09.11.2013 Views

Proceedings - Johannes Gutenberg-Universität Mainz

Proceedings - Johannes Gutenberg-Universität Mainz

Proceedings - Johannes Gutenberg-Universität Mainz

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Proceedings</strong> Volume 2, 2011<br />

Earthquake Geology and Archaeology:<br />

Science, Society and Critical Facilities<br />

Editors<br />

C. Grützner, R. Pérez-López, T. Fernández Steeger, I. Papanikolaou<br />

K. Reicherter, P.G. Silva, and A. Vött<br />

PROCEEDINGS<br />

2nd INQUA-IGCP 567 International Workshop on Active<br />

Tectonics, Earthquake Geology, Archaeology and Engineering<br />

19-24 September 2011<br />

Corinth (Greece)<br />

ISBN: 978-960-466-093-3


<strong>Proceedings</strong><br />

Earthquake Geology and Archaeology:<br />

Science, Society and Critical facilities<br />

2 nd INQUA-IGCP 567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering<br />

Editors<br />

C. Grützner, R. Pérez-López, T. Fernández-Steeger, I. Papanikolaou<br />

K. Reicherter, P.G. Silva and A. Vött<br />

This Volume of <strong>Proceedings</strong> has been produced for the 2 nd INQUA-IGCP 567 International Workshop<br />

on Active Tectonics, Earthquake Geology, Archaeology and Engineering held in Corinth (Greece),<br />

19-24 September 2011. The event has been organized jointly by the INQUA-TERPRO Focus Area on<br />

Paleoseismology and Active Tectonics and the IGCP-567: Earthquake Archaeology.<br />

This scientific meeting has been supported by the INQUA-TERPRO #0418 Project (2008-2011), the IGCP 567 Project, the Earthquake Planning<br />

and Protection Organization of Greece (EPPO – ) and the Periphery of the Peloponnese


Printed by<br />

The Natural Hazards Laboratory,<br />

National and Kapodistrian University of Athens<br />

<br />

Edited by<br />

INQUA-TERPRO Focus Area on<br />

Paleoseismology and Active Tectonics<br />

& IGCP-567 Earthquake Archaeology<br />

INQUA-IGCP 567 <strong>Proceedings</strong>, Vol.2<br />

© 2011, the authors<br />

I.S.B.N. 978-960-466-093-3<br />

PRINTED IN GREECE<br />

Vol. 2: Earthquake Geology and Archaeology: Science, Society and Critical facilities (C. Grützner, R. Pérez-López, T. Fernández-Steeger,, I. Papanikolaou, K.<br />

Reicherter, P.G. Silva & A. Vött, Eds.). ISBN 978-960-466-093-3. Printed in Greece, 2011. 2 nd INQUA-IGCP 567 International Workshop, Corinth (Greece).<br />

Vol. 1: Archaeoseismology and Palaeoseismology in the Alpine-Himalayan collisional Zone (R. Pérez-López, C. Grützner, J. Lario, K. Reicherter & P.G. Silva, Eds.).<br />

ISBN 978-84-7484-217-3. Printed in Spain, 2009. 1 st INQUA-IGCP 567 International Workshop. Baelo Claudia, Cádiz (Spain).


Preface<br />

After the very successful 1 st Workshop on Earthquake Archaeology and Paleoseismology held in the ancient roman site of Baelo<br />

Claudia (Spain, 2009), the INQUA Focus Group on Paleoseismology and Active Tectonics decided to elaborate a bi-annual<br />

calendar to support this joint initiative with the IGCP-567 “Earthquake Archaeology”. This second joint meeting moved to the<br />

eastern Mediterranean, a tectonically active setting within the Africa-Eurasia collision zone and located in the origins of the<br />

pioneer’s works on archaeoseismology. However, for the coming year 2012, at least a part of us will move also to the New World,<br />

where the 3 rd INQUA-IGCP 567 international workshop will take place in Morelia, Mexico in November 2012. It is planned to<br />

proceed with the meeting, so we are thinking of Aachen, Germany, to be the host in 2013, possibly together with Louvain, Belgium.<br />

The aim of this joint meeting is to stimulate the already emerging comparative discussion among Earthquake Environmental Effects<br />

(EEE) and Earthquake Archaeoseismological Effects (EAE) in order to elaborate comprehensive classifications for future<br />

cataloguing and parametrization of ancient earthquakes and palaeoearthquakes. One of the final goals our collaborative workshops<br />

is the integration of archaeoseismological data in Macroseismic Scales such as the Environmental Seismic Intensity Scale ESI-<br />

2007 developed within the frame of the International Union for Quaternary Research (INQUA). In this second workshop we offer<br />

again a multidisciplinary and cross-disciplinary approach and program, since there is an urgent necessity to share the knowledge<br />

and objectives among geologists, seismologists, geodesists, archaeologists and civil engineers in order to improve seismic hazard<br />

assessments and analyses in a near future. Also, we intend to sharpen geoscientists and their research more in the direction of<br />

critical facilities, which are of world-wide public and political interest after the dramatic catastroph in Fukushima, Japan.<br />

The last two years provided significant dreadful earthquake scenarios, which were in most of the cases oversized in relation to the<br />

data provided by the historical and instrumental seismicity. The Haiti Mw 7.0 (Haiti, Jan 2010), Malua Mw 8.8 (Chile, May 2010),<br />

Christchurch Mw 6.3 (New Zealand, Feb 2011), Tohoku Mw 9.0 (Japan, Mar 2011) and Lorca Mw 5.1 (Spain, May 2011) events<br />

illustrates that both extreme subduction earthquakes or moderate events can generate severe damage in relation to relevant<br />

secondary coseismic effects or Earthquake Environmental Effects (EEE). Most of these recent events have clearly demonstrated<br />

that the vibratory ground shaking is not the unique, or even most significant, source of direct damage, and it is by no means the<br />

only parameter that should be considered in seismic hazard assessments. The lessons offered by the aforementioned events<br />

corroborate once again the relevance of liquefaction, tsunamis, rockfalls, landslides, ground subsidence, uplift or failure as a major<br />

source of hazard. But this also underpins the need of re-evaluating the significance of macroseismic intensity as an empirical<br />

measurement of earthquake size. In fact, as highlighted in the last volume produced by the INQUA Focus Area (Serva et al., 2011),<br />

intensity is a parameter able to describe a complete earthquake scenario, based on direct field observation and suitable to be<br />

preserved in the geological, geomorphological and archaeological records.<br />

With this aim the INQUA TERPRO #0418 Project (2008-2011) has implemented a world-wide online EEE Catalogue based on<br />

Google Earth in order to promote the use of the ESI-2007 Scale for seismic hazard purposes<br />

www.eeecatalog.sinanet.apat.it/terremoti/index.php. On the other hand the IGCP-567 is promoting an interesting shared approach<br />

of EEE data and EAE data for the same purpose. Examples of this variety of original research coming from this collaborative<br />

approach are the Geological Society of London Special Volume 316 (2009) Paleoseismology: Historical and Prehistorical<br />

records of Earthquake Ground Effects for Seismic Hazard Assessment (K. Reicherter, A.M. Michetti & P.G. Silva, Eds.), the<br />

Geological Society of America Special Papers 471, Ancient Earthquakes (2010) (M. Sintubin, I.S. Stewart, T. Niemi & E. Altunel,<br />

Eds.) and the Special Volume of Quaternary International (2011) Earthquake Archaeology and Paleoseismology (P.G. Silva, M.<br />

Sintubin & K. Reicherter, Eds.). In the same way, this abstract volume contains more than 80 contributions from researchers of<br />

more than 27 different countries and illustrates the upgrading shared knowledge on palaeo-, ancient, historical and instrumental<br />

earthquakes and images an impressive growth of our community. Our workshop was co-ordinated through the newly established<br />

website www.paleoseismicity.org, where earthquake info and blogs are openly shared.<br />

Finally, we wish all participants a fruitful conference and workshop in the vicinity of the ancient sites of the Classical Greece around<br />

the Corinth Gulf, where earthquake science, wonderful landscapes, ancient cultures, amazing sunny days, fantastic “Greek<br />

cooking”, nice beaches, daily cool beers, wine tasting events and late night gin tonics mixed with hot discussions are waiting for all<br />

of us. A special “efharisto poli” goes to Christoph Grützner and Raul Pérez-López for their invaluable work with the organisation<br />

and the abstract handling.<br />

The Organizers of the 2 nd INQUA-IGCP 567 Workshop<br />

Dr. Ioannis Papanikolaou Prof. Klaus Reicherter Prof. Andreas Vött Prof. Pablo G. Silva<br />

Agricultural University of Athens (GRE) RWTH Aachen University (GER) University of <strong>Mainz</strong> (GER) University of Salamanca (ESP)<br />

Aon-Benfield Hazard Research Centre,<br />

University College London, (UCL) (UK)


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

FIRST PALEOSEISMIC EVIDENCES IN ECUADOR: THE PALLATANGA FAULT RECORD<br />

Baize, Stéphane (1, Laurence Audin (2), Thierry Winter (3), Alexandra Alvarado (4), Luis Pilatasig (5), Mercedes Taipe (4), Paul<br />

Kauffmann (1), Pedro Reyes (4)<br />

(1) Institut de Radioprotection et de Sûreté Nucléaire, BP 17, 92262 Fontenay-aux-Roses, France. stephane.baize@irsn.fr<br />

(2) Institut de Recherche pour le Développement, ISTERRE, Grenoble, France.<br />

(3) Bureau de Recherche Géologiques et Minières, Service Risques Naturels, Orléans, France.<br />

(4) Escuela Politécnica Nacional, Instituto de Geofísica, Quito, Ecuador.<br />

(5) Ministerio de Recursos Naturales No Renovables, Quito, Ecuador.<br />

Abstract (First paleoseismic evidences in Ecuador: the Pallatanga fault record): The Pallatanga fault (PF) is a prominent<br />

strike-slip fault of Central Ecuador. This structure is suspected to have hosted large earthquakes, including the 1797 Riobamba<br />

event (M~7.5). The scope of the study is to evaluate the paleoseismic history of the fault, together with enhancing the<br />

seismotectonic model of this part of the Andes and improving the seismic hazard assessment. From 3 trenches, we could infer that<br />

the PF experienced several strong events (M7.2 to 7.7) in the last 8500 years. According to a new mapping campaign, we also<br />

could evidence that the fault propagates north to the Riobamba outskirts, suggesting that faulting occurred nearby this big city.<br />

Key words: Paleoearthquakes, Holocene, Pallatanga fault, Ecuador<br />

INTRODUCTION<br />

The Pallatanga fault (PF) is a NNE-SSW segment of<br />

the Dolorès-Guayaquil Megashear. This megastructure<br />

is the large deformation zone<br />

accommodating the dextral displacement between<br />

the Northern Andean Block and the South America<br />

Plate with a rate of 6 to 8 mm/a. The PF is a 50 kmlong<br />

fault, for which a previous morphotectonic study<br />

validated its Holocene right-lateral motion (Winter et<br />

al., 1993). This fault is probably the source of one of<br />

the largest crustal earthquake in South-America,<br />

occurred in Riobamba in 1797 (M~7.5: Beauval et al.,<br />

2010) (figure 1). The “Riobamba Antigua” city was<br />

destroyed (25,000 casualties) and then replaced at<br />

its current place. It may also have generated a<br />

Mw6.1 event in 1911.<br />

The scope of the study is to (1) assess the<br />

occurrence of potential large paleoearthquakes and<br />

quantify their number, magnitudes and recurrence<br />

times, (2) improve the seismotectonic model of the<br />

area, and (3) enhance the seismic hazard<br />

assessment of this inhabited region (Riobamba city,<br />

200,000 inhabitants).<br />

PREVIOUS MORPHOTECTONIC STUDY OF PF<br />

A detailed topographic survey in the Pangor Rio<br />

valley (Winter et al., 1993) evidenced the fault plane<br />

characteristics (N40°E, 75°W) and allowed estimating<br />

the “near-field” displacement to be ~40m for the<br />

dextral component and ~8m for the reverse one.<br />

Winter et al. (1993) deduced a cumulated striae<br />

dipping slightly to the south (10°). They also<br />

proposed an average slip rate of 2.9 to 4.6 mm/yr,<br />

according to regional geomorphic correlation.<br />

RESULTS OF THE TRENCH SURVEY<br />

We focused on paleoseismological analyses at<br />

Rumipamba and on a detailed mapping of the active<br />

fault trace continuation towards Riobamba city.<br />

Trenches<br />

Figure 1: Location of the study area, with the main faults<br />

and historical seismicity. Figure extract from Beauval et al.<br />

(2010). Blue disks: historical events, with magnitude<br />

proportional to radius (ex. 1797: M=7.6; 1911: M=6.2; 1958:<br />

M=5). Specific faults: Pallatanga fault in blue and bounded<br />

by red arrows, Guamote-Huigra fault in green, Llanganates<br />

fault in cyan, Pucara fault in magenta. Pink lines: provinces<br />

boundaries. Names of provinces in pink; names of towns in<br />

black.<br />

At Rumipamba, the fault trace runs parallel to the<br />

mountain front and cuts perpendicularly the erosional<br />

features going downhill (figure 2). The mountain<br />

slope sedimentation is exclusively dominated by ash<br />

falls, with rare colluvium and alluvium. Unfortunately,<br />

3


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

we could not find any channel deposits in test<br />

trenches and it was thus impossible to find linear<br />

features that could account for lateral offsetting. In<br />

further developments, we thus used the<br />

“morphological slip vector” of Winter et al. (1993).<br />

In the three excavated trenches, ash falls are<br />

strongly pedogenetized in black organic soils<br />

(andisols), over thicknesses of 2 to 3.5 meters.<br />

last event in 1797 should then have a higher<br />

magnitude than mentioned in table 1. Moreover, the<br />

event #2, which is also recorded in trench #1 and<br />

well constrained by 14 C datings, is associated with 2<br />

successive and discordant colluvial wedges. If we<br />

assume this is not partly due to non-tectonic<br />

processes (climate degradation for ex.), then arises<br />

the issue of possible under-representation of<br />

earthquake number in trench #2 (and correlative<br />

overestimation of magnitude).<br />

Trenches<br />

#1<br />

#2 & 3<br />

Fault trace<br />

EQ<br />

Date<br />

approx.<br />

(cal y.)<br />

Vertical<br />

throw<br />

(cm)<br />

Total<br />

throw<br />

(cm)<br />

Mw<br />

SRL<br />

(km)<br />

5 1797 25 145 7.2 60<br />

4 ~1,000 60 345 7.5 100<br />

3 ~0 90 520 7.7 140<br />

2 ~ -1,000 90 520 7.7 140<br />

1 ~ -4,500 70 400 7.6 120<br />

Figure 2: South-eastward view of the mountain front, cut by<br />

the Rumipamba segment of the PF. Trenches were<br />

excavated in the the same area.<br />

Available 14 C datings of these soils (23 samples)<br />

range from 6650 cal BC to 1650 cal AD, scattered<br />

into 7 classes of ages (-6650, -5500, -3500, -2800, -<br />

1800, 800, 1650). They lie above the volcanic<br />

basement (Mesozoic-Tertiary).<br />

The trenches’ survey pointed out several interbeds<br />

with variable oxidized material contents -including<br />

basement fragments and root remnants- in the<br />

vicinity of the fault strands. Our basic assumption is<br />

that these “clastic layers” are stratigraphic markers<br />

of earthquakes (colluvial wedges) because they are<br />

associated with the erosion of a basement scarp<br />

appeared during surface faulting.<br />

Among the 3 trenches, the trench #2 is especially<br />

interesting in order to reconstruct the paleoseismic<br />

history of the fault. There, the stratigraphic series is<br />

more complete and stratigraphic units can be<br />

correlated on each side of the fault (figure 3). In<br />

trenches #1 and #3, the fault strands are obvious but<br />

correlations are difficult. In trench #2, the fault splits<br />

in 3 strands propagating from the basement fault<br />

gouge up to the Holocene deposits and the modern<br />

soil. The fault zone downthrows the eastern wall by<br />

about 3 m.<br />

By performing a retrodeformation of the trench, we<br />

could infer 5 strong events with individual vertical<br />

throws from 0.25 m up to 0.90 m (table 1). Assuming<br />

a slip vector dipping of 10° (Winter et al., 1993), the<br />

total throws are between 1.45 to 5.20 m, which<br />

provides magnitudes from 7.2 to 7.7 according to<br />

Wells & Coppersmith (1994) relations. To generate<br />

such earthquakes, the same empirical relations<br />

suggest that the fault rupture can be as long as 60 to<br />

140 km (which is larger than the mapped PF).<br />

Some uncertainties remain after trench survey. For<br />

example, the retrodeformation of trench #2 can lead<br />

to only 4 paleoearthquakes because arguments for<br />

separating events #4 and #5 (Table 1) are weak. The<br />

Table 1: synthetic table of the earthquake history along the<br />

Rumipamba segment of the Pallatanga fault. Vertical throw<br />

is the observed data; total throw is calculated from vertical<br />

throw and morphological slip vector; Mw is calculated from<br />

total throw with W&C (1994) relation (average<br />

displacement); SRL is estimated from total throw<br />

CONTRIBUTION OF MAPPING<br />

From its southern tip (Western Cordillera front) to the<br />

trench site (figure 1), the PF is mapped and reaches<br />

a length of about 60 km. Given the previous<br />

conclusion of large fault ruptures (SRL>60 km), it has<br />

been decided to investigate the “unknown” northern<br />

continuation of the Rumipamba segment where<br />

morphological features suggested its existence and<br />

where the rupture may have propagated during large<br />

past events. The interest is that section is also that it<br />

can help in improving the seismotectonic model of<br />

the region and in clarifying the seismic hazard<br />

assessment of this inhabited area.<br />

Before field mapping, fault traces were tracked down<br />

with analysis of remote-sensing images (SPOT5).<br />

Field control and mapping allow validating most of<br />

this preliminary work thanks to morphologic and<br />

geologic observations. The Rumipamba segment<br />

splits into 2 fault zones when entering the Inter-<br />

Andean valley (figure 4). The western branch<br />

(Rumipamba-Huacona and La Merced segments)<br />

seems to rotate into a N-S direction, with pure lateral<br />

kinematics, whereas the eastern ones changes to a<br />

NE-SW strike, with a more important vertical<br />

component (Cajabamba, Gatazo and Lican-<br />

Riobamba segments). This last segment cuts the<br />

Riobamba basin and seems to offset the surface<br />

deposits of the Upper Pleistocene (40 ka, Bernard et<br />

al., 2008) and later morphological features. The<br />

scarp is almost 100 m high and, locally, we could find<br />

a dextral offset of 50 m of a valley slope. To the<br />

north-east, some morphological features (mountain<br />

front spurs, stream profiles) and outcrops (normal<br />

faulting and tilting of recent sediments) suggest that<br />

active deformation propagates from the Riobamba<br />

basin north-eastwards along the Cordillera Reale<br />

front (Penipe area), along the Chambo segment.<br />

4


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Figure 3: Mosaic picture and sketch interpretation of trench #2. Ages are given in cal BC/AD.<br />

Figure 4: Sketch map of the northern extension of the PF within the Riobamba basin, and example of the morphological imprint of<br />

the fault. DS: segment with prevailing dextral component; N: with normal component (symbols on the downthrown block side); R:<br />

with reverse component; V: segment with undifferenciated vertical offset; VS: segment with vertical and dextral strike-slip<br />

component (symbols on the upthrown block side, for the 3 last).<br />

5


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

These preliminary conclusions need to be completed<br />

by further mapping, especially in the Chambo valley<br />

and along the western rim of the Inter-Andean valley<br />

(north of San Juan).<br />

All along the prospected area, several places have<br />

been quoted to be potential sites for further<br />

paleoseismic investigations.<br />

DISCUSSION<br />

For the first time in Ecuador, paleoseismic<br />

investigations prove the occurrence of 4 or 5 strong<br />

events which shook the Riobamba region in the past.<br />

We probably found the trace in sediments of the<br />

Riobamba event (1797) and then confirm that it has<br />

been generated by the Pallatanga fault. Magnitudes,<br />

inferred from geological signal and applying empirical<br />

relations, range between 7.2 and 7.7, which is in<br />

agreement with the independent estimation of the<br />

Riobamba earthquake by Beauval et al. (2010).<br />

Reccurrence times of these events seem variable<br />

(Table 1: 800 to 3,500 years) with a mean value<br />

around 1,000-1,500 years.<br />

Nonetheless, the trench survey and its interpretation<br />

are soiled by uncertainties and arises the issue of<br />

under or over-representation of the event number,<br />

which is a crucial point in seismic hazard assessment<br />

(Mc Calpin, 2009). In addition, the 3 trenches<br />

revealed variable tectono-sedimentary signals. These<br />

nuances weaken the above conclusions dealing with<br />

magnitudes of paleoearthquakes and hazard<br />

assessment for the PF, and we clearly need<br />

additional investigations to strengthen the results. In<br />

this first-step work, we assumed for instance during<br />

magnitude assessment that the observed coseismic<br />

offsets are “average” values of the event slip and this<br />

has clearly to be validated. The mapping clearly<br />

shows that the trenches were not performed at the<br />

northern tip of the fault and new trenching sites must<br />

be found out to increase the slip vs rupture length<br />

dataset.<br />

Dealing with the fault slip rate of the fault, our results<br />

based on radiometric datings outstandingly validate<br />

the rate published by Winter et al. (1993). Roughly,<br />

trenches show vertical displacement of about 3<br />

meters of recent soils -i.e. 20 meters following the<br />

hypothesis of a 10° slip vector- which have been<br />

produced during the last 5,000 years. This leads to a<br />

mean slip rate of ~4 mm/a. With respect to the total<br />

displacement rate (~6-8 mm/a) of the North-Andean<br />

Block, it suggests that other segments accommodate<br />

a significant part of the relative motion of North<br />

Andean Block wrt South America. Up to now, no<br />

other faults are potential candidates.<br />

What we know from our mapping campaign is that<br />

deformation probably propagates both to the north<br />

(along the western edge of the Inter-Andean valley)<br />

and to the north-east (towards the eastern rim of the<br />

Cordillera Reale). This last segment could be the<br />

structural link between the Pallatanga fault (to the<br />

south) and the Pucara/Llanganates faults (to the<br />

north), one of these being probably the source of the<br />

Pelileo destructive earthquake (1949) (Beauval et al.,<br />

2010). This propagation of deformation to the north<br />

also drastically increases the seismic hazard for<br />

Riobamba city, because rupture segments probably<br />

run all along the outskirts of this vulnerable big city.<br />

Acknowledgements: The authors give special thanks to<br />

the Indian communities of Rumipamba and neighboring<br />

villages, for their warm welcome and “technical” contribution<br />

to logistics and trench cleaning. Thanks also to the various<br />

authorities of Ecuador who make this research possible<br />

(prefecture and the civil safety services). We also benefit of<br />

the help and fruitful discussions with people from the<br />

Escuela Politecnica Nacional de Quito.<br />

References<br />

Beauval, C., H. Yepes H., W. H. Bakun, J. Egred, A.<br />

Alvarado & J.-C. Singaucho, (2010). Locations and<br />

magnitudes of historical earthquakes in the Sierra of<br />

Ecuador (1587–1996). Geophys. J. Int., doi:<br />

10.1111/j.1365-246X.2010.04569.x.<br />

Bernard, B., B. van Wyk de Vries, D. Barba, H. Leyrit, C.<br />

Robin, S. Alcaraz & P. Samaniego, (2008). The<br />

Chimborazo sector collapse and debris avalanche:<br />

Deposit characteristics as evidence of emplacement<br />

mechanisms. Journal of Volcanology and Geothermal<br />

Research 176, 36–43.<br />

Mc Calpin, J. P., (2009). Paleoseismology.<br />

INTERNATIONAL GEOPHYSICS SERIES, volume 95,<br />

Academic Press, 802 pages.<br />

Wells, D. L. & K. J. Coppersmith, (1994). New empirical<br />

relationships among magnitude, rupture length, rupture<br />

width, rupture area and surface displacement. Bull.<br />

Seism. Soc. America 84, 974-1002.<br />

Winter, T., J.P. Avouac & A. Lavenu, (1993). Late<br />

Quaternary kinematics of the Pallatanga strike-slip fault<br />

(Central Ecuador) from topographic measurements of<br />

displaced morphological features. Geophys. J. Int. 115,<br />

905-920.<br />

6


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SOME NOTES ON EARTHQUAKE AND FAULT RELATIONSHIPS FOR DIP-SLIP EVENTS<br />

Salvatore Barba (1) and Debora Finocchio (1, 2)<br />

(1) Istituto Nazionale di Geofisica e Vulcanologia, Via di Vigna Murata 605, Roma, Italy. salvatore.barba@ingv.it<br />

(4) Università di Urbino, Italy<br />

Abstract (Some notes on earthquake and fault relationships for dip-slip events): We developed a conceptual model for dipslip<br />

earthquakes to predict the coseismic surface throw and the dislocation as a function of depth along-dip (1-D averaged). By<br />

performing a Monte Carlo experiment in Matlab to test the conceptual model against earthquake data for years 1990-2006, we<br />

found that, for reverse faulting, the surface throw is better reproduced with the slip distributions which show a maximum slip closer<br />

to the surface. Our model predicts a surface throw for reverse faulting earthquakes that is larger than that contained in the Wells<br />

and Coppersmith (1994) relationships. The use of downward or upward directivity in the rupture changes dramatically the<br />

assessment and the perception of the earthquake hazard, as shown by the catastrophic 2011 Japan earthquake.<br />

Key words: Normal fault, thrust fault, coseismic rupture, Monte Carlo model.<br />

INTRODUCTION<br />

It is commonly accepted in the literature that<br />

deviatoric stress increases with depth (see, e.g., Das<br />

and Scholz, 1983; Henry and Das, 2001). The<br />

coseismic dislocation is predicted to initiate at the<br />

base of the seismogenic layer and to propagate<br />

upward. However, two classes of observations,<br />

coseismic throw and depth directivity, are hardly<br />

explained with such hypotheses in the case of dipslip<br />

earthquakes. Reverse-faulting earthquakes<br />

occurred in the past 20 years exhibited a much larger<br />

throw at the surface, with respect to normal-faulting<br />

earthquakes of similar magnitudes. Also, in certain<br />

cases of reverse-faulting earthquakes, seismological<br />

observations show a downward directivity of the<br />

rupture (Carminati et al., 2004).<br />

the USGS catalogue, and the depth distributions from<br />

regional earthquake catalogue. Based on these<br />

distributions, we generated a very large synthetic<br />

earthquake catalogue. For each of these synthetic<br />

earthquakes we set the average slip, the length and<br />

the width of the rupture based on a-priori<br />

relationships. Then, we predicted the surface throw<br />

based on two classes of slip distributions at depth:<br />

one class based on four a-priori distributions (our<br />

METHOD AND DATA<br />

In this work we developed a conceptual model for<br />

dip-slip earthquakes to predict the coseismic surface<br />

throw and the dislocation as a function of depth<br />

along-dip (1-D averaged). To this purpose, we<br />

determined relationships among the surface throw<br />

and the earthquake parameters (slip, dip, depth, and<br />

moment magnitude Mw) for two emblematic normal<br />

and reverse faults. Finally, assuming that the (1-D<br />

averaged) slip-distribution at depth may be related to<br />

the deviatoric stress, we discussed the difference<br />

between our model and Das e Scholz (1983) model.<br />

This work is thus developed in two parts: the<br />

generation of synthetic relationships between<br />

coseismic throw and magnitude, and the comparison<br />

of model predictions with a 20-years long earthquake<br />

dataset (Table 1). We made a Monte Carlo<br />

experiment in Matlab to test the conceptual model<br />

against data.<br />

The synthetic relationships are based on standard<br />

seismic catalogues. We determined the frequency<br />

distributions for dip, Mw and seismic moment from<br />

7<br />

Fig. 1: Cumulative percentage of earthquakes that<br />

ruptured the surface for reverse (TF) and normal faulting<br />

(NF). Lines represent the model prediction, the symbols<br />

represent the data.<br />

choice), and one class based on literature data,<br />

which have been grouped in three slip distributions at<br />

depth. These four (or three) distributions differ as it<br />

follows: (1) the maximum slip is in the center of the<br />

rupture, (2) it is closer to the bottom or (3) to the top<br />

edge of the rupture; (4) the slip is uniform. For each<br />

magnitude bin we computed the average and the


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

standard deviation of the predicted surface throw.<br />

Last, we compared the predicted maximum and<br />

average surface throw with the observed throw for<br />

earthquakes occurred in years 1990-2008. As a<br />

cross check, we computed the percentage of<br />

synthetic earthquakes that produce surface throw, as<br />

a function of magnitude (Fig. 1).<br />

RESULTS<br />

We found differences between reverse and normal<br />

faulting earthquakes. For reverse faulting, the surface<br />

throw is better reproduced with the slip distributions<br />

which show a maximum slip closer to the surface, or<br />

with the uniform slip (Fig. 2). For normal faulting, the<br />

best fit occurs when the maximum slip is closer to the<br />

bottom edge of the rupture.<br />

reverse faulting earthquakes that is larger than that<br />

contained in the Wells and Coppersmith (1994)<br />

relationships. The opposite occurs for normal faulting<br />

earthquakes. In the dataset of Wells and<br />

Coppersmith (1994), the fault dimensions are derived<br />

sometimes from seismological data, sometimes from<br />

geodetic data. As we used only earthquakes<br />

occurred in recent times, our dataset is based on<br />

seismological data and is therefore more<br />

homogeneous. Although we have fewer data, our<br />

result appears robust. In the hypothesis that the slip<br />

distribution can be used as a proxy for deviatoric<br />

stress, the maximum of the deviatoric stress is<br />

expected to be close to the upper edge of the fault<br />

rupture at least for earthquakes with low dip angle<br />

(=5.9).<br />

N Code Date Location Mw Depth (Km) Type Avg Throw (m)<br />

1 EV-022 1990/11/06 Iran 6.6 11.1 TF 2.6<br />

2 EV-073 1992/08/19 Turkey 7.2 17 TF 3<br />

3 EV-101 1993/05/17 California 6.1 7 NF 0.02<br />

4 EV-117 1993/09/29 India 6.2 14.1 TF 1.93<br />

5 EV-187 1995/05/13 Greece 6.5 14 NF 0.05<br />

6 EV-192 1995/06/15 Greece 6.5 14 NF 0.02<br />

7 EV-206 1995/10/01 Turkey 6.4 9 NF 0.2<br />

8 EV-282 1997/09/26 Italy 6 6 NF 0.03<br />

9 EV-330 1999/09/07 Greece 6 10 NF 0.08<br />

8


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

N Code Date Location Mw Depth (Km) Type Avg Throw (m)<br />

10 EV-331 1999/09/20 Taiwan 7.6 9 TF 5<br />

11 EV-405 2002/02/03 Turkey 6.5 5 NF 0.15<br />

12 EV-414 2002/06/22 Iran 6.5 11 TF 1<br />

13 EV-431 2002/11/03 Alaska 7.2 4.2 TF 4<br />

14 EV-518 2005/02/22 Iran 6.4 7 TF 0.8<br />

15 EV-532 2005/10/08 Pakistan 7.6 19.1 TF 3.7<br />

9


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

HOLOCENE COASTAL NOTCHES IN THE MEDITERRANEAN: PALAEOSEISMIC OR<br />

PALAEOCLIMATIC INDICATORS?<br />

Boulton, S. J. (1) and Stewart, I. S. (1)<br />

(1) School of Geography, Earth and Environmental Sciences, University of Plymouth, Plymouth, Devon, PL4 8AA, U.K.<br />

Sarah.boulton@plymouth.ac.uk<br />

Abstract (Holocene Coastal Notches in the Mediterranean: Palaeoseismic or Palaeoclimatic indicators?): Bioerosion and<br />

bioconstruction along rocky coastlines can lead to the development of coastal notches that are preserved when uplifted or<br />

submerged above or below the swash zone and thus can be used to quantify relative vertical coastal motions in tectonically active<br />

areas. There are two main models for the genesis of notch profiles. The first is tectonically driven, with notches forming during<br />

relative still-stands of sea-level; notches are subsequently raised (or lowered) relative to sea-level due to local seismic events. The<br />

second model favours a climatic origin for notch formation, where stable periods of Holocene climate allow enhanced erosion.<br />

Here, we explore these two models using a database of Eastern Mediterranean notches. We conclude that the spatial and<br />

temporal distribution of the notches favours a dominantly tectonic control on formation.<br />

Key words: notches, tectonics, climate, Holocene.<br />

INTRODUCTION<br />

Along rocky coastlines bioerosion and<br />

bioconstruction can lead to the development of<br />

coastal notches that can be preserved when uplifted<br />

or submerged above the swash zone (Stewart and<br />

Morhange, 2009). As these geomorphic features<br />

form at sea level, palaeoshorelines could be used to<br />

quantify relative coastal uplift and subsidence in<br />

tectonically active areas when the sea level history is<br />

known (e.g., Pirazzoli & Thommeret, 1977; Pirazzoli<br />

et al., 1981, 1982, 1989, 1991, 1994a, 1994b; Stiros<br />

et al., 1992; Stewart et al., 1997; Stiros et al., 2000;<br />

Antonioli et al., 2006; Shaw et al., 2008).<br />

Despite this large body of work there is some<br />

disagreement on the process of notch formation.<br />

The most commonly assumed theory is that notches<br />

form when the rate of sea level change matches the<br />

uplift (or subsidence) rate along a coastline resulting<br />

in a relative sea level still stand. Notches are then<br />

raised higher than sea level when a seismic event<br />

results in rapid coastal uplift preserving the feature<br />

from further erosion. However, in many other areas it<br />

is not clear whether notches reflect short-term<br />

seismic events or the long-term uplift rate averaged<br />

over many seismic cycles (Stewart & Vita Finzi,<br />

1996). Rarely, can individual notches be matched to<br />

documented earthquakes (Pirazzoli, 1994a).<br />

Recently, Cooper et al. (2007) have linked notch<br />

formation to Holocene periods of climatic stability.<br />

Specifically, when the rate of sea level rise is lower<br />

than the tectonic uplift rate but when climatic<br />

conditions favour high productivity, increased levels<br />

of bioerosion develop a notch. This enhanced<br />

erosion ceases during periods of rapid climate<br />

change (RCC), allowing notches to become<br />

emergent due to continuing coastal uplift. Cooper et<br />

al. (2007) discount any correlation between notches<br />

and individual seismic events, stating that individual<br />

earthquakes are too small to raise a notch clear<br />

above sea level and too numerous to explain the<br />

formation of the four notch intervals that they<br />

observed on the Perachora Peninsula (Greece). This<br />

is a dramatic reinterpretation of coastal notches<br />

which, if correct, means these phenomena cannot be<br />

used for palaeoseismic indicators and associated<br />

seismic hazard assessments. So what drives notch<br />

development: tectonic instability (earthquakes) or<br />

climatic stability (enhanced bioerosion)?<br />

If tectonic uplift is the dominant control in the<br />

Mediterranean, then notches will develop only in<br />

coasts where slow regional emergence is augmented<br />

by abrupt uplift events (earthquakes) and the age of<br />

notches will relate broadly to local palaeoseismic<br />

episodes. If climate is the dominant control on notch<br />

formation, then notches can also form on subsiding<br />

coasts and will date only to periods of climatic<br />

stability (no RCC) that are consistent across the<br />

region. Mayewski et al., (2004) propose five periods<br />

of stable climate dating to 8000-6000, 5400-4200,<br />

3800-3300, 2500-1200 and 1000-600 years BP. That<br />

implies that no more than five notches ought to be<br />

formed in any one place, each corresponding to an<br />

intervening periods of no RCC during the last 10,000<br />

yrs (Mayewski et al., 2004). Yet in some parts of the<br />

Mediterranean coast more than five notch levels are<br />

recorded; along the coast of Crete, nine to ten well<br />

preserved superimposed shorelines can be observed<br />

(Pirazzoli et al., 1982) and eight are present on<br />

Rhodes (Pirazzoli et al., 1989).<br />

10


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

a.<br />

Height (m)<br />

Frequency<br />

b.<br />

Sample age (recalibrated years B.P)<br />

Figure 1a) Histogram showing frequency of samples against age from notches around the Eastern Mediterranean in 100 year<br />

groups; b) Graph of radiocarbon age against height of all notch data for the Eastern Mediterranean region. Grey bars on both<br />

graphs indicate the periods of proposed Holocene rapid climate change.<br />

Such field observations seem inconsistent with the<br />

climate model. However, to appraise any relationship<br />

with climate in a more robust manner requires a<br />

systematic analysis of the Holocene palaeo-shoreline<br />

data for the Mediterranean region as a whole. To do<br />

that, we have compiled a dataset of palaeo-shoreline<br />

data that span several decades of published<br />

research in the region. The database itself comprises<br />

428 dated samples derived from 40 separate studies<br />

that were undertaken at locations from across the<br />

whole of the Eastern Mediterranean. These data are<br />

used to examine age and height relationships and<br />

test correlations with periods of RCC.<br />

The results of the database are summarised in figure<br />

1. Generally, with the exception of a few outliers, the<br />

initial phase of notch development in the<br />

Mediterranean occurs in the period 5000 – 6500<br />

years BP (Fig. 1). This period correlates with the mid-<br />

Holocene slowdown of global sea-level rise. The<br />

eustatic slowdown at this time ensures that along<br />

many tectonically uplifting coasts, rates of<br />

emergence and sea level rise are in unison and so<br />

notch development takes place. It is noteworthy that<br />

under the climate model this period is one of rapid<br />

climate change and so a time window in which notch<br />

formation ought to be muted.<br />

Indeed, overall, figure 1 shows that there appears to<br />

be no correlation between stable climatic periods and<br />

notch occurrence, with numerous notches dating to<br />

periods of RCC (grey bars). This strongly suggests<br />

that climate is not the controlling factor for notch<br />

formation. Instead, it appears that the complex age<br />

and height pattern of notches around the region more<br />

likely reflects local tectonic histories of slow<br />

interseismic crustal deformation overprinted by<br />

abrupt seismic movements of the coastline.<br />

Discriminating ambiguous palaeoseismic information<br />

from this complex shoreline record, however,<br />

remains a difficult task.<br />

Acknowledgements: We would like to thank Richard<br />

Martin for his help in constructing the notch database.<br />

11


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

References<br />

Antonioli, F., Ferranti, L., Lambeck, K., Kershaw, S.,<br />

Verrubbi, V., Dai Pra, G., (2006). Late Pleistocene to<br />

Holocene record of changing uplift rates in southern<br />

Calabria and northeastern Sicily (southern Italy, Central<br />

Mediterranean Sea). Tectonophysics, 422, 23-40.<br />

Cooper, F.J., Roberts, G.P., Underwood, C.J., (2007). A<br />

comparison of 10 3 -10 5 climate stability and the formation<br />

of coastal notches. Geophysics research letters, 34,<br />

L14310, doi:10.1029/2007GL030673<br />

Mayewski, P.A., Rohling, E., Stager, C., Karlén, K., Maasch,<br />

K., Meeker, L.D., Meyerson, E., Gasse, F., van Kreveld,<br />

S., Holmgren, K., Lee-Thorp, J., Rosqvist, G., Rack, F.,<br />

Staubwasser, M., and Schneider, R., (2004), Holocene<br />

climate variability, Quaternary Research 62, 243-255.<br />

Pirazzoli, P.A., Thommeret, J., (1977). Datation<br />

radiométrique d’une ligne de rivage à +2.5 m près de<br />

Aghia Roumeli, Crète, Grèce. Le Comte Rendu<br />

d’Academie Science, Paris, Series D, 97, 1255-1257.<br />

Pirazzoli, P.A., Thommeret, J., Thommeret, Y., Laborel, J.,<br />

Montaggioni, L.F. (1981). Les rivages emerges<br />

d’Antikythira (Cerigotto): correlations avec le Crete<br />

occidentale et implications cinematiques et geodynamics.<br />

Niveax Marins et Tectonique Quaternaires dans l’Aire<br />

Mediterraneene, Paris, UNRS and Universitie Paris I, pp.<br />

49-65.<br />

Pirazzoli, P.A., Thommeret, J., Thommeret, Y., Laborel, J.,<br />

Montaggioni, L.F. (1982). Crustal block movements from<br />

Holocene shorelines: Crete and Antikythira (Greece).<br />

Tectonophysics, 86, 27-43.<br />

Pirazzoli, P.A. 1986, Marine notches In: van de Plassche,<br />

O. (Ed), Sea-level Research: A Manual for the Collection<br />

and Evaluation of Data, Geo Books, Norwich pp. 361–<br />

400.<br />

Pirazzoli, P.A., Montaggioni, L.F., Saliege, J.F., Segonzac,<br />

G., Thommeret, Y., Vergnaud-Grazzini, C., (1989).<br />

Crustal block movements from Holocene shorelines:<br />

Rhodes Island (Greece). Tectonophysics, 170, 89-114.<br />

Pirazzoli, P.A., Laborel, J., Saliege, J.F., Erol, O., Kayan, I.,<br />

Person, A. (1991). Holocene raised shorelines on the<br />

Hatay coasts (Turkey): Palaeoecological and tectonic<br />

implications. Marine Geology, 96, 295-311.<br />

Pirazzoli, P.A., Stiros, S.C., Laborel, J., Laborel-Deguen, F.,<br />

Arnold, M., Papageorgiou, S., Morhange, C. (1994a).<br />

Late-Holocene shoreline changes related to<br />

palaeoseismic events in the Ionian Islands, Greece. The<br />

Holocene, 4, 397-405.<br />

Pirazzoli, P.A., Stiros, S.C., Arnold, M., Laborel, J., Laborel-<br />

Deguen, F., Papageorgiou, S., (1994b). Episodic uplift<br />

deduced from Holocene shorelines in the Perachora<br />

Peninsula, Corinth area, Greece. Tectonophysics, 229,<br />

201-209.<br />

Shaw, B., Ambraseys, N.N., England, P.C., Floyd, M.A.,<br />

Gorman, G.J., Higman, T.F.G., Jackson, J.A., Nocquet,<br />

J,-M., Pain, C.C., Piggott, M.D., (2008). Eastern<br />

Mediterranean tectonics and tsunami hazard inferred from<br />

the AD 365 earthquake. Nature Geoscience, 1, 268-276.<br />

Stewart, I.S., Vita-Finzi, C., (1996). Coastal uplift on active<br />

normal faults: The Eliki Fault, Greece. Geophysical<br />

Research Letters, 23, 1853-1856.<br />

Stewart, I.S., Morhange, C. (2009). Coastal geomorphology<br />

and sea-level change. In J.C. Woodward (Ed.). The physical<br />

geography of the Mediterranean (pp. 383–413). Oxford: Oxford<br />

University Press.<br />

Stewart, I.S., Cundy, A., Kershaw, S., Firth, C., (1997). Holocene<br />

coastal uplift in the Taormina area, northeastern Sicily:<br />

Implications for the southern prolongation of the Calabrian<br />

seismogenic belt. Journal of Geodynamics, 24, 37-50.<br />

Stiros, S.C., Arnold, M., Pirazzoli, P.A., Laborel, J., Laborel,,<br />

Papageorgiou, S., (1992). Historical coseismic uplift on Euboea<br />

Island, Greece. Earth and Planetary Science Letters, 108, 109-<br />

117.<br />

Stiros, S.C., Laborel, J., Laborel-Deguen, F., Papageorgiou, S.,<br />

Evin, J., Pirazzoli, P.A., (2000). Seismic coastal uplift in a<br />

region of subsidence: Holocene raised shorelines of Samos<br />

Island, Aegean Sea, Greece. Marine Geology, 170, 41-58.<br />

12


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ANALYSING THE LANDSLIDE SUSCEPTIBILITY WITH STATISTICAL METHODS IN<br />

MAILY-SAY, KYRGYZSTAN<br />

Braun, Anika (1), Tomas M. Fernandez-Steeger (1), Hans-Balder Havenith (2), Almaz Torgoev(2), Romy Schlögel (3)<br />

(1) Department of Engineering Geology and Hydrogeology, RWTH Aachen University, Lochnerstraße 4-20, 52064 Aachen,<br />

