GIA): Use of elastic foundations at material boundaries versus the geometrically non-linear formulation. Computers & Geosciences. 2019;122:1-14. DOI: 10.1016/j.cageo.2018.08.002
  21. 21. Arvidsson R. Fennoscandian earthquakes: Whole crustal rupturing related to postglacial rebound. Science. 1996;274:744. DOI: 10.1126/science.274.5288.744
  22. 22. Davenport CA, Ringrose PS. Deformation of Scottish quaternary sediment sequences by strong earthquake motions. In: Jones ME, Preston RM, editors. Deformation of Sediments and Sedimentary Rocks. Vol. 29. London, UK: Geological Society of London; 1987. pp. 299-314. DOI: 10.1144/GSL.SP.1987.029.01.23
  23. 23. Fjeldskaar W. The amplitude and decay of the glacial forebulge in Fennoscandia. Norsk Geologisk Tidsskrift. 1994;74(1):2-8. Available from: http://njg.geologi.no/images/NJG_articles/NGT_74_1_002-008.pdf
  24. 24. Stewart IS, Sauber J, Rose J. Glacio-seismotectonics: Ice sheets, crustal deformation and seismicity. Quaternary Science Reviews. 2000;19:1367-1389. DOI: 10.1016/S0277-3791(00)00094-9
  25. 25. Turpeinen H, Hampel A, Karow T, Maniatis G. Effect of ice sheet growth and melting on the slip evolution of thrust faults. Earth and Planetary Science Letters. 2008;269:230-241. DOI: 10.1016/j.epsl.2008.02.017
  26. 26. Hetzel R, Hampel A. Slip rate variations on normal faults during glacial–interglacial changes in surface loads. Nature. 2005;435:81-84. DOI: 10.1038/nature03562
  27. 27. Hampel A, Hetzel R. Response of normal faults to glacial-interglacial fluctuations of ice and water masses on Earth's surface. Journal of Geophysical Research: Solid Earth. 2006;111:B06406. DOI: 10.1029/2005JB004124
  28. 28. Hampel A, Hetzel R, Densmore AL. Postglacial slip-rate increase on the Teton normal fault, northern basin and range province, caused by melting of the Yellowstone ice cap and deglaciation of the Teton range? Geology. 2007;35:1107-1110. DOI: 10.1130/G24093A.1
  29. 29. Ellsworth WL. Injection-induced earthquakes. Science. 2013;341(6142):1225942. DOI: 10.1126/science.1225942
  30. 30. Suckale J. Induced seismicity in hydrocarbon fields. In: Advances in Geophysics. Vol. 51. Cambridge UK: Elsevier; 2009. pp. 55-106. DOI: 10.1016/S0065-2687(09)05107-3
  31. 31. Majer EL, Baria R, Stark M, Oates S, Bommer J, Smith B, et al. Induced seismicity associated with enhanced geothermal systems. Geothermics. 2007;36(3):185-222. DOI: 10.1016/j.geothermics.2007.03.003
  32. 32. Zoback MD, Gorelick SM. To prevent earthquake triggering, pressure changes due to CO2 injection need to be limited. Proceedings of the National Academy of Sciences. 2015;112(33):E4510-E4510. DOI: 10.1073/pnas.1508533112
  33. 33. Megies T, Wassermann J. Microseismicity observed at a non-pressure-stimulated geothermal power plant. Geothermics. 2014;52:36-49. DOI: 10.1016/j.geothermics.2014.01.002
  34. 34. Stark MA, Davis SD. Remotely triggered microearthquakes at the geysers geothermal field, California. Geophysical Research Letters. 1996;23(9):945-948. DOI: 10.1029/96GL00011
  35. 35. Yeck WL, Hayes GP, McNamara DE, Rubinstein JL, Barnhart WD, Earle PS, et al. Oklahoma experiences largest earthquake during ongoing regional wastewater injection hazard mitigation efforts. Geophysical Research Letters. 2017;44(2):711-717. DOI: 10.1002/2016GL071685
