Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 3735 2023-03-09 15:30:08 |
2 format Meta information modification 3735 2023-03-10 02:54:19 | |
3 format -1 word(s) 3734 2023-03-15 03:36:46 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Hinojosa, H.R. Geological Site Effects in Archaeoseismological Point of View. Encyclopedia. Available online: https://encyclopedia.pub/entry/42031 (accessed on 15 May 2024).
Hinojosa HR. Geological Site Effects in Archaeoseismological Point of View. Encyclopedia. Available at: https://encyclopedia.pub/entry/42031. Accessed May 15, 2024.
Hinojosa, Hector R.. "Geological Site Effects in Archaeoseismological Point of View" Encyclopedia, https://encyclopedia.pub/entry/42031 (accessed May 15, 2024).
Hinojosa, H.R. (2023, March 09). Geological Site Effects in Archaeoseismological Point of View. In Encyclopedia. https://encyclopedia.pub/entry/42031
Hinojosa, Hector R.. "Geological Site Effects in Archaeoseismological Point of View." Encyclopedia. Web. 09 March, 2023.
Geological Site Effects in Archaeoseismological Point of View
Edit

Earthquakes have and continue to, occur worldwide, though some places are affected more than others by earthquake-induced ground shaking and the same earthquake can cause more damage in one area than in nearby locations due to site-specific geological site conditions, also known as local site effects. Depending on the chronology of the earthquakes, various disciplines of seismology include instrumental and historical seismology, archaeoseismology, palaeoseismology and neotectonics, each focusing on using specific sources of information to evaluate recent or ancient earthquakes. Past earthquakes are investigated to expand the pre-instrumental and instrumental earthquake catalog and better evaluate a region’s seismic hazard. Archaeoseismology offers a way to achieve these goals because it links how ancient civilizations and their environment might have interacted and responded to past earthquake-induced ground motion and soil amplification. Hence, archaeoseismology explores pre-instrumental (past) earthquakes that might have affected sites of human occupation and their nearby settings, which have left their co-seismic marks in ancient manufactured constructions exhumed by archaeological excavations. However, archaeoseismological observations are often made on a limited epicentral area, poorly constrained dated earthquakes and occasionally on unclear evidence of earthquake damage. Archaeological excavations or field investigations often underestimate the critical role that an archaeological site’s ancient geological site conditions might have played in causing co-seismic structural damage to ancient anthropogenic structures. Nevertheless, the archaeological community might document and inaccurately diagnose structural damage by ancient earthquake shaking to structures and even estimate the size of past earthquakes giving little or no consideration to the role of geological site effects in addressing the causative earthquake. 

ancient earthquakes local site effects forward numerical modeling

1. Geological Site Effects

For a long time, earthquake records have shown that surface ground motions recorded at a given site can vary noticeably even over small inter-site distances [1][2][3][4][5][6]. Ground shaking and possibly induced structural damage to manmade structures are strongly influenced by the rupture mechanism of an earthquake source, the effects of the path traveled by seismic waves and the surface and underground structure of the site where the ground motion is recorded. Each of these three elements (i.e., source, path and site) is a seismological topic and has been investigated by experts in the field for many years (e.g., [3][6][7][8][9]).
When a geologic fault ruptures below the Earth’s surface, seismic energy radiates from the earthquake source in a spherical pattern; however, the radiation pattern of a shear rupture is non-spherical. These body waves are refracted and reflected when they reach the interface between geologic materials with different seismic wave velocities. Therefore, when the seismic rays reach the ground surface, multiple refractions have often bent the seismic rays to a nearly upright direction [10]. Even though seismic waves might travel through tens or hundreds of kilometers in the Earth’s crust and often less than 100 m of soil, the soil deposit strongly influences the characteristics of the ground surface motion [10].
The underground geologic structure, consolidation, variation of the groundwater table, variation of material mechanical properties in the near-subsurface, in addition to the presence of heterogeneities and discontinuities and surface topography can influence amplitude (may amplify or de-amplify motion), the frequency content (may shift to higher or lower) and the duration of strong shaking [6][10][11][12][13][14][15][16]. The amplification of seismic waves is due to crest or valley effects as well as the impedance contrast between horizontally layered sediments and overlying soils (lower impedance) and the underlying bedrock (higher impedance) [14][17]. Soil response depends on the soil’s type, thickness and stiffness. Recognized as the subject of intensive investigation for many years, this concept is referred to as “local site effects[3][6] or its equivalent term, “geological site effects”.
Seismic ground motion and related ground amplification are significant factors influencing the degree of damage to infrastructure [3][5][18][19][20]. A typical scenario of seismic wave amplification occurs during the seismic loading of soil deposits that overlie relatively more rigid bedrock [14][21][22][23].
Nowadays, earthquake engineering practice requires the estimation of the level of ground motion and ground amplification for a given site to assess the seismic vulnerability of infrastructure and the susceptibility of soils during future earthquakes [14][19][24][25][26][27][28]. Nevertheless, the evaluation of geological site effects is relatively sparse in quantitative archaeoseismology. Examples of archaeoseismic investigations considering local site effects include [29][30][31][32][33][34][35][36][37][38][39].