GERMANY. Email: braun@lih.rwth-aachen.de<br />

(2) Georisks and Environment, Department of Geology, University of Liège, B20 Sart Tilmann, 4000 Liège, BELGIUM.<br />

(3) School and Observatory of Earth Sciences, Institut de Physique du Globe de Strasbourg, 5 rue Descartes, F-67084<br />

Strasbourg Cedex<br />

Abstract (Analysing the landslide susceptibility with statistical methods in Maily-Say, Kyrgyzstan): A landslide<br />

susceptibility analysis was carried out for Maily-Say, a former uranium mining town in Kyrgyzstan. Numerous landslides threaten<br />

inhabitants, infrastructure and uranium tailings. Besides typical factors responsible in this region, like seismicity, geology,<br />

geomorphology and climatic conditions, land use plays an important role in Maily-Say. In order to predict landslide susceptibility<br />

and to clarify the interplay of different factors two statistical methods were implemented, a bivariate statistical method and a data<br />

mining approach based on a multi-temporal digital landslide inventory. Generally, with both methods areas could be mapped that<br />

show a high potential for future landslides. Furthermore, the correlation of landslides with the landform and loess deposits provide<br />

some information on the seismic effects on slope stability.<br />

Key words: landslide susceptibility, bivariate statistical analysis, data mining<br />

INTRODUCTION<br />

In this study a landslide susceptibility analysis in<br />

Maily-Say, Kyrgyzstan, was undertaken. In the<br />

vicinity of the former uranium mining and milling town<br />

more than 200 landslides were present in 2007.<br />

Besides frequent damage to houses and<br />

infrastructure, landslides already caused several<br />

fatalities. Landslides damming the main river during<br />

spring runoff lead to flooding (Havenith et al., 2006b).<br />

Numerous radioactive, partially instable uranium<br />

tailings and waste dumps are threatened by<br />

landslides. Since the main river leads to the<br />

Ferghana Valley, a densely populated and<br />

agricultural region covered mainly by Uzbekistan, the<br />

destabilisation of nuclear waste tailings in Maily-Say<br />

bears the potential of a major environmental<br />

catastrophe (Blacksmith Institute, 2006).<br />

SETTING<br />

Maily-Say is located in the western foothills of the<br />

seismically active Tien Shan high mountain belt, on<br />

the northern rim of the Ferghana Basin. The Tien<br />

Shan is an old orogenic belt from Variscan times<br />

which was reactivated during the collision of India<br />

and Eurasia 55 Ma ago and started to rise 10 Ma ago<br />

(Molnar & Tapponier, 1975; Bullen et al., 2001), see<br />

figure 1. The peaks exceed heights of 7000 m. The<br />

geology in Maily-Say is related to the transitional<br />

position between high mountains and a basin<br />

dominated by partially soft Jurassic, Cretaceous and<br />

Paleogene sedimentary rocks, see figure 4. A<br />

relatively weathering resistant formation is the<br />

Cretaceous limestone that was also mined for<br />

The term landslide susceptibility implies the spatial<br />

probability of occurrence of slope failures (Aleotti &<br />

Chowdury, 1999). By analysing geological and<br />

geomorphological situations that have lead to slope<br />

failures in the past, it is possible to predict the<br />

landslide susceptibility to a certain degree (Varnes,<br />

1984). In this study two different statistical methods<br />

were used for predicting the landslide susceptibility in<br />

Maily-Say, a bivariate statistical approach and data<br />

mining. Besides pointing out endangered areas, the<br />

study aimed at analysing the main factors causing<br />

slope failures in Maily-Say and their temporal<br />

development with the help of a multi-temporal<br />

landslide inventory.<br />

The idea behind this approach is to extract a<br />

maximum of information from a simple dataset<br />

(geology, digital elevation model, landslide inventory)<br />

to provide a first localisation of endangered areas<br />

without having been to the field.<br />

Fig. 1: Schematic tectonic map of Southern Central Asia,<br />

from Bossu et al. (1996) with the outline of figure 2 (a).<br />

13


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: Digital elevation model of Kyrgyzstan (SRTM data from Reuter et al. (2007)) with mapped landslide sites and Ms 5<br />

earthquakes, based on data of the Kyrgyz and world seismic catalogue.<br />

uranium. The landscape is characterised by a quiet<br />

rough relief with heights reaching from 700 m to<br />

4000 m. The climate is predominantly dry-continental<br />

with snowfall in winter and high run off in spring.<br />

Due to the tectonic, geological, geomorphological<br />

and hydrometerological conditions this region in<br />

general is highly prone to landslides (Roessner et al.,<br />

2005), see also figure 2). In Maily-Say also the<br />

uranium mining activities from 1946 to 1968 are<br />

supposed to have an important impact on the<br />

destabilisation of slopes. Actually, before 1946 there<br />

were only few landslides present (Havenith et al.,<br />

2006b). Then, landslide activity increased due to<br />

mining activities (Torgoev et al., 2002). Until 1962 the<br />

number of landslides reached 157. The direct link<br />

between mining activities and landslides may be<br />

explained by rock weakening because of the<br />

extraction works, collapse of underground mining<br />

galleries and rising groundwater levels in the<br />

abandoned galleries (Havenith et al., 2006b). But<br />

also indirect processes like changed land use due to<br />

population growth and increasing traffic may play a<br />

role. After the mining activities stopped in 1968, the<br />

increase of landslide activity was going on and even<br />

accelerating in the 90s. Large landslides, like the<br />

Koytash, the Tektonik and the Isolith landslides,<br />

involving up to 5 Million m 3 of rocks and soils formed<br />

during this period (Minetti, 2004). It is not clear,<br />

whether this are still long-term effects of the mining<br />

activities or if other factors, like seismic events or<br />

climatic trends contribute to this development. For<br />

instance a massive collapse of the Tektonik landslide<br />

occurred 7 weeks after a Ms 6,2 earthquake in 1992,<br />

at 20 km in the south of Maily-Say (Havenith et al.,<br />

2006b). Unfortunately documentation e.g. on<br />

earthquake damages are scarce, so it is not possible<br />

to draw clear connections between earthquake<br />

events and single landslides.<br />

14<br />

METHODS, RESULTS AND DISCUSSION<br />

A multi-temporal landslide and scarp inventory of the<br />

years 1962, 1984, 1996, 2002 and 2007 as well as<br />

geological and geomorphological data for the Maily-<br />

Say Valley were available and prepared in ArcGIS<br />

(ESRI). The landslide and scarp inventories were<br />

developed on the basis of existing digital inventories,<br />

aerial photographs from the 50s and 60s, satellite<br />

images from 2002 and 2007 and field observations.<br />

Bivariate statistical analysis<br />

For the bivariate statistical analysis landslide and<br />

scarp maps were compared to different factor maps<br />

on a pixel base using a GIS. According to the method<br />

of van Westen (1997), a weight index (W i ) was<br />

calculated expressing the probability of landslide<br />

occurrence for each raster cell according to the<br />

distinct parameter. Parameters investigated were e.g.<br />

slope angle, altitude, curvature, distance to rivers as<br />

erosion base, geology, distance to faults and loess<br />

deposits. Loess deposits in the Maily-Say valley<br />

mainly remained on plateaus which are due to the<br />

Fig. 3: Histogram showing the W i for scarps vs. loess<br />

boundary.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

lacking slope angle not very vulnerable to landslides.<br />

Often there are steeper slopes at the verge of these<br />

plateaus and the loess deposits. This is where loess<br />

landslides are likely to occur. Therefore, a 30 m<br />

buffer around the boundaries of loess was also<br />

analysed. The weight indices were compared for<br />

scarps and landslide bodies separately and between<br />

the different years. A landslide susceptibility map was<br />

created following the InfoVal method (Saha et al.,<br />

2005) based on the 1962 landslide inventory.<br />

In the following paragraph some interesting results<br />

are summarized. A seismic triggering is suggested by<br />

the correlation of the landslide scarps with convex<br />

landforms and with the boundaries of loess deposits.<br />

Basically, concave landforms are supposed to be<br />

more susceptible to landslides because of runoff<br />

convergence and lower depth to water table. On the<br />

other hand, in convex landforms the seismic waves<br />

are amplified when reflected at the surface. Havenith<br />

et al. (2006a) observed this phenomenon in the<br />

Suusamyr region, also within the Tien Shan. Hence,<br />

the correlation of landslide scarps with concave<br />

landforms indicates a seismic triggering. The<br />

correlation of landslide scarps with the boundaries of<br />

loess (see figure 3 and 4) is increasing over time and<br />

stagnating after 1996. Earthquakes can directly<br />

trigger loess landslides, e.g. by loess liquefaction as<br />

observed in China and Tajikistan (Wang et al., 2004)<br />

or indirectly by forming fractures that allow rapid<br />

infiltration and resulting collapse during the next<br />

rainfall event (Havenith & Bourdeau, 2010). The fact<br />

that this observed trend is rising abruptly between<br />

1984 and 1996 and then stagnating may indicate a<br />

connection to the 1992 Ms 6,2 earthquake event near<br />

Maily-Say. Not only a direct triggering but also a<br />

weakening effect of the aforementioned processes<br />

and a temporal delayed failure of slopes is possible.<br />

In addition to that, evidence for a climatic change<br />

was figured out. A shift of landslide correlation with<br />

slope aspect from S in 1962 and 1984 to NW in 1996<br />

indicates the contribution of a wetter climate to<br />

increasing slope failure.<br />

The method caused problems e.g. in implementing<br />

the geology (figure 4). Here landslides correlate<br />

strongly with the Sarybia formation, a Jurassic<br />

sandstone formation. This is because only a very<br />

small extend of this formation is included into the<br />

analysis extend. This small extend includes a very<br />

large landslide, which leads to an overestimation.<br />

The landslide susceptibility map created with this<br />

method based on the 1962 landslide inventory was<br />

compared to landslide inventories of the years after<br />

1962. The comparison showed agreement between<br />

regions mapped as highly susceptible and landslides<br />

that formed after 1962.<br />

Data Mining<br />

With data mining methods it is possible to detect<br />

landslides in a dataset using classification algorithms.<br />

This method aims at simulating the reasoning<br />

process, e.g., the one of a geologist as in this case<br />

(Fernandez-Steeger, 2002). An artificial neural<br />

network (ANN) and a Bayesian network were<br />

developed for analysing the multi-temporal landslide<br />

inventory. Since this method can handle large input<br />

Fig. 4: Geological map of the working area showing distribution of loess deposits and outlines of landslides in 2007. Geology after<br />

de Marneffe (2010).<br />

15


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

datasets, an abundance of geological, morphological<br />

and hydrological factors were used as input data.<br />

The dataset was prepared in a ArcGIS on a pixel<br />

base, while the main modelling was executed with<br />

the PASW Modeler 14 (SPSS Inc.), a user interface<br />

for common data mining algorithms as well as data<br />

pre- and post-processing The results of the<br />

classification were transferred back to the GIS for<br />

evaluation. While the ANN was adjusted quite<br />

optimal to the landslides and provided good<br />

classifications, the Bayesian networks developed the<br />

ability to identify also zones for possible future<br />

landslides. At this point of the work the contribution of<br />

each factor to the result is still not very transparent.<br />

Further development is needed here. Whereas ANNs<br />

have already successfully been implemented for<br />

landslide susceptibility analysis (Fernandez-Steeger,<br />

2002; Lee et al., 2003) there are no studies based on<br />

Bayesian networks. A comparison between the<br />

results of the InfoVal method and a Bayesian network<br />

is shown in figure 5<br />

Fig. 5: Comparison of the results of the statistical and the<br />

data mining approach, a) InfoVal method, b) Bayesian<br />

networks. Both based on landslide data from 1962.<br />

CONCLUSIONS<br />

Two statistical methods were applied to analyse<br />

landslide susceptibility in Maily-Say, Kyrgyzstan.<br />

Both methods provide the possibility of a remote<br />

landslide analysis based on relatively simple<br />

datasets. This is an advantage for remote places like<br />

Kyrgyzstan, where landslide sites are not easily<br />

accessible or experts are simply lacking. Especially<br />

the bivariate statistical method is a simple method<br />

that provides a useful first idea of landslide<br />

susceptibility. The data mining approach is also<br />

promising for the task of predicting landslide<br />

susceptibility, but the whole process of data<br />

preparation, modelling and validation requires more<br />

time and mathematical skills.<br />

References<br />

Aleotti, P., Chowdhury, R. (1999). Landslide hazard<br />

assessment: summary review and new perspectives.<br />

Bulletin of Engineering Geology and the Environment, 58,<br />

21–44.<br />

Blacksmith Institute (2006). The world’s worst polluted<br />

places-The Top Ten.<br />

Bossu, R., Grasso, J. R., Plotnikova, L. M., Nurtaev, B.,<br />

Fréchet, J., Moisy, M. (1996). Complexity of<br />

Intracontinental Seismic Faultings: The Gazli,<br />

Uzbekistan, Sequence. Bulletin of the Seismological<br />

Society of America, 86, 959–971.<br />

Bullen, M.E., Burbank, D.W., Garver, J.I., Abdrakhmatov,<br />

K.Y. (2001). Late Cenozoic tectonic evolution of the<br />

northwestern Tien Shan: New age estimates for the<br />

initiation of mountain building. Geological Society of<br />

America Bulletin, 113, 1544–1559.<br />

De Marneffe, C. (2010): Cartographie et modelisation 3D de<br />

la geologie de la vallee de Mailuu-Suu, Tien Shan,<br />

Master's Thesis, University of Liège, unpublished, 93p.<br />

Fernandez-Steeger, T. M. (2002). Erkennung von<br />

Hangrutschungssystemen mit Neuronalen Netzen als<br />

Grundlage für Georisikoanalysen. Ph.D. Thesis,<br />

Karlsruhe University.<br />

Havenith, H. B., Bourdeau, C. (2010). Earthquake-induced<br />

Landslide Hazards in Mountain Regions: A Review of<br />

Case Histories from Central Asia. Geologica Belgica, 13,<br />

135–150.<br />

Havenith, H. B., Strom, A., Caceres, F., Pirard, E. (2006a).<br />

Analysis of landslide susceptibility in the Suusamyr<br />

region, Tien Shan: statistical and geotechnical approach.<br />

Landslides, 3, 39–50.<br />

Havenith, H. B., Torgoev, I., Meleshko, A., Alioshin, Y.,<br />

Torgoev, A., Danneels, G. (2006b). Landslides in the<br />

Mailuu-Suu Valley, Kyrgyzstan-Hazards and Impacts.<br />

Landslides, 3, 137–147.<br />

Lee, S., Ryu, J.-H., Lee, M.-J., Won, J.-S. (2003). Use of an<br />

articial neural network for analysis of the susceptibility to<br />

landslides at Boun, Korea. Environmental Geology, 44,<br />

820–833.<br />

Minetti, L. (2004). Kyrgyz Republic proposed naturall<br />

disaster mitigation project, 1st IDA Preparation Mission.<br />

Ministry of Environment and Emergency (MEE), Bishkek,<br />

40p, inedit.<br />

Molnar, P., Tapponier, P. (1975). Cenozoic Tectonics of<br />

Asia: Effects of a continental collision. Science, 189,<br />

419–426.<br />

Reuter, H. I., Nelson, A., Jarvis, A. (2007). An evaluation of<br />

void filling interpolation methods for SRTM data.<br />

International Journal of Geographic Information Science.<br />

21:9, 983-1008.<br />

Roessner, S., Wetzel, H. U., Kaufmann, H., Sarnagoev, A.<br />

(2005). Potential of Satellite Remote Sensing and GIS for<br />

Landslide Hazard Assessment in Southern Kyrgyzstan<br />

(Central Asia). Natural Hazards, 35, 395–416.<br />

Saha, A. K., Gupta, R. P., Sarkar, I., Arora, M. K.,<br />

Csaplovics, E. (2005). An approach for GIS-based<br />

statistical landslide susceptibility zonation-with a case<br />

study in the Himalayas. Landslides, 2, 61–69.<br />

Torgoev, I. A., Alioshin, Y. G., Havenith, H. B. (2002).<br />

Impact of uranium mining and processing on the<br />

environment of mountainous areas of Kyrgyzstan. In:<br />

Merkel, Planer- Friedrich and Wolkersdorfer (eds),<br />

Uranium in the aquatic environment, pp. 93–98, Springer,<br />

Berlin Heidelberg New York.<br />

Wang, L., Wang, Y., Wang, J., Li, L., Yuan, Z. (2004). The<br />

Liquefaction Potential of Loess in China and its<br />

Prevention. 13th World Conference On Earthquake<br />

Engineering, Vancouver, B.C., Canada, 13p.<br />

Vandenhove, H. Q. H., Clerc, J.-J., Lejeune, J.-M., Sweeck,<br />

L., Sillen, X., Mallants, D., Zeevaert, T. (2003). Final<br />

report in frame of EC-TACIS Project N SCRE1/N 38<br />

remediation of uranium mining and milling tailing in<br />

Mailuu-Suu district Kyrgyzstan, Mol (Belgium).<br />

Van Westen, C. J. (1997). Statistical landslide hazard<br />

analysis. In: Application guide, ILWIS 2.1 for Windows.<br />

ITC, Enschede, The Netherlands, 73–84.<br />

Varnes, D. J. (1984). Landslide hazard zonation: a review of<br />

principles and practise. UNESCO, Paris, 63 p.<br />

16


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

TECTONIC INTERPRETATION OF THE 2008 WENCHUAN EARTHQUAKE: WHY IT ONLY<br />

PROPAGATED IN ONE DIRECTION - THE FUTURE?<br />

Burchfiel, B. C. (1), Royden, L. H. (1)<br />

(1) Department of Earth, Atmospheric and Planetary Sciences, 54-1010, Massachusetts Institute of Technology, Cambridge, MA<br />

02139, USA. EMAIL: bcbruch@mit.edu<br />

Abstract (Tectonic interpretation of the 2008 Wenchuan earthquake): The 2008 Wenchuan earthquake, (M = 7.9) occurred on<br />

a listric thrust fault that at the hypocenter (15-20 km) dipped ~30-40° NW and steepened upward to near vertical at the surface<br />

(Zhang et al., 2008) Maximum slip was 9 m vertical and 6 m right-slip. The fault propagated from the epicenter to the northeast<br />

for ~ 200 km with increasing right-slip eastward and broke across at least two fault segment boundaries. Although the earthquake<br />

occurs along the steep topographic slope of the Longmen Shan at eastern margin of the Tibet plateau, it was unexpected because<br />

data suggests that the recurrence interval on the fault zone was 2500 to 4000 years.<br />

Key words: Earthquake, Wenchuan, China, 2008<br />

INTRODUCTION<br />

The earthquake fault, Pengguan-Beichuan fault,<br />

broke in two segments, the main segment was along<br />

the east side of a major uplift of Precambrian<br />

basement rocks, the Pengguan massif, in the<br />

Longmen Shan at the eastern margin of the Tibet<br />

plateau, southwestern Sichaun, China, and<br />

propagated northeast into Paleozoic rocks. The<br />

Pengguan uplift is a northeast plunging anticline with<br />

its sedimentary cover still exposed at the north end.<br />

The east side of the anticline is the Pengguan-<br />

Beichuan fault. Because of the plunge to the north<br />

the east vergent thrust sheets are exposed also to<br />

the north and are cut by the Pengguan-Beichuan<br />

fault (figure 1). These geological relations were<br />

discussed by Burchfiel et al., (1995).<br />

The question of why the fault only propagated to the<br />

northeast and its long recurrence interval can be<br />

explained by the geological framework in the area<br />

that was known in 1995. The west side of Pengguan<br />

uplift and the Pengguan-Beichuan fault are cut off at<br />

the south end by the Wenchuan-Maowen fault, a<br />

steep fault that has active right-slip, but was not<br />

reactivated during the earthquake, that also cuts off<br />

the sedimentary cover and Mesozoic thrust sheets so<br />

that the southern Penguan Precambrian has no<br />

sedimentary cover and it is not clear how much it has<br />

been uplifted. How these two faults interact is critical,<br />

but is at present unknown. The Wenchuan-Beichuan<br />

fault has a low-grade mylonitic fabric that shows the<br />

west side down (a normal fault), that is consistent<br />

with the geology where the rocks west of the<br />

Penguan massif are metamorphosed Paleozoic that<br />

belong to the Mesozoic thrust sheets.<br />

DISCUSSION<br />

Why the earthquake did not propagate also to the<br />

southwest and what are the prospects of a large<br />

earthquake to the southwest can be explained by the<br />

geological framework of the area. To the south, the<br />

Precambrian is uplifted again in the Baoxing massif<br />

that also has the characteristics of an east vergent<br />

fold with a steep thrust along its eastern side (Fig. 1).<br />

The Wenchuan-Maowen fault appears to continue<br />

south and cuts off the west side of the Baoxing uplift<br />

and Paleozoic and even Triassic strata lie west of the<br />

fault against Precambrian rocks to the east.<br />

Regionally east of the Chengdu plain (the Quaternary<br />

area east of Pengguan uplift) is an single curving<br />

fold, the Longchaun anticline) that intersects the<br />

Longmen Shan obliquely in the north but trends away<br />

from the Longmen Shan to the south and curves<br />

south east of Chengdu and continue farther south<br />

east of Emei shan (Figure 1). Within the southern<br />

Chengdu plane several north-plunging anticlines<br />

appear and grow in amplitude southward so that the<br />

elevation to the south rises and rapidly becomes the<br />

eastern part of the Tibetan Plateau. These folds<br />

involve Precambrian rocks in the south. Thus what<br />

appears to be a horizontal decollement with the<br />

sedimentary section beneath the Chengdu plane in<br />

the north that connects the Longchaun anticline with<br />

the Pengguan-Beichuan fault, in the south the<br />

decollement must drop into the Precambrian rocks<br />

where the topography becomes part of the Tibetan<br />

Plateau.<br />

The geometry of active deformation suggests<br />

stresses focused mainly at the Pengguan uplift in the<br />

north, become distributed southward from the 2008<br />

epicenter across a broad of the area suggesting that<br />

the stress is being relieved in the southern area<br />

where is more distributed and being relieved by a<br />

broader and more active area of seismic activity.<br />

However, along the topographic eastern margin of<br />

plateau, mountains increase in elevation so that at<br />

the Siguniang Shan (4 girls) they reach more than<br />

6200 m and at Gonga Shan farther south they reach<br />

7550 m. The mountain front east of the 4 girls is very<br />

17


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

steep, suggesting again rapid uplift like along the<br />

Pengguan massif. This is similar to the conditions at<br />

the area of the Wenchuan earthquake, and needs to<br />

be tectonically and geomorphologically researched<br />

for recurrence interval in this area that may also be<br />

longer that recorded history. looked at in terms of the<br />

even though the stress appears to be more<br />

distributed to the east in this region. Like all geology<br />

relations there are ambiquities: the steepness of the<br />

topographic margin of the plateau in this area is a<br />

clue as to how structurally active the plateau margin<br />

is but the broader distribution of active structures and<br />

earthquakes suggests stresses are being distributed.<br />

Farther east there is second steep mountain front on<br />

the east side of Emei Shan with an active fault at its<br />

base. Another alarming relationship.<br />

The change in structure from north to south east of<br />

the Longmen Shan can be interpreted to be related<br />

to the geometery of the Xianshuihe fault, an arcuate<br />

NW-tending fault major active right-slip that<br />

separates two major crustal provinces in SW China:<br />

a province to the south that is part of a crustal<br />

fragment that rotates clockwise at 10-12 mm/yr<br />

around the Eastern Himalayan syntaxis, from the<br />

Longmen Shan province that moves NE ~parallel to<br />

the eastern margin the Tibetan Plateau with only a<br />

very slow convergence rate (1-2 mm/yr) with respect<br />

to South China to the east as shown by GPS data. In<br />

fact, the convergence rate is so slow that the eastern<br />

part of the Tibetan Plateau moves east nearly at the<br />

same rate as South China as first shown by King et<br />

al. (1997). The slow convergence rate across the<br />

topographic margin of Tibet Plateau explains the<br />

large recurrence interval of the Wenchuan<br />

earthquake, but does not easily explain the steep<br />

topographic margin and the rise in elevation of the<br />

front to the south.<br />

West of Kanding Xianshuihe fault curves more<br />

sharply that either to the NW or SE forming a<br />

restraining bend indicates there is a component of<br />

compression across the fault that forms the high<br />

topographic area east of the fault and where<br />

basement rocks are involved in the structure. At the<br />

bend the mountains reach 7550m where the large<br />

Cenozoic pluton that underlies the Gonga Shan. If so<br />

the it is also where to the east of the fault the Tibetan<br />

plateau was elevated and the folds that come from<br />

the southern part of the Sichuan basin involve<br />

Precambrian rocks. There are cross structures here;<br />

WNW trending thrust faults that involve Precambrian<br />

cut across the N-S trending folds from the southern<br />

Sichuan basin that also involve Precambrian. The<br />

decollement here is in the basement beneath the<br />

folds and probably beneath the huge NW-trending<br />

Danba antiform of Precambrian east of Kanding.<br />

This raises the question of how deep is the<br />

decollement and how does it interact with the<br />

Xianshuihe fault; the decollement is deep and the<br />

Xianshuihe fault may be a crustal feature.<br />

Fig. 1: Generalized map of the Longmen Shan and<br />

Southern Sichuan basin area showing structures of late<br />

Cenozoic to active age. Faults are in red decorated with<br />

barbs for thrust faults, arrows for strike-slip faults and<br />

double ticks for normal faults. Blue are the folds of late<br />

Cenozoic age, except in the southeast where they are of<br />

Middle Cretaceous age (dark blue). These structures<br />

overprint the early Mesozoic Longmen Shan thrust faults<br />

that moved the Songpan Ganze Unit (pale red) eastward<br />

above the Yangtze Unit (white).The area of the<br />

southwestern Sichuan basin shown in yellow, the Chengdu<br />

plain, has a thin cover of Pleistocene sediments derived<br />

from the Longmen Shan and ponded behind the active<br />

Longquan anticline (LQA). Large black dot at south end of<br />

Pengguan massif (PM) is location of the 2008 Wenchuan<br />

earthquake epicenter. 4G=Four Girls peaks, AF=Anninghe<br />

fault, BF=Beichuan fault, BSM=Baoxing massif, DA=Danba<br />

antiform, EA=Emei Shan anticline, GS=Gonga Shan,<br />

HF=Huya fault, PM=Pengguan massif, SF=Shimian fault,<br />

WMF=Wenchuan-Maowen fault, XF=Xianshuihe fault,<br />

XSP=Xushan platform, XPA=Xiong Po anticline.<br />

Acknowledgements: This work was supported by NSF<br />

Grant 000357.<br />

References<br />

Burchfiel, B. C., Z. Chen., Y. Liu, & L. H. Royden, (1995).<br />

Tectonics of the Longmen Shan and adjacent regions.<br />

International Geological Review, 37(8), 661-736.<br />

King, R. W., F. Shen, B. C. Burchfiel, Z. Chen, Y. Liu, L. H.<br />

Royden, E. Wang, X. Zhang, & J. Zhao, (1997).<br />

Geodetic measurement of crustal motion in southwest<br />

China. Geology, 25 (2), 179-182.<br />

Zhang, P-Z., X-Z. Wen, Z-K. Shen, & J-H. Chen, (2008).<br />

Oblique, High-Angle, Listric-Reverse Faulting and<br />

Associated Development of Strain: The Wenchuan<br />

Earthquake of May 12, 2008, Sichuan. Annual reviews<br />

of Earth and Planetary Sciences, 38, 353-382.<br />

18


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

CHARACTERIZATION OF LATE PLEISTOCENE-HOLOCENE EARTHQUAKE-INDUCED<br />

“HOMOGENITES” IN THE SEA OF MARMARA THROUGH MAGNETIC FABRIC.<br />

IMPLICATION FOR CO-SEISMIC OFFSETS DETECTION AND MEASUREMENTS<br />

Campos Corina (1)(2), Beck Christian (1), Crouzet Christian (1), Carrillo Eduardo (3)<br />

(1) Institut des Sciences de la Terre ISTerre, UMR CNRS 5275, Grenoble University, 73376 Le Bourget du Lac Cedex,<br />

France. Email : Corina.Guzman@etu.univ-savoie.fr<br />

(2) Departamento de Ciencias de la Tierra, Universidad Simón Bolívar, Sartenejas, Baruta, Venezuela.<br />

Email: campossc@usb.ve.<br />

(3) Instituto de Ciencias de la Tierra, Universidad Central de Venezuela, Av. Los Ilustres, Caracas, Venezuela.<br />

Email: eduardo.carrillo@ciens.ucv.ve.<br />

Abstract: The Marmara Sea is located in the eastern part of the Mediterranean region, in an area with strong seismic instability<br />

associated with the activity of North Anatolian Fault. In order to analyze the impact of seismicity on the sedimentation in the<br />

Marmara Sea we have studied three giant piston cores (27 to 37.7m long). They represent the last 20 kyr of sedimentation. The<br />

upper section (marine stage) is predominantly composed by fine grained terrigenous material (clay-silty) and in less percentage by<br />

the silty-sandy laminated intervals. The lower section (non marine) is composed by abundant fine grained terrigenous material,<br />

numerous turbidites sequences and levels with deformation structures (possible seismites). Some turbidites show an abrupt<br />

contact separating the coarse grain basal part (bed load) of the fine grain upper part (suspended load), this later are defined as<br />

“homogenites”<br />

Key words: Sea of Marmara, Earthquakes, Turbidites, Homogeneites.<br />

Introduction<br />

The Marmara Sea (Northwestern Turkey) is a pullapart<br />

basin developed along the North Anatolian<br />

Fault (Hancock and Barka, 1981). East-West<br />

elongated (200 km) it is composed of several aligned<br />

sub-basins (Tekirda, Orta, Kumburgaz and Çinarcic<br />

basin) (Fig.1). The North Anatolian Fault (N.A.F) is a<br />

1200-km-long dextral strike-slip fault (engör et al.,<br />

2004) and is considered as a major active boundary<br />

between Anatolia and Eurasia plates. (Armijo el al.,<br />

1999; McClusky et al., 2000; Flerit et al., 2003).<br />

The northern branch of the N.A.F. crosses the<br />

different deep sub-basins, where giant pistons cores<br />

(from 27 to 37,7 m) have been retrieved, with location<br />

based on high resolution seismic reflection imaging.<br />

This survey aimed to detect and characterize coseismic<br />

sedimentary episodes - especially<br />

“homogenites” – for a 20 000 yr-long period. Different<br />

previous studies have demonstrated the interest of<br />

sedimentary record (isolated marine basins and<br />

lakes) as archives of seismic activity (Hempton and<br />

Dewey, 1983; Calvo et al., 1998; Chapron et al.,<br />

1999; among many others). Movements of water<br />

masses (seiche effect, reflected turbidites) and mass<br />

wasting (triggered by major earthquakes) are inferred<br />

to combine themselves for specific depositions which<br />

locally compensate vertical offset of seafloor, as in<br />

the Sea of Marmara central basin (Beck et al, 2007;<br />

Stegmann et al., 2007; Strasser et al., 2006). Thus,<br />

the present work is dedicated to the Central (Orta)<br />

and eastern (Çinarcic) basins to detect, characterize,<br />

and correlate, these sedimentary “events”, as a<br />

contribution to pecise paleoseismic data (associated<br />

fault offset, chronology).<br />

Fig.1 Simplified geodynamic setting of the Sea of<br />

Marmara for the present day with coring locations (after<br />

Armijo et al., 1999; Beck et al., 2007). GPS velocity<br />

vector with fixed Eurasia Plate, from McClusky et al.,<br />

2000. The Anatolia Plate is bounded by major strike-slip<br />

faults systems (North and East) and subduction (South).<br />

The objective of this work is identify and characterize<br />

the impact of seismicity over the sedimentation and<br />

distinguish them from “normal” sedimentary<br />

processes as hemipelagic-type fallout and flooding,<br />

19


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

subsequently we hope to determine the earthquake<br />

recurrence and contribute to improve hazard<br />

estimates. We have studied 3 giant piston cores<br />

collected during the MARMACORE Cruise in August-<br />

Septembre 2001. The cores were taken in the<br />

Çinarcic basin (MD01-2425), and in the Orta basin<br />

(MD01-2429, MD01-2431), at depths between 1230<br />

and 1170m. The sedimentary record in theses cores<br />

represent the sedimentation of the last 20 kyr.<br />

METHODS<br />

One of the challenges is the distinction between finegrained,<br />

slow and continuous, hemipelagic<br />

sedimentation, from quite instantaneous resuspension<br />

and re-deposition. For this we combined<br />

different tools in order to characterize the textures:<br />

grain size (laser diffraction grain size analyzer,<br />

Malvern TM ), anisotropy of magnetic susceptibility<br />

(Kappabridges MFK1-FA AGICO) and X-ray imagery<br />

(D.G.O.’s SCOPIX). Compositions were controlled<br />

through magnetic susceptibility (BARTINGTON TM<br />

MS2 contact sensor), and microscopic analysis. The<br />

chronology was established from<br />

14 C (AMS<br />

measurements) derived from wood, plants and<br />

sapropelic muds. We calibrated the ages with Oxcal<br />

Program v4.1. The ages found represent the<br />

Holocene and part of the late Pleistocene.<br />

RESULTS<br />

The analyzed sections are composed of fine grained<br />

terrigenous material (clay-silty) intercalated with siltysandy<br />

laminated intervals, turbidites sequences and<br />

liquefaction features as ball and pillow. The upper<br />

marine part is predominantly compose by the claysilty<br />

slightly calcareous and in less percentage by the<br />

silty-sandy laminated intervals, theses intervals<br />

consist of milimetric’s lenticular and parallel planar<br />

beddings.<br />

In the lower non marine part the fine grained<br />

terrigenous material is abundant but we find<br />

numerous turbidites sequences with thicknesses<br />

ranging from centimeters to decimeters. This<br />

turbidites can show erosive bases, normal gradation<br />

and ripples, others can show an abrupt contact<br />

separating the coarse grain basal part (bed load) of<br />

the fine grain upper part (suspended load)(Fig.2), this<br />

later are defined as “homogenites”. In this non<br />

marine section we can also find in less proportion the<br />

slumps and the levels with deformation structures<br />

(possible seismites).<br />

Fig.2 Textural and compositional characterization of three homogenites present in the non marine section (core MD01-2425). The<br />

mean size, magnetic susceptibility and lineation magnetic don’t show differences between the hemipelagic normal sedimentation<br />

and the homogeneous coseismic sedimentation. The foliation magnetic (AMS) is higher in the homogenites than the hemipelagic<br />

deposits, this high foliation contrast has to be explained by different arrays of phyllosilicates, associate a specific settling<br />

conditions related to water mass oscillation.<br />

20


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The Magnetic Susceptibility signature in the marine<br />

part is in general lower than the non marine part<br />

(average in the marine part is10x10 -5 SI and 30x10 -5<br />

SI in the non marine part). In general this signature is<br />

strong in the silty-sandy laminated intervals and in<br />

the turbidites sequences. In the turbidites sequences<br />

this signature has a behavior similar to the one<br />

observed in the profiles of grain-size, being higher in<br />

the sand layers than the fine grain, and showing<br />

constants values in the homogenites. On the other<br />

hand, the foliation determined from the Anisotropy of<br />

Magnetic Susceptibility is much higher in the levels<br />

the fine grain size belonging to the homogenites<br />

(average of 1.12) than any other level of thin grain<br />

size (average 1.06) (Fig.2). As both types of levels<br />

(homogenites and hemipelagic deposits) have similar<br />

grain-sizes and not very different composition, this<br />

high foliation contrast has to be explained by different<br />

arrays of phyllosilicates, which, at their turn, cannot<br />

be explained by different compactions. Thus, we infer<br />

specific settling conditions related to water mass<br />

oscillation.<br />

In the study area others analyses are in processes<br />

such as the carbonates contents, clay minerals,<br />

analysis of terrigenous fraction (mineralogy), organic<br />

matter, etc.<br />

References<br />

Armijo, R., Meyer, B., Hubert, A., Barka, A. (1999).<br />

Westward propagation of the North Anatolian Fault into<br />

the northern Aegean: timing and kinematics. Geology 27<br />

(3), 267–270.<br />

Beck C., Mercier de Lepinay B., Schneider J.L., Cremer M.,<br />

Cagatay N., Wendenbaum E., Boutareaud S., Menot G.,<br />

Schmidt S., Weber O., Eris K., Armijo R., Meyer B.,<br />

Pondard N., Gutscher M.A., Turon J.L., Labeyrie L.,<br />

Cortijo E., Gallet Y., Bouquerel H., Gorur N., Gervais A.,<br />

Castera M.H., Londeix L., de Resseguier A., Jaouen A.<br />

(2007) Late Quaternary co-seismic sedimentation in the<br />

Sea of Marmara's deep basins, Sedimentary Geology,<br />

199, 65-89.<br />

Calvo, J.P., Rodriguez-Pascua, M., Martin-Velazquez, S.,<br />

Jimenez, S., De Vicente, G. (1998). Microdeformation of<br />

lacustrine laminate sequences from Late Miocene<br />

formations of SE Spain: an interpretation of loop bedding.<br />

Sedimentology 45, 279–292.<br />

Chapron, E., Beck, C., Pourchet, M., Deconinck, J.-F.<br />

(1999). 1822 AD earthquake-triggered homogenite in<br />

Lake Le Bourget (NW Alps). Terra Nova 11, 86–92.<br />

Flerit, F., Armijo, R., King, G.C.P., Meyer, B., Barka, E.<br />

(2003). Slip partitioning in the Sea of Marmara pull-apart<br />

determined from GPS velocity vectors. Geophysical<br />

Journal International 154, 1–7.<br />

Hancock, P.L., Barka, A.A. (1981). Opposed shear senses<br />

inferred from neotectonic mesofractures systems in the<br />

North Anatolian fault zone. J. Struct. Geol. 3, 383–392.<br />

Hempton, M.R., Dewey, J.F. (1983). Earthquake-induced<br />

deformational structures in young lacustrine sediments,<br />

East Anatolian Fault, southwest Turkey. Tectonophysics<br />

98, 7–14.<br />

McClusky, S., Bassalanian, S., Barka, A., Demir, C.,<br />

Ergintav, S., Georgiev, I., Gurkan, O., Hamburger, M.,<br />

Hurst, K., Hans-Gert, H.-G., Karstens, K., Kekelidze, G.,<br />

King, R., Kotzev, V., Lenk, O., Mahmoud, S., Mishin, A.,<br />

Nadariya, M., Ouzounis, A., Paradissis, D., Peter, Y.,<br />

Prilepin, M., Relinger, R., Sanli, I., Seeger, H., Tealeb, A.,<br />

Toksöz, M.N., Veis, G.(2000). Global positioning system<br />

constraints on plate kinematics and dynamics in the<br />

eastern Mediterranean and Caucasus. Journal of<br />

Geophysical Research 105, 5695–5719.<br />

Sengör, A.M.C., Tuysuz, O., Imren, C., Sakinc, M.,<br />

Eyidogan, H., Gorur, N., Le Pichon, X., Claude Rangin,<br />

C. (2004). The North Anatolian Fault. A new look. Ann.<br />

Rev. Earth Planet. Sci. 33, 1-75.<br />

Stegmann, S., Strasser, M., Anselmetti, F.S., Kopf, A.<br />

(2007). Geotechnical in situ characterisation of<br />

subaquatic slopes: The role of pore pressure transients<br />

versus frictional strength in landslide initiation.<br />

Geophysical Research Letters, 34, L07607<br />

Strasser, M, Anselmetti, F.S., Fäh, D., Giardini, D.,<br />

Schnellmann, M. (2006). Magnitudes and source areas of<br />

large prehistoric northern Alpine earthquakes revealed by<br />

slope failures in lakes. Geology, 34 (12)<br />

21


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PALEOSEISMOLOGICAL EVIDENCE FOR HISTORICAL SURFACE RUPTURE EVENTS IN<br />

S. MIGUEL ISLAND (AZORES)<br />

Carmo, Rita (1, José Madeira (2), Ana Hipólito (1), Teresa Ferreira (1)<br />

(1) Centro de Vulcanologia e Avaliação de Riscos Geológicos, Universidade dos Açores, Complexo Científico, 3º Piso – Ala Sul,<br />

Rua Mãe de Deus, 9500-321 Ponta Delgada, Açores, Portugal. Email: Rita.L.Carmo@azores.gov.pt<br />

(2) Dpto. Geologia, LATTEX/IDL, Faculdade de Ciências da Universidade de Lisboa, Edifício C6, Campo Grande, 1749-016<br />

Lisboa, Portugal. Email: jmadeira@fc.ul.pt<br />

Abstract (Paleoseismological evidence for historical surface rupture events in S. Miguel Island (Azores)):<br />

The Azores archipelago is located in the triple junction between the Eurasian, Nubian and North American<br />

lithospheric plates. The Achada das Furnas plateau, located in the central part of S. Miguel Island, between Fogo<br />

and Furnas volcanoes, is dominated by several basaltic cinder cones that define several WNW-ESE and E-W<br />

alignments. Two E-W trending scarps were identified by aerial photo analysis. Trenches were open across the<br />

scarps to confirm their tectonic nature exposing two active normal faults (the Altiprado Faults). At least four<br />

paleoearthquakes were deduced, three of which in historical times. Radiocarbon ages are in agreement with this<br />

interpretation.<br />

Key words: Azores, active faults, paleoseismology<br />

GEOLOGICAL SETTING<br />

The Azores archipelago is located at the Eurasia<br />

(Eu), Nubia (Nu) and North America (NA) triple<br />

junction (Fig. 1). The Mid Atlantic Ridge separates<br />

the NA from Eu and Nu plates, while the Azores-<br />

Gibraltar Fault Zone (Terceira Rift and Gloria Fault) is<br />

the boundary between Eu and Nu plates. The<br />

archipelago comprises nine islands distributed by<br />

three groups: the western islands lie on NA plate<br />

while the central and eastern groups are located<br />

along the western segment of the Azores-Gibraltar<br />

Fault Zone (Fig. 1). As a result of its complex tectonic<br />

setting, the Azores archipelago is subject to frequent<br />

seismic and volcanic activity.<br />

S. Miguel Island was settled in 1439-1443. It has<br />

three active explosive central volcanoes with summit<br />

calderas linked by zones of fissural volcanism. The<br />

main tectonic structures trend NW-SE to WNW-ESE,<br />

NNW-SSE and NE-SW.<br />

The Achada das Furnas plateau, located in the<br />

central area of the island, between Fogo and Furnas<br />

volcanoes, is dominated by several basaltic cinder<br />

cones defining WNW-ESE and E-W alignments (Fig.<br />

2). Aerial photo analysis identified the presence of<br />

two E-W trending scarps produced by the Altiprado<br />

Faults (AF1 and AF2, Fig. 2).<br />

Fig. 1: Azores tectonic setting. NA – North American<br />

plate; Eu – Eurasian plate; Nu – Nubian plate; MAR –<br />

Mid Atlantic Ridge; TR – Terceira Rift (s.l.); EAFZ –<br />

East Azores Fracture Zone; GF – Gloria Fault. World<br />

topography and bathymetry from ESRI (2008).<br />

Fig. 2: Main tectonic and volcanic structures of<br />

Achada das Furnas plateau.<br />

22


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE ALTIPRADO FAULTS<br />

Altiprado Fault 1<br />

The Altiprado Fault 1 (AF1) trace is marked by an<br />

835 m-long, 3 m-high, south-facing scarp, trending<br />

N87ºE (Fig. 3). To the east and to the west its trace<br />

becomes uncertain.<br />

A ~18 m-long trench exposed the fault, trending<br />

N87ºE and dipping 65ºS, displacing a stratigraphic<br />

succession that comprises 6 pumice fall deposits<br />

(units) separated by paleosols produced by eruptions<br />

of Fogo and Furnas volcanoes (Fig. 4): 1 - olivebrown<br />

paleosol containing basalt lapilli fragments,<br />

corresponding to Fogo A deposit (4520±90 years BP;<br />

Wallenstein, 1999); 2 – pumice fall deposit composed<br />

of alternating yellowish lapilli and light olive brown<br />

ash beds, probably related to Fogo C deposit; 3 -<br />

strong brown pumice ash fall deposit with some lapilli<br />

at the base (3a), and a yellowish pumice fall deposit<br />

in the central part (3b); probably Fogo D deposit; 4 -<br />

grey pumice ash deposit, corresponding to Furnas C<br />

deposit (1900 years BP; Guest et al., 1994), topped<br />

by a very dark grey soil (1440-1500 cal AD); 5 -<br />

stratified pumice fall deposit of alternating beds of<br />

fine to medium greyish white lapilli and ash<br />

containing sanidine crystals (from 1563 AD historical<br />

eruption in Fogo volcano), topped by a very dark grey<br />

paleosol rich in coal fragments (1660-1700 cal AD); 6<br />

- remobilised deposit from the underlying unit.<br />

The AF1 affects all stratigraphic units and the<br />

existing scarp is an uneroded free-face almost devoid<br />

of soil that corresponds to the projection of the fault<br />

to the surface (Fig. 4). Several WNW-ESE to E-W<br />

trending fractures and open cracks, sometimes filled<br />

with material from overlying units, and a colluvium<br />

composed of material of units 3 and 4 were also<br />

exposed in the trench (Fig. 4).<br />

Units 2 to 4 (soil 1440-1500 cal AD) are displaced by<br />

1.0m and units 5 (soil 1660-1700 cal AD) and 6 by<br />

0.38m, indicating two surface rupturing<br />

paleoearthquakes of Mw 6.7 and 6.4 (using the Wells<br />

& Coppersmith’s, 1994, M/MD correlation),<br />

respectively, accounting for an accumulated dip-slip<br />

of 1.38m.<br />

Two earthquakes in 571-511 years indicate a<br />

recurrence interval that ranges from 286 to 256<br />

years, yielding a slip rate of 2-3 mm/year.<br />

Altiprado Fault 2<br />

The Altiprado Fault 2 trace is marked by an almost<br />

imperceptible 1690 m-long and ~40 cm-high south<br />

facing scarp, trending N87ºE (Fig. 3). To the east its<br />

location becomes uncertain.<br />

Fig. 3: Vertical aerial photograph of Altiprado region, Achada das Furnas plateau, showing the geomorphic expression of<br />