  36. 36. Gupta HK. Reservoir induced earthquakes. In: Gupta HK, editor. Development in Geotechnical Engineering. Vol. 71. Amsterdam: Elsevier; 1992. pp. 1-364. DOI: 10.1016/B978-0-444-88906-5.50022-3
  37. 37. Talwani P. On the nature of reservoir-induced seismicity. Pure and Applied Geophysics. 1997;150:473-492. DOI: 10.1007/978-3-0348-8814-1_8
  38. 38. Talwani P. Seismotectonics of the Koyna-Warna area, India. Pure and Applied Geophysics. 1997;150:511-550. DOI: 10.1007/s000240050091
  39. 39. Tomic J, Abercrombie R, Do Nascimento A. Source parameters and rupture velocity of small M≤2.1 reservoir induced earthquakes. Geophysical Journal International. 2009;179:1013-1023. DOI: 10.1111/j.1365-246X.2009.04233.x
  40. 40. Hua W, Chen Z, Zheng S. Source parameters and scaling relations for reservoir induced seismicity in the Longtan reservoir area. Pure and Applied Geophysics. 2013;170:767-783. DOI: 10.1007/s00024-012-0459-7
  41. 41. Hua W, Chen Z, Zheng S, Yan C. Reservoir-induced seismicity in the Longtan reservoir, southwestern China. Journal of Seismology. 2013;17(2):667-681. DOI: 10.1007/s10950-012-9345-0
  42. 42. Simpson DW, Leith W, Scholz C. Two types of reservoir-induced seismicity. Bulletin of the Seismological Society of America. 1988;78:2025-2040. Link to full.pdf
  43. 43. Belferman M, Agnon A, Katsman R, Ben-Avraham Z. Identifying plausible historical scenarios for coupled lake level and seismicity rate changes: The case for the Dead Sea during the last 2 millennia. Natural Hazards and Earth System Sciences. 2022;22(8):2553-2565. DOI: 10.5194/nhess-22-2553-2022
  44. 44. Belferman M, Katsman R, Agnon A. Effect of large-scale surface water level fluctuations on earthquake recurrence interval under strike-slip faulting. Tectonophysics. 2018;744:390-402. DOI: 10.1016/j.tecto.2018.06.004
  45. 45. Hill RG, Weingarten M, Rockwell TK, et al. Major southern San Andreas earthquakes modulated by lake-filling events. Nature. 2023;618:761-766. DOI: 10.1038/s41586-023-06058-9
  46. 46. Brothers D, Kilb D, Luttrell K, et al. Loading of the San Andreas fault by flood-induced rupture of faults beneath the Salton Sea. Nature Geoscience. 2011;4:486-492. DOI: 10.1038/ngeo1184
  47. 47. Luttrell K, Sandwell D, Smith-Konter B, Bills B, Bock Y. Modulation of the earthquake cycle at the southern San Andreas fault by lake loading. Journal of Geophysical Research. 2007;112:B08411. DOI: 10.1029/2006JB004752
  48. 48. Tostrud MB. The Salton Sea, 1906-1996, computed and measured salinities and water levels. In: Draft Report. California, US: Colorado River Board of California; 1997
  49. 49. Freund R, Garfunkel Z, Zak I, Goldberg M, Weissbrod T, Derin B, et al. The shear along the Dead Sea rift. Philosophical Transactions for the Royal Society of London. Series A, Mathematical and Physical Sciences. 1970;267(1181):107-130. Available from: https://www.jstor.org/stable/73612
  50. 50. Le Beon M, Klinger Y, Amrat AQ, Agnon A, Dorbath L, Baer G, et al. Slip rate and locking depth from GPS profiles across the southern Dead Sea transform. Journal of Geophysical Research: Solid Earth. 2008;113(B11):B11403. DOI: 10.1029/2007JB005280
  51. 51. McClusky S, Reilinger R, Mahmoud S, Ben Sari D, Tealeb A. GPS constraints on Africa (Nubia) and Arabia plate motions. Geophysical Journal International. 2003;155(1):126-138. DOI: 10.1046/j.1365-246X.2003.02023.x