2. Archaeoseismology

Following Hinzen [40], archaeoseismology, also known as earthquake archaeology, is a subdiscipline of seismology that investigates pre-instrumental earthquakes that, by affecting sites of human occupation and their surroundings, have left their physical mark in ancient manufactured structures unearthed by archaeological excavations or on the monumental cultural heritage. These physical marks, relevant for archaeoseismic research, are occasionally (i) displacements along shear planes directly linked to the earthquake fault plane or its branches; (ii) off-fault-shaking effects including fractured building elements, tilted walls, a shift of building elements, lateral distorting, braking and overthrow of walls, rotations of vertically oriented objects; (iii) the secondary shaking effect’s lateral spreading, mass wasting and cyclic mobility as a consequence of soil liquefaction; and (iv) archaeologically detected abandonment of a site and evidence of repair and rebuilding.
Archaeoseismology brings together the efforts of seismologists, archaeologists, earthquake engineers, civil engineers, geologists, geoarchaeologists, architects and historians [41][42] towards the assessment of archaeoseismic evidence, the expansion of both the pre-instrumental and instrumental earthquake catalog and the assessment of the seismic hazard of a region [35][43][44]. Specific questions investigated by archaeoseismology are (i) how probable are seismic ground motions, or secondary earthquake effects, as the cause of damage observed in anthropogenic structures from the past; (ii) when did the damaging ground motion occur and (iii) what can be deduced about the nature of the causing earthquake [45].
Archaeoseismology utilizes data and techniques different from conventional seismology and earthquake geology, which rely on instrumental and historical records and structural data [44]. It is challenging to determine the precise cause of structural damage in archaeological records since various natural causes might yield similar-looking damage patterns and anthropogenic action can also create similar damage or permanent deformation [46][47]. Nonetheless, established qualitative archaeoseismic criteria have helped to distinguish seismic-induced structural damage to ancient structures from other natural and anthropogenic causes [35][40][44][45][48][49][50][51][52][53][54].
Nowadays, archaeological excavation-parallel [41][47][55] or non-excavation [38][56][57] three-dimensional (3D) laser scans of damaged archaeological structures accompanied by a quantitative damage analysis allow a fast and accurate identification, classification, quantification and testing of structural damage at a site and can assist archaeological work during or after archaeological excavation. Moreover, the 3D surface meshes derived from the same scan data can become the basis for developing virtual discrete element models of large and small anthropogenic structures of archaeological context such as rooms, aqueducts, wells, walls, terracotta vessels and figures [38][41][48][55][56][57][58][59]. The available discrete element models can then be used to test their stability using input ground motion signals (i.e., analytical, simulated earthquakes (assumed or historically documented), instrumental earthquakes, or strong motion records) to see if the structures topple or collapse [60], hence, allowing the determination of maximum upper ground motion bounds. Even the reconstruction of the slip velocities during ancient earthquakes based on faulted archaeological structures is now possible (i.e., [61]).
Archaeoseismic investigations have evolved from a qualitative (i.e., [44][50][52][62][63][64][65][66][67][68][69]) to a quantitative approach (i.e., [15][30][34][35][36][37][38][40][45][48][55][59][60][61][70][71][72][73]). The qualitative approach examines the typology of earthquake effects on architectural remains [52], sometimes including the landscape surrounding the site [64]. This kind of approach presents advantages and disadvantages. For instance, the criterion of Stiros [52] identifies earthquake-related structural damage to anthropogenic structures strictly from archaeological data, providing the elimination of natural and anthropogenic causes; however, the technique leaves various unanswered cases of destruction of architecture and abandonment of the site and it does not account for the effects of co-seismic morphological changes to the ground surface. For instance, the criterion of Rodriguez-Pascua et al. [64] utilizes the ground surface’s observed ‘seismic deformation pattern’ and the toppled patterns of archaeological artifacts to construct a theoretical strain ellipsoid for the archaeological site under investigation. However, it does not determine the source parameters of the causative fault. The significant assumptions are that the observed toppled pattern(s) is co-seismic and that the resulting surficial expression of the morphogenic fault has remained unaltered. Then, the systematically derived theoretical strain ellipsoid is compared with the historical-to-present tectonic stress field pattern, active faults, or nearby active seismic zones to gain a deeper insight into the potential earthquake source(s).
Conversely, quantitative archaeoseismic studies of toppled columns strongly suggest that it is not straightforward to deduce a reliable back azimuth toward the earthquake source based on the deformation and toppled patterns of manufactured structures [58][74]. Therefore, it is impossible to establish a direct link between the orientation of a fallen object and the tectonic stress field of a past earthquake. The method of Rodriguez-Pascua et al. [64] has somewhat limited quantitative applicability, so conclusive interpretations from their approach should be considered cautiously. Buck [51] provides a literature review and thoroughly examines the several qualitative methodologies adopted to appraise archaeoseismic damage. She concludes that, when using the universal identification criteria (e.g., ‘check-list’ approach), interpretations of qualitative observations are commonly subjective and lack the site’s human and physical context. Therefore, she proposes a project-specific interdisciplinary approach to assess archaeoseismic damage objectively.
Moreover, the systematically designed quantitative archaeoseismic approaches of Galadini et al. [40] and Hinzen et al. [48] test ‘archaeoseismic evidence’ before considering it reliable for quantitative comparison against the observed damage structures. These methods propose an analytical/numerical modeling procedure for archaeoseismic projects. The approach builds upon available upfront and newly collected geotechnical, geological, geophysical, geoarchaeological, archaeological and historical data. In most cases, newly collected field or laboratory data (e.g., geological, geophysical and geotechnical) is tailored to answer specific archaeoseismic questions [45]. Following the quantitative procedure, an archaeoseismic project is likely to become unique (cf. [51]). An up-to-date summary of archaeoseismological studies using advanced measuring methods and quantitative numerical modeling is given by Hinzen et al. [48]. Galadini et al. [40] discuss the methodologies and procedures in archaeoseismological research in detail.