Altiprado Faults (Aerial photo from Direcção Geral de Planeamento Urbanístico, 1974).<br />

Fig. 4: Map of the east wall of the Altiprado Fault 1 trench.<br />

23


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

A 29 m-long trench exposed two subvertical faults<br />

(AF2-1 and AF2-2) that display frequent changes in<br />

dip sense (N75-89ºE, generally dipping 62-88ºS) and<br />

an E-W trending paleo-channel (filled with unit 6)<br />

(Fig. 5). The exposed stratigraphic sequence is the<br />

same observed in AF1 trench, with exception of unit<br />

1 that is absent (Fig. 5). There are also open<br />

fractures trending ENE-WSW to E-W, sometimes<br />

filled with material from overlying units, and a<br />

colluvium, containing material of unit 3, deposited<br />

against the AF2-1 fault (Fig. 5).<br />

The faults produced differential vertical separations<br />

on units 2 and 3 (younger than ~4500 years BP), and<br />

on units 4 (~1900 years BP; soil 1440-1500 cal AD)<br />

and 5 (1563 AD): AF2-1 - accumulated dip-slip of<br />

33cm (26+7cm); AF2-2 - accumulated dip-slip of<br />

15cm (11+4cm). Assuming that the ruptures in AF2-1<br />

and AF2-2 were produced by the same earthquakes,<br />

displacement values of 0.37m (0.26+0.11cm) and<br />

0.11m (0.07+0.04cm), correspond to earthquakes of<br />

Mw of 6.4 and 6.0, respectively (using the Wells &<br />

Coppersmith’s, 1994, M/MD correlation). The<br />

accumulated displacement is 0.48m (0.37m+0.11m).<br />

The existence of an E-W trending paleo-channel<br />

suggest that it may have been developed at the base<br />

of a previous fault scarp, once the regional drainage<br />

system in this area is oriented N-S.<br />

Altiprado Faults evolution<br />

The Altiprado Faults evolution was deduced from<br />

geometric analysis of the trenches. As they are<br />

geographically close and affect the same<br />

stratigraphic succession, this allows correlating their<br />

evolution (Fig. 6):<br />

a) Deposition of units 1 (4520±90 years BP) to 3;<br />

b) Surface rupture at AF2-1 and AF2-2 with normal<br />

separations of 26cm and 11cm, respectively (Mw 6.4<br />

Fig. 5: Map of the west wall of the Altiprado Fault 2 trench.<br />

earthquake);<br />

c) Erosion truncating unit 3, with the formation of<br />

gullies in AF1 fault zone, and fault scarp retreat at<br />

AF2-1 with the formation of a colluvial wedge (C) with<br />

material from unit 3;<br />

d) Deposition of unit 4 (~1900 years BP) and soil<br />

development (1440-1500 cal AD);<br />

e) Surface rupture (Mw 6.7) at AF1 with normal<br />

separation of 1.0m;<br />

f) Erosion with minor fault scarp retreat with formation<br />

of a colluvial wedge (C) in AF1;<br />

g) Deposition of unit 5 (1563 AD – Fogo Volcano<br />

historical eruption) with sin-eruptive ruptures (Mw<br />

6.0?) in AF2-1 and AF2-2, without geomorphic<br />

expression, of 7cm and 4cm, respectively;<br />

h) Erosion truncating the top of unit 5 and formation<br />

of a new paleo-channel in AF2 fault zone;<br />

development of a soil (1660-1700 cal AD);<br />

i) Erosion truncating unit 5 and deposition of unit 6;<br />

j) Surface rupture (Mw 6.4) at AF1 with normal<br />

separation of 0.38m;<br />

k) Erosion truncating units 5 and 6; development of<br />

present top soil.<br />

DISCUSSION<br />

Analysing the geological history, the first surface<br />

rupture earthquake is associated to AF2, originating<br />

an accumulated dip-slip of 37cm in AF2-1 and AF2-2.<br />

It occurred in pre-historical times, after the deposition<br />

of unit 3 (Fogo D deposit,


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

earthquakes in the island and the youthful aspect of<br />

the scarp, this event is probably related to October<br />

22 nd , 1522 historical earthquake. This was one of the<br />

most destructive events that occurred in Azores,<br />

causing 5000 deaths. The epicentre was located<br />

inland, a few km southwards from AF1 (Fig. 7) and<br />

triggered several landslides and severe damage,<br />

mainly in the central eastern part of the island.<br />

Fig. 8: Epicentre distribution of 2005 seismic swarm<br />

(May 10 th<br />

to December; data from CVARG/CIVISA,<br />

2005).<br />

Fig. 7: Isoseismal map for October 22 nd , 1522<br />

earthquake (Silveira et al., 2003).<br />

The third rupture is associated with AF2. As it is not<br />

clear if the last displacement (11cm in AF2-1 and<br />

AF2-2) is affecting only the lower part of unit 5 (1563<br />

AD) or affects the entire unit, this could be associated<br />

with the intense seismic activity that accompanied<br />

the 1563 historical eruption of Fogo volcano. The<br />

earthquakes were felt in a wide region and caused<br />

severe damage in Ribeira Grande and Ribeira Seca<br />

villages. Otherwise it could have occurred after this<br />

volcanic event.<br />

The fourth earthquake is related with AF1 and<br />

occurred after the deposition of unit 6 (>1660-1700<br />

cal AD) with normal separation of 0.38m.<br />

If the last three earthquakes occurred in historical<br />

period, the recurrence interval for Altiprado Faults<br />

zone is 163 years.<br />

There is no evidence of a high magnitude earthquake<br />

(Mw 6.4) with inland epicentre in historical records,<br />

as well in the instrumental seismicity data that could<br />

justify the most recent displacement. However, there<br />

is a long time record of seismic swarms in this area,<br />

with earthquakes of low to moderate magnitude (e.g.<br />

1967, 1989 and 2005).<br />

The most recent seismic swarm started on May 10 th ,<br />

2005 and continued till the end of that year,<br />

registering more than 46000 earthquakes (Fig. 8).<br />

The strongest events occurred on September 20 th<br />

and 21 th and had ML 4.1 and 4.3, respectively,<br />

causing several landslides and ground cracking.<br />

GPS monitoring data allowed Trota (2008) to verify<br />

that there was also ground deformation associated<br />

with the seismic activity. Considering the seismicity<br />

pattern, the low seismic energy released and the<br />

GPS data, this activity has been related to a<br />

magmatic intrusion. The previous seismic swarms<br />

were not geodetic monitored and the seismic network<br />

was limited, so there are no evidences of previous<br />

episodes of crustal deformation and the seismic<br />

parameters may have significant errors associated.<br />

Nevertheless, one possibility is that the 0.38m<br />

displacement observed in AF1 trench could be an<br />

accumulated value as a consequence of several<br />

moderate earthquakes during the periods of seismic<br />

activity increments in this area that might have been<br />

associated with crustal deformation.<br />

Acknowledgements: Rita Carmo is supported by a PhD<br />

Grant from Fundo Regional da Ciência e Tecnologia.<br />

Fieldwork was financed by Centro de Informação e<br />

Vigilância Sismo-Vulcânica dos Açores (CIVISA).<br />

References<br />

Guest, J.E., Duncan, A.M., Cole, P.D., Gaspar, J.L.,<br />

Queiroz, G., Wallenstein, N., Ferreira, T. (1994) –<br />

Preliminary report on the volcanic geology of Furnas<br />

volcano, São Miguel, Azores. ESF Furnas Laboratory<br />

Volcano, Eruptive History and Volcanic Open File Report<br />

1, 24 p.<br />

Silveira, D., Gaspar, J.L., Ferreira, T., Queiroz, G (2003) -<br />

Reassessment of the historical seismic activity with major<br />

impact on S. Miguel island (Azores). Natural Hazards and<br />

Earth System Sciences 3, 615-623.<br />

Trota, A. (2008) - Crustal deformation studies in S. Miguel<br />

and Terceira islands (Azores). Volcanic unrest evaluation<br />

in Fogo/Congro area (S. Miguel). PhD thesis, Azores<br />

Univ., 281p.<br />

Wallenstein, N. (1999) – Estudo da história recente e do<br />

comportamento eruptivo do Vulcão do Fogo (S. Miguel,<br />

Açores). Avaliação preliminar do “hazard”. PhD thesis,<br />

Azores Univ., 266p.<br />

25


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

FAULT TECTONICS REGARDING THE NEOTECTONIC PERIOD AND INFLUENCE OF<br />

TECTONIC STRUCTURES ON GLACIAL PROCESS IN AREAS OF THICK QUATERNARY<br />

COVER<br />

Jolanta yžien<br />

Lithuanian Geological Survey. Konarskio, 35. 03123-Vilnius. LITHUANIA. Email: jolanta.cyziene@lgt.lt<br />

Abstract (Fault tectonics regarding the neotectonic period and influence of tectonic structures on glacial process in<br />

areas of thick Quaternary cover): Through the Quaternary period Lithuania has been covered by continental ice sheets<br />

originated in Fennoscandia which corresponds to all glaciations known so far in the Eastern Europe, this causing very complicated<br />

structure of the Quaternary cover. This paper discusses the coincidence of outlines of main morphological features with regional<br />

tectonic structures and also whether tectonic processes influence on Quaternary development. To understand the tectonic<br />

processes influence on Quaternary sediments, neotectonic structures and movements were estimated on the basis of studies of<br />

the sub-Quaternary relief, Quaternary succession, modern relief and drainage system. A comparison of sub-Quaternary relief and<br />

present surface with tectonic structure and neotectonic activity of studied territories has revealed a frequent coincidence of linear<br />

and areal geological and geomorphological objects.<br />

Key words: Quaternary, (neo)tectonics, Baltic, palaeoinsicions<br />

INTRODUCTION<br />

The territory of Lithuania and adjacent areas in the<br />

Baltic Sea Region, located in the south-western<br />

margin of the East European Craton, could be<br />

considered as a region of low seismic activity due to<br />

the Early Precambrian crust and distant location to<br />

the active tectonic regions. Nevertheless, more than<br />

40 seismic events (historical and instrumental) with a<br />

M4.5 and intensities up to VI-VII (MSK-64 scale)<br />

were reported in the Baltic countries and adjacent<br />

territories since year 1616 (Pasa, 2007), and two<br />

strong earthquakes that took place in the Kaliningrad<br />

region on 21 September 2004, the magnitudes being<br />

4.4 and 5.0. Recent seismic activity of Baltic Sea<br />

Region often related to glacio-isostatic rebound of the<br />

Fennoscandian Shield, but also could be triggered by<br />

plate-scale North Atlantic ridge-push forces (Pascal<br />

et al., 2010). The territory of Lithuania is one of<br />

classical regions with Quaternary cover formed<br />

during continental glaciations. The average thickness<br />

of Quaternary cover in Lithuania is approximately 130<br />

m and varies from 10-30 m in the northern part of<br />

country – the area of prevailing glacial erosion – up<br />

to 200-300 m in marginal highlands and the buried<br />

valleys or palaeoincisions. The processes of<br />

accumulation, erosion during the glaciations and icefree<br />

periods, developed under the influence of<br />

neotectonic movements, have created the wide<br />

variety of Quaternary sediments and landforms.<br />

These processes made great impact creating the<br />

present shape of the sub-Quaternary surface and<br />

modified tectonic structures. Structural units of the<br />

landscape generated by a combination of tectonic<br />

activity and climate (morphostructures) often coincide<br />

with deep pre-Quaternary structures (Šliaupa, 1998)<br />

The coincidence of the faults defined by geophysical<br />

and well data to morphotectonic lineaments was<br />

considered as strong evidence, determining its<br />

“activity” during the glacial and post-glacial times<br />

(Šliaupa, 2003). To understand the tectonic<br />

processes influence on Quaternary sediments,<br />

neotectonic structures and movements were<br />

estimated on the basis of studies of the sub-<br />

Quaternary relief, Quaternary succession, modern<br />

relief and drainage system. This paper presents<br />

discussions on the possibilities to detect<br />

(neo)tectonic characteristics in the regions with<br />

Quaternary cover formed during continental<br />

glaciations and assumptions for the morphotectonic<br />

evidence.<br />

GEOLOGICAL SETTING<br />

Lithuania is situated within the Baltic sedimentary<br />

basin. The sedimentary succession of Lithuania is<br />

subdivided into four major structural-sedimentary<br />

complexes: Baikalian, Caledonian, Hercynian and<br />

Alpine and consists of Upper Vendian to Quaternary<br />

sedimentary rocks resting on an Early Proterozoic<br />

crystalline basement. The thickness of the<br />

sedimentary cover ranges from 0.2 km in the East to<br />

2.3 km in west Lithuania. All of the complexes are<br />

separated by unconformities within the sedimentary<br />

succession that represents periods of non-deposition<br />

and erosion. A set of faults are recognized in the<br />

sedimentary cover that are most distinct in the<br />

Caledonian complex. The displacements of some<br />

reverse faults exceed 200 m in western Lithuania<br />

(e.g. Telšiai fault). The oldest traces of the tectonic<br />

activity are recorded in the sub-Jotnian rocks that<br />

were preserved in a small area in west Lithuania. A<br />

limited extensional faulting took place in Vendian and<br />

Cambrian times, as it was identified in seismic<br />

26


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

profiles in western and central Lithuania. This<br />

extensional event(s) is related to the initial stages of<br />

establishment of the passive margin due to breaking<br />

apart of the Rodinian supercontinent. In western<br />

Lithuania these basement blocks were draped by<br />

Cambrian rocks, and they are overlain by Vendian<br />

deposits in the eastern Lithuania. It was rather quiet<br />

during Ordovician and Silurian times, no structuring<br />

was documented except some gentle flexuring. The<br />

tectonic forces drastically increased in latest Silurian<br />

– earliest Devonian times relating to far-field stress<br />

transmission generated by Scandian orogeny in<br />

Scandinavian Caledonides. The dense family of<br />

compressional and transpersional faults was<br />

established in western Lithuania, while faulting was<br />

only scarce in the eastern half of Lithuania that is<br />

accounted to stronger lithosphere and longer<br />

distance the stress source. The W-E and WSW-ENE<br />

oriented faults show transpressional geometries,<br />

whereas the NE-SW striking faults are compressional<br />

features (Šliaupa et al., 2002). The morphology of<br />

faults is rather variable, ranging from single-plain<br />

fault to complex flower structure and terraced fault<br />

sets. The faulting associated to formation of local<br />

uplifts along the hanging walls of the faults. The<br />

Telšiai reverse fault, formed in transpressional<br />

regime, is the largest tectonic feature in the<br />

sedimentary cover of Lithuania. This fault trends W-E<br />

for a few hundred kilometres, it shows variable<br />

geometry along the strike that associate with<br />

changing abundance and scale of local uplifts. Faults<br />

are rather rare in the younger structural complexes<br />

where flexures are most common features of tectonic<br />

displacement. Structural-sedimentary complexes<br />

differ by their geological composition and<br />

independently by the structural pattern. Especially<br />

high variety of genesis and lithological composition of<br />

the Quaternary deposits is reflected in drastic<br />

changes both in vertical and lateral distribution of<br />

Quaternary sediments. Surface formations and<br />

geomorphological features in a large part of<br />

Lithuanian territory were formed during the Late<br />

Nemunas (Late Weichselian) Glaciation.<br />

ASSUMPTION FOR THE MORPHOTECTONIC<br />

EVIDENCE<br />

The studies focusing on mutual relationships<br />

between glacial landforms and tectonic structures in<br />

areas glaciated in the Pleistocene usually take into<br />

account two aspects. The first one concerns the<br />

influence exerted by pre-glacial tectonic structures on<br />

the behaviour of the ice-sheet, controlling thereby<br />

same glacial landforms. One of the issues under<br />

consideration is the influence of fault zones<br />

reactivated by ice-sheet load upon location and<br />

course of subglacial tunnels. The second aspect is<br />

related to post-glacial neotectonic movements, the<br />

most prominent manifestations of which are<br />

tectonically-controlled erosional scarps, either<br />

undermining or displacing glacial landforms.<br />

Essential question concerns coincidence of outlines<br />

of main morphological features with regional tectonic<br />

structures and also whether tectonic processes<br />

influence on Quaternary development, and in which<br />

way.<br />

Neotectonic movements of the Earth’s crust are<br />

reflected by deformation processes often<br />

accompanied by tectonic activity. Number of<br />

publications reported that the last deglaciation of<br />

northern Fennoscandia was accompanied by a high<br />

seismic activity. The earthquakes triggered landslides<br />

in glacial till, seismically-induced soft sediment<br />

deformation structures, “seismites”, are common in<br />

trench exposures in the vicinity of the faults in<br />

northern Sweden. Deformation of sandy-silty<br />

sediments, potentially caused by earthquakes, have<br />

also been encountered in central and southern<br />

Sweden (e.g., Mörner, 2004), but are less common<br />

than in the north. The deglaciation history in<br />

Lithuania is longer, but there is so far not recorded<br />

and no published evidence of paleoseismic events. It<br />

must be pointed out that the late- to postglacial fault<br />

scarps identified in northern Sweden are all<br />

developed in the Precambrian crystalline basement,<br />

and mainly in rocks of Proterozoic age.<br />

Morphologically prominent faults occur also in the<br />

Caledonian bedrock in the mountain range, but so far<br />

no recent fault movements along any of these<br />

features have been indicated (Lagerback & Sundh,<br />

2008). The geological setting of Lithuania is very<br />

different compared to Sweden. The study area is<br />

covered entirely by Quaternary sediments of glacial<br />

origin (average thickness of sediments is 130 m).<br />

Two major types of faults prevail in Lithuania, i.e. the<br />

oldest, defined only in the Precambrian crystalline<br />

basement and do not dissecting the sedimentary<br />

cover and younger that penetrate into the sediments<br />

overlying the crystalline basement. In comparison to<br />

Sweden, no faults so far have been detected in<br />

Quaternary succession. Fault tectonics regarding the<br />

neotectonic period in Lithuania, as well as in all Baltic<br />

region, is rather subtle problem. It is difficult to<br />

determine the influence of tectonic structures on<br />

glacial process in the Pleistocene and testing its<br />

influence on the process of the mass movements in<br />

glaciotectonic and neotectonic young-alpine<br />

structures. Tectonic deformations of the sub-<br />

Quaternary relief during Neogene-Quaternary or<br />

Quaternary period, i.e. a part of a tectonic and<br />

denudation factors imprinting the sub-Quaternary<br />

surface, are the key problem. Numerous evidences<br />

reported from the Baltic Region indicate that steps,<br />

elevations and depressions of the sub-Quaternary<br />

surface are partly of tectonic nature (Šliaupa &<br />

Popov, 1998).<br />

Large-scale landforms and neotectonic movements<br />

were investigated in close connection to structural<br />

features of the pre-Quaternary and Quaternary<br />

deposits. Those movements were predetermined at<br />

large by ancient pre-Quaternary tectonic structures<br />

showing inherited trends of movements. In Lithuania<br />

the commonly used term “neotectonic fault zone”<br />

concerns, in fact, mostly a system of linear elements<br />

defined by remote sensing, morphometric, and<br />

geomorphological-structural investigations. The<br />

influence of linear tectonic zones is shown in a sharp<br />

change of the composition and structure of the<br />

Quaternary cover. Based on available data of<br />

geological mapping at scale 1:50 000 (coverage c.a.<br />

27


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

48 % of the country), especially from the areas with<br />

dense network of boreholes, several factors having<br />

morphotectonic implications must be pointed out:<br />

- block structures of Quaternary cover;<br />

- palaeoincisions of pre-Quaternary surface and<br />

inside of Quaternary cover;<br />

- linear structures of present topography;<br />

- river valleys (tunnel valleys);<br />

- ravines;<br />

- palaeolacustrine basins and zones of distortions;<br />

- glaciodislocations, rafts of pre-Quaternary rocks;<br />

- intrusions of mineralised water and springs<br />

(yžien & Satknas, 2008).<br />

DISCUSSION<br />

A particular role belongs to deep palaeoincisions of<br />

sub-Quaternary surface and tunnel valleys of present<br />

topography. The palaeoincisions are distinct feature<br />

of pre-Quaternary surface and particularly are proper<br />

for the Baltic Highland reaching even 280 m in depth.<br />

Network of especially deep palaeoincisions is<br />

determined in the Moltai Lakeland (north of Vilnius).<br />

Genetically palaeoincisions of pre-Quaternary<br />

surface are analogous to the tunnel valleys, which<br />

were formed during subglacial erosion by meltwater<br />

under the glaciodynamic pressure. In the cases of<br />

clear correspondence of palaeoincisions with tunnel<br />

valleys, their morphotectonic implication could be<br />

concluded (yžien & Satknas, 2008). However,<br />

reliable determination of spatial form and presence of<br />

network of palaeoincisions requires very dense<br />

network of boreholes and is dependent on ways of<br />

interpretation and interpolation of topography of pre-<br />

Quaternary surface (e.g. Šliaupa et al., 1999).<br />

Therefore, different patterns of forms and networks of<br />

palaeoincisions are presented by different<br />

researchers.<br />

The problem of the genesis of palaeoincisions of sub-<br />

Quaternary surface in Lithuania is of great practical<br />

importance. The genesis of these particular forms is<br />

discussed taking into account all factors that could<br />

influence the formation of palaeoincisions in the<br />

environment of continental glaciations. Analysis of<br />

palaeohydrography during the interglacial periods<br />

shows that basis of erosion could not be the reason<br />

for incision of river valley 100-120 m below the<br />

present sea level. The basis of fluvial erosion at the<br />

end of the Neogene-beginning of the Quaternary<br />

most probably was not lower than 50-60 m above the<br />

present sea level and this could not affect incision or<br />

river valleys deeper than +50 m (Satknas, 2000).<br />

Bottom of deepest palaeoincisions of Vilnius environs<br />

occur at an altitude of 10-56 m below the present sea<br />

level. According to the age of tills, occurring in<br />

vicinities of the studied palaeoincisions, these forms<br />

have been formed during the Dainava (Elsterian) and<br />

Žemaitija (Saalian) glaciations. The lithology and<br />

rhythmic structure of the material filling the<br />

palaeoincisions show that they have been formed as<br />

a consequence of subglacial erosion under glacial<br />

pressure. Morphological similarity of palaeoincisions<br />

and ravines of present topography indicated the<br />

same origin of both these forms. The ravine Tapeliai<br />

(10 km north-east of Vilnius) was studied, using<br />

special borehole data. Sand and gravel (total<br />

thickness 44 m) filling the ravine has a rhythmic<br />

structure characteristic of the palaeoincisions of sub-<br />

Quaternary surface. This property also shows the<br />

general similarity of ravines of present topography<br />

and palaeoincisions (Satknas, 2008).<br />

Tunnel valleys of present topography are revealed on<br />

the geomorphological map (Guobyt, 2000) and their<br />

good correspondence with photo-lineaments are<br />

concluded (Guobyt, 1995). Besides that, the tunnel<br />

valleys in most cases are interpreted as<br />

neotectonically active linear zones (Šliaupa, 2005).<br />

The coincidence of palaeoincisions and the tunnel<br />

valleys in many cases still has to be confirmed by<br />

boreholes, which are generally absent in the tunnel<br />

valleys. Therefore, due to lack to the direct data the<br />

palaeoincisions in places seems hardly correlating<br />

with tunnel valleys (yžien & Satknas, 2008). On<br />

the other hand, it still remains underestimated<br />

presence of paleoincisions that do not reach the<br />

surface of pre-Quaternary rocks.<br />

A comparison of sub-Quaternary relief and present<br />

surface with tectonic structure and neotectonic<br />

activity of territories has revealed a frequent<br />

coincidence of linear and areal geological and<br />

geomorphological objects.<br />

Acknowledgements: Thanks to the Marie Curie<br />

Programme „Transfer of Knowledge“ in the Sixth<br />

Framework Programme for the support of the project<br />

"Morphotectonic Map of European Lowland Area”.<br />

References<br />

yžien, J., Satknas, J. (2008). Morphotectonic of<br />

Quaternary landforms, examples from Baltic region.<br />

Volume of abstracts of the third conference of MELA<br />

project Cartographical approach of the morphotectonic of<br />

European lowland area, 20-23.<br />

Guobyt, R. (1995). Geological mapping of the Moltai<br />

area, 1:50 000. Geological Report No. 5478, Lithuanian<br />

Geological Survey (LGT).<br />

Guobyt, R. (2000). Geomorphological map of Lithuania,<br />

1:200 000. Geological Report No. 4378, Lithuanian<br />

Geological Survey (LGT).<br />

Lagerbäck, R., Sundh, M. (2008). Early Holocene faulting<br />

and paleoseismicity in northern Sweden. Research Paper<br />

C 836, Geological Survey of Sweden, 80 p.<br />

Mörner, N.-A. (2004). Active faults and paleoseismicity in<br />

Fennoscandia, especially Sweden. Primary structures<br />

and secondary effects. Tectonophysics, 380, 139-157.<br />

Pasa, A. (2007). Seismological observation in Lithuania.<br />

Volume of abstracts of the international workshop<br />

Seismicity and seismological observations of the Baltic<br />

sea region and adjacent territories, September 10-12,<br />

2007, Vilnius, Lithuania, LGT, 66-69.<br />

Pascal, C., Roberts, D., Gabrielsen, R.H. (2010). Tectonic<br />

significance of present-day stress relief phenomena in<br />

formerly glaciated regions. Journal of the Geological<br />

Society, 167, 363-371.<br />

Satknas, J. (2000). Formation of palaeoincisions in the<br />

environment of continental glaciations: an example of<br />

Eastern Lithuania. Geologija, 31, 52-65.<br />

Satknas, J. (2008). Morphotectonic implication of tunnel<br />

valleys in areas with thick Quaternary cover. Volume of<br />

abstracts of the third conference of MELA project<br />

28


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Cartographical approach of the morphotectonic of<br />

European lowland area, 42-45.<br />

Šliaupa, A. (1998). Neotectonic structures of Lithuania and<br />

Adjacent Territories. Litosfera, 2, 37-46.<br />

Šliaupa, A. (2005). Neotectonic map of Lithuania and<br />

adjacent areas, 1:1 000 000. Evolution of geological<br />

environment in Lithuania, Vilnius, Lithuanian Geological<br />

Survey, Report, CD.<br />

Šliaupa, S., Popov, M. (1998). Linkage between Basement<br />

and Neotectonic Linear Structures in Lithuania. Litosfera,<br />

2, 23-36.<br />

Šliaupa, S. (2002). Kinematic features of the Telšiai fault in<br />

Western Lithuania: structural and permeability prognosis..<br />

Geologija, 38, 24-30.<br />

Šliaupa, S. (2003). Geodynamic evolution of Baltic<br />

sedimentary basin. Doctoral thesis, Geological Institute,<br />

Vilnius, 207 p.<br />

29


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PLIO – PLEISTOCENE TECTONIC ACTIVITY IN THE SOUTHWEST OF PORTUGAL<br />

Figueiredo, Paula M. (1, 2), Cabral, João (1,2), Rockwell, Thomas K. (3)<br />

(1) IDL, Instituto Dom Luiz, pmfigueiredo@fc.ul.pt<br />

(2) Geology Department, Science Faculty, Lisbon University<br />

(3) Geological Department, San Diego State University<br />

Abstract: Southwest Portugal, located close to the Eurasia-Núbia plate boundary, is characterized by moderate<br />

seismicity, although strong events may occur, as in 1755 (Mw8), 1969, (Mw 7.9), and more recently in 2007 (Mw<br />

6.3) and 2009 (Mw 6.1), all located in the offshore. No historical earthquakes with onshore rupture are known for<br />

this region. Inland, recent neotectonic and paleoseismological studies corroborate seismogenic activity during the<br />

Pleistocene, at the S.Teotónio–Aljezur–Sinceira Fault System, suggesting that these structures are active and are<br />

potential sources for moderate seismicity. At the coastline, several features such as old beach sediments, paleo<br />

abrasion platforms and paleo cliffs were recognized. A sequence of poorly preserved surfaces with thin deposits<br />

may correspond to Pleistocene marine terraces, suggesting a higher uplift rate than expected for this region.<br />

Key words: Pleistocene, Uplift, active tectonics, Portugal<br />

INTRODUCTION<br />

Southwestern Portugal is located close to the<br />

Eurasia-Nubia plate boundary, near the Azores-<br />

Gibraltar fracture zone. East of the Gloria transform<br />

fault, this boundary becomes complex and diffuse,<br />

where deformation is distributed across a few<br />

hundred kilometres wide zone, related to the NW-SE<br />

convergence of Iberia and Nubia at a rate of ca. 4-5<br />

mm/yr. In this area is located the inferred<br />

seismogenic source zone of the 1755 earthquake<br />

and tsunami (estimated Mw 8), and also of the Mw<br />

7.9 1969 event. (Zittelini et al., 2009).<br />

Fig. 1: Main Neotectonic structures located in southwern<br />

Portugal. Onshore structures: APF, Alentejo-Placencia<br />

Fault; STASFS, São Teotónio-Aljezur-Sinceira Fault<br />

System. Offshore Structures: PSF, Pereira de Sousa Fault;<br />

MPF, Marquês de Pombal Fault; HSF, Horseshoe Fault;<br />

GBF, Guadalquivir Bank Faults.<br />

and other faults trending close to E-W, such as the<br />

Guadalquivir Bank and SWIM faults (Figure 1). To<br />

understand and study the recent tectonic activity in<br />

this sector of Iberia, it is necessary to study the<br />

individual inland structures but also inland<br />

deformation evidences that may be related to some<br />

of the referred offshore active structures. This<br />

abstract depicts both perspectives.<br />

DISCUSSION<br />

The São Teotónio –Aljezur- Sinceira Fault System<br />

The São Teotónio–Aljezur–Sinceira fault system<br />

(STASFS) corresponds to the nearest inland brittle<br />

structure that may correlate to the ongoing plate<br />

boundary deformation in the offshore. It extends<br />

NNE-SSW for 50 km, parallel and close to the<br />

southwest Portuguese coast (Figure 1), and<br />

comprises left-lateral strike-slip faults, deforming a<br />

large regional abrasion platform ca. 10 km wide,<br />

considered of probable late Miocene age, and which<br />

was reoccupied during the Pliocene and the<br />

Pleistocene. Four small Cenozoic tectonic basins,<br />

filled with Miocene to Pleistocene sediments, occur<br />

along the STASFS.<br />

Post-Pliocene vertical displacements of up to 100 m<br />

may have occurred related to tectonic activity in the<br />

STASFS, but generally these only reach a few tens<br />

of meters. Dias (2001) estimated a slip-rate of ca.<br />

0.03-0.06 mm/yr, based on the vertical offset of<br />

morphological features. However, as the main slip on<br />

these structures is strike-slip, those values are a<br />

minimum estimative. Dias (2001), taking into account<br />

This seismogenic source zone is characterized by<br />

several structures trending NNE-SSW to NE-SW, as<br />

the Marquês de Pombal and the Horseshoe faults,<br />

30


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: South wall of trench ALF1, located at the Alfambras basin. Description of geology included in the figure.<br />

the length of the known fault also estimated a<br />

maximum magnitude of Mw 7, an average co-seismic<br />

displacement of 1 m and a recurrence period of<br />

19210 - 32017 yr, based upon Wells & Coppersmith<br />

empirical relationships.<br />

Recently, through detailed geomorphologic and field<br />

work studies, several paleoseismic sites were<br />

selected. At the Alfambras basin, two trenches were<br />

excavated: in the ALF-1 trench (Figure 2), one of the<br />

active (?) fault branches was identified, although no<br />

paleo-earthquakes were individualized. This fault<br />

deforms post-Miocene sediments that may be<br />

Pleistocene. OSL samples were collected to better<br />

constrain the ages. A paleosoil, probably 700 ka old,<br />

is also faulted showing a vertical displacement of ca.<br />

1 m. We saw no deformation within the upper soil<br />

units; therefore we could not state an obvious<br />

correlation between the fault trace and the<br />

topography which would imply a more recent activity.<br />

Since no piercing points were identified, no lateral<br />

displacement was quantified at this site.<br />

existence of several fault traces near the topographic<br />

surface.<br />

The Framangola site (at Alfambras basin) is located<br />

close to a feature recognized as a shutter ridge and<br />

two trenches were opened at the alluvial plain to<br />

investigate the Holocene sequence. Unfortunately,<br />

the very coarse sediments and the high water table<br />

did not allow us to safely investigate this trench site<br />

at a deeper level: a thick gravel sequence was<br />

recognized as likely to be related to several<br />

landslides triggered during the 1755 earthquake<br />

based upon archaeological artefacts.<br />

3<br />

2<br />

1<br />

Fig.4. Geoelectric tomography profile (using Schlumberger<br />

method) at Framangola site at the alluvial plain, strongest<br />

contrast coincides with previously identified fault trace; the<br />

trench units identified also correlate with the tomography (1)<br />

Clays and fine silts; (2) Very coarse Gravel (3) Silty sandy &<br />

fine conglomerate<br />

Fig.3: Geoelectric tomography profile (using<br />

Schlumberger method) at Monte Ferreiros site, strong<br />

contrast coincides with geomorphologic trace interpreted<br />

in the lower image.<br />

More paleoseismic sites were selected farther north<br />

in the Aljezur and Alfambras basins, where<br />

geoelectric tomography profiles (Figures 3 & 4) up to<br />

50 m depth were obtained corroborating the<br />

At its northern end, near São Teotónio, the STASFS<br />

joins the Alentejo-Placencia fault (APF), the iberian<br />

fault with higher length. In this area the tectonic<br />

deformation becomes distributed, suggesting than<br />

APF may have several splays along this intersection<br />

and it is difficult to scrutinize the relationship between<br />

both structures. No Holocene or late Pleistocene<br />

deposits were recognized in association with the<br />

identified fault segments, which increase the<br />

uncertainty concerning the recent activity. Presently it<br />

is still not clear whether the STASFS extend to the<br />

north of the APF.<br />

31


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Towards the south, at the southern sector of the<br />

STASFS, south of the Sinceira basin, the local<br />

morphology does not evidence any significant Plio-<br />

Pleistocene vertical deformation related with this fault<br />

system, and it seems that this fault system may splay<br />

into several faults. In fact, several parallel faults<br />

occur in the Mesozoic bedrock sediments where<br />

limestones dominate. Several karst pits are filled with<br />

Plio-Pleistocene sands (Faro-Quarteira sands)<br />

mainly at the coastal section Martinhal – Zavial<br />

where the STASFS should intersect the coastline.<br />

The elongated shape those karst sometimes present,<br />

suggested a structural control by a previous<br />

structural fabric on the karst development (Dias and<br />

Cabral, 2002).<br />

The Coastal region<br />

The southwestern Portuguese littoral is characterized<br />

by coastlines with two distinct morphologies: one<br />

trending N-S, presenting high cliffs in Paleozoic<br />

schists and greywackes that reach up to 100 m<br />

height, with several overhanging fluvial valleys and a<br />

narrow abrasion platform, and another trending E-W,<br />

formed mainly in Mesozoic limestones, forming an<br />

irregular and lower coastline, with karst wells filled<br />

with Plio-Pleistocene sands (generally of the Faro-<br />

Quarteira Sands regional stratigraphic unit) as<br />

referred above.<br />

abrasion platform considered late Miocene in age<br />

and reoccupied during Pliocene and Pleistocene<br />

times (figure 5).<br />

The highest beach sediments we were able to<br />

identify are at Fonte Santa, circa 350 m elevation,<br />

and underlying aeolionites considered to be Pliocene.<br />

In the western coast, north of Sagres, where the cliffs<br />

are very abrupt and the drainage is strongly incised,<br />

we recognized a raised beach consisting of a coarse,<br />

pebbly layer underlying beach sands at 76 m height<br />

(Figures 5 & 6). Two samples of sand were collected<br />

for OSL dating. The basal unconformity surface<br />

apparently dips gently southwards (~2º).<br />

Further to the north, at the Castelejo beach (Figure 5<br />

& 7), we identify an aeolionite with several paleosoils<br />

and colluviums that overlies a marine abrasion<br />

platform on Palaeozoic schists circa 2 m elevation,<br />

almost coincident with the modern one. We collected<br />

OSL samples at the base of this sequence with a<br />

resulting age of ca.64 ka. Underlying the aeolionite<br />

and overlying the abrasion platform two<br />

conglomerate beach deposits with a thin beach sand<br />

layer between them were identified and OSL samples<br />

were collected. We interpreted this to represent<br />

aeolian deposition during MIS 4, when sea level was<br />

lowered to expose the offshore sandy marine<br />

sediments, and the underlying platform to be the late<br />

stage 5 (MIS 5a) marine terrace. Correlative<br />

sediments were also identified further to the north, at<br />

Amoreira beach.<br />

ae<br />

Fig.5 Digital Terrain Model of the Sagres region, southwest<br />

STASFS showing the location of sites referred in the text.<br />

bc<br />

bs<br />

bc<br />

Fig. 6 Pleistocene beach sediments resting over a poorly<br />

preserved abrasion surface cut on Triassic sediments at 65-<br />

76 m height at Telheiro (adapted from Dias, 2001). Beach<br />

sands overlie a basal pebbly deposit. At the top of the<br />

sequence occur several eolionites apparently with distinct<br />

paleosoils, suggesting several eolic sedimentation events.<br />

Detailed field surveys were conducted in both<br />

coastlines in order to recognize paleo shore-line<br />

features and beach deposits in a wide raised<br />

Fig. 7 Pleistocene beach deposit lying over an abrasion<br />

platform cut on Carboniferous schists, at 2 m height (bc -<br />

beach conglomerate; bs – beach sand). Above this deposit<br />

outcrops a sequence of strongly carbonate cemented<br />

eolianites, likely to be MIS 4 (ae).<br />

At the E-W trending southern coast, detailed<br />

geomorphologic studies combined with field survey<br />

and paleosoils characterization, suggest that the<br />

culminant abrasion platform, which is well bc preserved<br />

near Sagres, dipping gently to the Southeast, may<br />

32


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

actually correspond to a sequence of closely spaced,<br />

poorly preserved, middle Pleistocene marine<br />

terraces.<br />

A raised abrasion platform cut on Mesozoic<br />

limestones, at ~12 m elevation, and the<br />

corresponding paleo-cliff were recognized at the<br />

Ingrina beach, though no beach sediments were<br />

encountered (Figure 8). This raised abrasion platform<br />

at Ingrina, probably corresponds to the MIS 5e<br />

abrasion surface.<br />

poorly expressed, early to middle Pleistocene<br />

terraces. OSL samples and marine shells were<br />

collected for dating at Telheiro, where beach deposits<br />

are present at 76 m elevation. In the same region at<br />

Castelejo, we dated an aeolionite that directly<br />

overlies a marine abrasion platform at 2 m elevation,<br />

with a resulting age of ca. 64 ka. We interpreted this<br />

to represent aeolian deposition during MIS 4, when<br />

sea level was lowered to expose the offshore sandy<br />

marine sediments, and the underlying platform to be<br />

the late stage 5 (MIS 5a) marine terrace.<br />

The 12 m terrace at Ingrina, cut across carbonate<br />

rocks, probably corresponds to the MIS 5e abrasion<br />

surface. If this uplift rate is applicable for the entire<br />

Quaternary, this implies a late Miocene to Pliocene<br />

age of the aeolionites and marine deposits at Fonte<br />

Santa, which lie at an elevation of nearly 350 m.<br />

These observations imply a long-term uplift rate of<br />

about 0.06 mm/yr.<br />

These are still preliminary results, to confirm with<br />

further studies.<br />

Fig. 8. Pleistocene raised abrasion platform at Ingrina<br />

beach, at 12m elevation.<br />

West of Ingrina beach, there is a surface slightly<br />

under 50 m height, that probably corresponds to an<br />

abrasion surface, with several residual reliefs up to 5<br />

m high. This surface is overlain by a thin clayey sand<br />

deposit, which is generally covered by abundant<br />

rounded quartz pebbles, thus suggesting that the<br />

surface may correspond to a marine terrace at 45-50<br />

m. Samples were collected for OSL dating. A poorly<br />

preserved surface around 60-65 m may correspond<br />

to the surface identified at Telheiro, but no beach<br />

sediments were recognized.<br />

All these surfaces are covered by several eolianite<br />

bodies, likely to have been reactivated trough time.<br />

CONCLUSIONS<br />

Trenches excavated across the São Teotónio–<br />

Aljezur–Sinceira fault system (STASFS), which<br />

extends NNE-SSW for 50 km parallel to the<br />

southwest Portuguese coast, exposed faulted<br />

alluvium that is inset below the regional marine<br />

abrasion surface. Sparse age control based on<br />

paleosol development, along with the geomorphic<br />

position of these alluvial terrace deposits relative to<br />

the marine deposits, suggests that this faulting<br />

occurred in the middle to late Pleistocene timeframe,<br />

although we have yet to find evidence of Holocene<br />

activity.<br />

We mapped a raised abrasion platform at an<br />

elevation of ~75 m near Sagres, which dips gently to<br />

the SE and may represent part of a sequence of<br />

We attribute the observed deformations to the<br />

continued NW-SE 4 to 5 mm/yr convergence of Iberia<br />

and Nubia, with the STASFS accommodating some<br />

of the ongoing plate boundary activity.<br />

Acknowledgements : This work was funded by Fundação<br />

da Ciência e Tecnologia, through a PhD scholarship<br />

(SFRH/BD/36892/2007) and Research Project<br />

“Paleoseismological study of active faults in Mainland<br />

Portugal” (PTDC/CTE-GIN/66283/2006), co-financed by<br />

FEDER.<br />

References<br />

Cabral, J. (1995). Neotectónica em Portugal Continental,<br />

Mem. Nº31, Inst.Geol.Min., 265 p.<br />

Carrilho, F. (2005). Estudo da Sismicidade do Sudoeste de<br />

Portugal Continental, Diss. Mestrado, Dep.Fisica, Univ.<br />

Lisboa, 172 p.<br />

Dias, R. (2001). Neotectónica da Região do Algarve,<br />

Dissertação de Doutoramento, Univ. Lisboa, 369 p.<br />

Dias, R. P., Cabral, J. (2002): Interpretation of recent<br />

structures in an area of cryptokarst evolution -<br />

neotectonic versus subsidence genesis. Geodinamica<br />

Acta, 15 (4), 233-248.<br />

Manuppella, G. (Coord.) (1992) - Carta Geológica da<br />

Região do Algarve, escala 1/100 000., Serviços<br />

Geológicos de Portugal.<br />

Wells, D., Coppersmith, K. (1994). New empirical<br />

relationships among magnitude, rupture length, rupture<br />

width, rupture area and surface displacement. Bulletin<br />

of the Seismological Society of America, Vol.84 (4),<br />

974-1002.<br />

Zitellini N., Gracia E., Matias L., Terrinha P., Abreu M.A.,<br />

DeAlteriis G., Henriet J.P., Danobeitia J.J., Masson<br />

D.G., Mulder T., Ramella R., Somoza L., Diez S. (2009)<br />

- The quest for the Africa–Eurasia plate boundary west<br />

of the Strait of Gibraltar, Earth and Planetary Science<br />

Letters Vol.280, 1-4, 15 April 2009, 13-50.<br />

33


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

GEODETIC EVIDENCE OF THE CONTROL OF A MAJOR INACTIVE TECTONIC<br />

BOUNDARY ON THE CONTEMPORARY DEFORMATION FIELD OF ATHENS (GREECE)<br />

Michael Foumelis (1, Ioannis Fountoulis (2), Ioannis D. Papanikolau (3, 4), Dimitrios Papanikolaou (2)<br />