  52. 52. Quennel AM. Tectonics of the Dead Sea Rift, Paper Presented at the 20th International Geological Congress. Africanos, Mexico: Assoc. Serv. Geol; 1956. p. 1959
  53. 53. Wilson JT. A new class of faults and their bearing on continental drift. Nature. 1965;207:343-347
  54. 54. Ben-Avraham Z, Garfunkel Z, Lazar M. Geology and evolution of the southern Dead Sea fault with emphasis on subsurface structure. Annual Review of Earth and Planetary Sciences. 2008;36:357-387
  55. 55. Weber FA, Voegelin A, Kretzschmar R. Multi-metal contaminant dynamics in temporarily flooded soil under sulfate limitation. Geochimica et Cosmochimica Acta. 2009;73(19):5513-5527. DOI: 10.1016/j.gca.2009.06.011
  56. 56. Hamiel Y, Amit R, Begin ZB, Marco S, Katz O, Salamon A, et al. The seismicity along the Dead Sea fault during the last 60,000 years. Bulletin of the Seismological Society of America. 2009;99:2020-2026. DOI: 10.1785/0120080218
  57. 57. Masson F, Hamiel Y, Agnon A, Klinger Y, Deprez A. Variable behavior of the Dead Sea fault along the southern Arava segment from GPS measurements. Comptes Rendus Geoscience. 2015;347(4):161-169. DOI: 10.1016/j.crte.2014.11.001
  58. 58. Sadeh M, Hamiel Y, Ziv A, Bock Y, Fang P, Wdowinski S. Crustal deformation along the Dead Sea transform and the Carmel fault inferred from 12 years of GPS measurements. Journal of Geophysical Research: Solid Earth. 2012;117:B08410. DOI: 10.1029/2012jB009241
  59. 59. Joffe S, Garfunkel Z. Plate kinematics of the circum Red Sea—A re-evaluation. Tectonophysics. 1987;141:5-22. DOI: 10.1016/0040-1951(87)90171-5
  60. 60. Bartov Y, Steinitz G, Eyal M, Eyal Y. Sinistral movement along the Gulf of Aqaba—Its age and relation to the opening of the Red Sea. Nature. 1980;285(5762):220-222
  61. 61. Ambraseys N. Earthquakes in the Mediterranean and Middle East: A Multidisciplinary Study of Seismicity Up to 1900. Cambridge, UK: Cambridge University Press; 2009
  62. 62. Guidoboni E, Comastri A. Catalogue of Earthquakes and Tsunamis in the Mediterranean Area from the 11th to the 15th Century. Rome, Italy: Istituto nazionale di geofisica e Vulcanologia; 2005. DOI: 10.1515/BYZS.2008.854
  63. 63. Guidoboni E, Comastri A, Traina G. Catalogue of Ancient Earthquakes in the Mediterranean Area Up to the 10th Century. Rome, Italy: Istituto nazionale di geofisica e Vulcanologia; 1994. DOI: 10.1163/182539185X01377
  64. 64. Kagan E, Stein M, Agnon A, Neumann F. Intrabasin paleoearthquake and quiescence correlation of the late Holocene Dead Sea. Journal of Geophysical Research: Solid Earth. 2011;116(B4):B04311. DOI: 10.1029/2010JB007452
  65. 65. Liu Y, Rubin AM. Role of fault gouge dilatancy on aseismic deformation transients. Journal of Geophysical Research - Solid Earth. 2010;115(B10):B10414. DOI: 10.1029/2010JB007522
  66. 66. Ellenblum R, Marco S, Agnon A, Rockwell T, Boas A. Crusader castle torn apart by earthquake at dawn, 20 may 1202. Geology. 1998;26:303-306. DOI: 10.1130/0091-7613(1998)026<0303:CCTABE>2.3.CO;2
  67. 67. Matmon A, Shaked Y, Porat N, Enzel Y, Finkel R, Lifton N, et al. Landscape development in an hyperarid sandstone environment along the margins of the Dead Sea fault: Implications from dated rock falls. Earth and Planetary Science Letters. 2005;240:803-817. DOI: 10.1016/j.epsl.2005.06.059