3. Geological Site Effects in an Archaeoseismological Context

The primary objectives of a quantitative archaeoseismic investigation are to estimate the ground motion that caused the damage [48] and obtain information about the earthquake source that caused the ground motion [40]. Hinzen et al. [37] point out that archaeoseismic observations are often limited to a small portion of the meso-seismal area and uncertainties often hinder the correlation of damage across several neighboring sites in dating the damaging events [40][68]. These factors can strongly bias the estimation of the strength of ancient earthquakes; therefore, the consideration and systematic assessment of local seismic site effects become critical in an archaeoseismic study [37].
In principle, neglecting ground amplification in archaeoseismological studies might lead to an overestimation of the size of an ancient damaging earthquake [45]. For seismic ground motion simulations, the use of only one horizontal component as an earthquake input signal in site response analysis can lead to a significant underestimation of seismic site response [17] and the dynamic soil properties (e.g., density, shear wave velocity, damping) should (preferably) be measured in situ [75]. Hence, if the goal is to estimate local site effects in archaeoseismology, it is appropriate to implement some quantitative tools used in earthquake engineering (cf. [40]). The estimation of surface ground motion can be carried out empirically with records of actual earthquakes (cf. [37]) or numerically with the stochastic or Green’s function methods [38][39][76]. Field tests and analytical/numerical models can assess the characteristics of seismic site amplification [15][37][77], recording and analyzing sites’ dynamic responses using active sources, ambient noise and actual earthquakes.
Analytical/numerical models are convenient in quantitative archaeoseismology because they can develop an understanding of seismic wave propagation characteristics of sedimentary basins when instrumentally recorded earthquake records and macro-seismic intensity data from historical records are absent [25][30][33][78][79][80]. These models require a conceptualized geotechnical model containing the geometry of all soil layers from bedrock to surface, their dynamic properties, the incident bedrock motions and ‘realistic’ synthetic earthquake records mainly obtained from rock sites. Synthetic earthquake ground motions are calculated based on carefully selected earthquake source parameters (i.e., rupture length, rupture width, seismic moment (Mo) and moment magnitude (Mw)) linked to a seismotectonic model representative of the region of interest. Posteriorly, these synthetic ground motions are used as the earthquake input signal to calculate site amplifications and the resulting site-specific surface ground motion.

4. Criteria for Forward Modeling Geological Site Effects in Archaeoseismology

From the seismic risk assessment point of view, an archaeological site within a seismically active region might be equivalent to a geotechnical site in a seismic region, even more so if the archaeological site records ancient structural damage to manmade structures. Generally, most archaeological sites worldwide occur in sedimentary basins or valleys; however, they also occur on topographic highs or slopes where topographic amplification can be an issue. The former observation is valid because most civilizations settled on accessible land with convenient environmental conditions that provided ecosystem services such as proximity to water, fertile arable land and canopy, among other natural features. A one-dimensional (1D) local site response (LSR) analysis is standard in geotechnical earthquake engineering. Its goal is to estimate the nonlinear cyclic response of soils subjected to earthquake-induced ground shaking, with either a nonlinear model or the equivalent linear model. A 1D LSR analysis can capture the essential aspects of surface ground response; however, it cannot model sloping, irregular ground surfaces, basin effects, topographic effects and embedded geologic structures, which are assessed adequately with 2D or 3D models. However, for most archaeological project aims and budgets, 2D or 3D modeling efforts are not cost-effective. Moreover, the 1D LSR analysis solves the problem of horizontally polarized vertically propagating shear waves with planar wavefronts from the bedrock into horizontally layered soils with frequency-independent damping (i.e., valley-like geology). The 1D LSR analysis considers the wave modification properties of layered, damped soil deposits overlying weathered or unweathered elastic bedrock.
Furthermore, following Zhang and Zhao [81], the use of 1D site models from sedimentary basins with a width-to-depth ratio (WDR) ≥ 6 is valid for a 1D LSR analysis. So for cases where the sedimentary basin’s width is much greater than its thickness, the 1D LSR analysis is justifiable. In addition, it is cost-effective for archaeological projects’ budgets.
The forward modeling of geological site effects through a 1D LSR analysis requires the proper knowledge of the site model parameters, including the conditions of the ancient ground surface and subsurface knowledge of the earth material properties (e.g., density, shear-wave velocity and seismic wave attenuation) and the computation of synthetic seismograms (i.e., surface acceleration time-series) using earthquake source parameters of hypothetical causative earthquake scenarios. However, the selected earthquake scenarios should fall under the seismotectonic context of the region under investigation. This information serves as input for the actual calculations of the site-specific 1D LSR analysis.

4.1. Seismotectonic Model

In archaeology, written records of ancient earthquakes devastating manmade structures are scarce to non-existent (cf. [72]). Therefore, researchers must hypothesize about the possible earthquake scenario that might be a causative earthquake and can explain the destruction patterns documented by archaeological field observations. This step requires insights from earthquake seismology, tectonics and geology to gather information. In this step, several earthquake scenarios must be developed and each one must contain realistic earthquake source parameters describing the possible physical conditions of the hypothetical causative rupturing fault. Of course, the fault geometry and faulting mechanism must stem from local field observations. The earthquake source parameters include moment magnitude (Mw), seismic moment (Mo), earthquake stress-drop (Δσ), surface rupture length (SRL), fault’s structural data (i.e., strike, dip and rake), hypocenter depth, rupture velocity and the reference depth (to fault’s upper edge. Values for these source parameters are found in a thorough literature search and are investigated and assigned by a seismologist. The source parameters become the input parameters to calculate synthetic earthquake-induced ground acceleration seismograms for reference sites, which excite the site-specific 1D geotechnical models.