(1) Department of Geography, Harokopio University of Athens, 70 El. Venizelou Str., Kallithea, 176 71, Athens, Greece,<br />

Email: mfoumelis@hua.gr<br />

(2) Department of Dynamics Tectonics & Applied Geology, Faculty of Geology and Geoenvironment, National and Kapodistrian<br />

University of Athens, Panepistimioupolis, Ilissia, 157 84, Athens, Greece<br />

(3) Laboratory of Mineralogy & Geology, Department of Geological Sciences and Atmospheric Environment, Agricultural<br />

University of Athens, 75 Iera Odos Str., 118 55, Athens, Greece<br />

(4) AON Benfield UCL Hazard Research Centre, Department of Earth Sciences, University College London, WC 1E 6BT,<br />

London, UK<br />

Abstract (Geodetic evidence for control of a major inactive tectonic boundary on contemporary deformation field of<br />

Athens (Greece)): A GPS-derived velocity field from a dense geodetic network established in the broader area of Athens is<br />

presented, whereas local variations of strain rates across a major inactive tectonic boundary separating metamorphic and nonmetamorphic<br />

geotectonic units are also highlighted. An apparent differentiation of the eastern part of Athens plain with negligible<br />

deformation rates, from the western part where relatively higher strain rates are observed, indicate its control of the above<br />

mentioned boundary on the contemporary deformation field of the region. These findings are in agreement with previous<br />

geological observations, however, due to the dense local GPS network it was fatherly possible to localize and quantify the effect of<br />

such a major inherited tectonic feature on the deformation pattern of the area.<br />

Key words: GPS velocities, strain rates, tectonics, Athens Basin<br />

INTRODUCTION<br />

Detailed instrumental observations of the tectonic<br />

movements in Athens Basin by geodetic or other<br />

methods are absent. The contribution of previous<br />

geodetic GPS studies to examine the kinematic field<br />

of Attica are limited to observations from regional<br />

networks, designed to monitor large-scale rather than<br />

local tectonic movements (Clarke et al., 1998; Veis et<br />

al., 2003). With a limited number of stations within<br />

the region, a general picture of the motion is gained,<br />

while changes within are hardly addressed.<br />

In the present study a comprehensive GPS-derived<br />

velocity field for the broader area of Athens is<br />

presented. Variations of strain rates across a major<br />

tectonic boundary occurring in the region are<br />

highlighted and implication on the contemporary<br />

kinematics and dynamics of the region are<br />

discussed.<br />

GEOLOGICAL SETTING<br />

The Athens basement belongs to alpine formations<br />

outcropping in the mountains and the hills of the<br />

area. Recent post-alpine sediments (syn-rift deposits)<br />

often cover the slopes of the mountains as well as<br />

areas of low altitude.<br />

The area presents a complex alpine structure<br />

comprising mainly by Mesozoic metamorphic rocks of<br />

Attica geotectonic unit, occurring at Pendeli and<br />

Hymmetus mountains and Mesozoic nonmetamorphic<br />

rocks of the Eastern Greece<br />

geotectonic unit, occurring at Parnitha, Poikilo and<br />

Aegaleo mountains. The boundary between the<br />

metamorphic and non-metamorphic geotectonic<br />

units, although generally accepted to be of tectonic<br />

origin, its exact geometric and kinematic<br />

characteristics are yet to be determined, since no<br />

direct geological mapping could be undertaken. The<br />

entire tectonic structure within the area is covered by<br />

an allochtonous system, called “Athens schists”,<br />

tectonically overlaid on the two previously mentioned<br />

units, as well as Neogene and Quaternary deposits.<br />

It is traced northwards from the Aegean coast of<br />

Southern Evia, through Aliveri to Kalamos in<br />

northeast Attica and continues to the southwest into<br />

the plain of Athens. Within the area of interest its<br />

locations coincide approximately with the riverbed of<br />

Kifissos R. (Papanikolaou et al., 1999; Mariolakos &<br />

Fountoulis, 2000; Xypolias et al., 2003) (Fig. 1), also<br />

confirmed by geophysical investigations at the<br />

northern part of the basin (Papadopoulos et al.,<br />

2007). Results of seismic tomography indicate the<br />

presence of abnormally high seismic velocities in the<br />

central part of the basin, most likely related to this<br />

major boundary, extending towards the southeast at<br />

Saronikos Gulf (Drakatos et al. 2005).<br />

According to Papanikolaou & Royden (2007) this<br />

boundary represents a broad extensional detachment<br />

with significant portion of dextral shear, whereas<br />

opinions of a right-lateral strike slip fault zone have<br />

also been reported (Mariolakos & Fountoulis, 2000;<br />

Krohe et al. 2009). Considering a depth of about 30<br />

km for the metamorphics (Lozios, 1993), it is clear<br />

that this tectonic boundary has accommodated more<br />

34


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

than 25 km of displacement. It was active throughout<br />

Late Miocene times and gradually became inactive<br />

during Early Pliocene (Papanikolaou & Royden,<br />

2007). However, it forms a major boundary that<br />

separates the E-W trending higher slip-rate active<br />

faults in the western part of Attica from the NW-SE<br />

trending lower slip-rate faults in the eastern part<br />

(Mariolakos & Papanikolaou, 1987; Papanikolaou et<br />

al. 2004).<br />

Fig. 1: Simplified neotectonic map of Attica showing the<br />

approximate location of the major tectonic boundary<br />

separating metamorphic and non-metamorphic alpine rocks<br />

(modified from Papanikolaou et al. 1999).<br />

each station for a period of at least four hours. In an<br />

effort to achieve optimal results, selected stations<br />

were occupied for several days per epoch<br />

(independent sessions). To avoid large tropospheric<br />

errors an initial elevation cut-off angle of 10 was<br />

used.<br />

Collected data were processed using Leica Geo<br />

Office v.1.1 and Bernese ver. 4.2 (Beutler et al.,<br />

2001). The realization of the reference frame was<br />

performed using the coordinates and velocity of<br />

Dionysos (DION) continuous GPS station, located on<br />

the metamorphic alpine basement. DION was tied to<br />

the ITRF2000 at epoch 2005.0 by almost a decade of<br />

observations from numerous sites of the EUREF<br />

permanent network (Prof. D. Paradissis, personal<br />

communication). It can be argued that connecting the<br />

local network to the ITRF through DION reference<br />

station would be sufficient, taking into account the<br />

network extend. Details on data collection and<br />

processing could be found in Foumelis (2009).<br />

GPS NETWORK ESTABLISHMENT<br />

Given the lack of previous instrumental observations,<br />

the design of the geodetic network was primarily<br />

focused on the investigation of the local tectonic<br />

regime. The minimum number of survey points<br />

required is imposed by the tectonic complexity of the<br />

region and the degree of fragmentation of the crust.<br />

The established Athens Geodetic Network (AGNET)<br />

consisted of a total number of 41 campaign GPS<br />

sites (Fig. 2) including already available benchmarks<br />

of the Hellenic Military Geographical Service<br />

(HMGS), as well as sites previously installed by the<br />

Hellenic Mapping and Cadastral Organization<br />

(HEMCO) and the National Technical University of<br />

Athens (NTUA). Continuous (real-time) GPS station<br />

operate in the region by Metrica company (MET0),<br />

National Observatory of Athens (NOA1) and National<br />

& Kapodistrian University of Athens (UOA1), and<br />

despite their relatively limited observations at the<br />

time, they where considered in the analysis as well.<br />

The network covers essentially both the basins of<br />

Athens and Thriassio as well as their bordering<br />

mountain ranges, showing a relatively uniform spatial<br />

distribution. With an average distance between<br />

stations of approximately 5 km, a sufficient sampling<br />

of local deformation field is accomplished.<br />

GPS MEASUREMENTS AND ANALYSIS<br />

GPS campaigns were carried out from 2005 to 2008<br />

(3.2 yr) following an annual re-occupation strategy.<br />

The benchmarks of the HMGS were first measured<br />

during network establishment and together with<br />

selected GPS sites once more on 2008.<br />

Measurements were conducted using LEICA<br />

geodetic GPS receivers equipped with SR299/399,<br />

AT202/302 and Ach1202Pro antennas. Carrier phase<br />

observations were recorded every 10 seconds from<br />

Fig. 2: Annual GPS velocities of broader Athens area,<br />

relative to DION, for the period 2005-2008. The error<br />

ellipses represent the 1-sigma confidence region. Velocities<br />

of E067 and G20 benchmarks from Veis et al. (2003), and<br />

NOA1 from the EUREF website, after transformation to the<br />

specific ITRF.<br />

Repeated campaign observations allow the<br />

determination of the displacement vector as a<br />

function of time. The estimation of velocities and the<br />

corresponding errors was carried out on a statistical<br />

basis, by analysis of time series of each individual<br />

component of motion, by least square adjustment.<br />

Uncertainties were determined using the average<br />

scatter of residuals of the linear regression, providing<br />

more realistic error estimates. The estimated GPS<br />

velocity field is presented in a local DION-fixed<br />

reference frame in order to allow the recognition of<br />

local scale displacement patterns (Fig. 2). Site<br />

velocities from previous geodetic studies (E067 &<br />

G20) as well as EUREF solutions (NOA1) were also<br />

considered for the sake of completeness.<br />

35


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

STRAIN RATES<br />

In order to provide results independent from the<br />

choice of the reference frame, strain analysis was<br />

performed by the grid_strain Matlab TM software<br />

package (Teza et al., 2008). It allows the definition of<br />

the deformation pattern by providing the intensity and<br />

direction of principal components of strain tensor<br />

together with corresponding errors, by means of a<br />

modified linear least-squares (LS) inversion, under<br />

the hypothesis of uniform strain field condition. Inputs<br />

for calculating strain were horizontal GPS annual<br />

velocities and their corresponding errors. In this<br />

sense results express the linear strain rates in the<br />

region.<br />

For the purpose of the analysis, GPS sites located on<br />

the mountains bordering the Athens Basin,<br />

specifically on the metamorphic basement of Pendeli<br />

and Hymettus mountains to the East (APR, ARG,<br />

HYM, NER and TAT) and on the non-metamorphic<br />

formations of eastern Parnitha Mt. and Aegaleo Mt.<br />

(E067, CHS, KOR, PKL and PRM) were selected.<br />

The analysis involved initially the calculation of a<br />

single strain tensor based on all selected stations<br />

and then, by gradual segmentation of the area for a<br />

more detailed investigation of spatial variations of the<br />

deformation regime. All calculations are referred to<br />

the center of mass of each set of sites considered.<br />

Fig. 3: Principle axes of the strain rate tensor for the area of<br />

interest, calculated from velocities of selected GPS sites, in<br />

background contour lines of 20m interval.<br />

From single strain tensor calculations, an extension<br />

of 0.27 0.06 strain/yr along a NNW–SSE direction<br />

(N 347) is shown, with a negative eigenvalue<br />

(compression) for the minimum principal axes (Fig.<br />

3). It is nevertheless evident that a single strain<br />

tensor is insufficient to express adequately the<br />

apparent heterogeneity of the local displacement<br />

field. Further examination of the strain field (Fig. 4)<br />

indicate negligible compressional rates at the<br />

southern part of the basin compared to the dominant<br />

extensional regime of relatively higher strain rates<br />

(0.91 0.09 strain/yr) at the northern part between<br />

Pendeli and Parnitha ranges.<br />

A more detailed consideration of the strain field<br />

between the two geotectonic units was performed by<br />

triangulation of the selected GPS sites (Fig. 5).<br />

Herein, it is interesting to note the major<br />

differentiation between the western and the eastern<br />

parts of Athens Basin with significantly lower strain<br />

rates in the latter. Moreover, the gradual increase of<br />

the extension rates at the western part of Athens<br />

plain moving to the North is clearly depicted, while a<br />

counterclockwise rotation of the maximum principle<br />

axis of the strain tensor is also observed. The<br />

compressional regime at the southeastern part of the<br />

basin should be underlined.<br />

DISCUSSIONS<br />

The stress field in the region is characterized by<br />

extension in a NNE–SSW direction, also confirmed<br />

by regional geodetic measurements (Veis et al.,<br />

2003). However, the 5-km spacing of the geodetic<br />

network allowed investigating local variations of the<br />

strain rates.<br />

Fig. 4: Principle axes of the strain rate tensors within Athens<br />

Basin. Dashed lines indicate local estimates around which<br />

GPS data are poorly distributed from a geometrical point of<br />

view.<br />

A differentiation of strain rates across the inactive<br />

tectonic boundary is evident with significantly higher<br />

rates at the western part of Athens Basin. Given its<br />

inactive characteristics, a passive control should be<br />

considered. Such behavior has also been mentioned<br />

during the Athens 1999 earthquake from SAR<br />

interferometric observations of the spatial expansion<br />

of the co- and post-seismic displacement field<br />

(Foumelis et al., 2009).<br />

36


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 5: Detailed strain analysis by different triangulations of<br />

selected GPS sites. Principle axes of the strain rate tensors<br />

are calculated at the center of mass of each triangle.<br />

The broader area is essentially a transitional zone<br />

between the Corinth Gulf and Beotia to the west,<br />

characterized by E-W trending active faults with<br />

significant seismic activity and those of southern<br />

Attica and Cyclades islands to the east, showing low<br />

deformation rates (Mariolakos & Papanikolaou, 1987;<br />

Papanicolaou & Lozios, 1990). Thus, the observed<br />

high strain rates at the northern part of the basin<br />

should be attributed to the high crustal velocities<br />

observed within Parnitha Mt. an area controlled<br />

mainly by E-W trending active fault zones (Ganas et<br />

al. 2005; Papanikolaou & Papanikolaou, 2007)<br />

although the role of NE-SW trending faults should be<br />

important as well (Mariolakos & Fountoulis, 2000).<br />

Acknowledgements: The authors would like to<br />

acknowledge Prof. E. Lagios for his support and supervision<br />

during GPS measurements, Prof. D. Paradissis for his<br />

suggestions, Dr. V. Sakkas for processing of UOA1 data<br />

and METRICA company for MET0 station data provision.<br />

References<br />

Beutler, G., H. Bock, E. Brockmann, R. Dach, P. Fridez, W.<br />

Gurtner, U. Hugenntobler, D. Ineichen, J. Johnson, M.<br />

Meindl, L. Mervart, M. Rothacher, S. Schaer, T. Springer<br />

& R. Weber, (2001). Bernese GPS Software Version 4.2.<br />

U. Hugentobler, S. Schaer & P. Fridez (eds.),<br />

Astronomical Institute, University of Berne, Switzerland,<br />

515p.<br />

Clark, P.J., R.R. Davies, P.C. England, B. Parsons, H.<br />

Billiris, D. Paradissis, G. Veis, P.A. Cross, P.H. Denys, V.<br />

Ashkenazi, R. Bingley, H.-G. Kahle, M.V. Müller, & P.<br />

Briole, (1998). Crustal strain in central Greece from<br />

repeated GPS measurements in the interval 1989-1997.<br />

Geophysical Journal International 135 (1), 195-214.<br />

Drakatos, G., V. Karastathis, J. Makris, J. Papoulia & G.<br />

Stavrakakis, (2005). 3D crustal structure in the<br />

neotectonic basin of the Gulf of Saronikos (Greece).<br />

Tectonophysics 400, 55-65.<br />

Foumelis, M., (2009). Study of crustal deformation in the<br />

broader area of Athens by means of Differential GPS and<br />

SAR Interferometry. Dissertation thesis, Department of<br />

Geology and Geoenvironment, University of Athens,<br />

358p (in Greek).<br />

Foumelis, M., I. Parcharidis, E. Lagios & N. Voulgaris,<br />

(2009). Evolution of post-seismic ground deformation of<br />

Athens 1999 earthquake observed by SAR<br />

interferometry. Journal of Applied Geophysics 69, 16-23.<br />

Ganas, A., S. Pavlides & V. Karastathis, (2005). DEMbased<br />

morphometry of range-front escarpments in Attica,<br />

central Greece and its relation to fault slip rates.<br />

Geomorphology 65, 301-319.<br />

Krohe, A., E. Mposkos, A. Diamantopoulos & G. Kaouras,<br />

(2009). Formation of basins and mountain ranges in<br />

Attica (Greece): The role of Miocene to recent low-angle<br />

normal detachment faults. Earth-Science Reviews 98 (1-<br />

2), 81-104.<br />

Lozios, S., (1993). Tectonic analyisis of the metamorphic<br />

rocks in NE Attica (in Greek). Dissertation thesis,<br />

Department of Geology, University of Athens, 299p (in<br />

Greek).<br />

Mariolakos, I. & I. Fountoulis, (2000). The Athens<br />

earthquake September 7, 1999 neotectonic regime and<br />

geodynamic phenomena. Ann. Geol. Pays Hellen. 38 (B),<br />

165-174.<br />

Mariolakos, I. & D. Papanikolaou, (1987). Deformation<br />

pattern and relation between deformation and seismicity<br />

in the Hellenic Arc. Bull. Geol. Soc. Greece 19, 59-76 (in<br />

Greek).<br />

Papadopoulos T.D., N. Goulty, N.S. Voulgaris, J.D.<br />

Alexopoulos, I. Fountoulis, P. Kambouris, V. Karastathis,<br />

C. Peirce, S. Chailas, J. Kassaras, M. Pirli, G. Goumas &<br />

E. Lagios, (2007). Tectonic structure of Central-Western<br />

Attica (Greece) based on geophysical investigations -<br />

Preliminary Results. Bull. Geol. Soc. Greece 40 (3),<br />

1207-1218.<br />

Papanikolaou, D., E. Bassi, H. Kranis & G. Danamos,<br />

(2004). Paleogeographic evolution of the Athens basin<br />

from upper Miocene to Present. Bull. Geol. Soc. Greece<br />

36 (2), 816-825 (in Greek).<br />

Papanikolaou, D., E. Lekkas, Ch. Sideris, I. Fountoulis, G.<br />

Danamos, Ch. Kranis, S. Lozios, I. Antoniou, E.<br />

Vassilakis, S. Vasilopoulou, P. Nomikou, I. Papanikolaou,<br />

E. Skourtsos & K. Soukis, (1999). Geology and tectonics<br />

of Western Attica in relation to the 7-9-99 earthquake.<br />

Newsletter of E.C.P.F.E., issue no. 3, 30-34.<br />

Papanikolaou, D. & S. Lozios, (1990). Comparative<br />

neotectonic structure of high (Korinthia-Beotia) and low<br />

rate (Attica-Cyclades) activity. Bull. Geol. Soc. Greece<br />

16, 47-65 (in Greek).<br />

Papanikolaou, D. & I. Papanikolaou, (2007). Geological,<br />

geomorphological and tectonic structure of NE Attica and<br />

seismic hazard implications for the northern edge of<br />

Athens plain. Bull. Geol. Soc. Greece 40 (1), 425-438.<br />

Papanikolaou, D. & L. Royden, (2007). Disruption of the<br />

Hellenic Arc: Late Miocene extensional detachment faults<br />

and steep Pliocene-Quaternary normal faults – or – What<br />

happened to Corinth ? Tectonics 26, TC5003,<br />

doi:10.1029/2006TC002007.<br />

Teza, G., A. Pesci & A. Galgaro, (2008). Grid_strain and<br />

grid_strain3: Software packages for strain field<br />

computation in 2D and 3D environments. Computers &<br />

Geosciences 34, 1142-1153.<br />

Xypolias, P., S. Kokkalas & K. Skourlis, (2003). Upward<br />

extrusion and subsequent transpression as a possible<br />

mechanism for the exhumation of HP/LT rocks in Evia<br />

island ( Aegean Sea , Greece ). Journal of Geodynamics<br />

35, 303-332.<br />

Veis et al. (2003). Tectonic displacements along the<br />

Aegean and the Alkyonides- Perachora-Parnitha triangle.<br />

Applied Research Project, Earthquake Planning and<br />

Protection Organization. Athens, 74p (in Greek).<br />

37


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

NEOTECTONICS AND COMPARISON OF THE ENVIRONMENTAL SEISMIC INTENSITY<br />

SCALE (ESI 2007) AND THE TRADITIONAL SCALES FOR EARTHQUAKE INTENSITIES<br />

FOR THE KALAMATA (SW GREECE) EARTHQUAKE (MS=6.2R, 13-09-1986)<br />

Fountoulis, Ioannis (1, Mavroulis, Spyridon (1)<br />

(1) National and Kapodistrian University of Athens, Faculty of Geology and Geoenvironment, Department of Dynamic Tectonic<br />

and Applied Geology, Panepistimiopoli, Zografou 15784 Athens, GREECE. Email: fountoulis@geol.uoa.gr , smavroulis@yahoo.gr<br />

Abstract: The Kalamata (13-09-1986, Ms=6.0R, SW Peloponnese) earthquake can be classified as a medium to small scale<br />

event based on the tectonic structures that triggered the earthquake and the effects caused on human, structural and natural<br />

environment. The aim of this paper is to present the geotectonic and seismotectonic regime of the earthquake affected region<br />

based on field data along the seismic fault zone and an attempt is made towards the: (i) estimation of the intensity values<br />

according to the European Macroseismic Scale (EMS 1998) and the Environmental Seismic Intensity Scale (ESI 2007) and the<br />

determination of their geographical distribution in a macroscale, (ii) interpretation of the intensity values data and their distribution<br />

according to the seismotectonic, geodynamic and geotechnical regime, and (iii) conduction of a comparative evaluation review on<br />

the application of both EMS 1998 and ESI 2007. The application of both EMS 1998 and ESI 2007 and the comparative evaluation<br />

of the results indicate that the estimated values of EMS 1998 and ESI 2007 were almost in agreement, despite the fact that the<br />

geographical locations of assessment data were different suggesting that the application and use of both scales appears to<br />

represent a useful and reliable tool for seismic hazard estimation.<br />

Key words: Kalamata, earthquake, neotectonics, environmental effects<br />

INTRODUCTION<br />

Kalamata is located very close (< 70km) to the<br />

Hellenic (Ionian) Trench region in which the<br />

subduction zone of the African plate beneath the<br />

European (Aegean) one exists and thus is one of the<br />

most seismically active areas of Europe (Figure 1).<br />

On 13 September 1986, a shallow depth (< 10km)<br />

earthquake struck the wider Kalamata area resulting<br />

in 20 casualties, extensive damages and many<br />

environmental effects. The epicenter of the main<br />

earthquake was located about 10km NNE of the city<br />

of Kalamata, and its magnitude was Ms=6.2<br />

(Papazachos, et al., 1988). Two days later, a second<br />

shock of Ms=5.4R (Papazachos, et al., 1988)<br />

occurred closer to the Kalamata city at the same<br />

depth. The focal mechanism of the main shock<br />

shows an E-W extension (Lyon-Caen et al., 1987;<br />

Papazachos, et al., 1988).<br />

Seismological studies of Papazachos et al. (1988),<br />

and Lyon-Caen et al. (1988) indicated that<br />

aftershocks defined two clusters and an about 450<br />

west-dipping fault plane. The foci depths of the<br />

seismic sequence were ranging between 11 and<br />

0.9km. Based on the variety of orientations and dips<br />

calculated for the sub-faults activated during the<br />

aftershock sequence, since the analysis of the<br />

northern cluster indicates the existence of two types<br />

of orientation, which are dipping in four different<br />

angles and the southern cluster is characterized by<br />

an almost uniform behaviour activated later in the<br />

sequence, Tselentis et al. (1989) concluded that the<br />

area is tectonically very complex which is in<br />

agreement with the neotectonic structure described<br />

by Mariolakos et al. (1989; 1992, 1993) and<br />

Mariolakos & Fountoulis (1998).<br />

Stiros and Kontogianni (2008) applied two first-order<br />

leveling traverses crossing the wider Kalamata area<br />

and measured subsidence of about 7cm NE of the<br />

Kalamata city in epicentral area of the southern<br />

cluster. The Kalamata earthquake produced a<br />

maximum intensity VIII+ on the IMM or EMS 1992<br />

scale (Elnashai et al., 1987; Gazetas et al., 1990),<br />

while Panou et al. (2004) based on building damages<br />

estimated the intensity up to IX - X for the city of<br />

Kalamata.<br />

The aim of this paper is to present the geotectonic<br />

and seismotectonic regime of the earthquake<br />

affected region based on field data along the seismic<br />

fault zone and an attempt is made towards the: (i)<br />

estimation of the intensity values in terms of the<br />

European Macroseimic Scale (EMS 1998; Grünthal,<br />

1998) and Environmental Seismic Intensity Scale<br />

(ESI 2007; Michetti et al., 2007) and the<br />

determination of their geographical distribution in a<br />

macroscale, (ii) interpretation of the intensity values<br />

data and their distribution according to the<br />

seismotectonic, geodynamic and geotechnical<br />

regime, and (iii) conduction of a comparative<br />

evaluation review on the application of both EMS<br />

1998 and ESI 2007.<br />

GEOLOGY - TECTONICS - NEOTECTONICS -<br />

FAULT ZONES - FAULTS<br />

In the broader Kalamata area the following four<br />

alpine geotectonic units from the lower to the upper<br />

occur (Psonis, 1986; Mariolakos et al., 1993): (a) the<br />

38


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Mani unit consisting mainly of marbles, (b) the Arna<br />

unit consisting of quartzites and phyllites, (c) the<br />

Tripolis unit which consists of neritic carbonates and<br />

flysch formation and (d) the Pindos unit consisting of<br />

thin-bedded pelagic carbonates and clastic<br />

formations. From the structural point of view, the four<br />

above-mentioned geotectonic units form a<br />

succession of three nappes. The Mani unit (slightly<br />

metamorphosed) is considered to be the relatively<br />

autochthonous one. The Arna unit overthrusts the<br />

Mani unit, the Tripolis unit (second nappe)<br />

overthrusts the Arna unit and the Pindos unit (third<br />

nappe) overthrusts the Tripolis unit (Figure 2). The<br />

Late Pliocene-Early Pleistocene marine deposits<br />

consist of marls, sandstone and conglomerates<br />

(Marcopoulou-Diacantoni et al. 1989; Mariolakos et<br />

al., 1993). The Middle-Late Pleistocene deposits<br />

consist mainly of red colored siliceous sandssandstones<br />

and conglomerates. Alluvial deposits,<br />

clastic material and talus represent the Holocene.<br />

were observed within this graben. The marginal fault<br />

zones consist of many faults, which are not<br />

continuous and differ on strike even when they<br />

belong to the same fault zone, as they form<br />

conjugate fault systems.<br />

Fig. 2: Simplified geological map showing the four alpine<br />

geotectonic units overthrust one on top of the other, as well<br />

as the post-alpine sediments of the region of the Kalamata<br />

area. 1: Holocene deposits; 2: Continental deposits; 3: Early<br />

Pleistocene marine deposits; 4: Pindos unit; 5: Gavrovo-<br />

Tripolis unit; 6: Arna unit; 7: Mani unit; 8: Overthrust; 9:<br />

Fault; 10: Detachment fault. The numbers in the black<br />

circles correspond to the smaller order neotectonic<br />

macrostructures of the Kato Messinia sub-graben 1:<br />

Asprochoma-Koutalas horst, 2: Dimiova-Perivolakia graben,<br />

3: Kalathion Mt. horst, 4: Altomyra semi-graben, 5: Kambos<br />

graben, 6: Vardia-Koka horst, 7: Kitries-Mantinia subgraben,<br />

XFZ:Xerilas Fault zone, NFZ:Nedon Fault Zone.<br />

Fig. 1: The second order neotectonic macrostructures within<br />

the first order neotectonic macrostructure of the Kalamata-<br />

Kyparissia graben. The numbers correspond to the<br />

following second order neotectonic macrostructures: 1: Kato<br />

Messinia graben, 2: Meligalas horst, 3: Ano Messinia<br />

graben, 4: Dorion basin, 5: Kyparissia-Kalo Nero graben<br />

The meizoseismal area is located at the eastern<br />

margin of the Kalamata - Kyparissia graben and<br />

constitutes the northward prolongation of the Gulf of<br />

Messinia (Figure 1). Large and composite fault zones<br />

define its margins and second order macrostructures<br />

are observed within as well as at the margins<br />

representing smaller grabens and horsts (Figure 1)<br />

(Mariolakos & Fountoulis, 1998). The E-W striking<br />

Dimiova - Perivolakia graben is bounded by the Kato<br />

Karveli - Venitsa fault zone to the north, by the<br />

Arahova to the east, by the Xerilas fault zone (XFZ)<br />

to the south and by the Nedon fault zone (NFZ) to the<br />

west (Figure 2). This macrostructure constitutes one<br />

of the most interesting minor order neotectonic<br />

macrostructures because of the occurrence of the<br />

Pindos unit. Mariolakos et al. (1989) interpreted the<br />

kinematic regime of this macrostructure suggesting<br />

that this graben rotates around an N-S axis located at<br />

the area of Arahova westwards. At the western part<br />

of the fault zone the total throw is more than 2.000m<br />

(Mariolakos et al., 1986; Mariolakos et al., 1989). The<br />

most of the environmental effects and damages<br />

caused during the seismic activity of September 1986<br />

SPATIAL DISTRIBUTION OF ENVIRONMENTAL<br />

EFFECTS<br />

During the above-mentioned seismic activity, fault<br />

reactivation (seismic faults), new faulting and seismic<br />

fracturing were observed (the latter distinguished by<br />

no displacement) (Figure 3). The reactivated faults<br />

strike in different directions (N-S, E-W, NNE-SSW)<br />

and the throw of the faults due to the reactivation is<br />

generally small (max=20cm) and of normal character.<br />

The maximum throw has been observed at a seismic<br />

fault caused by the main aftershock Ms=5.6 R.<br />

Numerous seismic ruptures trending N-S, NNE-SSW,<br />

NE-SW, E-W and NW-SE were mapped in the<br />

affected area, in most cases in en echelon<br />

arrangement (Mariolakos et al., 1989). These seismic<br />

fractures presented a vertical displacement of several<br />

mm up to 25-30cm and they often presented a<br />

horizontal component showing sinistral or dextral<br />

displacement.<br />

The majority of rock falls were observed in several<br />

sections along the slopes of the Tzirorema, Karveli<br />

and Xerilas streams and the Nedon river valleys as<br />

well as in the wider area of Eleochori, Karveli and<br />

Ladas villages (Figure 3). They were observed in<br />

areas characterized by steep slopes (> 50 per cent)<br />

and they were related almost everywhere to small or<br />

39


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

large faults with some of them reactivated during the<br />

earthquake and others not.<br />

because those areas belong to different neotectonic<br />

macrosturctures that were not reactivated during the<br />

earthquakes of 1986 (central region of the tectonic<br />

basin of Kato Messinia and tectonic horst of Kalathio<br />

Mt respectively).<br />

Fig. 3: The spatial distribution of the environmental effects<br />

observed during the Kalamata earthquake sequence (based<br />

on data from Mariolakos et al., 1992; Gazetas et al., 1990;<br />

Fountoulis, 2004; Stiros & Kontogianni, 2008).<br />

SPATIAL DISTRIBUTION OF DAMAGES<br />

The damages were limited to an area of triangular<br />

shape, which is defined to the south by the fault zone<br />

of the Xerilas River, to the east by the fault zone of<br />

Nedousa - Arahova and to the west by the fault zone<br />

of the Nedon River (Figure 3). No damages were<br />

recorded to the west of the Nedon fault zone and<br />

south of the Xerilas fault zone and especially in areas<br />

where the geological basement has the same<br />

seismo-geological behavior as those in the city of<br />

Kalamata and Eleohori village, which caused serious<br />

damage. Based on field observations, the damage is<br />

not determined only by the age, type, height and<br />

other characteristics of buildings. There were cases<br />

with two nearly identical constructions in the same<br />

area; one remained intact while the other was<br />

destroyed. In other cases the building destruction is<br />

linked to zones of seismic fracturing that were<br />

observed in the construction basement. Of course,<br />

this is not the rule. In many other cases the building<br />

destruction is linked to zones of seismic fracturing<br />

that were observed in the construction basement. Of<br />

course, this is not the rule.<br />

Fig. 4: (A) EMS 1998 intensity distribution of the Kalamata<br />

earthquake sequence (based on data from Gazetas et al.,<br />

1990; Panou et al., 2004). (B) EMS 1998 intensity<br />

distribution of the Kalamata earthquake sequence for<br />

Kalamata city (based on data from Panou et al., 2004).<br />

CONCLUSIONS<br />

Taking into account the aforementioned we can draw<br />

the following conclusions:<br />

The damages were limited to the area that can be<br />

regarded as a transitional area between the tectonic<br />

basin Kalamata - Kyparissia and the tectonic horsts<br />

of Asprohoma - Koutala to the north and the Kalathio<br />

Mt. to the south. On the contrary, in Messini and in<br />

Verga, damages of that size were not observed<br />

Figure 5: ESI 2007 Intensity distribution based on data of<br />

Figure 3.<br />

Rock falls were observed mainly in the tectonic<br />

graben that was activated and also north of it, at<br />

Tzirorema. On the other hand, on the steep slopes of<br />

the Kalathio Mt. that belong to the homonymous<br />

40


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

neotectonic macrostructure, which was not<br />

reactivated, no rock falls were observed.<br />

An important factor in the distribution of the damages<br />

and rock falls in the greater area was the reactivation<br />

of old faults or the creation of new soil ruptures. In<br />

this way, the fact that the destruction of buildings was<br />

observed in Giannitsanika (higher intensity in Figure<br />

4) and not near the coast can be explained, although<br />

the foundation ground - red siliceous clastic formation<br />

- in the first case theoretically presents better<br />

geotechnical characteristics in comparison to the<br />

loose coastal deposits.<br />

The ESI 2007 scale appears to fit better than the<br />

EMS scale in the neotectonic regime of the area as<br />

its boundaries coincide better with the boundaries of<br />

the activated graben and the observations we have<br />

done concerning the distribution of the environmental<br />

effects. The application of both EMS 1998 and ESI<br />

2007 and the comparative evaluation of the results<br />

indicate that the estimated values of EMS 1998 and<br />

ESI 2007 were almost in agreement, despite the fact<br />

that the geographical locations of assessment data<br />

were different suggesting that the application and use<br />

of both scales appears to represent a useful and<br />

reliable tool for seismic hazard estimation.<br />

References<br />

Elnashai, A., Pilakoutas, K., Ambraseys, N. & Lefas, I.<br />

(1987). Lessons learnt from the Kalamata (Greece)<br />

earthquake of 13 September 1986. European Earthquake<br />

Engineering, 1, 11-19.<br />

Fountoulis, I. 2004. The neotectonic macrostructures and<br />

the geological basement, the main factors controlling the<br />

spatial distribution of the damage and geodynamic<br />

phenomena resulting from the Kalamata (13 September<br />

1986) and Athens (7 September 1999) earthquakes. WIT<br />

Press: In Earthquake Geodynamics: Seismic Case<br />

Studies. Edited by: E. L. Lekkas, Series: Advances in<br />

Earthquake Engineering, vol. 12, p. 45-63.<br />

Gazetas, G., Dakoulas, P. & Papageorgiou, A. (1990). Local<br />

soil and source-mechanism effects in the 1986 kalamata<br />

(Greece) earthquake. Earthquake Engineering and<br />

Structural Dynamics, vol. 19, 431-456.<br />

Grünthal, G. 1998. (ed.): European Macroseismic Scale<br />

1998 (EMS-98). Cahiers du Centre Européen de<br />

Géodynamique et de Séismologie 15, Centre Européen<br />

de Géodynamique et de Séismologie, Luxembourg, 99<br />

pp.<br />

Lyon-Caen, H., Armijo, R., Drakopoulos, J., Baskoutas, J.,<br />

Delibassis, N., Gaulon, R., Kouskouna, V., Latoussakis,<br />

J., Makropoulos, K., Papadimitriou, P., Papanastassiou,<br />

D. & Pedotti, G. (1988). The 1986 Kalamata (South<br />

Peloponnesus) earthquake: Detailed study of a normal<br />

fault, evidences for east-west extension in the Hellenic<br />

arc, J. Geophys. Res. 93, 14.967–15.000.<br />

Marcopoulou-Diacantoni, A., Mirkou, M.R., Mariolakos, I.,<br />

Logos, E., Lozios, S. & Fountoulis, I. (1989). Stratigraphic<br />

observations in the post alpine deposits in Thouria-Ano<br />

Amfia area (Messinia Province, Greece) and their<br />

neotectonic interpretation. Bull. Geol Soc. Greece,<br />

XXIII/3, pp. 275-295, (in Greek).<br />

Mariolakos, I., Sabot, V., Alexopoulos, A., Danamos, G.,<br />

Lekkas, E., Logos, E., Lozios, S., Mertzanis, A. &<br />

Fountoulis, I. (1986). Microzonic study of Kalamata<br />

(Geology, Tectonics, Neotectonics, Geomorphology).<br />

Report of EPPO (Earth Planning Protection<br />

Organization), (in Greek).<br />

Mariolakos, I., Fountoulis, I., Logos, E. & Lozios, S. (1989).<br />

Surface faulting caused by the Kalamata (Greece)<br />

earthquakes (13.9.1986). Tectonophysics, 163, pp. 197-<br />

203.<br />

Mariolakos, I., Fountoulis, I. & Nassopoulou, S. (1992). The<br />

influence of the neotectonic macrostructures, fractures<br />

and the geological basement in the distribution of the<br />

damages in the Kalamata earthquake (13-9-1986). Proc.<br />

1st Greek Congress on Antiseismic Engineering and<br />

Technical Seismology Technical Chamber of Greece, 1,<br />

pp. 55-68, (in Greek).<br />

Mariolakos, I., Schneider, H., Fountoulis, I. & Vouloumanos,<br />

N. (1993). Paleogeography, sedimentation and<br />

neotectonic implications at the Kambos depression and<br />

Kitries Bay area (Messinia, Peloponnese, Greece). Bull<br />

Geol. Soc. Greece, VIII/1, pp. 397-413, 1993.<br />

Mariolakos, I. & Fountoulis, I. (1998). Is it safe to built on<br />

fault surfaces in a seismically active area? Proc. 8th<br />

International IAEG Congress, pp. 665-670, Balkema.<br />

Michetti A.M., Esposito, E., Guerrieri, L., Porfido, S., Serva,<br />

L., Tatevossian, R., Vittori, E., Audemard, F., Azuma, T.,<br />

Clague, J., Comerci, V., Gürpinar, A., Mccalpin, J.,<br />

Mohammadioun, B., Mörner, N.A., Ota, Y. & Roghozin, E.<br />

(2007). Environmental Seismic Intensity Scale 2007 - ESI<br />

2007. In: Guerrieri L. and Vittori E. (Eds.), Memorie<br />

Descrittive della Carta Geologica d’Italia, 74, 7-54,<br />

Servizio Geologico d’Italia – Dipartimento Difesa del<br />

Suolo, APAT, Roma, Italy.<br />

Panou, A., Theodoulidis, N. & Savvaidis, A. (2004).<br />

Correlation between (H/V) ratio and seismic damage: The<br />

case of the city of Kalamata. In SESAME project, Report<br />

of the WP04, H/V Technique: Empirical Evaluation,<br />

Project No. EVG1-CT-2000-00026.<br />

Papazachos, V., Kiratzi, A., Karacostas, B.,<br />

Panagiotopoulos, D., Scordilis, E. & Mountrakis, D.<br />

(1988). Surface fault traces, fault plane solution and<br />

spatial distribution of the aftershocks of the September<br />

13, 1986 earthquake of Kalamata (Southern Greece).<br />

PAGEOPH, vol. 126, No. 1, 55-68.<br />

Psonis, K. 1986. Geological map of Greece, Kalamata<br />

sheet, scale 1/50,000, Geological Survey of Greece:<br />

Athens.<br />

Stiros, S. & Kontogianni, V. (2008). Modelling of the<br />

Kalamata (SW Greece) earthquake faulting using<br />

geodetic data. Journal of Applied Geodesy, 2, 179-185;<br />

DOI 10.1515/JAG.2008.020<br />

41


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

QUANTIFICATION OF RIVER VALLEY MAJOR DIVERSION IMPACT AT KYLLINI<br />

COASTAL AREA (W. PELOPONNESUS, GREECE) WITH REMOTE SENSING<br />

TECHNIQUES<br />

Fountoulis Ioannis D. (1), Vassilakis Emmanuel (1), Mavroulis Spyridon (1), Alexopoulos John (2), Erkeki Athanasia (3)<br />

(1) School of Geology and Geoenvironment, Department of Dynamics, Tectonics and Applied Geology, National and Kapodistrian<br />

University of Athens, 15784, Greece. Email: evasilak@geol.uoa.gr<br />

(2) School of Geology and Geoenvironment, Department of Geophysics and Geothermy, National and Kapodistrian University of<br />

Athens, 15784, Greece.<br />

(3) School of Geology and Geoenvironment, Laboratory of Natural Hazards, National and Kapodistrian University of Athens,<br />

15784, Greece.<br />

Abstract: The effects of the geological, tectonic and neotectonic structure and the impact of the human presence and activity on<br />

the drainage network of Pineios river are presented here in order to determine the causes of its diversion and the implications to<br />

the shoreline. We used, analyzed and evaluated (a) geomorphological, geological, tectonic and neotectonic data of the study area,<br />

(b) historical information and archaeological findings from buried and eroded archaeological sites of the wider study area, (c)<br />

published data related to drill cores and radiocarbon dates, and (d) remote sensing datasets, as satellite and aerial photos of<br />

different capturing periods, as well as real-time kinematic differential GPS measurements for the definition of the current shoreline.<br />

It is concluded that the detected shoreline displacements and drainage diversions are the result of the combination of active<br />

tectonics and human activity during the last 100 kyrs.<br />

Key words: Kyllini peninsula, Pineios river, RTK DGPS, river evolution, coastal erosion<br />