  68. 68. Reches ZE, Hoexter DF. Holocene seismic and tectonic activity in the Dead Sea area. Tectonophysics. 1981;80:235-254. DOI: 10.1016/0040-1951(81)90151-7
  69. 69. Shaked Y, Agnon A, Lazar B, Marco S, Avner U, Stein M. Large earthquakes kill coral reefs at the north-West Gulf of Aqaba. Terra Nova. 2004;16:133-138. DOI: 10.1111/j.1365-3121.2004.00541.x
  70. 70. Kagan E, Agnon A, Bar-Matthews M, Ayalon A. Dating large infrequent earthquakes by damaged cave deposits. Geology. 2005;33:261-264. DOI: 10.1130/G21193.1
  71. 71. Migowski C, Agnon A, Bookman R, Negendank JF, Stein M. 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. 2004;222:301-314. DOI: 10.1016/j.epsl.2004.02.015
  72. 72. Karcz I, Kafri U, Meshel Z. Archaeological evidence for subrecent seismic activity along the Dead Sea-Jordan rift. Nature. 1977;269:234-235
  73. 73. Haase-Schramm A, Goldstein SL, Stein M. U-Th dating of Lake Lisan (late Pleistocene dead sea) aragonite and implications for glacial East Mediterranean climate change. Geochimica et Cosmochimica Acta. 2004;68:985-1005. DOI: 10.1016/j.gca.2003.07.016
  74. 74. Marco S, Stein M, Agnon A, Ron H. Long-term earthquake clustering: A 50,000-year paleoseismic record in the Dead Sea Graben. Journal of Geophysical Research: Solid Earth. 1996;101(B3):6179-6191. DOI: 10.1029/95JB01587
  75. 75. Waldmann N. The geology of the Samra Formation in the Dead Sea Basin. Unpublished [MSc thesis]. Jerusalem, Israel: The Hebrew University (in English, Hebrew summary). 2002
  76. 76. Marco S. Recognition of earthquake-related damage in archaeological sites: Examples from the Dead Sea fault zone. Tectonophysics. 2008;453(1–4):148-156. DOI: 10.1016/j.tecto.2007.04.011
  77. 77. Marco S, Rockwell TK, Agnon A, Heimann A, Frieslander U. Late Holocene slip of the Dead Sea transform revealed in 3D palaeoseismic trenches on the Jordan gorge fault. Earth and Planetary Science Letters. 2005;234(1–2):189-205. DOI: 10.1016/j.epsl.2005.01.017
  78. 78. Ambraseys NN, Jackson JA. Faulting associated with historical and recent earthquakes in the eastern Mediterranean region. Geophysical Journal International. 1998;133(2):390-406. DOI: 10.1046/j.1365-246X.1998.00508.x
  79. 79. Ambraseys NN, Melville CP, Adams RD. The Seismicity of Egypt, Arabia and the Red Sea: A Historical Review. Cambridge: Cambridge Univ; 1994. DOI: 10.1017/S1356186300007240
  80. 80. Amiran DH, Arieh E, Turcotte T. Earthquakes in Israel and adjacent areas: Macroscopic observations since 100 B.C.E. Israel Exploration Journal. 1994;44:260-305 Available from: http://www.jstor.org/stable/27926357
  81. 81. Enzel Y, Kadan G, Eyal Y. Holocene earthquakes inferred from a fan-delta sequence in the Dead Sea graben. Quaternary Research. 2000;53(1):34-48. DOI: 10.1006/qres.1999.2096
  82. 82. Ken-Tor R, Agnon A, Enzel Y, Stein M, Marco S, Negendank JF. High-resolution geological record of historic earthquakes in the Dead Sea basin. Journal of Geophysical Research-Solid Earth. 2001;106:2221-2234. DOI: 10.1029/2000JB900313
  83. 83. Niemi TM, Ben-Avraham Z. Evidence for the Jericho earthquakes from slumped sediments of the Jordan River delta in the Dead Sea. Geology. 1994;22:395-398. DOI: 10.1130/0091-7613(1994)022<0395:EFJEFS>2.3.CO;2