4.2. Site-Specific 1D Soil Models

When testing the plausibility of earthquake-induced destructions of an archaeological site, the evaluation becomes a geotechnical earthquake engineering problem. Therefore, retrospective site-specific 1D soil models must be developed for multiple locations throughout the archaeological site or complex.
Archaeological and geoarchaeological excavations, geophysical surveys and geological and geotechnical studies from the study area provide essential site-specific information required to estimate the geological site effects and define a site’s seismic response. Archaeological and geoarchaeological excavations provide information about the texture, density, type, age and thickness of the shallow soils and sediments that pre- and postdate the stratigraphic horizon of interest; however, these excavations rarely reach the soil–bedrock interface. The removal of the overburden (i.e., material that postdates the horizon of interest) and the depth to the soil–bedrock interface are required for a realistic and accurate estimation of geological site effects. Without deeper boreholes and geophysical surveys, these should be pursued to detect the soil–bedrock contact and gain information about possible soil and bedrock heterogeneities. In general, seismic methods (reflection or refraction) provide an in situ measurement of the P and S wave velocities, while geoelectrical and electromagnetic methods can detect and discriminate between fine-grained soils (cohesive) from coarse-grained soils (granular).
Inaccurate knowledge of the actual composition, thickness and dynamic properties of the subsurface materials can lead to the misrepresentation of the site and the inaccurate selection of strain-dependent shear modulus and damping values for individual material layers and uncertainties in the forward calculation of frequency-dependent surface amplifications, surface ground-motions and the estimation of the seismic site response (cf. [82]). The site class definition is essential for seismic site-specific response analysis. The geotechnical site classification scheme of Rodríguez-Marek et al. [82] is the most adequate for archaeoseismic research because it allows an accurate stratigraphic representation of the regolith column compared to the geologic and geophysical site classification schemes. Therefore, the dynamic response of ancient anthropogenic structures is estimated with better accuracy. The geotechnical site classification system is based on several observable parameters: the type of deposit (i.e., hard rock, competent rock, weathered rock, stiff soil, soft soil and potentially liquefiable sand), which automatically introduces a measure of the dynamic stiffness (Vs30) to the classification system; depth to bedrock defined by Vs30 > 760 m/s or to a significant seismic impedance contrast between surficial soil deposits and geologic material with a Vs ≈ 760 m/s; the depositional age of the soil(s) (i.e., Holocene or Pleistocene); and soil-type (i.e., cohesive or granular). The geotechnical site classification system breaks down sites traditionally grouped as “rock” into competent rock sites and weathered soft-rock/shallow stiff soil sites. This subdivision significantly reduces uncertainty in defining site-dependent surface ground motions and allows more accurate determination of proper model parameters and dynamic properties to individual material layers.
Conversely, the geologic site classification scheme is based on one or more parameters obtained from surficial geologic observations, namely geologic age-only, age-and-depositional environment, or age-and-sediment texture [83]. This classification system does not provide information about the bedrock’s depth and stiffness, a discriminating factor for a seismic site response analysis [82].
Moreover, the geophysical site classification scheme is based solely on the uppermost 30 m of the surface, Vs30 [84]. The use of the Vs30 has the advantage of uniformity within the 30 m depth range and correlates well with detailed surface geology (i.e., age-and-soil texture and age-and-weathering/fracture spacing for rock) [85][86]; however, it is still an oversimplification of most natural site conditions and, therefore, an indirect approach to define the actual composition and stratigraphy of the near-surface materials and to estimate the soil–bedrock interface.

4.3. Computational Tasks: Forward Modeling of Synthetic Seismograms and 1D LSR

This stage deals with the computation of synthetic acceleration seismograms of the reference site and the site-specific 1D LSR analysis. The synthetic seismograms from hypothetical earthquake scenarios are used as input signals in the 1D LSR analysis. The 1D LSR analysis deals with the ground surface’s dynamic loading during the earthquake’s duration. Each site-specific 1D LSR analysis yields a surface acceleration record and a seismic amplification factor.
The choice of computer codes to calculate synthetic seismograms and to forward model the equivalent-linear 1D site response is also essential, tasks that a seismologist or geotechnical earthquake engineer should perform. Nowadays, high-frequency synthetic seismograms are computed for modeling geological site effects [87][88]. Most high-frequency motions might be caused by direct P and S waves [89]. For modeling purposes, the computed synthetic seismogram should contain such body waves. However, not all algorithms can compute synthetic seismograms with distinct body waves, surface waves and high-frequencies (e.g., [90]). Nevertheless, the synthetic seismogram should be computed with an algorithm that allows the computation of high-frequency synthetic seismograms (i.e., near-field motions) with recognizable body and surface waves (e.g., [88]).
Several computer codes can model the nonlinear–elastic stress-strain behavior of a realistic 1D regolith column or geotechnical model [25][91][92][93][94]. The open-source code of Robinson et al. [93] used to forward the equivalent-linear 1D site-response model is an adequate code to forward model 1D local site effects in archaeoseismology. The code can incorporate uncertainties in the geotechnical model parameters, including density, shear-wave velocities and layer thickness. It also allows for defining thin material layers (<1 m thick), assigning strain-dependent shear modulus and degradation damping curves to each material layer and using either an average or a gradient shear-wave velocity model. While the shear-wave velocity of the subsurface materials within the archaeological site (e.g., 1D soil models) is measured in situ with near-surface seismic methods, the strain-dependent shear modulus (G/Gmax) and damping ratio (ξ) degradation curves of the same materials can and should be tested in a laboratory. It is worth mentioning that these elastic parameters are noticeably different for rocks (sedimentary, igneous, or metamorphic) and soft sediments or soils (clay, sand, gravel); hence it is crucial to understand the site’s stratigraphy fully.
The shear-wave velocity of the material layers is a fundamental model parameter for an equivalent-linear 1D site-specific response analysis. The use of an average shear-wave velocity (i.e., model A) for each material layer is an acceptable approach in the absence of depth-dependent (gradient) shear-wave velocities (model B). Nevertheless, the implementation of model A and model B yield similar results: model A typically produces a slightly higher amplification peak at a slightly lower frequency value than model B. The use of model B should be the first choice if available information (geologic and seismic) shows an increase in geologic age, density, consolidation and shear strength with increasing depth. In this way, the stiffness of the near-surface materials would be accurately represented. Conversely, model A should be adopted when the presence of homogenous material layers is demonstrated by data from geologic logs and archaeological and geoarchaeological excavations or when a gradient velocity model cannot be constrained for the depth interval of interest.