INTRODUCTION<br />

The Pineios River development and history takes<br />

place in one of the most tectonically and seismically<br />

active areas in Greece. The intense and continuous<br />

tectonic activity in the area is highly related to its<br />

location on the external part of the Hellenic arc and<br />

adjacent to the convergent boundary where African<br />

plate is subducted beneath the Aegean as well as the<br />

diapirism of near surface evaporitic domes. The<br />

highest seismicity levels recorded in the area<br />

(Hatzfeld et al., 1990) as well as the generation of<br />

many historic strong earthquakes confirm the<br />

neotectonic observations, which show that the area is<br />

undergoing a complicated tectonic deformation.<br />

d. 0,16 to 0,48 mm/yr for the eastern (inland) part of<br />

Kyllini peninsula (125 kyrs)<br />

The most important fault zones in the study area are<br />

the Panopoulo fault zone (Panopoulo FZ), Pineios<br />

fault zone (Pineios FZ) and the strike - slip fault zone<br />

that gave rise to the Andravida earthquake (08-06-<br />

2008, ML=6,5). These major faults form several<br />

neotectonic blocks in the study area including the<br />

Gastouni graben (hangingwall of Pineios fault zone),<br />

the uplifted area of Varda (footwall of Pineios fault<br />

zone) and the Kyllini horst (Fig. 1).<br />

GEOCHRONOLOGICAL INTERPERETATION<br />

In order to determine the effect of the ongoing active<br />

tectonics on the Pineios River diversion during the<br />

late 18 th or the early 19 th century, we calculated<br />

relative uplift rates for several sites of the study area<br />

based on<br />

230 Th/ 238 U dating of corals made by<br />

Stamatopoulos et al. (1988) and dating of marine<br />

deposits in Kyllini peninsula estimated by Mariolakos<br />

et al. (1988):<br />

a. 0,39 mm/yr for Psari area (103 kyrs)<br />

b. 0,50 mm/yr for Neapoli area (118 kyrs)<br />

c. 0,67 mm/yr for Aletreika area (209 kyrs)<br />

Fig. 1: Sketch map of the contemporary Pineios river deltaic<br />

area (at the hanging wall of the Pineios fault) and the former<br />

deltaic area at the footwall of the same fault. The estimated<br />

shorelines for the Roman and Neolithic periods are shown.<br />

The archaeological sites and the sampling sites of the<br />

geochronological analysis are also noted, along with the<br />

calculated uplift rates for the last 100kyrs.<br />

42


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The general conclusions after the interpretation of the<br />

geochronological data are:<br />

a. The maximum relative uplift rate (0,67 mm/yr)<br />

characterizes an area (Aletreika) located on the<br />

footwall side and very close to the Pineios FZ.<br />

b. The relative uplift rate of the Gastouni graben<br />

(hangingwall of Pineios FZ, 0,19 mm/yr) is less<br />

even than the lowest value of the Pineios FZ<br />

footwall relative uplift rate (0,39 mm/yr).<br />

c. The northeastern part of Kyllini peninsula has<br />

higher relative uplift rate (0,48 mm/yr) than the<br />

southeastern part (0,30 mm/yr)<br />

d. The maximum relative uplift rate of the footwall of<br />

Pineios FZ is significantly higher than the<br />

maximum relative rates of the eastern part of<br />

Kyllini peninsula and the Gastouni graben.<br />

Lagoon recorded by Kontopoulos and Koutsios<br />

(2010) we note that:<br />

a. the shoreline in the Pineios delta advanced by 3,5<br />

km into the sea in the 6.350 yrs period from<br />

Neolithic (8.500 yrs BP) to Roman (2.150 yrs BP)<br />

period, which shows a coastal progradation rate<br />

of the order of 0,55 m/yr, and<br />

b. the shoreline in the Pineios delta retreated by<br />

1,75 km in the 2.150 yrs period from Roman<br />

period (2.150 yrs BP) to present which shows a<br />

retrogradation rate as 0,81 m/yr from Roman<br />

period to present.<br />

HISTORICAL COASTLINE DATA<br />

It is more than obvious that the major percentage of<br />

the coastline displacements in the study area, during<br />

the last 8kyrs, are related to active structures and the<br />

tectonic instability as this affects the alongshore<br />

redistribution of sediments from the Pineios delta.<br />

After the organization of all the available geological<br />

and historical data we were able to estimate and<br />

reconstruct the paleo-coastline in several periods for<br />

the last 100kyrs (Fig. 2). It is quite easy to accept that<br />

during Tyrrhenian most of the area of Kyllini was<br />

under the water since the marine sediments were<br />

deposited.<br />

The palaeo-delta of Pineios River was developed N<br />

of Kyllini peninsula before and during Neolithic<br />

period. The Neolithic and Helladic shoreline was<br />

located 3,5 km onshore from the present shoreline.<br />

During the Roman period, Pineios River flowed<br />

directly S of the Kotychi lagoon forming a levee,<br />

which is now abandoned, eroded and stands as a<br />

low sea cliff. An acceleration of coastal deposition<br />

and consequently delta propagation took place. The<br />

Roman shoreline was 1,5 km seaward from the<br />

present shoreline. During the Othoman period,<br />

Pineios occupied the channel 5 km S of Kotychi<br />

lagoon forming another levee standing well above the<br />

floodplain at the shoreline and indicating coastal<br />

retreat. This channel is in the process of filling. The<br />

minimum age of this levee is about 200 yrs BP.<br />

The Pineios diversion to the south of Kyllini peninsula<br />

took place during the late 18 th century. Following this<br />

diversion, the pre-18 th -century-A.D. Pineios River<br />

delta shoreline in now undergoing marine<br />

transgression and intense coastal erosion, as is to be<br />

expected in a former delta now essentially starved of<br />

new sediment. The pre-18 th -century-A.D. northern<br />

channels of Pineios River and few smaller streams<br />

can still be seen in their courses to the northwest,<br />

now dry. The dominant geomorphic processes in the<br />

modern delta of Pineios River are progradation and<br />

aggradation with large volumes of river sediment.<br />

Based on the palaeogeographic reconstructions<br />

developed by Kraft et al. (2005) and the late<br />

Holocene environmental changes from Kotychi<br />

Fig. 2: Shoreline displacements in the study area during the<br />

last 8 kyrs.<br />

REMOTE SENSING CONTRIBUTION<br />

In order to determine whether or not progradation or<br />

retrogradation took place in Pineios former and<br />

current deltas in recent years, we initially mapped the<br />

shorelines at different times in the 27-year-period<br />

from 1972 to 1999 using (a) topographic maps at<br />

1:5.000 scale (1972), (b) two datasets of aerial<br />

photos (1987, 1996), (c) satellite images (1999).<br />

Then, these data were compared with the present<br />

shoreline (2011), which was traced with the use of<br />

real-time kinematic differential GPS.<br />

The initial phase was to collect the available remote<br />

sensing data and create a time series of images<br />

along the contemporary coastline. The oldest data<br />

available were the topographic maps acquired from<br />

the Geographic Agency of the Hellenic Army that was<br />

also based on photogrammetry techniques on<br />

previously acquired aerial photographs.<br />

Using 42 air photographs acquired during 1987 we<br />

generated an ortho-mosaic for the same year. During<br />

this photogrammetric procedure a high resolution (2-<br />

meters) DEM was produced, and used for the ortho-<br />

43


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

rectification of a 15-meter resolution Landsat-7<br />

ETM+, panchromatic image. All the data were<br />

registered with an ortho-mosaic produced by 1996<br />

aerial photographs (Fig. 3).<br />

Next, by using image interpretation techniques we<br />

traced the coastline in the different periods. The<br />

difficulty was to identify the exact points of contact<br />

between the seawater and the land. This was made<br />

by equalizing the image histogram and in some<br />

cases applying a threshold value. The use of the<br />

panchromatic part of the spectrum for all the<br />

collected remote sensing data provides the<br />

homogeneity of the methodology.<br />

there is no systematic progradation or retrogradation<br />

in these delta fronts according to the data covering<br />

the last 40-year-period from 1972 to 2011 (Fig. 5).<br />

Nevertheless, there are parts of the coastline,<br />

especially where the Roman and Othoman levees<br />

used to function, that most of 50 meters of the beach<br />

have been eroded.<br />

Fig. 5: Synthetic image with all the traced coastlines at a<br />

part of the study area where Pineios river used to flow into<br />

the sea before 1800’s.<br />

CONCLUSIONS<br />

Fig. 3: Parts of the digital data used for the interpretation of<br />

recent coastline displacement.<br />

Establishing 4 GPS bases along the shore and use<br />

the technology of real time kinematic GPS point<br />

acquisition completed the methodology. The<br />

accuracy of the present coastline was very good as<br />

the specifications of the equipment give less than<br />

10cm (Fig. 4).<br />

Fig. 4: Using high accurate RTK GPS measurements for the<br />

tracing of the present coastline.<br />

The combination of all the traced coastlines on the<br />

remote sensing data with the RTK GPS recorded<br />

coastline have shown that both the former and the<br />

current delta fronts of Pineios River are divided into<br />

various sub-areas characterized by different type,<br />

phase and rate of shoreline displacement. Moreover,<br />

It is obvious that the western part of Pineios drainage<br />

basin is developed in an area (Gastouni graben),<br />

which is uplifted with lower relative uplift rate in<br />

comparison with the other surrounding areas. Hence,<br />

the Lower Pineios River was and is forced to flow in<br />

this graben, close and parallel to Pineios FZ.<br />

Furthermore, the age of Pineios FZ initiation<br />

progressively decreases from E to W. A similar<br />

decrease from E to W is also observed in the throw<br />

of Pineios FZ. The throw of the western part was<br />

gradually increased until a critical point in time<br />

(probably during 18 th century A.D.) when the relative<br />

uplift rate of the Pineios FZ footwall was larger than<br />

the relative uplift rate of the hanging wall. Since then,<br />

Pineios River was blocked, not able to flow N-wards<br />

and over the morphology escarpment formed by the<br />

fault and consequently enforced to shift S-wards.<br />

Moreover, the combined uplift movement of the<br />

footwall of Pineios FZ in the E and the northeastern<br />

(inland) part of the Kyllini peninsula in the W resulted<br />

in the slightly uplifted margin of the northwestern part<br />

of Gastouni graben, the block of the northwards flow<br />

of Pineios River and the initiation of the southwards<br />

flow of the river.<br />

This natural trend of Pineios southward diversion<br />

during 18 th century was supported and enforced by<br />

the human activity in the study area and especially by<br />

the construction of the ancient retaining wall of<br />

Pineios River (Papaconstantinou, 1991) during the<br />

Hellenistic period (2.330-2.150 B.C.) in order to<br />

protect the northern banks from the destructive river<br />

action.<br />

44


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The study area is undergoing intense and differential<br />

tectonic deformation, which has continued since<br />

Pliocene.<br />

The areas of maximum thickness are constantly<br />

subsiding during the sedimentation phase and strictly<br />

related to the Pineios delta and river sediment loads<br />

and transport. Moreover, the southern area presents<br />

higher subsidence rates than the northern one.<br />

The study area is also divided to three subareas that<br />

uplift with different rates (Fig. 6): (i) the footwall of<br />

Pineios fault zone (0,39-0,67 mm/yr), (ii) the<br />

Gastouni graben (0,19 mm/yr), (iii) the eastern<br />

(inland) part of Kyllini peninsula (0,36-0,48 mm/yr).<br />

The western part of Pineios basin is corresponding to<br />

the subarea with the lowest relative uplift rate (0,19<br />

mm/yr, Gastouni graben).<br />

Fig. 6: Areas of uplift and subsidence around the Pineios<br />

river former and contemporary deltas.<br />

The diversion of Pineios River to the south of Kyllini<br />

peninsula during the 18 th century is a case of fluvial<br />

antecedence upon the slightly uplifted margin of the<br />

Gastouni graben and is the result of the gradually<br />

increase of the throw along the western part of the<br />

Pineios fault zone during historic times marked by<br />

strong and destructive earthquakes during the late<br />

18 th or early 19 th century A.D. This natural process<br />

and trend is supported and enforced by the human<br />

activity in the study area during historic times as it is<br />

revealed by significant human constructions in the<br />

area.<br />

References<br />

Fountoulis, I., 1994. Neotectonic evolution of the Central-<br />

Western Peloponnese. Ph.D Thesis, Faculty of Geology,<br />

National and Kapodistrian University of Athens, GAIA 7,<br />

386pp. (in Greek, abridged English version)<br />

Hatzfeld, D., Pedotti, G., Hatzidimitriou, P., Makropoulos,<br />

K., 1990. The strain pattern in the western Hellenic arc<br />

deduced from a microearthquake survey: Geophys. J.<br />

Int., v. 101, p. 181-202.<br />

Kamberis, E., 1987. Geology and oil-geologic study of the<br />

post alpine sediments of NW Peloponnese. PhD Thesis.<br />

National Technical University of Athens, Greece (in<br />

Greek).<br />

Kontopoulos, N., Koutsios, A. 2010. A late Holocene record<br />

of environmental changes from Kotihi lagoon, Elis,<br />

Northwest Peloponnesus, Greece, Quaternary<br />

International, v. 225, 2, p. 191-198.<br />

Kraft, J.C., Rapp, (Rip), G., Gifford, J.A., Aschenbrenner,<br />

S.E., 2005. Coastal change and Archaeological Setting in<br />

Elis. Hesperia 74, 1-39.<br />

Lekkas, E., Papanikolaou, D., Fountoulis, I., 1992.<br />

Neotectonic Map of Greece, Pyrgos - Tropaia sheets,<br />

scale 1:100.000, Research project of the University of<br />

Athens, Department of Geology, Division of Dynamic,<br />

Tectonic, Applied Geology, Athens<br />

Lekkas, E., Papanikolaou, D., Fountoulis, I. 1995. The<br />

Pyrgos earthquake - The geological and geotechnical<br />

conditions of the Pyrgos area (W. Peloponnese, Greece).<br />

XV Congress of the Carpatho-Balcan Geological<br />

Association, Seminar on active faults, Geol. Soc. Greece,<br />

42-46, Athens.<br />

Mariolakos, I., Lekkas, E., Danamos, G., Logos, E.,<br />

Fountoulis, I., Adamopoulou, E., 1988. Geological -<br />

Tectonic Study of the Earthquake affected areas in Elis<br />

Prefecture (Kyllini Peninsula). Research project of the<br />

University of Athens, Department of Geology, Division of<br />

Dynamic, Tectonic, Applied Geology, 108 p., Athens<br />

1989 (in Greek).<br />

Mariolakos, I., Papanikolaou, D., Lagios, E. 1985. A<br />

Neotectonic Geodynamic Model of Peloponnesus Based<br />

on Morphotectonics, Repeated Gravity Measurements<br />

and Seismicity. Geol. Jb., v. B 50, p. 3-17.<br />

Mavroulis, S. 2009. Fault activity assessment in NW<br />

Peloponnesus – The Andravida Earthquake<br />

(08/06/2008). Msc Thesis, Inter-University Post-Graduate<br />

Studies Programme on “Prevention and Management of<br />

Natural Disasters”, Department of Geology and Geo-<br />

Environment of the National and Kapodistrian University<br />

of Athens, and Department of Geoinformatics and<br />

Topography of the Technological Educational Institute of<br />

Serres. 622 p.<br />

Mavroulis, S., Fountoulis, I., Lekkas, E., 2010.<br />

Environmental effects caused by the Andravida (08-06-<br />

2008, ML = 6.5, NW Peloponnese, Greece) earthquake.<br />

In: A. Williams, G. Pinches, C. Chin, T. McMorran and C.<br />

Massey, Editors, Geologically Active: 11 th IAEG<br />

Congress, Taylor & Francis Group, Auckland, New<br />

Zealand (2010), pp. 451-459.<br />

Papaconstantinou, E., 1991. Architectonic elements of the<br />

ancient Agora of Elis in the retaining wall of the Pineios<br />

River. First International Symposium on Achaia and Elis<br />

in Antiquity, Athens, p. 329-334 (in Greek).<br />

Papanikolaou, D., Fountoulis, I., Metaxas, C., 2007. Active<br />

faults, deformation rates and Quaternary paleogeography<br />

at Kyparissiakos Gulf (SW Greece) deduced from<br />

onshore and offshore data, Quatern. Int. 171-172, pp. 14-<br />

30.<br />

Stamatopoulos L., Voltaggio, M., Kontopulos, N., Cinque,<br />

A., La Rocca, S., 1988. 230 Th/ 238 U dating of corals from<br />

Tyrrhenian marine deposits of Varda area (North-western<br />

Peloponnesus), Greece. Geogr. Fis. Dinam. Quat., v. 11,<br />

p. 99-103.<br />

45


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

COULD LARGE PALAEOEARTHQUAKES BREAK GIANT STALACTITES IN<br />

CACAHUAMILPA CAVE? (TAXCO, CENTRAL MÉXICO)<br />

Garduño-Monroy VH. (1), R. Pérez-López (2)*, MA Rodríguez-Pascua (2), J. García Mayordomo (2)<br />

I. Israde-Alcántara(1) and J. Bischoff (3)<br />

(1) Universidad Michoacana San Nicolás de Hidalgo. Morelia. México. Email: vgmonroy@umich.mx<br />

(2) IGME – Instituto Geológico y Minero de España, Área de Investigación en Riesgos Geológicos. Madrid, España. Email:<br />

r.perez@igme.es , m.rodríguez@igme.es<br />

(3) United States Geological Survey. 345 Middlefield Road, MS 211, Menlo Park, CA 94025. EE.UU. jbischoff@usgs.gov<br />

* Corresponding author<br />

Abstract (Could large palaeoearthquakes break giant stalactites in Cacahuamilpa cave? (Taxco, central Mexico)): the<br />

Cacahuamilpa cave, an outstanding karstic system located in the nearby of Taxco, Central Mexico, shows several damaged<br />

dripstones: stalagmites and stalactites. This cave is determined by large E-W horizontal phreatic galleries, with large dripstones,<br />

mainly stalagmites, stalactites and travertine columns. The size of these speleothems reaches a maximum value of 30 m for<br />

stalagmites (The Champagne Bottle) and 10 m for stalactites. Oriented and fallen stalagmites and broken stalactites were dated<br />

by U-Th technique, and we have obtained a probable age of 1540 BP. We also have modelled the earthquake magnitude for<br />

breaking these stalactites and a minimum magnitude M7 was estimated. The cave is located 250 km away from the Middle<br />

American Trench, responsible of large earthquakes as the Michoacán, M8.1 in 1985. Besides these large subduction zone<br />

earthquakes, there are also large active intraplate normal and strike slip faults, but with low reoccurrence intervals that can<br />

generate large earthquakes in the region.<br />

Key words: speleoseismology, stalagmites, earthquake, Mexico<br />

THE CACAHUAMILPA CAVE SYSTEM<br />

Cacahuamilpa cave (CC) is located within the<br />

Guerrero State of Mexico, near of Taxco city, wellknown<br />

for its silver mining industry. This karst system<br />

is located within the Ixtapan Valley, a NW-SE<br />

elongated valley, 60 km long and 40 km wide. The<br />

Nevado de Toluca volcano (4558 m asl) is the<br />

highest peak and the karst is determined by two main<br />

fluvial channels: the San Jeronimo (running N-S) and<br />

the Chontalcoatlán (E-W). Both rivers meet under the<br />

La Corona hill and both appear outside of the<br />

carbonitic massif and from their hypogeum tunnelling<br />

at Dos Bocas. The lowest topographic point<br />

corresponds with the cave entrance of<br />

Cacahuamilpa, 1000 m asl. The Amacuzac River<br />

rises from the cave pit (Fig. 1). The CC belongs to<br />

the La Estrella Karstic System (LAKS). LAKS is<br />

configured by several caves in the surrounding of<br />

Cacahuamilpa: Cuevas Pacheco, Cueva Agua<br />

Brava, Gruta de Acuitlapán and Cuevas de La<br />

Estrella, among other minor caves.<br />

The geology of the cave is described as the Morelos<br />

Unit, a stratified limestone and dolostone deposit of<br />

Lower Cretaceous (Fries, 1960), with a maximum<br />

thickness of 900 m. Furthermore appears Albian<br />

limestone (Xochicalco and Cuautla Units) and the<br />

Mezcala Unit, formed by sandstone with interlayered<br />

black limestone (Fries, 1960). Quaternary deposits<br />

are volcanic andesitic rocks and basalts and the<br />

youngest ones are Holocene travertines.<br />

Hydrothermal activity related with active volcanoes<br />

(i.e. Nevado de Toluca) is described in Ixtapan and<br />

Tonatico (35º-40ºC)(Fries, 1960).<br />

The cave topography is mainly determined by a large<br />

horizontal phreatic tube (Fig. 2)(Bonet, 1971), with<br />

semi circular cross section. Large dripstones appear<br />

along the cave: (1). El Chivo, travertine; (2) and (3):<br />

La Tortuga and El Guerrero, cracked and toppled<br />

flowstone and columns. (4) El Guerrero dripstones.<br />

(5) The Botella de Champán, the greatest stalagmite<br />

36 m high, (6) The Calendario Azteca (Sun Stone)<br />

and the Volcán (7), a cone-shaped large flowstone.<br />

This work aims to demonstrate the possibility that<br />

large earthquakes (with magnitude M>7) could break<br />

giant speleothemes, and giving the fingerprint to<br />

obtain large palaeoearthquakes in the geological<br />

record that affected the area.<br />

Fig. 1: Overview of the carbonatic massif of Cacahuamilpa.<br />

Main fluvial channels run from NW- to SE. The topography<br />

of the cave is related with the hydraulic gradient. Yellow<br />

arrows indicate the hydraulic gradient.<br />

DAMAGED SPELEOTHEMS<br />

Spelaeoseismology is a new branch of<br />

palaeoseismology, which studies ancient<br />

earthquakes from the karstic record in caves (Lacave<br />

and Koller, 2004; Kagan et al., 2005; Pérez-López et<br />

46


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

al., 2009). Earthquakes are well recorded in caves by<br />

multiple damage affecting dripstones, travertine, also<br />

by ceiling collapse and faulting. Furthermore, the age<br />

of damage could be well constrained by using<br />

techniques such as U-Th disequilibrium of<br />

speleothems (e.g. Kagan et al., 2005).<br />

suggest earthquakes larger than M 7 close to the<br />

cave (50 km away). We have used the technique<br />

described in Maestro et al. (2011). In both cases the<br />

fallen structures have the same lithological<br />

composition (carbonates) and similar geometric<br />

relationships (H/D factor). However, in this case no<br />

strong water-flows are related to the toppled<br />

stalagmites. Nevertheless, the natural vibration of<br />

such large dripstone could be the responsible of the<br />

broken pieces (Cadorin et al. 2001). Assuming<br />

gravitational instability, the stalagmites would be<br />

oriented to the slip side instead of perpendicular<br />

(Fig.3).<br />

SEISMIC SOURCES<br />

Fig. 2. Cave topography in plant of Cacahuamilpa.<br />

Numbers indicate sites where broken speleothems are<br />

described. Red stars show those sites where sampled<br />

were collected. 1. El Chivo, 2 y 3. La Tortuga. 4. El<br />

Guerrero. 5. Botella de Champán. 6. Calendario Azetca o<br />

piedra del Sol. 7. El volcán. After cave topography from<br />

Bonet, 1971.<br />

Different damaged speleothems (seismothems)<br />

appear along the cave: broken stalagmites,<br />

stalactites, cracked columns and travertines. The<br />

Calendario Azteca stalactite, or Sun Stone,<br />

represents one of the most spectacular dripstone<br />

ever described in caves around the world. The total<br />

length estimated is 13 m approximately, and it shows<br />

an outstanding size for a hanging structure, 3.5 m of<br />

diameter and estimated longitude of 12 m. This E-W<br />

trending stalactite appears broken in six slices of<br />

concentric carbonate mega stones from the soda<br />

straw line (cm). Despite that this stalactite represents<br />

a national natural monument several samples were<br />

collected with the aim to obtain the age and the<br />

carbonate precipitation rate for dripstones in this time<br />

interval. At this moment data are still processing at<br />

the laboratory.<br />

The convergence between the Cocos plate and the<br />

North American plate at 53 mm/yr (GPS velocity, De<br />

Mets et al. 1990; Pardo & Suarez, 1995), configures<br />

and active area dominated by subduction<br />

earthquakes along the Middle American Trench<br />

(MAT) (Fig. 4). During the last century, thirteen<br />

instrumental earthquakes with magnitude greater<br />

than 7 have been recorded (1967-2011) in the trench<br />

zone of the MAT, and in a direction perpendicular to<br />

the cave (ca 250 km away). Moreover, the great<br />

Oaxaca earthquake of 1931 (M 8.1), the historic<br />

earthquake of Michoacán (1858) of 7.5 (Singh et al,.<br />

1985) and the Acambay earthquake of M7 (1912),<br />

suggest a complex relationship between subduction<br />

earthquakes and intraplate earthquakes due to<br />

normal lithospheric faulting.<br />

ABUNDANCE g/gsample Activity (dpm/s)<br />

234U 6.90E-11 ±1.4845E-13 0.95341 ±0.0020526<br />

238U 9.52E-07 ±1.7109E-09 0.7105 ±0.0012766<br />

230Th 2.95E-13 ±8.7190E-15 0.013429 ±0.00039744<br />

232Th 1.66E-09 ±4.9300E-11 0.00040293 ±0.000012<br />

Activity Ratio<br />

234U/238U 1.34E+00 ±0.0037630 AGE yr BP 1548±47<br />

230Th/234U 1.41E-02 ±0.00041796<br />

230Th/232Th 3.30E+01 ±1.4<br />

232Th/238U 5.67E-04 ±1.6921E-05<br />

230Th/238U 1.89E-02 ±5.6E-04<br />

Table 1. Radiometric dating by U-Th disequilibrium of<br />

the growth stalagmite post breaking (see Fig. 3 down).<br />

The sample wt is 0.52 g. Sample CAVO11-09.<br />

Other spectacular seismothems and fallen<br />

stalagmites are similarly oriented and could be<br />

related with the same phenomenon (Fig.2, 7 th site, El<br />

Volcán)(Fig.3). In this site, 4 stalagmites ranging<br />

between 4.2 and 1.7 m, with a diameter between<br />

0.57 and 0.4 m, appear broken and oriented E-W.<br />

We have dated the breaking by U-Th on the<br />

subsequent growing post the oriented toppling (Table<br />

1).<br />

The natural period of vibration of stalagmites for H/D<br />

ranging between 3 and 4 and acceleration measures<br />

Fig. 3. Up: Broken and oriented stalagmites located at El<br />

Volcán (see Fig. 2). Down: Photo interpretation of the<br />

fallen speleothems. The axis orientation is E-W,<br />

approximately. Red dots indicate sample collected points<br />

and blue ones points to be sampled in the future field<br />

work. Black and grey triangles indicate the maximum<br />

slope direction.<br />

47


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The Michoacán earthquake of the 19 th of September,<br />

1985 was the more devastated with a magnitude of<br />

8.1. This earthquake caused 9.500 fatalities, about<br />

30.000 people were injured and more than 100.000<br />

people were homeless, whereas severe damage was<br />

caused in buildings of Mexico DF and in different<br />

states of central Mexico also<br />

(http://earthquake.usgs.gov/earthquakes/world/event<br />

s/1985_09_19.php).<br />

Therefore, both the MAT subduction earthquakes as<br />

the large intraplate normal and strike-slips faults (e.g.<br />

the Morelia-Acambay Fault System, The Oaxaca<br />

Fault zone)(Garduño-Monroy et al., 2009), could<br />

generate large earthquakes to affect large<br />

speleothems within the Cacahuamilpa cave. It is<br />

interesting to describe large palaeoearthquakes<br />

triggered by intraplate faults in this zone due to, at<br />

present there is no instrumental record for such sized<br />

earthquakes. Therefore, infrequent large intraplate<br />

earthquake could be hibernating and palaeoseismic<br />

studies in the surroundings will be worthy to fill up the<br />

gap in the historical records.<br />

RESULTS<br />

Preliminary results suggest a potential<br />

palaeoearthquake dated c.a.1548 ±48BP, obtained<br />

from U-Th disequilibrium dating technique (Table 1.).<br />

The magnitude of the potential earthquake has been<br />

speculated from the geometry and mechanical<br />

properties of the stalagmites. Results suggest M>7.<br />

The geometry factor is defined by the relationship<br />

between the high and width (H/D), and the<br />

mechanical properties assuming a crystalline<br />

carbonate. Hence, we can conclude that:<br />

(1) The broken and oriented stalagmites within the<br />

Cacahuamilpa cave could be related with<br />

palaeoearthquakes.<br />

(2) The large size of damaged speleothems could<br />

be related with large earthquakes (M>7), in<br />

agreement with the tectonic framework of the<br />

Pacific side of Mexico: the convergence between<br />

Cocos and North American plates.<br />

(3) Palaeoearthquakes in this zone could be related<br />

to either subduction earthquakes or large<br />

intraplate earthquakes but with long<br />

reoccurrence intervals.<br />

The next step is to assign the potential seismic<br />

source. More accurate data are required to<br />

reconstruct the complete history of the broken<br />

speleothems in Cacahuamilpa. However, the<br />

earthquake hypothesis appeared as the strongest<br />

cause for this large damage. The interest of this work<br />

is for finding large shallow earthquakes (>M7) in a<br />

zone which is located 250 km far from the Middle<br />

American Trench.<br />

Fig. 4. Tectonic frame of Mexico. Most of earthquakes are related to the subduction between the Cocos plate and North<br />

American plate. Earthquakes represented here were obtained from Harvard online catalog (http://www.globalcmt.org/). After<br />

Pérez-López et al. (2011). TMVB: Trans Mexican Volcanic Belt<br />

48


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Acknowledgements: We are very grateful to Parque<br />

Nacional Grutas de Cacahuamilpa to give us all<br />

permissions to access to all parts of the cave, for extract<br />

samples and provide people to support us during the<br />

underground field work. Thanks are given to the CONACYT<br />

ministerial Mexican project entitled: Estudio tectónico, de<br />

paleosismología y de efectos sísmicos arqueológicos en<br />

lagos del Holoceno y actuales del Cinturón Volcánico<br />

Transmexicano y del bloque Jalisco.<br />

References<br />

Bonet, F. (1971). Espeleología de l aRegión de<br />

Cacahuamilpa. Universidad Autónoma Nacional<br />

Mexicana Instituto Geológico. Boletín nº 90. 98 p.<br />

Cadorin, J.-F., Jongmans, D., Plumier, A., Camelbeeck, T.,<br />

Delaby, S. and Quinif, Y. (2001). “Modelling of<br />

speleothems failure in the Hotton cave (Belgium). Is the<br />

failure earthquake induced?” Netherlands Journal of<br />

Geosciences 80(3–4): 315–321.<br />

DeMets, C., R.G. Gordon, D.F. Argus, and S. Stein,<br />

(1990).Current plate motions. Geophysics Research<br />

Journal International 101: 425-478.<br />

Garduño-Monroy V. H., R. Pérez-López, I. Israde-Alcantara,<br />

M. A. Rodríguez-Pascua, E. Szynkaruk, V. M.<br />

Hernández-Madrigal, M. L.García-Zepeda, P. Corona-<br />

Chávez, M. Ostroumov, V. H. Medina-Vega, G. García-<br />

Estrada, O. Carranza, E. Lopez-Granados and J. C. Mora<br />

Chaparro. (2009).Geometry and Paleoseismology of the<br />

Southwestern Part of the Morelia-Acambay Fault System,<br />

Central Mexico. Geofisica Internacional 48(3): 319- 335.<br />

Kagan, E. J., Agnon, A., Bar-Matthews, M. & Ayalon, A.<br />

(2005). Dating large, infrequent earthquakes by damaged<br />

cave deposits. Geology 33(4), 261–264.<br />

Fries, C. (1960). Geología del Estado de Morelos y de<br />

partes adyacentes de México y Guerrero, región central<br />

meridional de México. Instituto Geológico de México,<br />

Boletín nº 60(9), 236p.<br />

Lacave C. & M. G. Koller . (2004). What Can Be Concluded<br />

About Seismic History From Broken And Unbroken<br />

Speleothems? Journal of Earthquake Engineering, Vol. 8,<br />

No. 3 (2004) 431–455.<br />

Maestro A, G. Jané, J. García-Mayordomo, B. Fernández-<br />

Revuelta, M.A. Rodríguez-Pascua, J.J. Martínez-Díaz &<br />

R. Pérez-López. (2011). Cause Of The Rupture And<br />

Distribution Of Broken Submarine Carbonate Chimneys<br />

In The Gulf Of Cádiz (Southwestern Spain). Quaternary<br />

Int. In Press.<br />

Pardo, M. & Suárez, G. (1995). Shape of the subducted<br />

Rivera and Cocos plates in southern Mexico: seismic and<br />

tectonic implications. Journal of Geophysical Research<br />

100: 12,357-12,373.<br />

Pérez-López R., Rodríguez-Pascua, M.A., Giner-Robles,<br />

J.L., Martínez-Díaz, J.J., Marcos-Nuez, A., Silva, P.,<br />

Bejar, M. and Calvo, J.P. (2009). Spelaeoseismology<br />

and palaeoseismicity of the “Benis Cave” (Murcia, SE of<br />

Spain): coseismic effects of the 1999 Mula earthquake<br />

(mb 4.8). Geological Society of London, Special<br />

Publications, 316: 207-216.<br />

Pérez-López R., D. Legrand, V.H. Garduño-Monroy, M.A.<br />

Rodríguez-Pascua , J.L. Giner-Robles. (2011). Scaling<br />

laws of the size-distribution of monogenetic volcanoes<br />

within the Michoacán-Guanajuato Volcanic Field<br />

(Mexico). Journal of Volcanology and Geothermal<br />

Research 201: 65–72.<br />

Singh, S. K., Gerardo Suárez & T. Domínguez. (1985). The<br />

Oaxaca, Mexico, earthquake of 1931: lithospheric normal<br />

faulting in the subducted Cocos plate. Nature 317:56-58.<br />

49


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THREE-DIMENSIONAL INVESTIGATION OF THE AD 1621 PEDRO MIGUEL FAULT<br />

RUPTURE FOR DESIGN OF THE PANAMA CANAL’S BORINQUEN DAM<br />

Gath, Eldon M. (1 and Tania Gonzalez (1)<br />

(1) Earth Consultants International, 1642 E. 4 th Street, Santa Ana, California, 92701, USA. Email: gath@earthconsultants.com<br />

Abstract (Three-dimensional investigation of the AD 1621 Pedro Miguel fault rupture fro desing of the Panama canal`s<br />

Bronquen dam): Using a series of trenches, we excavated the Pedro Miguel fault in 3-D to measure the displacement magnitude<br />

and kinematics of the AD 1621 rupture. The purpose of the study was to use the Most Recent Event event as a proxy for a future<br />

rupture of the fault through the foundation of Borinquen Dam, a major new component of the Panama Canal Expansion. The study<br />

used a small, fault-affected, cobble-filled channel as the target for displacement measurements. Hundreds of ground survey<br />

points were obtained for contacts, faults, the channel thalweg and margins. The channel is offset 3.0±0.2 m right-laterally, and<br />

0.5±0.5 m reverse-vertically, with the vertical component occurring only within a few meters of the main fault tip. The fault rupture<br />

is expressed as a low-angle, one-sided, transpressive flower structure, exploiting weak bedding planes to propagate an enechelon<br />

stepping rupture across the landscape. Mitigation of this rupture will be an important requirement for the dam designers.<br />

Key words: Panama, paleoseismology, dam design, fault rupture<br />

Introduction<br />

As part of the Panama Canal’s Expansion Project, a<br />

four-segment earthen dam is being designed to form<br />

an ~7 km-long waterway to bypass the existing<br />

Miraflores and Pedro Miguel Locks (Fig. 1).<br />

Borinquen Dam will retain the Gatun Lake water<br />

elevation ~11 m above the Miraflores Lake elevation,<br />

and as such, must be designed to resist seismic<br />

loads. Extensive prior paleoseismic investigations<br />

(Rockwell et al., 2010) of the Pedro Miguel fault have<br />

shown the dam must also be designed to resist fault<br />

rupture.<br />

500±100 years, and its Most Recent Event (MRE) on<br />

May 2, 1621. Attempts to constrain the<br />

displacements from the MRE resulted in only<br />

minimum values of ~2 m near the dam site, but a well<br />

constrained 2.8 m at the fault’s northern end where it<br />

severed the Camino de Cruces (Gath and Rockwell,<br />

2009). The purpose of this latest investigation (ECI,<br />

2010) was to attempt to reconcile these two<br />

displacement results at the dam’s location.<br />

Fig. 1: Aerial view, looking NW, of the Pacific approach to<br />

the Panama Canal, showing the proposed location of the<br />

new locks, channel, and Borinquen Dam. The northerlytrending<br />

Pedro Miguel fault cuts through the proposed dam<br />

foundation about midway between the existing Miraflores<br />

and Pedro Miguel locks.<br />

The Pedro Miguel fault is a right-lateral strike-slip<br />

fault that passes through the planned Borinquen<br />

Dam’s foundation (Fig. 1). In earlier work, we (ECI,<br />

2007, 2008, 2010) determined that the fault has had<br />

multiple Holocene ruptures, a late Quaternary slip<br />

rate of ~5 mm/yr, a Holocene recurrence interval of<br />

Fig. 2: The investigation’s target was a small channel that<br />

appeared to be offset 3-4 meters. The fault (red line) was<br />

inferred to extend through the area, right-laterally offsetting<br />

a channel that approaches from the photo’s upper right<br />

corner and exits along the bottom left (shown by the<br />

geologists and the blue lines).<br />

From earlier studies and recent construction<br />

exposures, we knew the fault location to within a few<br />

meters (Fig. 2). The purpose of this paper is to<br />

present the technical details of the study’s findings,<br />

and to also present and discuss the methodology of<br />

the investigation, including the planning, execution,<br />

findings, and modifications that were made along the<br />

way, because there were many.<br />

50


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Investigation<br />

The investigation began with a simple plan (Fig. 3) to<br />

expose the fault on both sides of the small<br />

geomorphically defined area, and then characterize<br />

the channel geometry on both sides of the fault.<br />

Once the site was delineated, the channel margins<br />

could be slowly excavated towards the fault trace<br />

using a series of thin slices. Unfortunately, the initial<br />

fault-perpendicular trench (T-48 in Fig. 4) did not<br />

expose the fault where expected, and we failed to<br />

recognize the significance of a fault that was exposed<br />

elsewhere in the trench. The second fault-locator<br />

trench (C in Fig. 4) immediately filled with water from<br />

a sudden storm and was abandoned for four days<br />

until it could be pumped out. The two fault-parallel<br />

trenches (A and left part of B in Fig. 4) intended to<br />

define the channel geometry into and out of the fault<br />

did not expose any channel deposits or channel<br />

morphology. After four initial trenches, our<br />

investigation was definitely in trouble, with no fault<br />

and no channels to show for the work done to that<br />

point.<br />

excavations did not remove the geologic data and<br />

relationships that were vital to the measurement of<br />

the channel displacements. This was accomplished<br />

by excavating from the outer edges of the site<br />

inward, and by always keeping a mental map of the<br />

site and the goal, a preserved and measurable<br />

channel offset.<br />

N<br />

N<br />

Fig. 4: Schematic layout of our final trenching study.<br />

Trenches are shown as rectangles, the faults are in red, and<br />

the channel structure is shown by the blue lines, trending<br />

across the middle of the study area. The complex nature of<br />

the fault rupture pattern meant that the initially simple<br />

geomorphic offset inferred from the pre-trenching landscape<br />

was incorrect. The channel was effectively trapped within<br />

the fault zone, and each transpressive “petal” of the fault<br />

offset the channel progressively. The channel that we<br />

trenched first appears to have been man-made, to facilitate<br />

surface drainage to a culvert under Borinquen Road.<br />

Fig. 3: The original plan was to locate the fault (red line) on<br />

opposite sides of the displaced channel (blue arrows) by<br />

trenching perpendicular to the fault (large gray rectangles),<br />

then locate the channel margins by trenching parallel to the<br />

fault and perpendicular to the channel form (small gray<br />

rectangles). Using hand excavations and continuous survey<br />

control, we would then excavate the channel margins<br />

progressively closer to the fault, until they were in faultcontact<br />

on both sides of the fault.<br />

Fig. 4 diagrammatically shows the trenches that were<br />

finally excavated as we tried to sort out the details of<br />

the site and salvage some data for use by the dam<br />

designers. Once T-48 and 48-C failed to expose the<br />

faults where expected (Fig. 4), and trenches 48-A<br />

and B failed to expose the channels where expected,<br />

we lengthened 48-B until we found both the fault and<br />

the channel (Fig. 6). Fortunately, the channel was<br />

still fully contained on the hanging wall of the fault,<br />

and was not yet in fault contact, so we had not<br />

removed that important interaction point with our<br />

excavation.<br />

Fig. 5 shows a modification of Fig. 3 to reflect the<br />

pattern of the faults and channels at the site, as<br />

defined by the final trenches. The challenge was to<br />

continue the excavations but be careful that the<br />

N<br />

Fig. 5: Pedro Miguel fault trenching site immediately south<br />

of the old Borinquén Road (base of photo), following brush<br />

removal, but before trenching started. The fault and channel<br />

locations, as interpreted from the geomorphology, are<br />

shown with the lighter, dashed lines, whereas the actual<br />

fault and channel locations found after trenching are shown<br />

diagrammatically with the bold and solid lines.<br />

Fig. 7 shows the channel on the hanging wall above<br />

the fault, whereas Fig. 8 shows the structural<br />

complexity of the fault zone that forced us to evolve<br />

the initial investigation plan to accommodate the<br />

unexpected. The extreme low angle of the fault<br />

acted as a bulldozer of the surface soils pushing<br />

them out and over the channel alluvium, but this also<br />

51


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

served to bury, and thereby preserve, the channel<br />

deposits under the fault petal.<br />

The purpose of digging 48-K (Fig. 9) was to continue<br />

the exposure of 48-B up-dip of the fault to get as<br />

close as possible to the spot where the fault first cut<br />

the base of the channel. With the fault dipping to the<br />

west and the channel flowing to the east, this could<br />

occur suddenly. In Fig. 9 it appears that the deepest<br />

part of the channel is touching the fault, but there are<br />

still 3-5 cm of separation. In 48-M (Fig. 10) however,<br />

the base of the channel is the fault, and the SE<br />

channel margin is completely removed by the fault.<br />

Thus, our channel margin’s northern piercing point<br />

lies between 48-K and 48-M (±2 m), and the<br />

thalweg’s piercing point lies within 48-M (±0.5 m).<br />

Fig. 6: Trench 48-B - the discovery trench, showing the fault<br />

(right) and the channel deposits (left), looking out over the<br />

Panama Canal in the background.<br />

Fig. 9: Interpreted image of the end of Trench 48-K, with 48-<br />

B (Fig. 7) ~ 1 meter on the other side (to the south) of the<br />

trench. This image shows the channel deposits still above<br />

the fault trace.<br />

Fig. 7: Interpreted image of part of Trench 48-B showing the<br />

low-angle fault and the channel deposits on the hanging<br />

wall. The area shown as a “clay extrusion” is interpreted to<br />

be a weathered mole track from the MRE as it is intruded<br />

into, and deformed, the modern surface soils..<br />

Fig. 10: Looking easterly at the interpreted image of the end<br />

of Trench 48-M, excavated ~1 m to the left (east) of 48-K<br />

(Fig. 7). This photo shows the base of the channel deposits<br />

now in fault contact in the head of the trench, and truncated<br />

by the fault on the right side.<br />

Fig. 8: Interpreted image of Trench 48-B, looking back<br />

towards Fig. 7, showing the low-angle fault petals where<br />

they have broken upwards to the surface, and the lack of<br />

alluvial deposits on the eastern (right) wall.<br />

In addition to the fault complexity shown in Fig. 8, it is<br />

important to note that there were no alluvial deposits<br />

visible on the fault’s footwall in the northern wall of<br />

the trench. However, trenches 48-F, G, H, J, & M all<br />

exposed channel deposits and cobbles (Figs 4, 10<br />

and 11). This is because the channel margin on the<br />

footwall lies 0.5-1.0 m north of the north face of<br />

Trench 48-B. Thus Trench 48-B missed taking out<br />

52


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

the southern margin of our target channel by less<br />

than 1 m. With the southern margin so tightly<br />

constrained, the most accurate offset measurements<br />

came from that side.<br />

channel by the most recent earthquake. Using 15<br />

trenches and hundreds of surveyed data points, we<br />

were able to constrain the MRE rupture to 3.0±0.2 m<br />

of right-lateral displacement, and 0.5±0.5 m of<br />

localized reverse-slip uplift at the surface tip of the<br />

fault. Because the fault is expressed through the<br />

dam as an en-echelon stepping, transpressional<br />

flower structure that exploits the weak bedding<br />

planes of the near-surface strata, it will be a difficult<br />

fault to mitigate in the design of Borinquen Dam (Fig.<br />

13).<br />

N<br />

Fig. 11: Interpreted image of Trench 48-H, with 48-B (Fig. 8)<br />

exposed through the window at the end of the trench. This<br />

image shows the limits of the channel deposits on the<br />

footwall side of the fault, and shows a secondary fault petal<br />

on the right wall vertically truncating the cobble deposits<br />

within the excavated width of the trench.<br />

Fig. 13: Map of the Pedro Miguel and Miraflores faults<br />

through the Borinquen Dam area (hachured). Areas where<br />

we have conducted paleoseismic trenching are shown with<br />

the green squares; the green square directly over the dam<br />

location is the location of this study.<br />

Acknowledgements<br />

Thanks to Kay St. Peters and Barrett Salisbury for excellent<br />

field assistance and data mapping. Appreciation is due to<br />

the Autoridad del Canal de Panama for permission to<br />

conduct this study and to Ms. Pastora Franceschi for<br />

arranging all the details.<br />

Fig. 12: Geologically interpreted map of the channel offsets<br />

based on hundreds of survey points collected. The yellow<br />

area reflects the full margins of the sandy channel deposits<br />

whereas the blue area defines the cobble-filled channel<br />

thalweg. The northern margin of the channel is more poorly<br />

constrained (±2 m), while the southern margin and thalweg<br />

margins are constrained to less than ±1 m. The right-lateral<br />

offset across the fault is measured at 3.0±0.2 m. A 0.5 m<br />

vertical offset occurred at the tip of the main fault petal,<br />

resulting in the upward bowing and erosional removal of the<br />

channel thalweg, but this uplift is localized to only a few<br />

meters from the fault tip and is not present away from the<br />

fault. Farther east, the fault tip over-rides and protects the<br />

alluvial channel deposits.<br />

Conclusions<br />

Although the study was much more complicated than<br />

initially planned, we were successful in locating the<br />

fault, in exposing a young channel that was offset<br />

across the fault, and in measuring the offset of that<br />

References<br />

Earth Consultants International (ECI), 2007, Paleoseismic<br />

Trenching of the Pedro Miguel Fault in Cocolí, Located<br />

Immediately Southwest of the Panamá Canal, Panamá;<br />

consulting report for the Autoridad del Canal de Panamá<br />

(ACP), Project No. 2614.02.<br />

ECI, 2008, Quantitative Characterization of the Pedro<br />

Miguel Fault, Determination of Recency of Activity on the<br />

Miraflores Fault, and Detailed Mapping of the Active<br />

Faults Through the Proposed Borinquén Dam Location;<br />

consulting report for the ACP, Project No. 2708.05.<br />

ECI, 2010, Additional Trenching of the Pedro Miguel and<br />

Miraflores faults in Cocoli, immediately southwest of the<br />

Panamá Canal; consulting report for the ACP, Project No.<br />

3012.<br />

Gath, E.M. and Rockwell, T.K., 2009, Coseismic offset of<br />

the Camino de Cruces confirms the Pedro Miguel fault as<br />

the cause of the AD 1621 Panamá Viejo earthquake; in<br />

Pérez-Lopez, R. and others (eds), Archeoseismology and<br />

Palaeoseismology:: 1 st<br />

INQUA-IGCP-567 Int. Workshop<br />

on Earthquake Archaeology and Palaeoseismology,<br />

Baelo Claudia, Spain, p. 32-34.<br />

Rockwell, T. and others, 2010, Neotectonics and<br />

paleoseismology of the Limón and Pedro Miguel faults in<br />

Panamá: Earthquake Hazard to the Panamá Canal: B. of<br />

the Seis. Soc. America, v. 100, p. 3097-3129.<br />

53


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

VELOCITY FIELD IN BULGARIA AND NORTHERN GREECE FROM GPS CAMPAIGNS<br />

SPANNING 1993-2008<br />

Georgiev, Ivan (1, Dimitar Dimitrov (1), Pierre Briole (2), Emil Botev (3)<br />

(1) Department of Geodesy, National Institute of Geophysics, Geodesy and Geography, Bulgarian Academy of Sciences, Sofia<br />