  84. 84. Russell KW. The Earthquake Chronology of Palestine and Northwest Arabia from the Second through the Mid-8th Century AD. Cambridge, Mass: Bulletin of the American Schools of Oriental Research; 1985. pp. 37-59. DOI: 10.2307/1356863
  85. 85. Ben-Menahem A. Four thousand years of seismicity along the Dead Sea rift. Journal of Geophysical Research. 1991;96:20195-20216. DOI: 10.1029/91JB01936
  86. 86. Hough SE, Avni R. The 1170 and 1202 CE Dead Sea rift earthquakes and long-term magnitude distribution of the Dead Sea fault zone. Israel Journal of Earth Sciences. 2011;58:295-308. DOI: 10.1560/IJES.58.3-4.295
  87. 87. Darvasi Y, Agnon A. Calibrating a new attenuation curve for the Dead Sea region using surface wave dispersion surveys in sites damaged by the 1927 Jericho earthquake. Solid Earth. 2019;10(2):379-390. DOI: 10.5194/se-10-379-2019
  88. 88. Enzel Y, Bookman(Ken-Tor) R, Sharon D, Gvirtzman H, Dayan U, Ziv B, et al. Late Holocene climates of the near east deduced from Dead Sea level variations and modern regional winter rainfall. Quaternary Research. 2003;60:263-273. DOI: 10.1016/j.yqres.2003.07.011
  89. 89. Frumkin A, Magaritz M, Carmi I, Zak I. The Holocene climatic record of the salt caves of Mount Sedom Israel. The Holocene. 1991;1:191-200. DOI: 10.1177/09596836910010
  90. 90. Klein C, Flohn H. Contributions to the knowledge of the fluctuations of the Dead Sea level. Theoretical and Applied Climatology. 1987;38:151-156. DOI: 10.1007/BF00868099
  91. 91. Lisker S, Vaks A, Bar-Matthews M, Porat R, Frumkin A. Stromatolites in caves of the Dead Sea fault escarpment: Implications to latest Pleistocene lake levels and tectonic subsidence. Quaternary Science Reviews. 2009;28:80-92. DOI: 10.1016/j.quascirev.2008.10.015
  92. 92. Migowski C, Stein M, Prasad S, Negendank J, Agnon A. Holocene climate variability and cultural evolution in the near east from the Dead Sea sedimentary record. Quaternary Research. 2006;66(3):421-431. DOI: 10.1016/j.yqres.2006.06.010
  93. 93. Stein M. The sedimentary and geochemical record of Neogene- quaternary water bodies in the Dead Sea Basin – Inferences for the regional paleoclimatic history. Journal of Paleolimnology. 2001;26:271-282
  94. 94. Stein M, Starinsky A, Katz A, Goldstein SL, Machlus M, Schramm A. Strontium isotopic, chemical, and sedimentological evidence for the evolution of Lake Lisan and the Dead Sea. Geochimica et Cosmochimica Acta. 1997;61:3975-3992
  95. 95. Torfstein A, Goldstein SL, Stein M, Enzel Y. Impacts of abrupt climate changes in the Levant from last glacial Dead Sea levels. Quaternary Science Reviews. 2013;60:1-7. DOI: 10.1016/j.quascirev.2013.02.015
  96. 96. Bartov Y, Stein M, Enzel Y, Agnon A, Reches Z. Lake levels and sequence stratigraphy of Lake Lisan, the late Pleistocene precursor of the Dead Sea. Quaternary Research. 2002;57:9-21. DOI: 10.1006/qres.2001.2284
  97. 97. Belmaker R, Lazar B, Beer J, Christl M, Tepelyakov N, Stein M. 10Be dating of Neogene halite. Geochimica et Cosmochimica Acta. 2013;122:418-429. DOI: 10.1016/j.gca.2013.08.033
  98. 98. Starinsky A. Relationship between Ca-chloride brines and sedimentary rocks in Israel [PhD thesis]. The Hebrew Univ. of Jerusalem; 1974