References

  1. Reid, H.F. The Mechanics of the earthquake, v. II of Lawson, A.C., chairman. In the California Earthquake of April 18, 1906: Report of the State Earthquake Investigation Commission No. 87; Carnegie Institution of Washington Publication: Washington, DC, USA, 1910; 192p.
  2. Hough, S.E.; Friberg, P.A.; Busby, R.; Field, E.F.; Jacob, K.H.; Borcherdt, R.D. Sediment induced amplification and the collapse of the Mimitz freeway. Nature 1990, 344, 853–855.
  3. Aki, K. Local site effects on weak and strong ground motion. Tectonophysics 1993, 218, 93–111.
  4. Lermo, J.; Chávez-García, F.J. Site effect evaluation at Mexico city: Dominant period and relative amplification from strong motion and microtremor records. Soil Dyn. Earthq. Eng. 1994, 13, 413–423.
  5. Chávez-García, F.J.; Cuenca, J. Site effects in Mexico city urban zone. A complementary study. Soil Dyn. Earthq. Eng. 1996, 15, 141–146.
  6. Boore, D.M. Can site response be predicted? J. Earthq. Eng. 2004, 8, 1–41.
  7. Lachet, C.; Hatzfeld, D.; Bard, P.Y.; Theodulidis, N.; Papaioannou, C.; Savvaidis, A. Site effects and microzonation in the city of Thessaloniki (Greece) comparison of different approaches. Bull. Seismol. Soc. Am. 1996, 86, 1692–1703.
  8. Thompson, E.M.; Baise, L.G.; Tanaka, Y.; Kayen, R.E. A taxonomy of site response complexity. Soil Dyn. Earthq. Eng. 2012, 41, 32–43.
  9. Boore, D.M.; Thompson, E.M. Path durations for use in the stochastic-method simulation of ground motions. Bull. Seismol. Soc. Am. 2014, 104, 2541–2552.
  10. Kramer, S.L. Geotechnical earthquake engineering. In Prentice-Hall International Series in Civil Engineering and Engineering Mechanics, 1st ed.; Prentice-Hall: Hoboken, NJ, USA, 1996; 653p.
  11. Boore, D.M. A note on the effect of simple topography on seismic SH waves. Bull. Seismol. Soc. Am. 1972, 62, 275–284.
  12. Davies, L.T.; West, L.R. Observed effects of topography on ground motion. Bull. Seismol. Soc. Am. 1973, 63, 283–298.
  13. Geli, L.; Bard, P.Y.; Jullien, B. The effect of topography on earthquake ground motion: A review and new results. Bull. Seismol. Soc. Am. 1988, 78, 42–63.
  14. Şafak, E. Local site effects and dynamic soil behavior. Soil Dyn. Earthq. Eng. 2001, 21, 453–458.
  15. Bensalem, R.; Chatelain, J.L.; Mechane, D.; Oubaiche, E.H.; Hellel, M.; Guiller, B.; Djeddi, M.; Djadia, L. Ambient vibration techniques applied to explain heavy damages caused in Corso (Algeria) by the 2003 Boumerdes earthquake: Understanding seismic amplification due to gentle slopes. Seismol. Res. Lett. 2010, 81, 928–940.
  16. Chen, Q.S.; Gao, G.Y.; Yang, J. Dynamic response of deep soft soil deposits under multi-directional earthquake loading. Eng. Geol. 2011, 121, 55–65.
  17. Cornou, C.; Bard, P.Y. Site-to-bedrock over 1D transfer function ratio: An indicator of the proportion of edge-generated surface waves? Geophys. Res. Lett. 2003, 30, 1453.
  18. Atkinson, G.M.; Boore, D.M. Ground-motion relations for eastern North America. Bull. Seismol. Soc. Am. 1995, 85, 17–30.
  19. Kawase, H. Site effects on strong ground motions. In International Handbook of Earthquake & Engineering Seismology: Part B; William, H.K.L., Kanamori, H., Jennings, P.C., Kisslinger, C., Eds.; International Association of Seismology and Physics of the Earth’s Interior (IASPEI): Tallinn, Estonia, 2003; pp. 1013–1030.
  20. Fritsche, S.; Fäh, D. The 1946 magnitude 6.1 earthquake in the Valais: Site-effects as contributor to the damage. Swiss J. Geosci. 2009, 102, 423–439.
  21. Mohraz, B. A study of earthquake response spectra for different geological conditions. Bull. Seismol. Soc. Am. 1976, 66, 915–935.
  22. Bard, P.Y.; Bouchon, M. The two-dimensional resonance of sediment-filled valleys. Bull. Seismol. Soc. Am. 1985, 75, 519–541.
  23. Fletcher, J.B.; Boatwright, J. Site response and basin waves in the Sacramento-San Joaquin Delta, California. Bull. Seismol. Soc. Am. 2013, 103, 196–210.
  24. Raptakis, D.; Chávez-García, F.J.; Pitilakis, K. Site effects at Euroseistest–I. Determination of the valley structure and confrontation of observations with 1D analysis. Soil Dyn. Earthq. Eng. 2000, 19, 1–22.
  25. Hashash, Y.M.A.; Park, D. Nonlinear one-dimensional seismic ground motion propagation in the Mississippi embayment. Eng. Geol. 2001, 62, 185–206.
  26. Sørensen, M.B.; Oprsal, I.; Bonnefoy-Claudet, S.; Atakan, K.; Mai, P.M.; Pulido, N.; Yalciner, K. Local site effects in Ataköy, Istanbul, Turkey, due to a future large earthquake in the Marmara Sea. Geophys. J. Int. 2006, 167, 1413–1424.
  27. Koçkar, M.K.; Akgün, H. Evaluation of the site effects of the Ankara basin, Turkey. J. Appl. Geophys. 2012, 83, 120–134.