1113, Acad. Bonchev Str., Bl. 3, BULGARIA, E-mail: ivan@bas.bg<br />

(2) Ecole Normale Superieure, Paris 75230, 45 rue d’Ulm, FRANCE<br />

(3). Department of Seismology, National Institute of Geophysics, Geodesy and Geography, Bulgarian Academy of Sciences, Sofia<br />

1113, Acad. Bonchev Str., Bl. 3, BULGARIA<br />

Abstract (Velocity Field in Bulgaria and Northern Greece from GPS Campaigns Spanning 1993-2008): We<br />

use GPS observations during the period from 1993 to 2008 to obtain the velocities of about 100 points in Bulgaria<br />

measured in more than 30 campaigns. Along with data from 21 points in Northern Greece, measured between 1999<br />

and 2008, the results constrain the present-day motion in the region. Points in FYROM and Albania are also<br />

included in the processing to outline the picture of recent motions from Black to Adriatic Sea. The velocity field<br />

shows overall motion to the south relative to stable Eurasia and increasing motion from north to south confirming<br />

the extensional regime in Southwest Bulgaria and Northern Greece. The results can be used to constrain the north<br />

boundary of South Balkan extensional region.<br />

Key words: GPS, velocity field, Bulgaria and Nortern Greece, South Balkan Extensional Region<br />

MOTIVATION<br />

The region of South Bulgaria, especially Southwest<br />

Bulgaria and the Rhodopes Mountains, is the most<br />

active tectonic and seismotectonic area of the<br />

country with proved recent active tectonic structures<br />

and crustal movements. The strongest earthquake in<br />

continental Europe in the last two centuries occurred<br />

in the Krupnik-Kresna region. The territory belongs to<br />

the southern part of the Central-Balkan neotectonic<br />

region – a zone with recent extension of the crust<br />

and with complex interaction between the horizontal<br />

and vertical movements of the tectonic structures<br />

(Zagorchev, 2001). The geological and geophysical<br />

data confirm the recent activity of the fault structures<br />

formed during the Late Neogene and the Quaternary.<br />

The area is located in the northern part of the North<br />

Aegean region and is strongly affected by its<br />

tectonics and high seismicity. This is the main reason<br />

for establishing in the early 2000 a relatively dense<br />

geodynamic GPS network in Southwest Bulgaria for<br />

long-term monitoring of recent crustal movements.<br />

The network is repeatedly measured in the period<br />

2001 - 2008. Along with GPS data from the new<br />

National GPS network of Bulgaria and a few EUREF<br />

campaigns spanning the period 1993 - 2008 we<br />

obtained velocities of about 100 points in Bulgaria.<br />

We use also GPS data from several campaigns in<br />

Northern Greece, few points in FYROM and Albania<br />

to obtain the velocity field and to constrain tectonics<br />

in the region. The obtained results can be used to<br />

constrain the northern boundary of the South-Balkan<br />

extensional province.<br />

GPS DATA<br />

GPS data used in the solution include campaigns on<br />

the territory of Bulgaria, Northern Greece, FYROM<br />

and Albania and comprise 158 points. In Bulgaria<br />

GPS data span the period 1993 – 2008. The data<br />

from the nineties are mainly from the EUREF<br />

campaigns in Bulgaria. Most of the data are from<br />

extensively measured geodynamic network in<br />

Southwest Bulgaria and points from the new National<br />

GPS network. The typical duration for each point<br />

measurement in the successive campaigns is 48<br />

hours. The points with long observation history are<br />

98. In Northern Greece we analyzed data from 21<br />

points collected in 1999, 2000 and 2008. The data<br />

from 2008 are from joint Bulgarian-Greek GPS<br />

campaign with 48 to 72 hours measurements. Data<br />

from 5 points in FYROM are included in the solution<br />

from campaigns in 1996, 2000 and 2008. The data<br />

from 1996 and 2000 are measured in two EUREF<br />

campaigns, courtesy given by the Agency of<br />

Cadastre and the 2008 campaign is performed jointly<br />

by Department of Geodesy, Sofia, and Faculty of<br />

Civil Engineering, University “Ss. Cyril and Metodius”,<br />

Skopje. Albanian data comprise 34 stations in four<br />

campaigns from 2003 till 2009 and are committed by<br />

the Laboratory of Alpine Belts Geodynamics,<br />

University of Grenoble, France.<br />

All the data are processed and analyzed by the<br />

Bernese 5.0 software. The reference frame is<br />

ITRF2005 and the processing follows the EUREF<br />

standards (see for example Boucher and Altamimi,<br />

2007). The reference and kinematic frame are<br />

defined by points of the European Permanent<br />

Network (EPN) included in the solution. The details of<br />

54


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

the processing strategy can be found in Georgiev<br />

(Georgiev et al., 2007).<br />

RESULTS AND DISCUSSION<br />

Fig. 1 shows the obtained velocities of all 158 points<br />

in Bulgaria, Northern Greece, FYROM and Albania<br />

relative to (stable) Eurasia. The results filled in the<br />

gap in velocity field in the Eastern Mediterranean to<br />

the north of the extensively studied Marmara and<br />

Aegean regions. Results from other authors are also<br />

presented on the figure (Reilinger et al., 2006,<br />

Hollenstein et al., 2008, Floyd et al., 2010) to show<br />

the position of Bulgarian and Northern Greece<br />

territory in the East Mediterranean geodynamic<br />

settings. The velocities of the Albanian stations and<br />

points in Bulgaria (red vectors, points mainly from the<br />

National GPS network) show the velocity field in<br />

east-west direction from Black to Adriatic Sea.<br />

Fig. 1: Velocity field in Bulgaria, Northern Greece and surrounding areas relative to Eurasia. The South Balkan extensional region<br />

is outlined according to Burchfiel et al. (Burchfiel at al., 2006)<br />

mm/y in Central-west Bulgaria (Balkan Mountain) to<br />

10 mm/y for the region of Chalkidiki. On the northsouth<br />

profile from the Balkan Mountain till the<br />

Chalkidiki peninsula the north component of the<br />

estimated velocities are drawn (Fig. 2). The<br />

increasing velocities from north to south are clearly<br />

seen on this 400 km long profile confirming the northsouth<br />

extensional regime in the region (see also<br />

McClusky et al., 2000, Hollenstein et al. 2008).<br />

Fig. 2: North-south velocity gradient from Central-west<br />

Bulgaria (Balkan Mountain) to the region of Chalkidiki<br />

The obtained velocities in Southwest Bulgaria and<br />

Northern Greece indicate an overall motion to the<br />

south. The results show an increasing rate from 2<br />

One point deviates from the overall south motion in<br />

Southwest Bulgaria and it is clearly seen on the<br />

velocity gradient profile (Fig. 2). This is a point from<br />

the local geodynamic network of the Department of<br />

Geodesy around the Krupnik fault, located on its<br />

northern edge. The result is in good agreement with<br />

the dipping of the fault obtained by geologic and<br />

seismotectonic data.<br />

The points in northern and eastern part of the<br />

Bulgarian territory (red vectors) are not dense<br />

enough to draw unquestionable conclusions. But<br />

from the obtained results it is clear that North<br />

Bulgarian territory, north from the Balkan Mountain, is<br />

55


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

part of Eurasia with velocities practically near to zero.<br />

On the Fig. 1 are shown the slightly changed borders<br />

of the South Balkan extensional region proposed by<br />

Burchfiel et al. (Burchfiel at al., 2006). According to<br />

this model the northern border of the extensional<br />

region follow approximately the Balkan Mountain.<br />

Our data as a whole, although not sufficiently dense<br />

in Southeast Bulgaria, seems to confirm this<br />

hypothesis. The question which remains is the<br />

behavior of few points located on the southeastern<br />

Black sea coast showing north-northeast motion. We<br />

need a denser network in this region to clear these<br />

two questions – to confirm the northern border of the<br />

South Balkan extensional region and to resolve the<br />

motion along the Black Sea coast.<br />

Acknowledgements: We gratefully acknowledge the<br />

support of all agencies in the countries for the<br />

measurements: Military Geographyc Service and Agency of<br />

Geodesy, Cartography and Cadastre, Sofia, Bulgaria;<br />

Faculty of Civil Engineering, University “Ss. Cyril and<br />

Metodius”, Skopje, FYROM; Aristotle University of<br />

Thessaloniki, Greece and Laboratory of Alpine Belts<br />

Geodynamics, University of Grenoble, France.<br />

References<br />

Boucher, C., Altamimi, Z. (2007). Memo: Specifications for<br />

reference frame fixing in the analysis of a EUREF GPS<br />

campaign, Version 27-03-2007.<br />

Burchfiel, B.C., King, R.W., Todosov, A., Kotzev, V.,<br />

Durmurdzanov, N., Serafimovski, T., Nurce, B. (2006),<br />

GPS results for Macedonia and its importance for the<br />

tectonics of the Southern Balkan extensional regime.<br />

Tectonophys 413 (3-4), 239-248.<br />

Floyd, M. A., et al. (2010), A new velocity field for Greece:<br />

Implications for the kinematics and dynamics of the<br />

Aegean. J. Geophys. Res., 115 (10), B10403,<br />

doi:10.1029/2009JB007040.<br />

Georgiev I., Gabenski, P., Gladkov, G., Tashkov, T.,<br />

Danchev, P., Dimitrov D. (2006). NATIONAL GPS<br />

NETWORK: Processing the observations from the<br />

Main order. Geodesy N 18, Special edition, Military<br />

Geographic Service of the Bulgarian Army, 209p.<br />

Georgiev I., (2009). National and Permanent GPS networks<br />

of Bulgaria – processing of the observation, analysis and<br />

application in geodynamic Thesis to obtain the degree<br />

Doctor of Science”, Department of Geodesy, National<br />

Institute of Geophysics, Geodesy and Geography, Sofia,<br />

Bulgaria, 258p.<br />

Hollenstein et al. (2008). Crustal motion and deformation in<br />

Greece from a decade of GPS measurements, 1993–<br />

2003. Tectonophysics 449, 17–40.<br />

Matev, K. (2011). GPS constrains on current tectonics of<br />

Southwest Bulgaria, Northern Greece and Albania.<br />

Thesis to obtain the degree “Doctor” of University of<br />

Grenobe, Laboratory of Alpine Belts Geodynamics,<br />

University of Grenoble, France, Department of Geodesy,<br />

National Institute of Geophysics, Geodesy and<br />

Geography, Sofia, Bulgaria, 203p.<br />

McClusky, S., Balassanian, S., Barka, A. Demir, C.<br />

Ergintav, S., Georgiev, I., Gurcan, O., Hamburger,<br />

O., Hurst, K, Kahle, H., Kastens, K., Kekelidze, G.,<br />

King, R,, Kotzev, V., Lenk, O., Mahmoud, S.,<br />

Mishin, A., Nadariya, M., Ouzounis, A.. Paradissis,<br />

D., Peter. Y., Prilepin, M., Reilnger, R., Sanli, I.,<br />

Seeger, H., Tealeb, A., Toksoz, M., Veis, G. (2000).<br />

Global Positioning System Constraints on Plate<br />

Kinematics and Dynamics in the Eastern<br />

Mediterranean and Caucasus. Journal of<br />

Geophysical Research - Solid Earth, Vol. 105, No.B35,<br />

5695-5719.<br />

Reilinger, R., et al. (2006), GPS constraints on continental<br />

deformation in the Africa-Arabia-Eurasia continental<br />

collision zone and implications for the dynamics of plate<br />

interactions, J. Geophys. Res., 111, B05411,<br />

doi:10.1029/2005JB004051.<br />

Zagorchev, I. (2001). South Western Bulgaria, Geological<br />

Guidebook. Sofia, Bulg. Acad. Sci., 98p.<br />

56


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ACROCORINTH - GEOLOGICAL HISTORY AND THE INFLUENCE OF<br />

PALEOSEISMIC EVENTS TO RECENT ARCHAEOLOGICAL RESEARCH<br />

Gielisch, Hartwig (1)<br />

(1) EurGeol. Dr. Hartwig Gielisch, DMT GmbH & Co. KG, Am Technologiepark 1, 45307 Essen Germany<br />

Phone, +49 201 172 1888, email: hartwig.gielisch@dmt.de<br />

Abstract (Acrocorinth - Geological history and the influence of paleoseismic events to recent archaeological research):<br />

The Acrocorinth is the result of two major tectonical events. From Upper Cretaceous to Eocene the Acrocorinth limestone, origins<br />

from the Mesozoic carbonate platforms, were mixed with cherts of the Böotian Ocean in an accretionary wedge of the Central<br />

Hellenic Melange Belt. The melange units were covered by Neogene Acrocorinth conglomerates. Pliocene Corinthian marls cover<br />

all older units. In the Pleistocene the hill was pushed up very rapidly through the Tertiary units. Pleistocene marine terraces occur<br />

in 220 m a. s. l. at Acrocorinth of Pre-Paleotyrrhean age, thus verifying an uplift of more than 200 m in the last 400,000 years. All<br />

sediments are covered by the several sheets? of debris, the results of several major paleoseismic events. These breccias contain<br />

as components Mesozoic and Cainozoic rocks occurring at the Acrocorinth and cover probably? the remaining nine temples<br />

described by Pausanias. In order to find the missing temples it is essential to know, which part of the sediments is in-situ or are<br />

part of landslides triggered by paleoseismic events.<br />

Key words: conglomerates, rapid uplift, debris breccias, covered temples<br />

INTRODUCTION<br />

Acrocorinth is located in the northern Argolis south of<br />

the Corinthian golf and directly south of the ruins of<br />

ancient Corinth. The geological history of the hill is<br />

subdivided in two major tectonical events: the<br />

genesis of a melange during a continent-continent<br />

collision in the lower Tertiary and a rapid uplift in the<br />

younger Pleistocene. Several paleoseismic events<br />

form the actual shape of the hill.<br />

GEOLOGICAL HISTORY OF THE ACROCORINTH<br />

The geological history of Acrocorinth starts in the<br />

middle Mesozoic. Neritic lime- and dolostones<br />

formed on the Parnassus-carbonate platform as part<br />

of the Centralhellenic Systems, while in the center of<br />

the Böotian Ocean, a sub basin of the Pindos Ocean,<br />

pelagic cherts and ophiolitic sequences were built.<br />

During the closing of the Böotian Ocean carbonates,<br />

cherts and ophiolites were tectonically mixed as part<br />

of an accretion in the subduction trench thus forming<br />

a mélange belt between the Parnassus- and<br />

Pelagonian-plates (GIELISCH, 1993, 1994). The<br />

youngest components of these mega-breccias are of<br />

Eocene age. The alpine orogenesis of Greece is<br />

completed in the Lower to Middle Miocene.<br />

During the Upper Miocene the alpine Hellenides were<br />

fragmented by a phase of intense brittle faulting. The<br />

major systems of the Greek intra-mountainous<br />

graben-systems developed during the Upper<br />

Miocene and the Lower Pliocene.<br />

During and directly after these phases of brittle<br />

faulting the Acrocorinth conglomerates have been<br />

deposited over the Mesozoic units. Components of<br />

the conglomerates comprise mainly local and<br />

regional geological units (carbonates, cherts,<br />

serpentinites of the ophiolites, quartzites) (RICHTER<br />

et al., 1992). Exotic clasts like olivine-gabbros,<br />

gap<br />

limestones<br />

(Triassic)<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

v<br />

flyschsandstones<br />

(Cretaceous-<br />

Paleogene)<br />

gap<br />

gap<br />

v<br />

v<br />

v<br />

ophiolites<br />

(Jurassic)<br />

Acrocorinth<br />

limestones<br />

(U. Jurassic-<br />

L. Cretaceous)<br />

v<br />

v<br />

v<br />

Sheet of debris<br />

(Quarternary)<br />

Corinthian Marls<br />

(Neogene)<br />

Acrocorinth<br />

conglomerates<br />

(Lower Neogene)<br />

radiolarites,<br />

cherts<br />

pyroclastic<br />

intercalations<br />

(Jurassic)<br />

Fig. 1: Schematic profile of sequences in the Corinthian<br />

area showing the stratigraphical position of the main<br />

geological units<br />

57


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

marbles and other metamorphic rocks show a<br />

connection to source rocks, which are nowadays<br />

located further away. Following after the Acrocorinth<br />

conglomerates, the Corinthian Marls formed in the<br />

Neogene and cover all other geological units (s. fig.<br />

1).<br />

In the beginning of the Pliocene the Graben of<br />

Corinth started to sink. As compensation for the<br />

deepening of the graben the Parnassus-Ghiona<br />

Mountains rose in the north, while in the south the<br />

Trapazona Mountains and isolated areas north,<br />

including Acrocorinth have been up-lifted pushing<br />

through the Neogene and Quaternary sediments. At<br />

Acrocorinth marine terraces are developed at 220 m<br />

a.s.l.. The age of the terraces is Pre-Paleotyrrhean<br />

(460,000 to 400,000 a) thus verifying an uplift of the<br />

hill of some 200 m in the last 400,000 years (SEDAT,<br />

1986).<br />

Today the components of the Mesozoic mélange<br />

form the top of the Acrocorinth hill as well as a<br />

smaller hill west of Acrocorinth, (Penteskufion).<br />

Acrocorinth conglomerates occur east and west of<br />

the hill, pushed up during the up-lift of Acrocorinth,<br />

with circular dip directions outwards on the flanks.<br />

West of the Penteskufion the conglomerate also<br />

occurs, dipping also into a westerly direction. In the<br />

eastern and western areas the conglomerates are<br />

covered by Corinthian Marls. In the south of<br />

Acrocorinth, conglomerates do not outcrop. Here the<br />

Corinthian Marls are directly in contact with Mesozoic<br />

units. On the northern flank conglomerates do also<br />

not outcrop. The outcropping units on the northern<br />

flank of the hill are interpreted to be breccias with<br />

angular components and not conglomerates.<br />

last 2000 years. Following PHILLIPSON (1892) and<br />

v. FREYBERG (1952) three bigger earth quick<br />

destroyed the area in historical times: 420 B.C., 77<br />

A.C. and 551 A.C.<br />

One of the biggest was the earthquake of 551 A.D.<br />

This earthquake affected mainly Acrocorinth and<br />

thereafter the acropolis on the Acrocorinth remained<br />

inhabited until the 8th century. The northern flanks of<br />

the hill collapsed and a mixture of natural rocks and<br />

cultural debris formed talus deposits, which cover the<br />

main parts of the northern flanks (s. Fig. 2). These<br />

slide breccias have been cemented in the following<br />

centuries and can therefore, be misinterpreted as<br />

conglomerates. However, the angular to sub-angular<br />

character of the clasts point to an origin as breccias<br />

in contrast to the well-rounded clasts of the<br />

Acrocorinth conglomerates. In summary, it was<br />

possible it identify four different generations of slide<br />

breccias:<br />

I. Very coarse sheet of debris: components sharpedged,<br />

up to 50 cm diameter, lightly cemented with<br />

powdery, carbonate matrix. This breccia shows a<br />

fining upwards, the components become smaller to<br />

the top of the unit. The unit is older than unit II. No<br />

other dating was possible.<br />

PALAEO-EARTHQUAKES<br />

In ancient times, beginning from the Mycenaean<br />

period until the Turkish invasion, the Acrocorinth was<br />

the acropolis and fortress of the ancient city of<br />

Corinth. Nearly every civilization left archaeological<br />

remains on the top and the flanks of the hill.<br />

Following PAUSANIAS (II. 4.6) in Roman times (A.D.<br />

160) ten smaller temples and sanctuaries<br />

(BOOKIDIS & STROUD (1987) were located on the<br />

northern flank of the hill:<br />

“:::[2.4.6] As you go up this Acrocorinthus you see<br />

two precincts of Isis, one if Isis surnamed Pelagian<br />

(Marine) and the other of Egyptian Isis, and two of<br />

Serapis, one of them being of Serapis called “in<br />

Canopus.” After these are altars to Helius, and a<br />

sanctuary of Necessity and Force, into which it is not<br />

customary to enter…..” PAUSANIAS (II. 4.6)<br />

The Demeter and Persephone temple, one of these<br />

sanctuaries, has been excavated in the 1970ies and<br />

80ies by the American School of Classical Studies.<br />

Locations of the remaining 9 temples or sanctuaries<br />

are unknown. But where are the ruins of the<br />

sanctuaries?<br />

The Graben of Corinth is a seismic high risk area.<br />

Several major earthquakes affected this area in the<br />

Fig.2: Sheets of debris at the Northern flank of the<br />

Acrocorinth: a: base of unit IV; b: base of unit I<br />

II. Fine-coarsed sheet of debris: components angular<br />

rounded with diameters of 5 to 10 cm. The<br />

components are firmly cemented by carbonate matrix<br />

and building pedocretes. The unit is older than 700<br />

B.C.. The small theatre in the temple of Demeter was<br />

cut into this unit and this temple was built around 700<br />

B.C..<br />

III. Medium-coarsed sheet of debris: components<br />

angular of up to 20 cm diameter. The components<br />

are firmly cemented by a carbonate matrix. Also here<br />

pedocrete genesis is possible to observe. This<br />

breccia is young the unit II, because it overlies unit II.<br />

The breccia contains in addition to the natural<br />

components bricks and human processed stones.<br />

This unit could be the sheet of debris of the earth<br />

quick from 420 B.C., because it overlies the theatre<br />

of the temple of Demeter.<br />

58


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

IV. Fine-to medium coarsed sheet of debris:<br />

components sharp-edged, up to 10 cm diameter,<br />

lightly cemented with terra fusca and terra rossa. The<br />

base of the breccia shows bricks, destroyed pottery<br />

and traces of cinders (s. Fig. 3). The breccia contains<br />

a lot of terrestrial Stylommatophora shells. In the year<br />

146 B.C. the Romans destroyed Corinth fighting<br />

against the Archaean League. The city was<br />

completely razed and Corinth was nearly 100 year’s<br />

uninhabitated. Julius Ceasar refounded the city as<br />

Colonia Laus Iulia Corinthiensis in 44 B.C. shortly<br />

before his assassination. This horizon could be the<br />

level of the Roman ravage in the year 146. B.C.. In<br />

this case the unit IV is younger than 146 B.C..<br />

Pausanias visited the area around 160 A.D., 83 years<br />

after the major earth quake from 77 A.D. The unit IV<br />

conforms to a typical rock fall landslide and based on<br />

this study it is the final bigger unit becoming exist as<br />

result of such an event. Hence unit IV could be the<br />

sheet of debris of the 551 A.D. seismic event.<br />

Following another earthquake on the 21st February<br />

1858 the city was given up and rebuilt 6 km<br />

northwest of the ancient settlement.<br />

CONCLUSIONS<br />

BOOKIDIS & STROUD (1987) describe the northern<br />

flank sediments as conglomerates. These<br />

conglomerates at Acrocorinth are of a Neogene age<br />

and excavations in these sediments will probably not<br />

lead to any successes looking for remains from the<br />

Roman period.<br />

Recent studies indicate that the the northern flanks of<br />

Acrocorinth are covered by breccias resulting from<br />

palaeoseismic events. Sedimentary characteristics<br />

do indicate an origin as cemented talus deposits,<br />

some of which triggered by paleo-earthquakes.<br />

These breccias are most probably younger than the<br />

Roman period because of cultural debris. A Neogene<br />

age can be excluded.<br />

Archaeological excavations in and below these<br />

recent sediments may locate the remaining of<br />

sanctuaries described by PAUSANIAS 160 A.D.<br />

References<br />

Fig. 3: Unit IV a: shells of Stylommatophora; b: level of the<br />

Roman ravage; c: conglomerates of the Corinthian Marls<br />

Corinth and parts of the fortifications on Acrocorinth<br />

were rebuilt during the rule of Justinian I, Caesar of<br />

Byzantium. In this time the Byzantine Empire was<br />

nearly 200 years converted to Christianity and there<br />

was no reason to rebuild ancient heathen temples. In<br />

order to locate the remaining ruins of the 9 missing<br />

temples it is necessary to identify the level between<br />

unit III and unit IV and start excavations at this level.<br />

Bookidis, N. & Stroud, R. (1987). Demeter and Persephone<br />

in ancient Corinth. 33 p., American School of Classical<br />

Studies, ISBN 0876616716, Princton, New Jersey.<br />

Freyberg, B. v. (1952). Der Bau des Isthmus von Korinth.-<br />

Ann. geol. Pays Hell., 4, p. 157-188, Athens.<br />

Gielisch, H. Dragastan, O. & Richter, D.K. (1993). Lagoonal<br />

to tidal carbonate sequences of Upper Jurassic/Lower<br />

Cretaceous age in the Corinthian area: Melange blocks of<br />

the Parnassos Zone.- Bull. Geol. Soc. Greece, Vol. XXVI-<br />

II/3, 663676, Athen.<br />

Gielisch, H. (1994). Mikrofazies und stratigraphische<br />

Stellung des Jura-Kreide-Übergangsbereichs der<br />

Parnass-Kiona-Zone zwischen Mittelgriechenland und<br />

Argolis.- Dissertation, 223 p., Bochumer geol. u.<br />

geotechn. Arb., Bd. 43, Bochum.<br />

Pausanias (II, 4.6) from http://www.theoi.com/<br />

Text/Pausanias1A.html<br />

Philippson, A. (1892). Der Peloponnes – versuch einer<br />

KLandeskunde auf geologischer Grundlage. 642 p.,<br />

Berlin.<br />

Sedat,R (1986). Das Stufenland von Korinth<br />

(Griechenland).- 183 p., Diss., Bochum.<br />

Richter, D.K., Dragastan, O. & Gielisch, H. (1992):<br />

Microfacies, diagensis and biostratigraphy of the<br />

Jurrassic/Cretaceous lagoonal Acrocorinth-Limestone<br />

(Parnassos Zone NE-Peloponnesos, Greece).-<br />

Bochumer geol. u. geotechn. Arb., 39, 1-70, Bochum.<br />

59


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

INTERPRETING OFFSHORE SUBMERGED TSUNAMI DEPOSITS:<br />

AN INCOMPLETELY COMPLETE RECORD<br />

Beverly N. Goodman-Tchernov (1)<br />

(1) Moses Strauss Department of Marine Geosciences, Leon Charney School of Marine Sciences, University of Haifa, Mt.<br />

Carmel, Israel 31905. bgoodman@univ.haifa.ac.il<br />

Abstract (Interpreting Offshore Submerged Tsunami Deposits: An Incompletely Complete Record): The complications and<br />

challenges of understanding and interpreting offshore submerged tsunami deposits is presented here and discussed in relation to<br />

recently and previously published tsunami studies. Regular annual storm activity impacts the shallow uppershelf and littoral zone,<br />

and rarer larger storms can impact deeper water depths. These events combined, with subwater surface currents, can regularly or<br />

periodically alter the seabottom, manipulating and reworking previously laid tsunamigenic horizons. Because of this, offshore<br />

deposits must be carefully considered within their meteorological and coastal morphological contexts. The differentiation between<br />

storm and tsunami deposits has long been a primary concern for understanding the sedimentological signature of paleotsunami<br />

deposits. Here, examples of paleotsunami deposits offshore on the Mediterranean coast of Israel will be presented with a<br />

discussion of how to address the issue of ‘completeness’ of the offshore record in order to properly place these evidence in the<br />

overall context of tsunami records.<br />

Key words: tsunami, sedimentology, uppershelf, Mediterranean<br />

INTRODUCTION<br />

Recent research offshore from Caesarea, Israel<br />

(Reinhardt et al. 2006, Goodman-Tchernov et al.<br />

2009, Fig. 1) argues for the presence of preserved<br />

tsunamites from multiple paleotsunami events. The<br />

specific horizons were identified based on the<br />

presence of tsunamigenic characteristics anchored in<br />

granulometry, shell taphonomy, sedimentological<br />

structures, erosional features, and terrestrial and<br />

marine mixing. While these horizons are well<br />

documented and laterally extensive, there are<br />

variations in the thickness and some events<br />

mentioned in historical documentation that are not<br />

present. The question arises whether the physical<br />

record is incomplete, or the historical records<br />

fabricated. The offshore record is highly problematic<br />

due to regularly occurring high-energy storm events<br />

which can disturb and rework remains from tsunami<br />

events (Dawson 2007, Shanmugam 2011). In<br />

addition, this reworking varies by a multitude of<br />

factors including depth, seafloor sedimentology and<br />

morphology, ultimately leaving a heterogeneous<br />

footprint. A deposit of one composition might survive<br />

intact due to width, material hardness, or mechanical<br />

strength, while another is easily erased. This<br />

absence and presence is seen when comparing<br />

across cores. In Caesarea, current research is aimed<br />

at tackling the complicated story in the near offshore<br />

environment through detailed analysis of sediment<br />

cores and the comparative study of modern storm<br />

impact and deposits.<br />

STORM IMPACTS<br />

In mid-December 2010, a large storm hit the Israeli<br />

coastline with measured wave heights of 13.7<br />

meters, just after a comparative storm/non-storm<br />

study was launched. A storm of this magnitude is<br />

only known to have a reoccurrence cycle of 2-3 times<br />

a century. The impact of the storm on the coastline<br />

was significant, dismantling and transporting large<br />

sections of concrete from a modern wave breaker,<br />

scouring the base of the ancient aqueduct, stripping<br />

the upper layer of sand from the seabottom in many<br />

parts of the harbor, and scraping 1-2 meters off of the<br />

face of coastal cliffs (see Fig. 2). A small underwater<br />

excavation survey and sediment study was<br />

completed in the summer and fall of 2010, preceding<br />

the storm. Following the storm, the same areas were<br />

Fig. 1: Caesarea, Israel. Aerial photograph of the<br />

coastline with the submerged ancient harbor. Tsunamis<br />

in antiquity struck the harbour, likely contributing to the<br />

harbour’s ultimate demise.<br />

60


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

revisited. This rare large storm had a great effect on<br />

the areas studied, and the deposits give a glimpse<br />

into the degree of impact that large storms may have<br />

had in antiquity, providing a reference point for<br />

differentiating between normal sea conditions, large<br />

storms, and tsunamis.<br />

Fig. 2: December 15, 2010, Caesarea, Israel. Two days<br />

after height of major decadal storm. Rubble eroded from<br />

buildings and concrete slabs moved from the outer<br />

breakwater cover a structure on the harbour pier.<br />

Normal, storm, and tsunami conditions were defined<br />

in sediment cores with the use of particle size<br />

distribution contour mapping (Goodman-Tchernov et<br />

al. 2009, see Fig. 3). Now, the recent large storm has<br />

provided a modern analog for comparison.<br />

DISCUSSION<br />

Offshore sedimentological research presents a<br />

unique preservation scenario, one in which there is<br />

no one-for-one record as one might find in certain<br />

varved lake sediments. The finds offshore of<br />

Caesarea are surprising, in the regard that any<br />

record remains in 20 meters water depth and<br />

shallower. Observations from the recent large storm<br />

clearly demonstrate that storms are capable of<br />

altering the record substantially, and it is therefore<br />

significant that finds demonstrate the presence of<br />

some preservation. What does remain, could be<br />

argued, is telling about the types of materials that are<br />

capable of withstanding later storm impacts. It could<br />

also be argued that the presence of events in the<br />

historical record that lack physical correlation in the<br />

offshore sediments should not be dismissed. Such<br />

absences might relate to the size or characteristic of<br />

the tsunami deposit—that it was not substantial<br />

enough to survive later storms. Future work might<br />

investigate cores from deeper waters, beyond any<br />

storm influence (~30m water depth) to see whether<br />

indications of those ‘missing’ events are present.<br />

Offshore tsunami deposits, particularly in the zone of<br />

storm-influence, do not represent a complete<br />

depositional sequence, rather, they are a<br />

representation of specific sets of circumstances<br />

dictated by deposit composition, thickness, and<br />

surrounding geomorphological conditions that<br />

nonetheless provide an important and independent<br />

window into paleotsunami impacts.<br />

Fig. 3: Particle size distribution contour mapping used to<br />

differentiating between storm and tsunami events (from<br />

Goodman-Tchernov et al. 2009). ‘S’ indicates storms, ‘E’<br />

indicates tsunami events. The lack of storm events in core 2<br />

is interpreted as the product of water depth and distance<br />

from the shore.<br />

Acknowledgements: Appreciation expressed to Israel<br />

Antiquities Authority (K. Sharvit), Hatter Laboratory for<br />

Coastal Archaeology (M. Artzy), Israel Science Foundation,<br />

Israel Ministry of Infrastructure, and H. Dey.<br />

References<br />

Dawson, A. G., and I. Stewart (2007), Tsunami deposits in<br />

the geological record, Sediment.Geol., 200(3-4), 166-<br />

183.<br />

Goodman-Tchernov, B. N., H. W. Dey, E. G. Reinhardt, F.<br />

McCoy, and Y. Mart (2009), Tsunami waves generated<br />

by the Santorini eruption reached Eastern Mediterranean<br />

shores, Geology, 37(10), 943-946.<br />

Reinhardt, E. G., B. N. Goodman, J. I. Boyce, G. Lopez, P.<br />

van Hengstum, W. J. Rink, Y. Mart, and A. Raban<br />

(December, 2006), The tsunami of 13 December A.D.<br />

115 and the destruction of Herod the Great's harbor at<br />

Caesarea Maritima, Israel, Geology, 34(12), 1061-1064.<br />

Shanmugam, G. (2011), Process-sedimentological<br />

challenges in distinguishing paleo-tsunami deposits,<br />

Natural Disasters.<br />

61


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKE ENVIRONMENTAL EFFECTS, INTENSITY AND SEISMIC HAZARD<br />

ASSESSMENT: THE EEE CATALOGUE (INQUA PROJECT #0418)<br />

Guerrieri, Luca (1, Anna Maria Blumetti (1), Elisa Brustia (1), Eliana Esposito (2), Mauro Lucarini (1), Alessandro M. Michetti (3),<br />

Sabina Porfido (2), Leonello Serva (1) & Eutizio Vittori (1), & the INQUA TERPRO Project #0811 Working Group<br />

(1) Dipartimento Difesa del Suolo, Servizio Geologico d’Italia, ISPRA, Via V. Brancati 48, 00144 Roma, ITALY. Email:<br />

luca.guerrieri@isprambiente.it<br />

(2) Istituto per l’Ambiente Marino e Costiero, CNR, Calata Porta di Massa - 80133 Napoli, ITALY<br />

(3) Dipartimento di Scienze Chimiche e Ambientali, Università dell'Insubria, Via Valleggio, 11, 22100 Como, ITALY<br />

Abstract (Earthquake Environmental Effects, intensity and seismic hazard assessment: the EEE Catalogue (INQUA<br />

Project #0418): Earthquake Environmental Effects (EEE) are the effects produced by an earthquake on the natural environment,<br />

either directly linked to the earthquake source or triggered by the ground shaking. These include surface faulting, regional uplift<br />

and subsidence, tsunamis, liquefaction, ground resonance, landslides, and ground failure phenomena.<br />

The EEE catalogue is a data collection of Earthquake Environmental Effects from modern, historical and paleoseismic<br />

earthquakes compiled at global level by the INQUA TERPRO Project #0811 WG.<br />

The damages caused by recent catastrophic seismic events have been mostly linked to the vulnerability of physical environment<br />

enhancing the crucial role of EEEs, including tsunamis, for seismic hazard purposes. Therefore, these events have confirmed that<br />

the EEE Catalogue is an essential tool to complete traditional SHA based on PGA maps, since it allows to identify the natural<br />

areas most vulnerable to earthquake occurrence and to objectively compare in time and in space the earthquake intensity through<br />

the ESI scale.<br />

Key words: Earthquake Environmental Effects, ESI intensity scale, seismic hazard.<br />

INTRODUCTION<br />

Earthquake Environmental Effects (EEE) are the<br />

effects caused by an earthquake on natural<br />

environment, including surface faulting, regional uplift<br />

and subsidence, tsunamis, liquefactions, ground<br />

resonance, landslides and ground failure, either<br />

directly linked to the earthquake source or provoked<br />

by the ground shaking (Michetti et al., 2007).<br />

Most of the damage resulting from moderate to large<br />

earthquakes is typically related with the vulnerability<br />

of the physical environment. This point has been<br />

sadly and dramatically confirmed by the two relevant<br />

seismic events occurred in the last months in<br />

countries with strong economy and modern building<br />

codes, i.e., the February 22, 2011, Mw 6.3<br />

Christchurch and the March 11, 2011, Mw 9.0 East<br />

Japan earthquakes.<br />

The March 11, 2011, earthquake occurred in the<br />

Pacific Ocean near the coast of NE Japan. Most of<br />

the damage in terms of dead toll (more than 20,000<br />

people) and destruction was caused by the huge<br />

tsunami (run up values larger than 38 m; more than 5<br />

km inland penetration in the Sendai coastal plain).<br />

Comparatively, the amount of damage induced by<br />

the vibratory ground motion itself was modest. The<br />

size of the 2011 tsunami was fairly larger than those<br />

recorded in the affected area in the last century, but<br />

comparable with the tsunamis affecting the same<br />

areas in 869 A.D. recently well documented by<br />

means of geological and paleoseismic studies (Sawai<br />

et al., 2007; HERP, 2009).<br />

Nevertheless, the relevance of EEE has been also<br />

shown by the moderate-size event occurred on<br />

February 22, 2011, very close to the town of<br />

Christchurch, New Zealand.<br />

Fig. 1: – Geological record of past tsunamis at Watari and<br />

Yamamoto, Myagi prefecture (Sawai et al., 2007). The<br />

characteristics and spatial distribution of these deposits<br />

allowed to identify ancient tsunamis comparable with the<br />

March 11, 2011 event.<br />

62


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

This was basically an aftershock of the 2010<br />

September 4, Mw 7.0 event. However, while the main<br />

shock did not cause victims or large damage, the<br />

2011 event caused the death of at least 75 people<br />

and the destruction of several districts in the<br />

Christchurch area. This scenario of bad damage was<br />

mainly linked to local effects of site amplification,<br />

which depends largely on the stratigraphic<br />

characteristics (geological history) of the recent<br />

deposits over which the town is built; as well, the<br />

same sediments are particularly susceptible to<br />

liquefaction. As a consequence, also houses<br />

designed in agreement with the local seismic codes<br />

have collapsed, due to liquefaction within the<br />

foundation soils.<br />

Both events, although not comparable in size and<br />

geodynamic setting, have confirmed once again i) the<br />

relevance of earthquake environmental effects as a<br />

major source of seismic hazards, in addition to<br />

vibratory ground motion; ii) the need of re-evaluating<br />

the significance of macroseismic intensity as<br />

unequivocal measurement of the final earthquake<br />

impact on the local natural and built environment,<br />

and iii) the relationships between epicentral intensity,<br />

earthquake size, and source geometry. As a matter<br />

of fact, intensity is a parameter able to describe a<br />

complete earthquake scenario, based on direct field<br />

observation. Thus, the knowledge of the<br />

characteristics and spatial distribution of EEE<br />

induced by past earthquakes will strongly improve<br />

standard seismic hazard assessments, that typically<br />

consider only the vibratory ground motion hazard.<br />

THE EEE CATALOGUE<br />

Nowadays, a significant amount of data about<br />

Earthquake Environmental Effects is available for a<br />

very large number of recent, historical and paleoearthquakes.<br />

However available information is<br />

located in several different sources (scientific papers,<br />

historical documents, professional reports), and often<br />

difficult to access.<br />

The EEE Catalogue has been promoted with the aim<br />

to properly retrieve the available information about<br />

EEE at global level and archive it into a unique<br />

database, in order to facilitate their use for seismic<br />

hazard purposes. Its implementation has been<br />

endorsed at global level by the INQUA TERPRO<br />

Project #0811, through a Working Group coordinated<br />

by ISPRA - Geological Survey of Italy.<br />

The EEE catalogue collects the characteristics, size<br />

and spatial distribution of coseismic effects on nature<br />

in a standard way from modern, historical and<br />

paleoearthquakes. For each event, we have<br />

assessed epicentral and local intensities based on<br />

EEE data through the ESI 2007 scale (Michetti et al.,<br />

2007), that integrates and completes the traditional<br />

macroseismic intensity scales, allowing to assess the<br />

intensity parameter also where buildings are absent<br />

or damage-based diagnostics saturates. This<br />

procedure has allowed an objective comparison in<br />

terms of earthquake intensity, for events occurred in<br />

different areas and/or in different periods.<br />

The information is collected at three levels of<br />

increasing detail (Earthquake, Locality, Site). Also<br />

available imagery documentation (photographs,<br />

video, sketch maps, stratigraphic logs) can be<br />

uploaded into the database.The quality of the<br />

database in terms of completeness, reliability, and<br />

resolution of locations is strongly influenced by the<br />

age of the earthquake so that it is expected to be<br />

very variable.<br />

Nevertheless, even where the information is less<br />

accurate (historical earthquakes), the documented<br />

effects are typically the most relevant i.e. most<br />

diagnostic for intensity assessment. Similarly, the<br />

information from paleoseismic investigations,<br />

although poorly representative of the entire scenario,<br />

still includes significant data (i.e. local coseismic fault<br />

displacements) very helpful of a minimum size of the<br />

earthquake.<br />

A first official release of the EEE Catalogue has been<br />

done in the frame of the XVIII INQUA Congress, held<br />

in Bern in July 2011. However, the implementation of<br />

the EEE catalogue is always in progress at<br />

http://www.eeecatalog.sinanet.apat.it/login. Data can<br />

be explored on a public interface (Fig. 2) based on<br />

Google Earth at http://www.eeecatalog.sinanet.<br />

apat.it/terremoti/index.php. Earthquake records<br />

validated by the Scientific Committee of the Project<br />

can be also downloaded from the site.<br />

THE ADDED VALUE<br />

The major added value of the EEE Catalogue in<br />

terms of seismic risk is the possibility to explore the<br />

scenarios of environmental effects induced by past<br />

earthquakes and therefore identify the areas where<br />

the anthropic settlements and infrastructures are<br />

more exposed to this source of potential hazard. To<br />

this end, a good accuracy of EEEs location becomes<br />

crucial. Typically, EEEs from recent earthquakes are<br />

mapped with good accuracy immediately after the<br />

event. Nevertheless, even for some historical<br />

earthquakes it is possible to retrieve with very good<br />

detail this information. It is the case of the December<br />

28, 1908 Messina Straits earthquake and consequent<br />

tsunami (Fig. 3), where the EEE Catalogue allows to<br />

locate the earthquake/tsunami effects over the<br />

present urban texture with a spatial resolution of a<br />

few meters, pointing out the areas characterized by<br />

the highest risk. Furthermore, the EEE Catalogue<br />

allows to reveal possible trends in the spatial<br />

distribution of primary and secondary effects. For<br />

example, Fig. 4 shows the spatial distribution of<br />

EEEs induced by the October 8 2005, Muzaffarabad,<br />

Pakistan, earthquake (Ali et al., 2009): it is quite<br />

evident that the location and amount of surface<br />

faulting is consistent with the spatial distribution of<br />

coseismic slope movements, in terms of both areal<br />

density and size.<br />

A similar result is shown by the spatial distribution of<br />

EEEs induced by the 1811-1812 New Madrid,<br />

Missouri, earthquakes, mapped in Fig. 5. Indeed, the<br />

most relevant primary and secondary effects are<br />

located along the Mississippi valley near New<br />

63


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Madrid, consistently with the surface projection of the<br />

causative faults, and unquestionably provide<br />

diagnostic elements for assessing an epicentral<br />

intensity equal to XI.<br />

Fig. 2: The public interface of the EEE Catalogue, developed on Google Earth<br />

http://www.eeecatalog.sinanet.apat.it/terremoti/index.php.<br />

Fig. 3: EEEs induced by the December 28, 1908 Messina Straits, Italy, earthquake in the area of the Messina harbour. If<br />

information from contemporary sources is very precise, it is possible to use the EEE Catalogue also for local seismic<br />

microzonation.<br />

FINAL REMARKS<br />

The recent catastrophic earthquakes occurred in<br />

Japan and New Zealand have clearly pointed out that<br />

traditional seismic hazard assessment based only on<br />

vibratory ground motion data need to be integrated<br />

with information about the local vulnerability of the<br />

territory to earthquake occurrence.<br />

The collection of Earthquake Environmental Effects<br />

provided by the EEE Catalogue aims at identifying<br />

the areas most vulnerable to earthquake occurrence.<br />

This information must complement traditional SHA<br />

based on PGA maps.<br />

Moreover, based on EEE characteristics, size and<br />

spatial distribution it is possible i) to assess the<br />

earthquake intensity through the ESI scale, and ii) to<br />

objectively compare the earthquake intensity of<br />

events occurred in different areas and/or in different<br />

periods.<br />

64


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 4 :Surface faulting and slope movements induced by the October 8, 2005 Muzaffarabad earthquake.<br />