  99. 99. Stein M, Starinsky A, Agnon A, Katz A, Raab M, Spiro B, et al. The impact of brine-rock interaction during marine evaporite formation on the isotopic Sr record in the oceans: Evidence from Mt. Sedom, Israel. Geochimica et Cosmochimica Acta. 2000;64:2039-2053. DOI: 10.1016/S0016-7037(00)00370-7
  100. 100. Steinitz G, Bartov Y. The Miocene-Pliocene history of the Dead Sea segment of the rift in light of K–Ar ages of basalts. Israel Journal of Earth Sciences. 1991;40:199-208
  101. 101. Zak I. The Geology of Mount Sedom [PhD thesis]. Jerusalem: The Hebrew University; 1967
  102. 102. Neev D, Emery KO. The Dead Sea; Depositional Processes and Environments of Evaporites. Jerusalem: Geological Survey; 1967
  103. 103. Torfstein A, Haase-Schramm A, Waldmann N, Kolodny Y, Stein M. U-series and oxygen isotope chronology of the mid-Pleistocene Lake Amora (Dead Sea basin). Geochimica et Cosmochimica Acta. 2009;73:2603-2630. DOI: 10.1016/j.gca.2009.02.010
  104. 104. Waldmann N, Stein M, Ariztegui D, Starinsky A. Stratigraphy, depositional environments and level reconstruction of the last interglacial Lake Samra in the Dead Sea basin. Quaternary Research. 2009;72:1-15
  105. 105. Bartov Y. The geology of the Lisan Formation at Massada plain and Lisan Peninsula [MSc thesis]. The Hebrew University of Jerusalem; 1999
  106. 106. Begin ZB, Broecker W, Buchbinder B, Druckman Y, Kaufman A, Magaritz M, et al. Dead Sea and lake Lisan levels in the last 30,000 years. Geological Survey of Israel. 1985;85:1-18
  107. 107. Yechieli Y, Magaritz M, Levy Y, Weber U, Kafri U, Woelfli W, et al. Late quaternary geological history of the Dead Sea area, Israel. Quaternary Research. 1993;39:59-67. DOI: 10.1006/qres.1993.1007
  108. 108. Bartov Y, Goldstein SL, Stein M, Enzel Y. Catastrophic arid episodes in the eastern Mediterranean linked with the North Atlantic Heinrich events. Geology. 2003;31(5):439-442
  109. 109. Torfstein A, Goldstein SL, Kagan EJ, Stein M. Integrated multi-site U–Th chronology of the last glacial Lake Lisan. Geochimica et Cosmochimica Acta. 2013a;104:210-231. DOI: 10.1016/j.gca.2012.11.003
  110. 110. Bookman R, Bartov Y, Enzel Y, Stein M. Quaternary lake levels in the Dead Sea basin: Two centuries of research. In: Enzel Y, Agnon A, Stein M, editors. New Frontiers in Dead Sea Paleoenvironmental Research. Geological Society of America Special Paper. Vol. 401. Colorado, US. 2006. pp. 155-170. DOI: 10.1130/2006.2401(10)
  111. 111. Bookman R, Enzel Y, Agnon A, Stein M. Late Holocene lake levels of the Dead Sea. Geological Society of America Bulletin. 2004;116:555-571. DOI: 10.1130/B25286.1
  112. 112. Kaufman A, Yechieli Y, Gardosh M. Reevaluation of Lake-sediment chronology in the Dead Sea basin, Israel, based on new 230Th/U dates. Quaternary Research. 1992;38:292-304. DOI: 10.1016/0033-5894(92)90039-L
  113. 113. Schramm A, Stein M, Goldstein SL. Constancy of 234u/Z38u ratios in late Pleistocene aragonitic sediments—Lake Lisan, paleo-Dead Sea. Terra Nova. 1995;7:325
  114. 114. Stein M, Goldstein SL, Ron H, Marco S. Precise TIMS 23 Th-234U ages and magnetostratigraphy of Lake Lisan sediments (paleo-Dead Sea.). EOS Trans. AGU 73, (Spring Meet. Suppl. 154). 1992
  115. 115. Bartov Y, Agnon A, Enzel Y, Stein M. Late quaternary faulting and subsidence in the Central Dead Sea basin. Israel Journal of Earth Sciences. 2006;55:17-31. DOI: 10.1560/E3RK-0RDY-3CUP-LN5A