  28. Maufroy, E.; Chaljub, E.; Hollender, F.; Kristek, J.; Moczo, P.; Klin, P.; Priolo, E.; Iwaki, A.; Iwata, T.; Etienne, V.; et al. Earthquake ground motion in the Mydgonian basin, Greece: The E2VP verification and validation of 3D numerical simulation up to 4 Hz. Bull. Seismol. Soc. Am. 2015, 105, 1398–1418.
  29. Hinojosa-Prieto, H.R. Local Site Effects in Archaeoseismology: Examples from the Mycenaean Citadels of Tiryns and Midea (Argive Basin, Peloponnese, Greece). Ph.D. Thesis, Universität zu Köln, Cologne, Germany, 2016. Available online: http://kups.ub.uni-koeln.de/id/eprint/6650 (accessed on 20 December 2022).
  30. Hinzen, K.G. The use of engineering seismological models to interpret archaeoseismological findings in Tolbiacum, Germany: A case study. Bull. Seismol. Soc. Am. 2005, 95, 521–539.
  31. Fäh, D.; Steimen, S.; Oprsal, I.; Ripperger, J.; Wössner, J.; Schatzmann, R.; Kästil, P.; Spottke, I.; Huggenberger, P. The earthquake of 250 AD in Augusta Raurica, a real event with a 3D effect? J. Seismol. 2006, 10, 459–477.
  32. Harbi, A.; Maouchea, S.; Vaccari, F.; Aoudia, A.; Oussadou, F.; Panza, G.F.; Benouar, D. Seismicity, seismic input and site effects in the Sahel—Algiers region (North Algeria). Soil Dyn. Earthq. Eng. 2007, 27, 427–447.
  33. Bottari, C.; Bottari, A.; Carveni, P.; Saccà, C.; Spigo, U.; Teramo, A. Evidence of seismic deformation of the paved floor of the decumanus at Tindari (SE, Sicily). Geophys. J. Int. 2008, 174, 213–222.
  34. Hinzen, K.G.; Weiner, J. Testing a Seismic Scenario for the Damage of the Neolithic Wooden Well of Erkelenz-Kückhoven, Germany; Special Publications 316; Geological Society: London, UK, 2009; pp. 189–205.
  35. Caputo, R.; Hinzen, K.G.; Liberatore, D.; Schreiber, S.; Helly, B.; Tziafalias, A. Quantitative archaeoseismological investigation of the Great Theatre of Larissa, Greece. Bull. Earthq. Eng. 2010, 9, 347–366.
  36. Hinzen, K.G.; Kehmeier, H.; Schreiber, S. Quantitative archaeoseismological study of a Roman Mausoleum in Pinara (Turkey)-testing seismogenic and rockfall damage scenarios. Bull. Seismol. Soc. Am. 2013, 103, 1008–1021.
  37. Hinzen KGHinojosa-Prieto, H.R.; Kalytta, T. Site Effects in Archaeoseismic Studies at Mycenaean Tiryns and Midea. Seismol. Res. Lett. 2016, 87, 1060–1074.
  38. Hinzen, K.G.; Maran, J.; Hinojosa-Prieto, H.; Damm-Meinhardt, U.; Reamer, S.K.; Tzislakis, J.; Kemna, K.; Schweppe, G.; Fleischer, C.; Demakopoulou, K. Reassessing the Mycenaean Earthquake Hypothesis: Results of the HERACLES Project from Tiryns and Midea, Greece Reassessing the Mycenaean Earthquake Hypothesis. Bull. Seismol. Soc. Am. 2018, 108, 1046–1070.
  39. Hinojosa-Prieto, H.R. Estimation of the moment magnitude and local site effects of a postulated Late Bronze Age earthquake: Mycenaean citadels of Tiryns and Midea, Greece. Ann. Geophys. 2020, 63, SE331.
  40. Galadini, F.; Hinzen, K.G.; Stiros, S. Archaeoseismology: Methodological issues and procedure. J. Seismol. 2006, 10, 395–414.
  41. Schreiber, S.; Hinzen, K.G.; Fleischer, C.; Schütte, S. Excavation-parallel laser scanning of a medieval cesspit in the archaeological zone Cologne, Germany. ACM J. Comput. Cult. Herit. 2012, 5, 1–22.
  42. Jusseret, S. Contextualizing the birth of Mediterranean archaeology. Antiquity 2014, 88, 964–974.
  43. Caputo, R.; Helly, B. The use of distinct disciplines to investigate past earthquakes. Tectonophysics 2008, 453, 7–19.
  44. Bottari, C.; Stiros, S.C.; Teramo, A. Archaeological evidence for destructive earthquakes in Sicily between 400 BC and AD 600. Geoarchaeology 2009, 24, 147–175.
  45. Hinzen, K.G. Archaeoseismology. In Encyclopedia of Solid Earth Geophysics; Gupta, H.K., Ed.; Springer: Berlin/Heidelberg, Germany, 2011; pp. 11–15.
  46. Nikonov, A.A. Reconstruction of the main parameters of old large earthquakes in Soviet central Asia using the palaeoseismological method. Tectonophysics 1988, 147, 297–312.
  47. Hinzen, K.G.; Schreiber, S.; Fleischer, C.; Reamer, S.K.; Wisona, I. Archaeoseismic study of damage in Roman and Medieval structures in the center of Cologne, Germany. J. Seismol. 2012, 17, 399–424.
  48. Hinzen, K.G.; Fleischer, C.; Reamer, S.K.; Schreiber, S.; Schütte, S.; Yeril, B. Quantitative methods in archaeoseismology. Quat. Int. 2011, 242, 31–41.
  49. Karcz, I.; Kafri, U. Evaluation of supposed archaeoseismic damage in Israel. J. Archaeol. Sci. 1978, 5, 237–253.
  50. Rapp, G., Jr. Assessing archaeological evidence for seismic catastrophes. Geoarchaeology 1986, 1, 365–379.
  51. Buck, V. Archaeoseismology in the Atalanti Region, Central Mainland Greece: Theories, Methods and Practice; BAR International Series 1552; Archaeopress: Oxford, UK, 2006; 110p.