Fig. 5 : EEE induced by the December 16 1811 New Madrid, Missouri, earthquake. Primary and secondary effects indicative of<br />

intensity XI in the ESI 2007 scale are located in the epicentral area along the Mississippi Valley.<br />

Acknowledgements:<br />

Fabio Baiocco and Antonio Scaramella are acknowledged<br />

for their crucial support in building the infrastructure of the<br />

EEE Catalogue. We wish also to acknowledge all the<br />

people who has participated to the implementation of the<br />

EEE Catalogue in the frame of the #0418 INQUA TERPRO<br />

Project.<br />

References<br />

Ali Z., Qaisar M., Mahmood T., Ali Shah M., Iqbal T.,, Serva<br />

L., Michetti A.M.. and Burton P.W. (2009) - The<br />

Muzaffarabad, Pakistan, earthquake of 8 October 2005:<br />

surface faulting, environmental effects and macroseismic<br />

intensity Geological Society, London, Special<br />

Publications 2009, 316:155-172; doi:10.1144/SP316.9.<br />

HERP (2009). Long-term assessment on seismic activity<br />

along Japan trench offshore Sanriku and offshore Boso<br />

Peninsular.<br />

http://www.jishin.go.jp/main/chousa/09mar_sanriku/index.<br />

htm [in Japanese with many figures].<br />

Michetti, A. M., E. Esposito, L. Guerrieri, S. Porfido, L.<br />

Serva, R. Tatevossian, E. Vittori, F. Audemard, T.<br />

Azuma, J. Clague, V. Comerci, A. Gurpinar, J. McCalpin,<br />

B. Mohammadioun, N. A. Morner, Y. Ota, and E.<br />

Roghozin (2007). Intensity Scale ESI 2007. In Memorie<br />

Descrittive Carta Geologica d’Italia L. Guerrieri and E.<br />

Vittori (Editors), APAT, Servizio Geologico d’Italia—<br />

Dipartimento Difesa del Suolo, Roma, Italy, 74, 53 pp.<br />

Sawai Y., Shishikura M., Okamura Y., Takada K., Matsu’ura<br />

T., Tin Aung T., Komatsubara J., Fuji Y., Fujiwara O.,<br />

Satake K., Kamataki T., and Sato N. (2007). A study on<br />

paleotsunami using handy geoslicer in Sendai Plain<br />

(Sendai, Natori, Iwanuma, Watari and Yamamoto),<br />

Miyagi, Japan. N. 7, p. 47-80, 2007<br />

http://unit.aist.go.jp/actfaulteq/seika/h18seika/pdf/sawai.pdf<br />

[In Japanese with<br />

English abstract and captions].<br />

65


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ACTIVE TECTONICS OF ALBANIA INFERRED FROM FLUVIAL TERRACES GEOMETRIES<br />

Guzman, Oswaldo (1,2), Jean-Louis Mugnier (1), Rexhep Koçi (3), Riccardo Vassallo (1), Julien Carcaillet (4), Francois Jouanne (1),<br />

Eric Fouache (5)<br />

(1) Institut de sciences de la Terre. Université de Savoie, Bâtiment des Belledonnes, Campus Scientifique, 73376 Le Bourget du<br />

Lac Cedex, France. E-mail : Oswaldo-Jose.Guzman-Guitierrez@etu.univ-savoie.fr, Jean-Louis.Mugnier@univ-savoie.fr,<br />

Riccardo.Vassallo@univ-savoie.fr, Francois.Jouanne@univ-savoie.fr<br />

(2) Dpto. Ciencias de La Tierra, Universidad Simon Bolivar, Edificio FE-II. Apartado 89000, Caracas 1081-A, Venezuela<br />

(3) Institut de Sciences de la Terre. Université Joseph Fourier. Maison des Géosciences, 1381 rue de la Piscine, 38400 Saint<br />

Matin d’Hères, France. E-mail : Julien.Carcaillet@ujf-Grenoble.fr<br />

(4) Institute of Seismology of the Academy of Science, Tirana, Albania. E-mail: rexhep.koci@yahoo.com<br />

(5) University of Paris 10, Nanterre, UMR 8591, EA435 Gecko, France. E-mail: eric.g.fouache@wanadoo.fr<br />

Abstract (Active Tectonics of Albania Inferred from Fluvial Terraces Geometries): We studied four rivers of the southern<br />

Albanides in order to quantify the geodynamic control on the incision of alluvial terraces. The Vjosa, Osum, Devoll and Shkumbin<br />

rivers flow from north-western Greece and southern Albania across the active Dinaro-Hellenic Alpine fold belt to the west. From<br />

mapping and dating of terrace abandonment, we reconstructed spatial and temporal variations of incision rates along those rivers.<br />

These reconstructions allow a quantification of the fluvial evolution at a higher resolution than a glacial-inter-glacial cycle, and<br />

therefore an estimation of the effect of geodynamic uplift on river incision. We identified ten preserved terraces units developed<br />

since the “MIS-6” up to historic time. Uplift seems to produce two distinct effects: an overall increase of the incision rate from west<br />

to east related to regional bulging and local pulses of increasing incision generated by activation of faults.<br />

Key words: Albania, Active Tectonics, Alluvial Terraces.<br />

INTRODUCTION<br />

Active tectonics, eustatic and climatic variation are<br />

the forcing that controls the formations of fluvial<br />

terraces. In a tectonically active setting characterised<br />

by long term uplift, Pazzaglia (submitted) suggests<br />

that the incision rate equals to the long-term uplift<br />

rate. Nonetheless, the morphology of the rivers in this<br />

context is strongly climate-dependent: a change in<br />

downstream base level could have a profound impact<br />

on long profile shape (Merrits et al., 1994), and<br />

hydrologically-driven changes in discharge and<br />

sediment load induce periodically lateral incision<br />

processes that widen the channel bottom producing<br />

strath fluvial terraces (Hancock & Anderson 2002) or<br />

aggradation leading to the development of filled<br />

terraces (Ouchi, 1985)<br />

The Albanide domain is rather small (less than 5. 10 4<br />

km 2 ). We consider that the climatic evolutions of the<br />

different Albanide catchments are rather similar.<br />

Therefore differences in the evolution of the studied<br />

catchments would not be explained by local climate<br />

changes.<br />

The active tectonics pattern in Albania is quite<br />

variable (Fig. 1), with a NE-SW shortening in the<br />

western part of the Albania, and a NW-SE extension<br />

in the inner part of the chain, it leads to spatially<br />

variable long term uplift (Aliaj, 2000). Therefore the<br />

comparison of the different catchments evolution<br />

would help to estimate the uplift rate in Albania and<br />

to outline the role of the variation of the tectonics rate<br />

in the sedimentation/erosion behaviour of the<br />

catchments.<br />

GEOLOGICAL SETTING<br />

The Dinaro-Hellenic Alpine fold belt constitutes a<br />

segment of the wide Circum-Mediterranean Peri-<br />

Tethyan thrust belt. This belt as a result of the Dinaric<br />

subduction. The Albanian foothills have been thrust<br />

westwards over the Adriatic foredeep during the<br />

Alpine orogeny (Roure et al., 2004).<br />

The presenty-day geodynamic deformation results<br />

from the subduction of the Adriatic Sea floor beneath<br />

the foreland domain. This geodynamic setting<br />

produces contrasting relief with a flexural basin filled<br />

with Plio-Quaternary deposits and forming a flat<br />

costal plain in the foreland, a thin skinned fold and<br />

thrust belt in the midland, and a large basin and<br />

range zone in the hinterland, resulting from<br />

synorogenic Neogene-Quaternary grabens that<br />

crosscut the thrust system (Roure et al., 2004,<br />

Niewland, et al., 2001).<br />

The area (Fig.1) is still tectonically active and<br />

produces permanent microseismicity and frequent<br />

earthquakes reaching intensities of IX. GPS data<br />

suggest that the western Albania is being affected by<br />

southwestward motions relative to Adriatic<br />

microplate, illustrating the ongoing collision of<br />

external Albanides, whereas inner Albanides present<br />

southward motion, increasing from north to south,<br />

relative to both Adriatic and stable Eurasia (Jouanne<br />

et al., submitted).<br />

The neotectonics are controlled by NE-SW<br />

compressive stress in the western part connected to<br />

subduction of the Apulian lithosphere and extensional<br />

stress in the internal zones (Goldsworthy et al., 2002;<br />

Roure et al., 2004). The major extension direction<br />

varies from E-W close to the Ohrid and Prespa lakes<br />

66


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

to NNW-SSE at the southern end of the Ersekë<br />

basins (Jouanne et al., submitted). On the regional<br />

scale the Albanides are governed by long-term uplift<br />

(Aliaj, 2000).<br />

We concentrate our study on a zone located between<br />

two major transverse structures which cut the area:<br />

the Vlore-Fier-Elbasani-Dibra transfer zone in the<br />

41.5°<br />

Durres<br />

Erzen<br />

Tirane<br />

central part of Albania (Roure et al., 2004), and the<br />

prolongation of the Keffalinia strike-slip zone at the<br />

northwestern border of Greece (Baker et al., 1997).<br />

The Vjosa, Osum, Devoll and Skhumbin rivers<br />

constitute four of the six main rivers of Albania. They<br />

are close to each other; therefore, their climatic<br />

settings are comparable.<br />

N<br />

Librazdhi<br />

41°<br />

Karavatsas<br />

Lagoon<br />

C.p.<br />

S hkumbin<br />

Ar.A.<br />

Seman<br />

S.A.<br />

?<br />

?<br />

Lushnje<br />

Elbasan<br />

Berat<br />

E.G.<br />

Pa.D.<br />

T.T.<br />

Gramsh<br />

Progradec<br />

Devo l<br />

Ohrid<br />

Lake<br />

N.N.F.<br />

W.G.N.F.S.<br />

Prespa<br />

Lake<br />

Korçë<br />

L.T.T.<br />

40.5°<br />

Nartës<br />

Lagoon<br />

Ap.A.<br />

Vlorë<br />

O<br />

s u m<br />

W.E.N.F.S.<br />

Vithkuq<br />

Tepelenë<br />

V j<br />

o<br />

s a<br />

Ersekë<br />

Permët<br />

40°<br />

0 50 km<br />

42°<br />

41°<br />

40°<br />

19° 20° 21°<br />

Tirana<br />

Ionian<br />

Sea<br />

Reverse Fault<br />

Normal Fault<br />

39.5°<br />

19° 20°<br />

Sarandë<br />

Strike-slip Fault<br />

Anticline Fold<br />

K.N.F.S.N.N.F.<br />

Aristi<br />

Ioannina<br />

Borders<br />

Konitsa<br />

P.W.N.F.<br />

K.N.F.<br />

Kipi<br />

Scale<br />

0 20km<br />

Figure 1. Neo-tectonic map of the Southern Albania and North-western Greece (Modified from Aliaj et al., 2000 and<br />

Carcaillet, et al, 2009). Areas located within 200 m above Sea level are in light grey. Dark grey stars indicate published data<br />

(Lewin et al., 1991; Hamlin et al., 2000; Woodward et al., 2001; Carcaillet et al., 2009) and red stars indicate data computed<br />

in the present study. Fault and fold types are described in the picture caption and the overall current tectonic deformation is<br />

represented by black arrows (from Jouanne et al., submitted). (C.p.) Coastal pop-up; (Ar.A.) Ardenica Anticline; (Ap.A.)<br />

Apollonia Anticline; (S.A.) Shkumbin Anticline; (T.T.) Tomorrica Thrust; (L.T.T.) Lushnje - Tepelenë Thrust; (B.N.F.) Bulcar<br />

Normal Fault; (E.G.) Elbasan Graben; (N.N.F.) Nerotrivi Normal Fault; (K.N.F.S.) Konitsa Normal Fault Systems; (P.W.N.F.)<br />

Papingo West Normal Fault; (K.N.F.) Kipi Normal Fault; (W.G.N.F.S.) West Graben Normal Fault Systems; (W.E.N.F.S.)<br />

West Ersekë Normal Fault Systems; (PaD) PaleoDevoll.<br />

21°<br />

METHODS AND RESULTS<br />

The terraces has been mapped from field work,<br />

satellite image and 1/25000 topo sheets. Terrace<br />

thicknesses and elevations above the present-day<br />

river bed have been measured using a laser range<br />

distancemeter. Terrace age control was obtained<br />

from data previously published by Lewin et al. (1991),<br />

Hamlin et al. (2000), Woodward et al. (2001) and<br />

Carcaillet et al. (2009) as well as from in situ<br />

67


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Incision rate (mm/yr)<br />

produced<br />

10 Be and radiocarbon ( 14 C) dating<br />

specifically performed for this study.<br />

We dated organic residues (thirteen 14 C ages) and<br />

quartz rich pebbles collected into layers developed<br />

during the last flooding event and considered the<br />

value as the age of abandonment of the alluvial<br />

terrace. This implies that the age of the terrace<br />

abandonment is contemporaneous to the age of<br />

deposition of the highest terrace layer. The samples<br />

for cosmonuclide (in situ produced<br />

10 Be) age<br />

determinations have been collected on siliceous rich<br />

pebbles along depth-profiles of two terraces of the<br />

lower Skhumbin and one on the top of the upper<br />

Osum.<br />

West<br />

5<br />

4.5<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Post LGM<br />

Pre LGM<br />

Linear regression Post LGM<br />

by tectonic blocks<br />

120 145 170<br />

Distance from the Sea (km)<br />

Figure 3. Evolution of the incision rate<br />

along the river profile of the Vjosa. bed. In<br />

the river profile are located the Lushnje -<br />

Tepelenë Thrust (L.T.T.), Nerotrivi Normal<br />

fault (N.N.F.), Konitsa Normal Fault<br />

Systems (K.N.F.S.) Papingo West Normal<br />

Fault (P.W.N.F.).<br />

West<br />

T.T.<br />

28.6 m<br />

1.7 m/ka<br />

Incision<br />

Incision<br />

rate<br />

rate<br />

(mm/yr)<br />

(m/ka)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

W.E.N.F.S.<br />

35 m<br />

2.1 m/ka<br />

195 220<br />

West<br />

1.2<br />

Post LGM<br />

1 Pre LGM<br />

Incision rate has been calculated by dividing the<br />

elevation of the terraces above the river by the age of<br />

the upper surface of the terraces; the vertical motions<br />

of the faults have been deduced from the local<br />

change of the incision rates.<br />

Our study has allowed estimating the vertical slip rate<br />

of actives structures in southern Albania ( Fig, 2, 3<br />

and 4, Table 1). These results confirm an important<br />

uplift rate, up to 2 mm/year (Carcaillet, et al. 2009),<br />

for the central Albania and the current activity of the<br />

Plio-Quaternary folds of external Albanides. They<br />

also show significant Quaternary displacement rates<br />

along normal faults of inner Albanides. Our results<br />

are in agreements with the E-W to N-S extension<br />

deduced from the GPS network.<br />

East<br />

1000<br />

500<br />

0<br />

Elevation above<br />

Sea level (m)<br />

3m<br />

0.2 m/ka<br />

Figure 2. Evolution of the incision rate along the<br />

river profile of the Osum. Left axis is the<br />

calculated incision rate; right axis is the elevation<br />

of the present-day river bed. Colour of the<br />

diamonds refers to the different terraces units.<br />

Into the lower box, appear the actual river profile<br />

where are located the actives faults. Bold lines<br />

represent normal faults belong of West Ersekë<br />

Normal Fault Systems (W.E.N.F.S.) producing<br />

surface displacements and the dashed lines<br />

represent the Tomorrica Thrust (T.T.) which<br />

would not influence in the evolution of the river<br />

profile.<br />

Konitsa Basin<br />

K.N.F.S.<br />

East<br />

L.T.T.<br />

N.N.F. P.W.N.F.<br />

0<br />

60 85 110 135 160 185<br />

>0.4m/ka<br />

1000<br />

500<br />

Elevation above<br />

Sea level (m)<br />

East<br />

2.0<br />

Pre LGM<br />

Incision rate (mm/yr)<br />

Incision rate (m/ka)<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0<br />

Linear regression Pre LGM<br />

by tectonic blocks<br />

23 m<br />

0.9 m/ka<br />

5m<br />

0.2 m/ka<br />

Elbasan Basin<br />

>1 m/ka<br />

11 m<br />

0.4 m/ka<br />

9m<br />

0.3 m/ka<br />

?<br />

?<br />

> 0.8 m/ka<br />

Korça Basin<br />

L.T.T.<br />

Lower Shkumbin PaleoDevoll Devoll<br />

S.A.<br />

E.G.<br />

T.T. B.N.F. W.G.N.F.S.<br />

25 50 75 100 125 150 175<br />

Figure 4. Evolution of the incision rate along the combined river profile of the lower Shkumbin and Devoll. In the river profile are<br />

located the Lushnje - Tepelenë Thrust (L.T.T.), the Shkumbin Anticline (S.A.), Tomorrica Thrust (T.T.), the Burcal Normal Fault<br />

(B.N.F.) and the West Graben Normal Fault Systems (W.G.N.F.S.) producing surface displacements.<br />

800<br />

400<br />

0<br />

Elevation above<br />

Sea level (m)<br />

68


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Table 1. Kinematic of active structures active analyzed from the incision of the terraces levels of the Vjosa, Osum, Devoll and<br />

Shkumbin rivers.<br />

Name Structure River Vertical slip rate (mm/yr)<br />

Coastal pop - up PaleoSeman ~ 0.1<br />

Shkumbin Anticline Shkumbin ~ 0.2<br />

Elbasan Graben Shkumbin > 1<br />

Lushnje - Tepelenë Thrust Shkumbin ~ 0.9<br />

Lushnje - Tepelenë Thrust Vjosa ~ 0.2<br />

Tomorrica Thrust Devoll ~ 0.4<br />

Burcal Normal Fault Devoll ~ 0.3<br />

Nerotrivi Normal Fault Vjosa > 0.4<br />

Konitsa Normal Fault Systems Vjosa > 0.4<br />

Papingo West Normal Fault Vjosa ~1.8 between 25 and 17 Ka (Carcaillet, et al, 2009)<br />

West Graben Normal Fault Systems Devoll > 0.8<br />

West Ersekë Normal Fault Systems Osum 1.7 – 2.1<br />

SEISMICITY, PRESENT-DAY DEFORMATION AND<br />

ACTIVE TECTONICS<br />

Current tectonics of Albania is characterized by an<br />

important microseismicity, small and medium size<br />

earthquake and a few large events as shown by the<br />

occurrence of earthquakes with magnitudes<br />

exceeding 6 during the last century (Jouanne et al.,<br />

submitted), (e.g. 1905, Shkodra earthquakes Ms=6.6,<br />

1911 Ohrid lake earthquake Ms = 6.7, 1920 Tepelena<br />

event Ms=6.4, 1926 Durres earthquake Ms=6., 1967<br />

Dibra earthquake Ms=6.6and 1979 Montenegro<br />

earthquake Ms=6.9).<br />

The strongest earthquakes occur in three well-defined<br />

seismic belts: a) The Ionian-Adriatic coastal<br />

earthquake belt at the eastern margin of the Adria<br />

microplate, which trends NW-SE represents the most<br />

seismically active zone in the country; b) The<br />

Peshkopia-Korça earthquake belt, which trends N-S,<br />

and c) The Elbasani-Dibra earthquake belt, which<br />

trends N-E, (Sulstarova et al. 2000; Aliaj, 2000).<br />

Focal mechanisms (Sulstarova et al., 2000, Aliaj,<br />

2000) and geodetic data (Jouanne et al., submitted),<br />

fit the geodynamics setting with a NE-SW<br />

compression across external Albanides and E-W to<br />

NNW-SSE extension in the internal Albanides.<br />

The comparison of our results and the deformation<br />

deduced from the GPS network (jouanne et al.,<br />

submitted) indicates that:<br />

a) The horizontal shortening linked to the active thrust<br />

faults exceeds 2 mm/yr (assuming a mean dip of 30°<br />

for the thrusts) and is close to the deformation rates<br />

deduced from GPS ; b) the extension linked to normal<br />

fault system would be in the order of 1 mm/yr<br />

(assuming a dip of 60° for the normal faults) and is<br />

smaller than the deformation rates deduced from<br />

GPS; c) The normal faulting linked to the East<br />

Albanian graben systems seems more active than the<br />

normal faulting that affects the North West part of<br />

Greece.<br />

Acknowledgements: The authors thank the NATO SFP<br />

977993 and the Science for Peace team to have supported<br />

this work.<br />

References<br />

Aliaj, S. (2000). Neotectonics and seismicity in Albania, in:<br />

Geology of Albania, edited by: Meco, S., Aliaj, S., and<br />

Turku, I., Beitr. Regional. Geol. Erde, 28, 135–178.<br />

Baker, C., Hatzfeld, D., Lyon-Caen, H., Papadimitriou, E.,<br />

Rigo, A. (1997). Earthquake mechanisms of the Adriatic<br />

Sea and Western Greece: implications for the oceanic<br />

subduction-continental collision transition. Geophysical<br />

Journal International 131, 559-594.<br />

Carcaillet, J., Mugnier, J.L., Koçi, R., Jouanne, F. (2009).<br />

Uplift and active tectonics of southern Albania inferred<br />

from incision of alluvial terraces. Quaternary Research<br />

71, 465-476.<br />

Goldsworthy, M., Jackson, J., Haines, J. (2002). The<br />

continuity of active fault systems in Greece. Geophysical<br />

Journal International 148(3), 596-618.<br />

Hancock, G., Anderson, R. (2002). Numerical modeling of<br />

fluvial strath-terrace formation in response to oscillating<br />

climate. Geological Society of America Bulletin 114(9):<br />

1131-1142<br />

Hamlin, R.,Woodward, J., Black, S., Macklin, M.G. (2000).<br />

Sediment fingerprinting as 580 a tool for interpreting longterm<br />

river activity: the Voidomatis basin, NWGreece. In:<br />

581 Foster, I.D.L. (Ed.), Tracers in Geomorphology.<br />

Wiley, Chichester, 473–501.<br />

Jouanne, F., Bushati, S., Mugnier, J.L., Shinko, I., Pasha,<br />

M., Koci, R. (Submitted). GPS constrains on current<br />

tectonics of Albania. Geophysical Research Letters.<br />

Lewin, J., Macklin, M.G., Woodward, J.C. (1991). Late<br />

Quaternary fluvial sedimentation in the Voidomatis basin,<br />

Epirus, Northwest Greece. Quaternary Research 35, 103-<br />

115.<br />

Merritts, D.J., Vincent, K.R., Wohl, E.E. (1994). Long river<br />

profiles, tectonism, and eustasy: a guide to interpreting<br />

fluvial terraces. Journal of Geophysical research, 99,<br />

14031-14050.<br />

Niewland, D.A., Oudemayer, B.C., Valbona, U. (2001). The<br />

tectonic development of 616 Albania: explanation and<br />

prediction of structural styles. Marine and Petroleum 617<br />

Geology 18, 161–177.<br />

Ouchi, S. (1985) Response of alluvial rivers to slow active<br />

tectonics movements, Geol. Soc. Am. Bull., 96, 504-515.<br />

Pazzaglia, F. J. River Terraces, in Wohl, E., ed., Treatise of<br />

Geomorphology, Elsevier. Submitted<br />

Sulstarova, E., Peçi, V., Shuteriqi, P. (2000). Vlora-<br />

Elbasani-Dibra (Albania) transversal fault zone and its<br />

seismic activity. Journal of Seimology 4, 117-131.<br />

Roure, F., Nazaj, S., Mushka, K., Fili, I., Cadet, J.P.,<br />

Bonneau, M. (2004). Kinematic evolution and petroleum<br />

systems - An appraisal of the Outer Albanides. In: Mc<br />

Clay K.R. (Ed.), Thrust tectonics and hydrocarbon<br />

systems. AAPG memoir 82, 474-493.<br />

69


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

LECHAION, THE ANCIENT HARBOUR OF CORINTH (PELOPONNESE, GREECE)<br />

DESTROYED BY TSUNAMIGENIC IMPACT<br />

Hadler, Hanna (1, Andreas Vött (1), Benjamin Koster (2), Margret Mathes-Schmidt (2), Torsten Mattern (3), Konstantin<br />

Ntageretzis (1), Klaus Reicherter (2), Dimitris Sakellariou (4), Timo Willershäuser (1)<br />

(1) Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong> <strong>Mainz</strong>, Johann-Joachim-Becher-Weg 21, 55099 <strong>Mainz</strong>, GERMANY.<br />

E-mail: hadler@uni-mainz.de<br />

(2) Neotectonics and Natural Hazards Group, RWTH Aachen, Lochnerstr. 4-20, 52056 Aachen, GERMANY.<br />

(3) Faculty of Classical Archaeology, University of Trier, <strong>Universität</strong>sring 15, 54296 Trier, GERMANY.<br />

(4) Institute of Oceanography, Hellenic Centre for Marine Research, 47 km Athens-Sounio Ave., 19013 Anavyssos, GREECE.<br />

Abstract: Lechaion, the harbour of ancient Corinth, is situated at the south-eastern extension of the Gulf of Corinth (Peloponnese,<br />

Greece). Due to extensive fault systems dominating the gulf, seismic activity is frequent and often related to landslides or<br />

submarine mass movements. Thus, the study area is highly exposed to tsunami hazard. By means of geo-scientific studies<br />

comprising geomorphological, sedimentological and geophysical methods, evidence of multiple palaeotsunami impact was<br />

encountered at the Lechaion harbour site and the surrounding coastal area. The detected tsunami signatures include<br />

allochthonous marine sediments intersecting quiescent harbour deposits, extensive units of tsunamigenic beachrock and geoarchaeological<br />

destruction layers. Our results suggest that the harbour at Lechaion was finally destroyed in the 6 th century AD by<br />

strong tsunami impact.<br />

Key words: Palaeotsunami, beachrock, ancient harbours, Lechaion<br />

INTRODUCTION<br />

Situated in the eastern Mediterranean, the Gulf of<br />

Corinth belongs to one of the seismically most active<br />

regions in the world. Due to ongoing continental<br />

rifting with high extension rates up to 16 mm/yr, this<br />

half-graben structure is surrounded by active onshore<br />

and submarine faults (Papazachos & Dimitriu<br />

1991; Sachpazi et al., 2003; Avallone et al., 2004).<br />

Strong earthquakes, often accompanied by<br />

coseismic displacement of the seafloor and/or<br />

inducing submarine slides are therefore common<br />

throughout the gulf. Water depths of maximum 900<br />

m, steep submarine slopes and a narrow shelf<br />

additionally enhance the potential of strong tsunamis<br />

(Hasiotis et al., 2002; Stefatos et al., 2006).<br />

Accordingly, numerous tsunami catalogues for the<br />

Mediterranean and especially for Greece report on<br />

historical tsunami events that occurred within the Gulf<br />

of Corinth (Soloviev et al., 2000; Papadopoulos,<br />

2003; Ambrasey & Synolakis, 2010). Solely for the<br />

20 th century, six tsunamis have been recorded; the<br />

most destructive occurred near Aeghio in 1963<br />

(Papadopoulos, 2003).<br />

The main objectives of our study were to (i) detect<br />

allochthonous high-energy event layers within the<br />

sedimentary record of the harbour basin at Lechaion<br />

and (ii) to reconstruct the late-Holocene coastal<br />

evolution of the study area.<br />

LECHAION – THE<br />

CORINTH<br />

ANCIENT HARBOUR OF<br />

Lechaion, the western harbour of ancient Corinth, is<br />

situated at the Isthmus of Corinth, which connects<br />

the Peloponnese with mainland Greece. The harbour<br />

was most probably founded around 600 BC when the<br />

Corinthians expanded their military and trading<br />

activities (Stiros et al., 1996). According to ancient<br />

sources, Lechaion served as a naval base from the<br />

4 th century BC and was connected to the city of<br />

Corinth by a harbour road protected by strong walls<br />

(Xen. Hell. 4.4.6-8 after Brownson, 1918). With the<br />

devastation of Corinth in 146 BC, Lechaion was<br />

abandoned but re-activated under Roman supremacy<br />

in 44 BC. In Roman times, the harbour underwent<br />

different phases of reconstruction and mainly served<br />

as trading base (Strab. Geogr. 8.6.20-23 after<br />

Hamilton & Falconer, 1903). The final abandonment<br />

of Roman Lechaion is reported to be associated with<br />

the destruction of ancient Corinth by a series of<br />

strong earthquakes in 521 or 551 AD. Though<br />

occasionally re-used in medieval times, the harbour<br />

never regained its former importance (Rothaus,<br />

1995).<br />

Today, certain harbour installations are still visible,<br />

including two outer moles, an entrance channel<br />

leading to an elongated inner harbour basin and the<br />

remains of an ancient quay wall (Paris, 1915). The<br />

present day topography is dominated by large<br />

mounds of sediments dredged from the harbour<br />

basin (Fig. 1). Between the inner harbour basin and<br />

the present beach the remains of an early Christian<br />

basilica, dating to the late 5 th century AD, were<br />

70


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

excavated. The basilica, 186 m long, represents the<br />

largest from this period and was completely<br />

destroyed by the 521 or 551 AD earthquake series<br />

(Krautheimer, 1989; Rothaus, 1995).<br />

Fig. 1: (a) Location of study area in Greece. (b)<br />

Overview of the Corinthia. Red box indicates harbour<br />

site of Lechaion, red dashes indicate beachrock. (c)<br />

Map of the Lechaion harbour basin and adjacent area.<br />

Vibracores LEC 1-3 are indicated by red dots, ERTmeasurements<br />

by red dashed lines and GRPmeasurements<br />

by white lines (maps modified after<br />

Google Earth images, 2009).<br />

TSUNAMI IMPACT AT THE LECHAION HARBOUR<br />

SITE<br />

Sedimentary evidence from the harbour basin<br />

As previous studies show, tsunami impact may be<br />

documented by layers of allochthonous coarsegrained<br />

sediment intersecting quiescent near-shore<br />

deposits (Vött et al. 2009). Low-energy environments<br />

such as harbour basins therefore represent excellent<br />

geological archives for palaeotsunami research.<br />

Sedimentological, geochemical and micropalaeontological<br />

studies were carried out at the Lechaion<br />

harbour site (see Fig. 1 for location of vibracores).<br />

The local stratigraphy comprises shallow marine<br />

(pre-harbour) deposits overlain by harbour-related,<br />

quiescent lagoonal sediments followed by limnic to<br />

terrestrial deposits accumulated when the harbour<br />

was already out of use (Fig. 2). The stratigraphical<br />

record of all vibracores is repeatedly interrupted by<br />

layers of allochthonous coarse-grained sediment and<br />

shell debris, partly characterized by fining upward<br />

sequences. These ex situ-layers extend up to 450 m<br />

inland. Erosional unconformities at the base (Fig. 2b)<br />

and immediately re-established quiescent conditions<br />

on top of the event layers indicate short term highenergy<br />

interference of the Lechaion harbour site.<br />

Fig. 2: (a) Vibracore LEC 2 (N 37° 55’ 58.6 E 22° 53’<br />

03.5’’, -0.04 m a.s.l.) was drilled in the central<br />

Lechaion harbour basin. Here, two coarse-grained<br />

tsunami layers were found, intersecting the<br />

autochthonous, predominantly shallow-marine to<br />

lagoonal facies. Youngest tsunami deposit as<br />

encountered in core LEC 3 and covering early<br />

Christian basilica is dredged at site LEC 2 (see ).<br />

14<br />

C-samples are marked by *. (b) Detailed view of the<br />

younger event layer. The quiescent harbour<br />

environment was abruptly covered by allochthonous<br />

high-energy deposits (Photo by T. Willershäuser,<br />

2010).<br />

The event-stratigraphical correlation of all vibracore<br />

profiles reveals three distinct tsunami layers. The<br />

geochronological framework is based on radiocarbon<br />

dating and age determination of diagnostic ceramic<br />

fragments. Tsunamigenic impact could be timebracketed<br />

by radiocarbon dating to around 760 cal<br />

BC and 50 cal AD and to the 6 th century AD by<br />

geoarchaeological findings. The youngest event<br />

correlates to the geoarchaeological destruction layer<br />

presented below.<br />

Geoarchaeological destruction layers<br />

Large sediment mounds adjacent to the inner<br />

harbour basin are generally referred to as natural<br />

dunes or sediments obtained from dredging activity<br />

(Frazer, 1965; Rothaus, 1995; Stiros, 1996). They<br />

consist of sand and gravel, intermingled with<br />

numerous ceramic fragments and marine<br />

macrofossils. The sediment composition and grain<br />

size distribution thus exclude natural dune formation.<br />

However, gravel is also completely atypical for the<br />

siltation of a quiescent harbour basin environment<br />

(Marriner & Morhange 2007). According to their grain<br />

size, the sediments provide evidence of high-energy<br />

influence. Incorporating marine as well as<br />

terrigeneous material, they indicate strong<br />

tsunamigenic inundation and backflow affecting the<br />

coastal plain at Lechaion. Since associated to<br />

dredging activities, the mounds therefore document<br />

tsunami influence burying the Lechaion harbour<br />

basin.<br />

Widespread burial of the harbour site by high-energy<br />

influence must also be assumed regarding the ruins<br />

71


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

of the 5 th century AD Christian basilica. The entire<br />

foundation like the adjacent area is completely<br />

covered by an up to 2 m thick sediment layer. The<br />

surrounding topography as well as the sediment<br />

composition, however, exclude a colluvial, fluvial or<br />

mass denudative burial of the site. The deposits are<br />

also not related to the archaeological excavation of<br />

the site. A former door lintel used as a threshold after<br />

the destruction of the church attests that the entrance<br />

level was already elevated during historic times.<br />

Made up of a sandy matrix incorporating abundant<br />

gravel, numerous ceramic fragments and marine<br />

macrofossils, the high-energy sediment layer again<br />

reflects tsunami backflow.<br />

probable cause for the final destruction and<br />

abandonment of the harbour.<br />

Beachrock as indicator for tsunami impact<br />

Along the Corinthian coastline, extensive beachrock<br />

formations dominate the present coastal<br />

geomorphology. As recently described by Vött et al.<br />

(2010), beachrock deposits may be of tsunamigenic<br />

origin. At the Corinth Canal at a distance of 7 km<br />

from Lechaion, the beachrock complex extends up to<br />

300 m inland (Fig. 4a), while the maximum extension<br />

of the recent beach is less than 25 m. The deposits<br />

show a clearly laminated structure with multiple fining<br />

upward sequences reaching from gravel to fine sand.<br />

The inner structure reveals a landward orientated<br />

imbrication of the gravel components (Fig. 4b)<br />

documenting landward flow dynamics as may be<br />

induced during tsunami inundation. According to their<br />

sedimentary structure, these beachrock deposits<br />

must not automatically be regarded as lithified beach<br />

but may rather represent calcified tsunamigenic<br />

deposits (for further discussion see Vött et al. 2010).<br />

The “diolkos”, an ancient slipway across the Isthmus<br />

of Corinth, is partly covered by beachrock; used until<br />

the early 1 st century AD, the diolkos represents a<br />

terminus post quem for the tsunamigenic impact.<br />

Fig. 3: (a) Excavated remains of the 5 th century AD<br />

Christian basilica erected to the north of the former<br />

harbour basin. (b) The present day ground surface<br />

overtops the basilica’s original ground level by up to 2<br />

m (in photo: ca. 1.2 m). (c) The sedimentary cover<br />

consists of sand and gravel with abundant ceramic<br />

fragments and marine shell debris.<br />

Several earth resistivity transects were carried out<br />

around the ancient harbour basin (Fig. 1c). They<br />

revealed a sharp boundary between the allochthonous<br />

high-energy deposits and the underlying<br />

autochthonous fine-grained harbour sediments as<br />

well as a clear thinning landward of the event layer<br />

up to 450 m inland. Furthermore, ground penetrating<br />

radar profiles (Fig. 1c) revealed channel-like<br />

structures at their base. Orientated perpendicular<br />

towards the coastline, these incised channels<br />

indicate strong linear erosion in land-/seaward<br />

direction. Similar structures, created by strong<br />

backflow processes, were observed along the<br />

Chilean coast after the February 2010 Chile tsunami<br />

(Bahlburg & Spiske 2010).<br />

In a summary view, our results give evidence for the<br />

influence of tsunami impact on the Lechaion harbour.<br />

According to Rothaus (1995), only few ceramic<br />

fragments younger than late Roman to early<br />

Byzantine times were found in the vicinity of the<br />

harbour basin and basilica. Thus, we consider<br />

tsunami landfall in the 6 th century AD as the most<br />

Fig. 4: (a) Extensive beachrock deposits at the<br />

Corinth Canal. (b) The internal structure shows a<br />

fining upward sequence as well as imbricated pieces<br />

of gravel.<br />

CONCLUSION<br />

Based on our studies, multiple tsunami impact was<br />

determined for the Lechaion harbour site and<br />

adjacent coastal areas. At least three distinct event<br />

layers were identified, the youngest and obviously<br />

most destructive event dating to the 6 th century AD.<br />

The spatial distribution and geomorphological<br />

variability of the encountered tsunami deposits<br />

require an event-stratigraphical approach to<br />

understand and reconstruct the chronology of events.<br />

Based on the combination of historical accounts with<br />

geomorphological, sedimentological, geophysical and<br />

geoarchaeological data we conclude that the ancient<br />

harbour of Lechaion, though influenced by tsunamis<br />

before, was finally destroyed by tsunami impact,<br />

most probably triggered during the 521 or 551 AD<br />

earthquake series. The present day topography of<br />

the harbour site is due to a latter re-activation of the<br />

72


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

harbour in medieval or pre-modern times and thus<br />

only partly represents the ancient Corinthian harbour.<br />

Acknowledgement: We acknowledge funding of the<br />

project by the German Research Foundation, DFG (Bonn,<br />

Gz. VO 938/3-1). Sincere thanks are due to the 37 th<br />

Ephorate of Prehistoric and Classical Antiquities, Corinth,<br />

for support during field work and discussion.<br />

References<br />

Ambrasey, N. & C. Synolakis, (2010). Tsunami catalogs for<br />

the eastern Mediterranean, revisited. Journal of<br />

Earthquake Engineering 14, 309-330.<br />

Avallone, A., P. Briole, A.M. Agatza-Balodimou, H. Billiris,<br />

O. Charade, C. Mitsakaki, A. Nercessian, K. Papazissi,<br />

D. Paradissis & G. Veis, (2004). Analysis of eleven years<br />

of deformation measured by GPS in the Corinth Rift<br />

Laboratory area. Competes Rendus Geosciences 336 (4-<br />

5), 301-311.<br />

Bahlburg, H. & M. Spiske, (2010). The February 27, 2010<br />

Chile Tsunami. Sedimentology of runup and backflow<br />

deposits at Isla Mocha. American Geophysical Union,<br />

Abstract #OS42B-06.<br />

Brownson, C.L. (ed.), (1918). Xenophon, Hellenica. Harvard<br />

University Press. Cambridge.<br />

Frazer, J.G., (1965). Pausanias’s description of Greece.<br />

Biblo and Tannen. New York. 652 p.<br />

Hamilton, H.C. & W. Falconer, (1903). The Geography of<br />

Strabo. George Bell & Sons. London.<br />

Hasiotis, T., G. Papatheodorou, G. Bouckovalas, C. Corbau<br />

& G. Ferentinos, (2002). Earthquake-induced coastal<br />

sediment instabilities in the western Gulf of Corinth,<br />

Greece. Marine Geology 186, 319-335.<br />

Krautheimer, R., (1989). Early Christian and Byzantine<br />

Architecture. Penguin Books. London. 556 p.<br />

Marriner, N. & C. Morhange, (2007). Geoscience of ancient<br />

Mediterranean harbours. Eart-Science Reviews 80, 137-<br />

194.<br />

Papadopoulos, G.A., (2003). Tsunami hazard in the eastern<br />

Mediterranean: strong earthquakes and tsunamis in the<br />

Corinth Gulf, Central Greece. Natural Hazards 29, 437-<br />

464.<br />

Papazachos, B.C. & P.P. Dimitriu, (1991). Tsunamis in and<br />

near Greece and their relation to the earthquake focal<br />

mechanism. Natural Hazards 4, 161-170.<br />

Paris, J., (1915). Contributions à l’étude des portes antiques<br />

du mond grec. Notes sur Lechaion. Bulletin de<br />

Correspondance Hellénique 39 (1), 5-16.<br />

Rothaus, R., (1995). Lechaion, western port of Corinth: a<br />

preliminary archaeology and history. Oxford Journal of<br />

Archaeology 14 (3), 293-306.<br />

Sachpazi, M., C. Clément, M. Laigle, A. Hirn & N. Roussos,<br />

(2003). Rift structure, evolution, and earthquakes in the<br />

Gulf of corinth, from reflection seismic images. Earth and<br />

Planetary Science Letters 216, 243-257.<br />

Soloviev, S.L., O.N. Solovieva, C.N. Go, K.S. Kim & N.A.<br />

Shchetnikov, (2000). Tsunamis in the Mediterranean Sea<br />

2000 BC – 2000 AD. Kluwer Academic Publishers.<br />

Dordrecht. 237 p.<br />

Stefatos, A., M. Charalambakis, G. Papatheodorou & G.<br />

Ferentinos, (2006). Tsunamigenic sources in an antive<br />

European half-graben (Gulf of Corinth, Central Greece).<br />

Marine Geology 232, 35-47.<br />

Stiros, S., P. Pirazolli, R. Rothaus, S. Papageorgiou, J.<br />

Laborel & M. Arnold, (1996). On the date of construction<br />

of Lechaion, western harbor of Ancient Corinth, Greece.<br />

Geoarchaeology: An International Journal 11 (3), 251-<br />

263.<br />

Vött, A., H. Brückner, S.M. May, D. Sakellariou, O. Nelle, F.<br />

Lang, V. Kapsimalis, S. Jahns, R. Herd, M. Handl & I.<br />

Fountoulis, (2009). The Lake Voulkaria (Akarnania, NW<br />

Greece) palaeoenvironmental archive – a sediment trap<br />

for multiple tsunami impact since the mid-Holocene.<br />

Zeitschrift für Geomorphologie N.F., Suppl. Issue 53 (1),<br />

1-37.<br />

Vött, A., G. Bareth, H. Brückner, C. Curdt, I. Fountoulis, R.<br />

Grapmayer, H. Hadler, D. Hoffmeister, N. Klasen, F.<br />

Lang, P. Masberg, S.M. May, K. Ntageretzis, D.<br />

Sakellariou & T. Willershäuser, (2010). Beachrock-type<br />

calcarenitic tsunamites along the shores of the eastern<br />

Ionian Sea (western Greece) – case studies from<br />

Akarnania, the Ionian Islands and the western<br />

Peloponnese. Zeitschrift für Geomorphologie N.F., Suppl.<br />

Issue 54 (3), 1-50.<br />

73


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

STRUCTURAL CHARACTERISTICS AND EVOLUTION OF THE YANGSAN-ULSAN FAULT<br />

SYSTEM, SE KOREA<br />

S.-R. Han (1), M. Lee (1), J. Park (2), Y.-S., Kim (1*)<br />

(1) Dept. of Environmental Geosciences, Pukyong National University, Busan 608-737, Korea. *Email: ysk7909@pknu.ac.kr<br />