  116. 116. Waldmann N, Starinsky A, Stein M. Primary carbonates and Ca-chloride brines as monitors of a paleo-hydrological regime in the Dead Sea basin. Quaternary Science Reviews. 2007;26:2219-2228
  117. 117. Kadan G. Evidence of Dead-Sea level fluctuations and neotectonic events in the Holocene fan-delta of Nahal Darga. M.Sc., Ben Gurion University of the Negev. 1997. p. 54
  118. 118. Ben-Avraham Z, Niemi T, Heim C, Negendank J, Nur A. Holocene stratigraphy of the Dead Sea: Correlation of high-resolution seismic reflection profiles to sediment cores. Journal of Geophysical Research. 1999;104(B8):617-625. DOI: 10.1029/1999JB900084
  119. 119. Heim C, Leroy S, Negendank J. Near east desertification—Evidence from the Dead Sea. Naturwissenschaften. 1997;84:398-401
  120. 120. Torfstein A, Gavrieli I, Katz A, Kolodny Y, Stein M. Gypsum as a monitor of the paleo-limnological–hydrological conditions in Lake Lisan and the Dead Sea. Geochimica et Cosmochimica Acta. 2008;72:2491-2509. DOI: 10.1016/j.gca.2008.02.015
  121. 121. Pandey AP, Chadha RK. Surface loading and triggered earthquakes in the Koyna–Warna region, western India. Physics of the Earth and Planetary Interiors. 2003;139:207-223. DOI: 10.1016/j.pepi.2003.08.003
  122. 122. Durá-Gómez I, Talwani P. Reservoir-induced seismicity associated with the Itoiz reservoir, Spain: A case study. Geophysical Journal International. 2010;181(1):343-356. DOI: 10.1111/j.1365-246X.2009.04462.x
  123. 123. Roeloffs EA. Fault stability changes induced beneath a reservoir with cyclic variations in water level. Journal of Geophysical Research - Solid Earth. 1988;93:2107-2124. DOI: 10.1029/JB093iB03p02107
  124. 124. Wang H. Theory of Linear Poroelasticity with Applications to Geomechanics and Hydrogeology. Princeton: University Press; 2000
  125. 125. Simpson DW. Seismicity changes associated with reservoir loading. Engineering Geology. 1976;10(2–4):123-150. DOI: 10.1016/0013-7952(76)90016-8
  126. 126. Anderson EM. The Dynamics of Faulting and Dyke Formation with Applications to Britain. Edinburgh: Oliver and Boyd; 1951
  127. 127. Bell ML, Nur A. Strength changes due to reservoir-induced pore pressure and stresses and application to Lake Oroville. Journal of Geophysical Research - Solid Earth. 1978;83(B9):4469-4483. DOI: 10.1029/JB083iB09p04469
  128. 128. Snow DT. Hydrogeology of induced seismicity and tectonism: Case histories of Kariba and Koyna. In: Narasimhan TN, editor. Recent Trends in Hydrogeology. Geological Society of America Special Paper. Vol. 189. Colorado, US. 1982. pp. 317-360. DOI: 10.1130/SPE189-p317
  129. 129. Rao NP, Shashidhar D. Periodic variation of stress field in the Koyna–Warna reservoir triggered seismic zone inferred from focal mechanism studies. Tectonophysics. 2016;679:29-40. DOI: 10.1016/j.tecto.2016.04.036
  130. 130. Rajendran K, Thulasiraman N, Sreekumari K. Microearthquake activity near the Idukki reservoir, South India: A rare example of renewed triggered seismicity. Engineering Geology. 2013;153:45-52. DOI: 10.1016/j.enggeo.2012.11.004
  131. 131. Crone TJ, Wilcock WS. Modeling the effects of tidal loading on mid-ocean ridge hydrothermal systems. Geochemistry, Geophysics, Geosystems. 2005;6(7):Q07010. DOI: 10.1029/2004GC000905
  132. 132. Fletcher JB, Sykes LR. Earthquakes related to hydraulic mining and natural seismic activity in western New York state. Journal of Geophysical Research. 1977;82:3767-3780. DOI: 10.1029/JB082i026p03767