  52. Stiros, S. Identification of earthquakes from archaeological data: Methodology, criteria and limitations. In Archaeoseismology; Stiros, S., Jones, R.E., Eds.; Fitch Laboratory Occasional Paper No. 7 Athens; Short Run Press: Exeter, UK, 1996; pp. 129–152.
  53. Ambraseys, N.N. Archaeoseismology and neocatastrophism. Seismol. Res. Lett. 2005, 76, 560–564.
  54. Marco, S. Recognition of earthquake-related damage in archaeological sites: Examples from the Dead Sea fault zone. Tectonophysics 2008, 453, 148–156.
  55. Hinzen, K.G.; Schreiber, S.; Yerli, B. The Lycian sarcophagus of Arttumpara, Turkey: Testing seismogenic and anthropogenic damage scenarios. Bull. Seismol. Soc. Am. 2010, 100, 3148–3164.
  56. Hinzen, K.G.; Schwellenbach, I.; Schweppe, G.; Marco, S. Quantifying earthquake effects on ancient arches, example: The Kalat Nimrod Fortress, Dead Sea fault zone. Seismol. Res. Lett. 2016, 87, 751–764.
  57. Hinzen, K.G.; Meghraoui, M.; Bahrouni, N.; Reamer, S.K. Testing the earthquake damage and vulnerability of the Cherichira aqueduct bridge, Kairouan (Tunisia) with discrete element modeling. Mediterr. Geosci. Rev. 2022, 4, 495–516.
  58. Hinzen, K.G. Sensitivity of earthquake toppled columns to small changes in ground motion and geometry. Isr. J. Earth Sci. 2012, 58, 309–326.
  59. Hinzen, K.G.; Vetters, M.; Kalytta, T.; Reamer, S.K.; Damm-Meinhardt, U. Testing the response of Mycenaean terracotta figures and vessels to earthquake ground motions. Geoarchaeology 2015, 30, 1–18.
  60. Schweppe, G.; Hinzen, K.G.; Reamer, S.K.; Fischer, M.; Marco, S. The ruin of the Roman Temple of Kedesh, Israel; example of a precariously balanced archaeological structure used as a seismoscope. Ann. Geophys. 2017, 60, S0444.
  61. Schweppe, G.; Hinzen, K.G.; Reamer, S.K.; Marco, S. Reconstructing the slip velocities of the 1202 and 1759 CE earthquakes based on faulted archaeological structures at Tell Ateret, Dead Sea Fault. J. Seismol. 2021, 25, 1021–1042.
  62. Di Vita, A. Archaeologist and earthquakes: The case of 365 AD. Ann. Geofis. 1995, XXXVIII, 971–976.
  63. Guidoboni, E.; Bianchi, S.S. Collapse and seismic collapses in archaeology: Proposal for a thematic atlas. Ann. Geofis. 1995, XXXVIII, 1013–1017.
  64. Rodríguez-Pascua, M.A.; Pérez-López, R.; Giner-Robles, J.L.; Silva, P.G.; Garduño-Monroy, V.H.; Reicherter, K. A comprehensive classification of earthquake archaeological effects (EAE) in archaeoseismology: Application to ancient remains of Roman and Mesoamerican cultures. Quat. Int. 2011, 242, 20–30.
  65. Sintubin, M. Archaeoseismology: Past, present and future. Quat. Int. 2011, 242, 4–10.
  66. Garduño-Monroy, V.H.; Macías, J.L.; Oliveros, A.; Hernández-Madrigal, V.M. Progress in seismic and archaeoseismic studies in the zone of Mitla, Oaxaca. Earthquake Geology and Archaeology: Science, society and seismic hazard. In Proceedings of the 3rd INQUA-IGCP 567 International Workshop on Earthquake Geology, Palaeoseismology and Archaeoseismology, Morelia, Mexico, 19–24 November 2012; Volume 3, pp. 43–46.
  67. Garduño-Monroy, V.H.; Benavente-Escóbar, C.; Oliveros, A.; Rodríguez-Pascua, M.A.; Pérez-López, R.; Giner, J.L. Evidence of past seisms in Cusco (Peru) and Tzintzuntzan (Mexico): Cultural Relations. Earthquake Geology and Archaeology: Science, society and Seismic hazard. In Proceedings of the 3rd INQUA-IGCP 567 International Workshop on Earthquake Geology, Palaeoseismology and Archaeoseismology, Morelia, Mexico, 19–24 November 2012; Volume 3, pp. 47–50.
  68. Jusseret, S.; Langohr, C.; Sintubin, M. Tracking earthquake archaeological evidence in Late Minoan IIIB (~1300–1200 BC.) Crete (Greece): A proof of concept. Bull. Seismol. Soc. Am. 2013, 103, 3026–3043.
  69. Stiros, S.C.; Pytharouli, S.I. Archaeological evidence for a destructive earthquake in Patras, Greece. J. Seismol. 2014, 18, 687–693.
  70. Tertulliani, A.; Graziani, L.; Esposito, A. How Historical Seismology can Benefit from Bureaucracy: The Case of the “Lettere Patenti” in the City of Rome in 1703. Seismol. Res. Lett. 2020, 91, 2511–2519.
  71. Suter, M. Macroseismic Study of the Devastating 22–23 October 1749 Earthquake Doublet in the Northern Colima Graben (Trans-Mexican Volcanic Belt, Western Mexico). Seismol. Res. Lett. 2019, 90, 2304–2317.
  72. Triantafyllou, I.; Koukouvelas, I.; Papadopoulos, G.A.; Lekkas, E. A Reappraisal of the Destructive Earthquake (Mw 5.9) of 15 July 1909 in Western Greece. Geosciences 2022, 12, 374.
  73. Tendürüs, M.; van Wijngaarden, G.J.; Kars, H. Long-Term Effect of Seismic Activities on Archaeological Remains: A Test Study from Zakynthos, Greece, in Ancient Earthquakes; Sintubin, M., Stewart, I.S., Niemi, T.M., Altunel, E., Eds.; Geological Society of America Special Paper 471; The Geological Society of America: Boulder, CO, USA, 2010; pp. 145–156.
  74. Hinzen, K.G. Simulation of toppling columns in archaeoseismology. Bull. Seismol. Soc. Am. 2009, 99, 2855–2875.
  75. Meng, J. Earthquake ground motion simulation with frequency-dependent soil properties. Soil Dyn. Earthq. Eng. 2007, 27, 234–241.
  76. Somerville, P.; Moriwaki, Y. Seismic hazard and risk assessment in engineering practice. In International Handbook of Earthquake & Engineering Seismology: Part B; William, H.K.L., Kanamori, H., Jennings, P.C., Kisslinger, C., Eds.; International Association of Seismology and Physics of the Earth’s Interior (IASPEI): Tallinn, Estonia, 2003; pp. 1065–1080.
  77. Chávez-García, F.J.; Rodríguez, M.; Field, E.H.; Hatzfeld, D. Topographic site effects. A comparison of two non-reference methods. Bull. Seismol. Soc. Am. 1997, 87, 1667–1673.
  78. Berilgen, M.M. Evaluation of local site effects on earthquake damages of Fatih mosque. Eng. Geol. 2007, 91, 240–253.
  79. Karastathis, V.K.; Karmis, P.; Novikova, T.; Roumelioti, Z.; Gerolymatou, E.; Papanastassiou, D.; Lakopoulos, S.; Tsombos, P.; Papadopoulos, G.A. The contribution to geophysical techniques to site characterization and liquefaction risk assessment: Case study of Nafplion city, Greece. J. Appl. Geophys. 2010, 72, 194–211.
  80. Karastathis, V.K.; Papadopoulos, G.A.; Novikova, T.; Roumelioti, Z.; Karmis, P.; Tsombos, P. Prediction and evaluation of nonlinear site response with potentially liquefiable layers in the area of Nafplion (Peloponnesus, Greece) for a repeat of historical earthquakes. Nat. Hazards Earth Syst. Sci. 2010, 10, 2281–2304.
  81. Zhang, J.; Zhao, J.X. Response spectral amplification ratios from 1- and 2-dimensional nonlinear soil site models. Soil Dyn. Earthq. Eng. 2009, 29, 563–573.
  82. Rodríguez-Marek, A.; Bray, J.D.; Abrahamson, N.A. An empirical geotechnical seismic site response procedure. Earthq. Spectra 2001, 17, 65–87.
  83. Stewart, J.P.; Liu, A.H.; Choi, Y. Amplification factors for spectral acceleration in tectonically active regions. Bull. Seismol. Soc. Am. 2003, 93, 332–352.
  84. Borcherdt, R.D. Estimates of site-dependent response spectra for design (methodology and justification). Earthq. Spectra 1994, 10, 617–653.
  85. Wald, D.J.; Allen, T.I. Topographic slope as a proxy for seismic site conditions and amplification. Bull. Seismol. Soc. Am. 2007, 97, 1379–1395.
  86. Stewart, J.P.; Klimis, N.; Savvaidis Al Theodoulidis, N.; Zargli, E.; Athanasopoulos, G.; Pelekis, P.; Mylonakis, G.; Margaris, B. Compilation of a local VS profile database and its application for interface of VS30 from geologic- and terrain-based proxies. Bull. Seismol. Soc. Am. 2014, 104, 2827–2841.
  87. Friederich, W.; Dalkolmo, J. Complete synthetic seismograms for a spherically symmetric earth by a numerical computation of the Green’s function in the frequency domain. Geophys. J. Int. 1995, 122, 537–550.
  88. Wang, R. A simple orthonormalization method for stable and efficient computation of Green’s functions. Bull. Seismol. Soc. Am. 1999, 89, 733–741.
  89. Spudich, P.; Frazer, L.N. Use of ray theory to calculate high-frequency radiation from earthquake source having spatially variable rupture velocity and stress drop. Bull. Seismol. Soc. Am. 1984, 74, 2061–2082.
  90. Beresnev, I.A.; Atkinson, G.M. FINSIM–A FORTRAN program for simulating stochastic acceleration time histories from finite faults. Seismol. Res. Lett. 1998, 69, 27–32.
  91. Seed, H.B.; Idriss, I.M. Influence of soil conditions on ground motions during earthquakes. ASCE J. Soil Mech. Found. Div. 1969, 95, 99–137.
  92. Stewart, J.P.; On-Lei Kwok, A.; Hashash, Y.M.A.; Matasovic, N.; Pyke, R.; Wang, Z.; Yang, Z. Benchmarking of Nonlinear Geotechnical Ground Response Analysis Procedures; PEER Report 2008/04; Pacific Earthquake Engineering Research Center, College of Engineering, University of California: Berkeley, CA, USA, 2008; 205p.
  93. Robinson, D.; Dhu, T.; Schneider, J. SUA: A computer program to compute regolith site-response and estimate uncertainty for probabilistic seismic hazard analyses. Comput. Geosci. 2006, 32, 109–123.
  94. Hashash, Y.A.; Groholski, D.R.; Phillips, C. Recent advances in nonlinear site response analysis. In Proceedings of the Fifth International Conference in Recent Advances in Geotechnical Earthquake Engineering and Soil Dynamics and Symposium in Honor of Professor I. M. Idriss, San Diego, CA, USA, 24–29 May 2010; Paper No. OSP 4. pp. 1–22.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 288
Revisions: 3 times (View History)
Update Date: 22 Mar 2023
1000/1000