  133. 133. Healy JH, Rubey WW, Griggs DT, Raleigh CB. The Denver earthquake. Science. 1968;161:1301-1310. DOI: 10.1126/science.161.3848.1301
  134. 134. Raleigh CB, Healy JH, Bredehoeft JD. An experiment in earthquake control at Rangely, Colorado. Science. 1976;191:1230-1237. DOI: 10.1126/science.191.4233.1230
  135. 135. Aldersons F, Ben-Avraham Z, Hofstetter A, Kissling E, Al-Yazjeen T. Lowercrustal strength under the Dead Sea basin from local earthquake data and rheological modeling. Earth and Planetary Science Letters. 2003;214:129-142. DOI: 10.1016/S0012-821X(03)00381-9
  136. 136. Godey S, Bossu R, Guilbert J, Mazet-Roux G. The Euro-Mediterranean Bulletin: A comprehensive seismological Bulletin at regional scale. Seismological Research Letters. 2006;77:460-474. DOI: 10.1785/gssrl.77.4.460
  137. 137. Shamir G. The active structure of the Dead Sea depression. In: Enzel Y, Agnon A, Stein M, editors. New Frontiers in Dead Sea Paleoenvironmental Research. Geological Society of America Special Paper. Vol. 401. Colorado, US. 2006. pp. 15-32. DOI: 10.1130/2006.2401(02)
  138. 138. Dor O, Ben-Zion Y, Rockwell T, Brune J. Pulverized rocks in the Mojave section of the San Andreas fault zone. Earth and Planetary Science Letters. 2006;245:642-654. DOI: 10.1016/j.epsl.2006.03.034
  139. 139. Terzaghi K. Theoretical Soil Mechanics. New York: Wiley; 1942. DOI: 10.1002/9780470172766
  140. 140. Jaeger JC, Cook NG, Zimmerman R. Fundamentals of Rock Mechanics. Oxford: Blackwell; 2009
  141. 141. Weiss J, Pellissier V, Marsan D, Arnaud L, Renard F. Cohesion versus friction in controlling the long-term strength of a self-healing experimental fault. Journal of Geophysical Research - Solid Earth. 2016;121(12):8523-8547. DOI: 10.1002/2016JB013110
  142. 142. Byerlee J. Friction of rocks. In: Rock Friction and Earthquake Prediction. Basel, Switzerland: Birkhäuser; 1978. pp. 615-626. DOI: 10.1007/978-3-0348-7182-2_4
  143. 143. Agnon A, Reches Z. Frictional rheology: Hardening by rotation of active normal faults. Tectonophysics. 1995;247:239-254. DOI: 10.1016/0040-1951(94)00178-C
  144. 144. Stern O. Geochemistry, Hydrology and Paleo-Hydrology of Ein Qedem Spring System, Geological Survey of Israel, Jerusalem, Israel, Report GSI/17/2010. 2010. Available from: https://www.gov.il/he/Departments/publications/reports/stern-report-201 [Accessed: August 2, 2022] (in Hebrew)
  145. 145. Mörner NA. Paleoseismicity and geodynamics in Sweden. Tectonophysics. 1985;117:139-153. DOI: 10.1016/0040-1951(85)90242-2
  146. 146. Calais E, Freed AM, Van Arsdale R, Stein S. Triggering of New Madrid seismicity by late-Pleistocene erosion. Nature. 2010;466:608-611. DOI: 10.1038/nature09258
  147. 147. Heifetz E, Agnon A, Marco S. Soft sediment deformation by kelvin Helmholtz instability: A case from Dead Sea earthquakes. Earth and Planetary Science Letters. 2005;236(1–2):497-504. DOI: 10.1016/j.epsl.2005.04.019
  148. 148. Wetzler N, Marco S, Heifetz E. Quantitative analysis of seismogenic shear-induced turbulence in lake sediments. Geology. 2010;38(4):303-306. DOI: 10.1130/G30685.1

Written By

Mariana Belferman and Amotz Agnon

Submitted: 27 June 2023 Reviewed: 03 October 2023 Published: 15 November 2023