(2) GeoGeny Consultants Group Inc. 807-2 Bangbae-Dong Seocho-Gu, Seoul 137-831, Korea.<br />

Abstract (Structural characteristics and evolution of the YangSAN-Ulan fault system, SE Korea): The Yangsan-Ulsan Fault<br />

System (YUFS) is a major fault system in Korea. More than 40 Quaternary faults have been discovered around the YUFS. To<br />

understand the evolution and structural characteristics of the YUFS, we performed numerical modelling for the evolution of the<br />

YUFS and fault zone analysis on the eastern part of the Ulsan fault. The result of the modelling shows that the YUFS evolved into<br />

-fault, a low angle merging fault system. In a fault zone analysis, the hanging wall block of the fault shows relatively higher<br />

damage with higher fracture density and deformation structures compared with the footwall. This indicates that the hanging wall is<br />

more susceptible to deformation than the footwall during faulting events; this result is consistent with damage patterns in other<br />

Quaternary faults. Epicenters of recent earthquakes around the study area are also concentrated on the hanging wall of the Ulsan<br />

fault. This kind of study can help us to evaluate hazards resulting from future potential earthquakes in and around active fault<br />

systems.<br />

Key words: Yangsna-Ulsan fault system,<br />

-fault , Quaternary faults<br />

INTRODUCTION<br />

The Korean Peninsula is generally considered as a<br />

relatively safe region considering earthquake<br />

hazards, because it is located away from plate<br />

boundaries. However, many moderate and minor<br />

earthquakes occur around the Yangsan-Ulsan fault<br />

system (YUFS), in the southeastern part of the<br />

Korean Peninsula (Lee and Yang, 2006). Recently,<br />

more than 40 Quaternary faults were discovered near<br />

the YUFS (Jin et al., 2011).<br />

There are two nuclear power plant sites near the<br />

YUFS (Fig.1). To prevent nuclear catastrophe due to<br />

a potential large scale earthquake, it is necessary to<br />

understand the deformation characteristics and future<br />

evolution of the YUFS. For this purpose, a numerical<br />

modelling related to the evolution of the YUFS and a<br />

detailed field investigation on a reactivated fault were<br />

carried out in this study.<br />

STUDY AREA<br />

The study area is located in the Cretaceous<br />

Gyeongsang Basin, SE Korea (Fig. 1). The basement<br />

of the study area is composed of Cretaceous<br />

sedimentary rocks with later igneous and volcanic<br />

rocks. The Yangsan fault is a NNE-SSW trending<br />

right-lateral dominant strike-slip fault, and the NNW-<br />

SSE trending Ulsan fault is interpreted as a strike-slip<br />

fault overprinted by later reverse-slip. Recently, more<br />

than 40 Quaternary faults have been discovered near<br />

the Yangsan and Ulsan faults. Historical records<br />

show many earthquakes in this region, with an<br />

earthquake producing over 100 casualties recorded<br />

Fig. 1: Regional geologic map of Gyeongsang basin,<br />

the SE part of the Korean peninsula and Quaternary<br />

faults (circles) around the Yangsan and Ulsan faults.<br />

Two Nuclear Power plants are located near the YUFS<br />

(modified from Lee, 2000)<br />

in A.D. 779 (estimated M= 6.7), near the intersection<br />

of the two faults (Lee and Yang, 2006).<br />

74


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

<br />

Fig. 2: Comparison between coulomb stress changes, distribution of Quaternary faults and epicenters of earthquakes<br />

around the Yangsan-Ulsan fault system (Han et al., 2009). (a) Coulomb stress changes are concentrated in the eastern<br />

and southeastern region of the Yangsan fault and eastern region of the Ulsan fault in Quaternary. (b) Locations of<br />

Quaternary faults around the Yangsan-Ulsan fault system. (c) Earthquake distributions around the southeastern part of the<br />

Korean peninsula (1994-1998). Green circles indicate the epicenters of the recent earthquakes. Earthquakes are<br />

concentrated in the eastern part (hanging wall) of the Ulsan fault.<br />

EVOLUTION MODELLING OF THE YANGSAN-<br />

ULSAN FAULT SYSTEM<br />

Analysis of lineaments and geomorphology of the<br />

YUFS indicate that the YUFS is analogous to the<br />

characteristics of -fault. The concept of -fault was<br />

suggested by Du and Aydin (1995). This is the<br />

geometric relationship between two intersecting<br />

strike-slip faults. When a strike-slip fault propagates<br />

into a pre-existing strike-slip fault, the orientation of<br />

the propagation depends on the angle between the<br />

maximum principal stress and propagating fault. To<br />

understand the evolution of the fault system, a<br />

numerical modelling on the YUFS with regards to<br />

stress changes with time was performed, using the<br />

USGS open program Coulomb 3<br />

(http://earthquake.usgs.gov). This modelling was<br />

carried out based on a previously suggested tectonic<br />

evolution model for the YUFS (Park, 2004). The<br />

result of this numerical modelling indicates that the<br />

YUFS is a -fault. During the propagation of the<br />

Ulsan fault into the Yangsan fault in the late Miocene,<br />

coulomb stress increases at the north-western tip of<br />

the Ulsan fault. After the Ulsan fault connected to the<br />

Yangsan fault in Pliocene, a new segment was<br />

initiated at the northern part of the Ulsan fault due to<br />

stress changes. Since the penetration of the Ulsan<br />

fault into the Yangsan fault in Quaternary, coulomb<br />

stress increased in the north-eastern region of the<br />

Yangsan fault and the eastern region of the Ulsan<br />

fault. These regions are well consistent with the<br />

distribution of Quaternary faults and locations of<br />

epicenters of recent earthquakes around the YUFS<br />

(Fig. 2).<br />

FAULT GEOMETRY AND SLIP ANALYSIS OF A<br />

FAULT ZONE IN THE WEOLSEONG AREA<br />

In the SE part of Korea, two nuclear power plant<br />

(NPP) sites are located near the YUFS. In the new<br />

Weolseong NPP site located in the southeastern<br />

coast of Korea, an east dipping, N-S trending fault<br />

was discovered during construction. The main fault<br />

zone shows various different coloured fault gouge<br />

bands and shear fabrics indicating opposite slips (Fig.<br />

3). These features indicate that there were discrete<br />

stages of normal and reverse faulting events along<br />

this fault zone. The analyzed kinematic indicators,<br />

such as lineations, cleavages, and slickenlines in the<br />

fault zone suggest that the fault was initiated as a<br />

normal fault and was reactivated as a reverse fault<br />

under NW-SE compression. The hanging wall block<br />

of the fault shows relatively higher damage indicating<br />

higher deformation in the hanging wall than the<br />

footwall.<br />

CONCLUSION<br />

It is common that earthquakes occur by reactivation<br />

of pre-existing faults (e.g. Chen, 2002; Ota et al.,<br />

2004; Ota et al., 2005). It is well known that most of<br />

the damage from earthquake is concentrated on the<br />

hanging wall block during the faulting (Ota et al.,<br />

2005). A N-S trending, east dipping fault zone was<br />

discovered in the construction site of the new<br />

Weolseong NPP.Detailed analysis was performed on<br />

this fault zone and the result indicates multiple<br />

faulting events. According to the analysis, this fault<br />

zone initiated as a normal slip fault and it was<br />

75


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 3: Photomosaic (a), sketch (b) and sense indicators (c) of reactivated fault zone located at the new Weolseong nuclear<br />

power plant (Fig. 1).<br />

reactivated as a reverse fault. The damages in the<br />

wall rocks of this fault were concentrated on the<br />

hanging wall block rather than footwall block. This<br />

result indicates that the hanging wall block is subject<br />

to higher deformation than footwall block during<br />

faulting. This result is well consistent with the result<br />

from larger scale deformation characteristics around<br />

the YUFS producing intense distribution of<br />

Quaternary faults and earthquakes in the hanging<br />

wall of the Ulsan fault. Therefore, detailed analysis of<br />

fault zone and fault evolution modelling can<br />

contribute to understanding of the deformation<br />

characteristics of a fault zone and the evaluation of<br />

earthquake hazards.<br />

Therefore, the fault geometries and kinematics<br />

are important factors to evaluate conditions of<br />

important sites. Furthermore, fault damage zones<br />

(Kim et al., 2004) around faults are very important<br />

indications for the seismic hazard assessment<br />

due to secondary fractures and aftershocks (Kim<br />

and Sanderson, 2008).<br />

Acknowledgements: This work was funded by Korea<br />

Institute of Nuclear Safety under Grant M20702070001-<br />

08M0207-00110.<br />

References<br />

Lee, J.I. (2000) Provenance and thermal maturity of the<br />

lower Cretaceous Gyeongsang Supergroup, Korea. Ph.D.<br />

Thesis. Seoul National University.<br />

Lee, K. and Na, S.H. (1983) A study of microearthquake<br />

activity of the Yangsan fault. Journal of the Geological<br />

Society of Korea, 19, 127-135.<br />

Jin, K., Lee, M., Kim, Y.-S., Choi, J.-H. (2011).<br />

Archaeoseismological studies on historical heritage sites<br />

in the Gyeongju area, SE Korea. Quaternary International,<br />

doi:10.1016/j.quaint.2011.03.055.<br />

Lee, M. S., Kang, P. C. (1964) Geological map of Yangsan<br />

area. Geological Survey of Korea. ( In Korean with<br />

English abstract )<br />

Lee, K., Yang, W.-S. (2006). Historical seismicity of Korea.<br />

Bulletin of the Seismo-logical Society of America 96,<br />

846e855.Stiros, S. 1988b. Archaeology, a tool to study<br />

active tectonics – The Aegean as a case study. Eos,<br />

Transactions, American Geophysical Union 13, 1636-<br />

1639.<br />

Ota, Y., Watanabe, M., Suzuki, Y., Sawa, H. (2004).<br />

Geomorphological identification of pre-exiting active<br />

Chelungpu Fault in central Taiwan, especially its relation<br />

to the location of the surface rupture by the 1999 Chichi<br />

earthquake. Quaternary International 115–166, 155– 166.<br />

Ota, Y., Chen, Y.-G., Chen, W.-S., (2005). Review of<br />

paleoseismological and active fault studies in Taiwan in<br />

the light of the Chichi earthquake of September 21, 1999.<br />

Tectonophysics 508, 63-77.<br />

Park. J. (2004) The structural characteristics of the Tertiary<br />

basin in Gyeongju region, Korea. Master thesis. Seoul<br />

National University. ( In Korean with English abstract )<br />

Du, Y. and Aydin, A. (1995) Shear fracture patterns and<br />

connectivity at geometric complexities along strike-slip<br />

faults. Journal of Geophygical Research, 100, 18093-<br />

18102.<br />

Han, S.-R., Park. J., Kim, Y.-S. (2009) Evolution modelling<br />

of the Yangsan-Ulsan fault system with stress changes.<br />

Journal of the Geological Society of Korea. 45, 361-377. (<br />

In Korean with English abstract )<br />

Chen, Y.-G., Chen, W.S., Wang, Y., Lo, P.W., Lee, J.C., Liu,<br />

T.K.(b) (2002) Review of paleoseismological and active<br />

fault studies in Taiwan in the light of the Chichi<br />

earthquake of September 21, 1999.Tectonophysics 508,<br />

63-77.<br />

76


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

WHERE LANDSLIDES REPRESENT THE MOST IMPORTANT<br />

EARTHQUAKE-RELATED HAZARDS: THE MOUNTAIN AREAS OF CENTRAL ASIA<br />

Havenith, Hans-Balder (1)<br />

(1) Georisks and Environment group, Department of Geology, University of Liege. B18. 4000 Liege. BELGIUM. Email:<br />

HB.Havenith@ulg.ac.be<br />

Abstract (Where landslides represent the most important earthquake-related hazards: the mountain areas of Central<br />

Asia): The goal of this paper is to show which areas are particularly prone to earthquake-induced slope failure. In this regard, the<br />

paper considers the short- and long-term effects of geological, tectonic, climatic and morphological conditions. Case histories<br />

related to five earthquake events in the Tien Shan and Pamir Mountains are outlined: the earthquakes of Kemin in 1911, Sarez in<br />

1911, Khait in 1949, Gissar in 1989 and Suusamyr in 1992. The potential impact of seismically triggered mass movement affecting<br />

loess deposits is further pointed out through comparison with the 1920 Haiyuan (China) earthquake. One conclusion is that in<br />

semi-arid mountain regions marked by a strong seismic activity, such as those in Central Asia, seismogenic landslides and related<br />

long-term effects may represent the most important geohazards. Further, the susceptibility to seismic slope instability is highest<br />

along active fault zones and on convex slopes made of soft or fractured materials.<br />

Key words: landslides, co- and post-seismic effects, dynamics<br />

INTRODUCTION: THE PERCEPTION OF SEISMIC<br />

LANDSLIDE HAZARDS<br />

During the last ten years, after a series of disastrous<br />

earthquake events in mountain regions in Taiwan<br />

(1999), El Salvador (2001), Pakistan (2005) and<br />

China (2008), increasing attention has been<br />

addressed to landslides triggered by earthquakes.<br />

Previously, landslides have been considered as<br />

minor effects of earthquakes compared to the impact<br />

of the ground shaking itself. Schuster and Highland<br />

(2001) partly attributed the perception of the relatively<br />

small impact of earthquake-triggered mass<br />

movements to the fact that many related losses are<br />

often referred to as direct consequences of the<br />

earthquake. However, for the earthquake that hit the<br />

Kashmir mountains on October 8, 2005, Petley et al.<br />

(2006) estimated that about 30% of the total number<br />

of killed people (officially 87350), i.e. 26500, had<br />

been victims of co-seismic landslides. Less than<br />

three years later, on May 12, 2008, the Wenchuan<br />

earthquake hit the Sichuan and neighbouring<br />

provinces of China and caused ‘more than 15000<br />

geohazards in the form of landslides, rockfalls, and<br />

debris flows, which resulted in about 20000 deaths’<br />

(Yin et al. 2009). These casualties represent again<br />

almost 30 % of the official total number of fatalities of<br />

69197 of this event.<br />

In the remote areas of Central Asian mountains, M>6<br />

earthquakes generally do not cause catastrophes –<br />

but if they do, they do it through mass movements,<br />

like in the case of the M=8.2 Kemin, 1911, M=7.4<br />

Sarez, 1911, M=7.4 Khait, 1949, M=5.5 Gissar, 1989<br />

and M=7.3 Suusamyr, 1992 earthquakes (see<br />

location in Fig. 1). These case histories show that<br />

both rapid flows (mainly) in loess deposits and<br />

massive rock avalanches created the greatest<br />

disasters. To highlight the important issue related to<br />

landslides triggered in loess deposits, a comparison<br />

is made with one of the worldwide most catastrophic<br />

seismic events involving landslides: the M=7.8<br />

Haiyuan (or Gansu, 1920) earthquake in China.<br />

Special attention must also to be paid to post-seismic<br />

increase of landslide activity. In this regard, we have<br />

to take into consideration so-called ‘secondary or<br />

tertiary effects’ of earthquakes, such as landslide<br />

dams and related flooding impacts. The 2008<br />

Wenchuan earthquake clearly marked the<br />

importance of such effects – which could have killed<br />

another few thousands of people if efficient mitigation<br />

measures had not been taken by Chinese authorities.<br />

GEOLOGICAL, TOPOGRAPHIC AND CLIMATIC<br />

SETTING OF EARTHQUAKE-INDUCED SLOPE<br />

FAILURE<br />

Sassa (1996) observed that seismic landslide<br />

occurrence is strongly dependent on the proximity of<br />

the fault rupture. This observation is confirmed by<br />

many others made after the 1999 Taiwan, 2005<br />

Pakistan and 2008 Sichuan earthquakes.<br />

The geological factors have been extensively<br />

analysed by Keefer (1984) who concluded that any<br />

geological material with low geomechanical strength,<br />

be it soil or rock, may be susceptible to earthquakeinduced<br />

slope instability. The low strength may be<br />

related to weak cementation, intense weathering or<br />

fracturing, high water saturation or poor compaction.<br />

Concerning the structural elements (bedding,<br />

foliation, sediment-rock contact), no clear trend can<br />

be outlined. Especially for seismic rock slope failure,<br />

e.g., it is not clear if cross-bedding joints or bedding<br />

contacts support the development of a sliding surface<br />

77


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

or not. Here, clearly more research based on<br />

structural geology mapping is needed.<br />

Further, earthquake-induced landslides may be<br />

triggered from any surface morphology, in flat areas,<br />

such as lateral spreads, or on steep cliffs, such as<br />

rock falls. Still, several particularities can be outlined.<br />

From observations after the 2005 Pakistan<br />

earthquake, Sato et al. (2007) inferred that ‘there was<br />

a slight trend that large landslides occurred on<br />

vertically convex slopes rather than on concave<br />

slopes; furthermore, large landslides occurred on<br />

steeper (30° and more) slopes than on gentler<br />

slopes’. From my experience of earthquake-induced<br />

landslides in the Tien Shan (see below), it can be<br />

concluded that mainly the surface curvature has an<br />

influence on seismic slope stability at a global scale;<br />

particularly, hillcrests, higher parts of slopes and<br />

convex surface morphologies are prone to seismic<br />

slope failure. The influence of slope angle on seismic<br />

mountain regions marked by a high seismicity – such<br />

as the Tien Shan and the Pamir, but also the Hindu-<br />

Kush and others. Why semi-arid? If the climate is<br />

wet, the overall landslide hazards increase more than<br />

the specific hazard related to seismically triggered<br />

slope failure (e.g. Himalayas). On the contrary, if the<br />

climate is very dry – arid, the seismic hazards with<br />

direct impacts on building stability prevail (e.g.<br />

mountains in the Middle East, Southern Altai) since<br />

slope instability – also under seismic conditions -<br />

requires a certain amount of humidity in the top-soils<br />

to develop. In the Tien Shan, we observed that<br />

earthquakes, which occurred at the end of the dry<br />

summer season, caused relatively few slope failures<br />

(e.g., Ms=7.3 Suusamyr earthquake on August 19,<br />

1992; Ms=6.7 Nura earthquake on October 5, 2008);<br />

while, for instance, the Ms=7.4 Khait earthquake of<br />

July 10, 1949 – at the beginning of the dry highmountain<br />

summer - in northern Tajikistan triggered<br />

widespread slope instability.<br />

Fig. 1: Map of Tien Shan and Pamir Mountains in Central Asia with location of major faults and earthquakes (white filled<br />

circles show all recorded M>=7 earthquakes with the year of occurrence; the magnitude is indicated for analysed events) and<br />

related major mass movements (stars).<br />

slope stability is not clear; in some cases, especially<br />

in rocks, steeper slopes are more prone to instability;<br />

in others, especially in soft sediments, gentle slopes<br />

produce most of the mass movements, indicating that<br />

the combined effect of slope and geology has to be<br />

taken into consideration.<br />

The contribution of the climate to seismic slope<br />

failure has not yet been well investigated. According<br />

to our experience, the strongest contribution of<br />

earthquake-induced landslides to the total seismic<br />

and landslide hazards can be observed in semi-arid<br />

CASE HISTORIES FROM CENTRAL ASIA<br />

COMPARED WITH THE 1920 CHINA EVENT<br />

Nadim et al. (2006) assessed landslide and<br />

avalanche occurrence probabilities worldwide on the<br />

basis of morphological, geological, meteorological<br />

and seismological data. They clearly showed that all<br />

landslide hotspots are located in seismically active<br />

mountain ranges. For Central Asia, they estimate that<br />

global landslide hazard can be rated as medium to<br />

very high. They further noted that some areas in<br />

78


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Tajikistan are marked by highest mortality risk due to<br />

landslides.<br />

The following case histories document the landslide<br />

hazard and risk triggered by earthquakes in the Tien<br />

Shan and the Pamir Mountains. A comparison will be<br />

made with the 1920 Haiyuan (China) event to outline<br />

the most important factors contributing to landslide<br />

hazard and risk related to the presence of thick loess<br />

deposits in Central Asia.<br />

The Kemin earthquake, 1911<br />

The Kemin Ms=8.2 earthquake of 1911 (January 3) is<br />

one of the strongest events ever recorded in the Tien<br />

Shan. The earthquake caused extensive landsliding<br />

along the activated fault segments over a length of<br />

200 km. The ‘Ananevo’ rockslide located in the north<br />

of lake Issyk Kul (at some 80 km east of the<br />

presumed epicentre) is one of the most prominent<br />

feature produced by the Kemin earthquake (Havenith<br />

et al., 2002, see Fig. 2). Failure took place at the<br />

southern end of a mountain ridge, just above the<br />

discontinuous Chon Aksu fault also activated by the<br />

1911 Kemin earthquake. This section of the Chon<br />

Aksu fault is a thrust gently dipping towards the<br />

northeast into the collapsed slope. Evidence of the<br />

presence of the fault is the related scarp with a height<br />

of almost 10 m at 12 km to the WNW. On the site<br />

itself, outcrops at the foot of the southwest-oriented<br />

slope show particularly disintegrated and weathered<br />

granitic rocks within a 100-200 m thick fault zone.<br />

thousands of inhabitants (Fig. 3) - the exact number<br />

of fatalities will never be known. This rock avalanche<br />

had been triggered from Borgulchak mountain at an<br />

altitude of about 2950 m and travelled more than 6<br />

km before reaching the inhabited valley at an altitude<br />

of 1550 m. The volume is about 40*10 6 m 3 . A<br />

significant part of the mass movement was made of<br />

loess, which probably contributed to the mobility of<br />

the initial rockslide. In the Yasman valley, opposite to<br />

the Khait rock avalanche, massive loess earth-flows<br />

are believed to have buried about 20 villages. In total,<br />

the Khait rock avalanche and loess earth-flows are<br />

likely to have killed more than 5000 people during the<br />

1949 event.<br />

Fig. 3: Khait rock avalanche; view towards the East from<br />

Yasman valley (unpublished photograph of 2005 provided<br />

by A. Ischuk). The length of the scarp is about 1 km.<br />

The Gissar earthquake, 1989<br />

South of Dushanbe, in Gissar, Tajikistan, a Ms=5.5<br />

earthquake on January 23, 1989 had triggered a<br />

series of earth-flows in loess. At least 200 people<br />

were killed and hundreds of houses were buried.<br />

According to Ishihara et al. (1990), those slides were<br />

all related to extensive liquefaction, which had<br />

developed for a horizontal acceleration of about<br />

0.15g. Ishihara et al. (1990) associated the<br />

liquefaction to the ‘collapsible nature’ of the highly<br />

porous loess material (a silt-sized deposit with an<br />

average content of clay of 15 % and a low plasticity).<br />

Fig. 2: Photograph of the Ananevo rockslide. The horsefarm<br />

building (lower left corner) is 60 by 60 m.<br />

The Sarez earthquake, 1911<br />

The Sarez earthquake, Ms=7.6, struck the central<br />

Pamir Mountains, Tajikistan, on February 18, 1911.<br />

Such an earthquake is likely to have triggered<br />

hundreds or thousands of mass movements, but only<br />

one is well documented: the giant Usoi rockslide,<br />

which fell from a 4500 m high mountain down to an<br />

elevation of 2700 m in the valley (Schuster and<br />

Alford, 2004). This rockslide has formed a dam with a<br />

volume of about 2*10 9 m 3 on Murgab River.<br />

The Khait earthquake, 1949<br />

According to Leonov (1960), the M=7.4 Khait<br />

earthquake that struck Northern Tajikistan on July 10,<br />

1949, produced a very destructive mass movement<br />

that had buried the villages of Khait and Kusurak with<br />

The Suusamyr earthquake, 1992<br />

The most recent large seismic event hitting Central<br />

Asian mountain regions was the Ms=7.3 Suusamyr<br />

earthquake on August 19, 1992, triggering various<br />

types of ground failures in the Northern-Central Tien<br />

Shan (Bogachkin et al., 1997).<br />

Most of the 50 people killed in the remote areas were<br />

victims of mass movements. Korjenkov et al. (2004)<br />

described a series of ground failures and also a great<br />

variety of gravitation cracks. Ground instability could<br />

be observed along the crest and southern slope of<br />

the Chet-Korumdy ridge – here, most landslides had<br />

developed from previously existing ground<br />

instabilities. The largest mass movement, a rock<br />

avalanche, had formed a landslide dam that partly<br />

failed in 1993, causing a long-runout debris flow and<br />

widespread flooding downstream.<br />

79


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The Haiyuan or Gansu earthquake, 1920<br />

On December 16, 1920, a M=8.5 earthquake<br />

occurred near Ganyan Chi, Haiyuan County of the<br />

Ningxia Hui Autonomous Region in China (Zhang,<br />

1995). Several hundreds of thousands of houses<br />

collapsed and officially 234.117 people died. Zhang<br />

(1995) noticed that particularly high intensities were<br />

recorded over areas covered by thick loess deposits<br />

and that in those deposits ‘landslides were not only<br />

controlled by the intensity of the earthquake, but by<br />

the structure of the subsoil’. Zhang and Wang (2007)<br />

reported that about 100000 people were killed just by<br />

landslides in loess deposits. They observed that<br />

loess earth-flows triggered by the Haiyuan<br />

earthquake had developed on relatively gentle slopes<br />

compared to those triggered by rainfall in the same<br />

region. These observations highlight the particular<br />

susceptibility of loess areas to ground failure, such as<br />

it was clearly shown by Derbyshire et al. (2000)<br />

analysing geological hazards affecting the loess<br />

plateau of China.<br />

CONCLUSIONS<br />

A series of case histories of earthquake-induced<br />

landslides in Central Asia have been presented.<br />

These show that the most disastrous mass<br />

movements in Central Asia are long runout and rapid<br />

(>20 m/s) rock avalanches and loess earth-flows.<br />

While the giant rockslides are almost exclusively<br />

triggered by large magnitude seismic events (M7) in<br />

Central Asian mountain regions, loess earth-flows<br />

may also be triggered by smaller earthquakes – or<br />

even by climatic factors. Here, I presented some<br />

examples of fatal loess landslides triggered directly<br />

by a M=5.5 earthquake in Tajikistan. The comparison<br />

with the Haiyuan earthquake event of 1920 shows<br />

that such loess landslides can be very disastrous.<br />

The importance of mid- and long-term effects is<br />

outlined for both rockslides and loess landslides.<br />

Several case histories showed that one important – if<br />

not the most important – long-term consequence of<br />

massive rockslides can be the formation of a dam<br />

and the impoundment of a natural reservoir. Actually,<br />

the largest still existing rockslide dam on earth had<br />

formed in 1911 in the Pamir Mountains.<br />

I also wanted to show that similar earthquakes may<br />

not necessarily trigger the same number of<br />

landslides, due to different climatic conditions and<br />

groundwater level at the time of the earthquake.<br />

Finally, it is important to note that landslides are not<br />

only instantaneous effects of earthquakes – some<br />

had already developed before the seismic shock and<br />

some continued or started moving well after the<br />

shaking. To better assess the short- to long-term<br />

effects earthquakes on slopes, landslides need to be<br />

monitored by geophysical, seismological and<br />

geotechnical systems, coupled to multi-temporal<br />

satellite imagery and numerical modeling of multievent<br />

scenarios. In the frame of new projects on<br />

landslide problems in Central Asia, focus will be on<br />

the installation of such coupled monitoring–modeling<br />

systems.<br />

Acknowledgements: This study was supported by the<br />

NATO science for Peace and Security Project LADATSHA<br />

983289, 2009-2012.<br />

References<br />

Derbyshire, E., Meng, X. & T.A. Dijkstra, (2000). Landslide<br />

in the Thick Loess Terrain of North-West China. John<br />

Wiley, Chichester, U.K.<br />

Havenith, H.B., Jongmans, D., Faccioli, E., Abdrakhmatov,<br />

K. & P.Y. Bard, (2002). Site effects analysis around the<br />

seismically induced Ananevo rockslide, Kyrgyzstan.<br />

Bulletin of the Seismological Society of America, 92,<br />

3190-3209.<br />

Ishihara, K., Okusa, S., Oyagi, N. & A. Ischuk, (1990).<br />

Liquefaction-induced flow slide in the collapsible deposit<br />

in the Soviet Tajik. Soils and Foundations, 30: 73-89.<br />

Keefer, D.K., (1984). Landslides Caused by Earthquakes,<br />

Geological Society of America Bulletin, 95: 406–421.<br />

Korjenkov, A.M., Mamyrov, E., Omuraliev, M., Kovalenko,<br />

V.A. & S.F. Usmanov, 2004. Rock Avalanches and<br />

Landslides Formed in Result of Strong Suusamyr (1992,<br />

M=7,4) Earthquake in the Northern Tien Shan. Test<br />

Structures for Mapping of Paleoseismic Deformations by<br />

Satellite Images. In High Mountain Remote Sensing<br />

Cartography VII (HMRSC VII), M. F. Buchroithner (ed.),<br />

Institute for Cartography of the Dresden University of<br />

Technology. Kartographische Bausteine, Band 23,<br />

Dresden, 19 p.<br />

Leonov, N.N., (1960). The Khait, 1949 earthquake and<br />

geological conditions of its origin. In Proc. of Academy of<br />

Sciences of the USSR, Geophysical Series, 3: 409-424<br />

(in Russian).<br />

Nadim, F., Kjekstad, O., Peduzzi, P., Herold, C. & C.<br />

Jaedicke, (2006). Global landslide and avalanche<br />

hotspots. Landslides, 3: 159–173.<br />

Petley, D., Dunning, S., Rosser, N. & A.B. Kausar, (2006).<br />

Incipient Landslides in the Jhelum Valley, Pakistan<br />

Following the 8th October 2005 Earthquake. Disaster<br />

Mitigation of Debris Flows, Slope Failures and<br />

Landslides, Universal Academy Press Inc./Tokyo, Japan:<br />

47–55.<br />

Sassa, K., (1996). Prediction of earthquake induced<br />

landslides. In Proc. Landslides, Senneset (ed.), 115-131.<br />

Sato, H.P., Hasegawa, H., Fujiwara, S., Tobita, M., Koarai,<br />

M., Une, H. & J. IWAHASHI, (2007). Interpretation of<br />

landslide distribution triggered by the 2005 Northern<br />

Pakistan earthquake using SPOT5 imagery. Landslides,<br />

4:113–122.<br />

Schuster, R.L., L.M. Highland, (2001). Socioeconomic and<br />

environmental impacts of landslides in the western<br />

hemisphere. USGS Open-File Report 2001-0276.<br />

Schuster, R.L., D. ALFORD, (2004). Usoi Landslide Dam<br />

and Lake Sarez, Pamir Mountains, Tajikistan. Environmental<br />

& Engineering Geoscience, X, 2: 151–168.<br />

Yin, Y., Wang, F. & P. Sun, (2009). Landslide hazards<br />

triggered by the 2008 Wenchuan earthquake, Sichuan,<br />

China. Landslides. DOI 10.1007/s10346-009-0148-5.<br />

Zhang, Z., (1995). Geological Disasters in Loess Areas<br />

during the 1920 Haiyuan Earthquake, China. GeoJournal,<br />

36.2, 3: 269-274.<br />

Zhang, D., Wang, G., (2007). Study of the 1920 Haiyuan<br />

earthquake-induced landslides in loess (China).<br />

Engineering Geology, 94: 76–88.<br />

80


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

A CASE STUDY OF EARTHQUAKES AND ROCKFALL -<br />

INDUCED DAMAGE TO A ROMAN MAUSOLEUM IN<br />

PINARA, SW TURKEY<br />

Hinzen, Klaus-G. (1, Kehmeier, Helen (1), Schreiber, Stephan (1), Reamer, Sharon K. (1)<br />

(1) Earthquake Geology and Archaeoseismology Group, Cologne University, Vinzenz-Pallotti-Str. 26, 51429 Bergisch Gladbach.<br />

GERMANY. Email hinzen@uni-koeln.de<br />

Abstract (A Case Study of Earthquakes and Rockfall): A Roman mausoleum located in the ancient city of Pnara, southwest<br />

Turkey, shows clear signs of damages due to dynamic loading. Considering the seismotectonic potential of the area, earthquake<br />

ground motions are a possible cause of the damages. However, the building is located at the foot of a 90 m high cliff with a<br />

significant rock fall hazard. We present a 3D discrete element model of the mausoleum based on a 3D laser scan. The range of<br />

impact velocities of blocks of different size and form and the actual slope of the cliff have been incorporated into 2D model<br />

calculations. The deformations caused by simulated rock impacts are compared to the actual displacements of blocks quantified<br />

from the laser scan. In addition, analytic ground motions signals are used to study the principal reaction of the building. The<br />

second damage scenario using earthquake strong ground motions shows that the damage is more likely caused by an earthquake<br />

than impacting rocks.<br />

Key words: Quantitative Methods, Rockfall, Strong Ground Motion, Archaeoseismology.<br />

INTRODUCTION<br />

A Roman mausoleum (Fig. 1) in the southwestern<br />

Turkey in the ancient city Pnara shows clear signs of<br />

dynamic loading that deformed the building. Its<br />

simple block structure and the fact that most parts of<br />

the building are still standing makes it an excellent<br />

test case for a quantitative archaeoseismic analysis.<br />

While collapsed structures usually allow at most an<br />

estimate of a minimum ground motion threshold,<br />

deformed but standing structures offer a deeper<br />

insight into the causes of the deformations.<br />

(A)<br />

(C)<br />

(B)<br />

(D)<br />

The suitability of the ancient city of Pnara in SW<br />

Turkey for archaeoseismic studies has been proven<br />

in several previous studies (Sintubin et al., 2003;<br />

Yerli et al., 2010; Yerli et al. 2009). Yerli et al. (2011)<br />

and Hinzen et al. (2010) provide detailed information<br />

about the seismicity and tectonic setting of the site.<br />

The pronounced topography gives Pnara an unique<br />

character, but also introduces a rockfall risk for<br />

several exposed structures, including the Roman<br />

mausoleum.<br />

ROMAN MAUSOLEUM<br />

(E)<br />

(F)<br />

Laserscan Model<br />

The mausoleum was surveyed with a 3D phaselaserscanner<br />

(Fleischer et al. 2010; Schreiber et al.,<br />

2010, Schreiber et al. 2009); nine individual scans<br />

(from in- and outside positions) were combined into a<br />

virtual model of 79 million 3D points (Fig. 1B-D). This<br />

model was used as the basis for a discrete element<br />

model (Fig. 1E and F) of the undamaged mausoleum<br />

and also to quantify the deformation of the still<br />

standing part of the structure.<br />

Fig. 1: (A) Photo with a view from SE of the Roman<br />

mausoleum in Pnara; (B) laserscan from the same<br />

perspective; (C) and (D) ortho-views of the scan of<br />

the west and south wall, respectively; (E) and (F)<br />

wireframe and rendered view of the discrete element<br />

model.<br />

81<br />

A clear increase of the amount of deformation from<br />

bottom to top rows exists. Some parts of the sill and<br />

the pediment have fallen down. This also applies to<br />

the columns that are no longer found in situ. A part of<br />

the ceiling of the colonnade fell down and broke a


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

massive block of the front platform (Fig. 1 A and B).<br />

No fallen blocks are found inside the structure. On<br />

the western (back-) side of the structure the fallen<br />

pediment blocks are located in close proximity to the<br />

wall.<br />

Discrete Element Model<br />

Descriptions by Vitruvius (1796) and intact examples<br />

of a Roman mausoleum were used to reconstruct the<br />

missing front section of the Pnara mausoleum<br />

(shown in dark colour in Fig. 1F). The size of each<br />

construction block was measured from the 3D scan<br />

and transferred into a model of discrete rigid blocks<br />

(Fig. 1E). The model contains 180 blocks and with a<br />

density of the local conglomerate of 2.87 Mgm -3 it<br />

has a total mass of 1.8x10 5 kg.<br />

Rockfall<br />

2D Cliff Model<br />

A 2D model of the cliff and the slope on which the<br />

mausoleum is built was used to estimate impact velocities<br />

of rockfall material of different size and form.<br />

Boulders currently resting on the slope with sizes of<br />

several cubic meters and fresh fracture faces on the<br />

cliff indicate the persisting rockfall hazard at the site.<br />

Figure 2 shows the distribution of impact velocities of<br />

falling material on the mausoleum summarizing the<br />

results from numerous simulation calculations.<br />

Impact to Mausoleum<br />

Scenarios with rocks measuring 0.5 to 2.0 m with<br />

impact velocities between 10 m/s and 35 m/s and<br />

impacting the mausoleum at different height levels<br />

were tested.<br />

Velocity (m/s)<br />

Elevation a.s.l. (m)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

600<br />

550<br />

500<br />

450<br />

400<br />

Blocksize (m)<br />

0 1 2 3 4 5<br />

350<br />

200 250 300 350 400 450 500<br />

Distance (m)<br />

160<br />

120<br />

80<br />

40<br />

0<br />

Velocity (km/h)<br />

rows of blocks; however the rest of the structure is<br />

not significantly deformed (Fig. 3).<br />

30 m/s<br />

G<br />

F<br />

E<br />

D<br />

C<br />

B<br />

A<br />

Z<br />

Y<br />

X<br />

W<br />

log 10 Displacement (m)<br />

1<br />

0<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

X Z B D F<br />

Layer<br />

Fig. 3: (left) View of the mausoleum from south after<br />

the impact of a rock with 1 m edge length and 30 m/s<br />

impact velocity. Block layers of the structure are<br />

labelled with capital letters. (right): Boxplots of the<br />

distribution of the log of displacements of block corners<br />

in five layers. The upper boxes (hatched downwards<br />

and blue) show the displacements measured<br />

from the laserscan of the building; the lower boxes<br />

(hatched ascending and red) are the displacements<br />

after the rock impact.<br />

Fig. 4: Reaction of the Roman mausoleum to an<br />

analytic ground motion signal, a Morlet wavelet with<br />

1 Hz main frequency and 10 s duration.<br />

GROUND MOTION<br />

Analytic Signals<br />

Before using full 3D earthquake ground motions, a<br />

series of tests was carried out with analytic signals in<br />

the form of Morlet wavelets (Goupillaud et al., 1984).<br />

The colonnade of the structure is highly vulnerable to<br />

ground motion frequencies around 1 Hz (Fig. 4). At<br />

signals with main frequencies of 2 Hz and above, the<br />

typical corner expulsions and block shifts are<br />

observed.<br />

Synthetic Seismograms<br />

Site-specific Green’s functions and an arbitrary number<br />

of rectangular dislocation planes were used to<br />

calculate synthetic seismograms using the method<br />

described by Wang (1999). A model of the local active<br />

faults is based on the work by Yerli et al. (2011).<br />

Fig. 2: (top) Red dots give median values of impact<br />

velocities of differently sized blocks at the location of<br />

the Roman mausoleum. (bottom) 2D model of the<br />

slope with trajectories of rocks of different size and<br />

form falling from the top of the cliff; a photo of which is<br />

shown in the insert.<br />

Impacting rocks cause localized damage on the west<br />

wall of the structure. With rocks of 1 m size at velocities<br />

above 20 m/s blocks are being pushed inside<br />

the building. Above velocities of 30 m/s strong<br />

damage occurs if the rock impacts into the upper<br />

0 km 5 km 10 km<br />

S3<br />

S7<br />

30.5°N<br />

S6<br />

S8S9<br />

S2<br />

S5<br />

S10<br />

S4<br />

S1<br />

Fig. 5: Active faults in<br />

the vicinity of Pnara.<br />

Fault segments for<br />

different earthquake<br />

scenarios are labelled<br />

S1 to S10 (map<br />

from Yerli et al.,<br />

2011).<br />

29.3°E<br />

82


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Impact to Mausoleum<br />

An example of the deformation from a synthetic<br />

earthquake record is shown in Figure 6. In this<br />

scenario activation of segments S1, S10, S9, and S3<br />

was assumed. The M W 6.0 example earthquake did<br />

not topple the colonnade; however, the calculated<br />

displacements are very close to the actual ones with<br />

the exception of an underestimated F-layer. Test<br />

calculations with measured ground motions from the<br />

2009 L’Aquila earthquake did also destroy the<br />

colonnade. As part of the ongoing work we will<br />

1<br />

0<br />

Modeling of rock fall impacts and diverse earthquake<br />

scenarios indicate that the damages and displacements<br />

of building blocks of the Pnara mausoleum<br />

are more likely the result of an earthquake than being<br />

formed by rockfall. So far only a single earthquake<br />

has been used for the modeling. Further calculations<br />

are planned to test for the possibility of repeated<br />

earthquake action over the millennia since the<br />

construction of the mausoleum.<br />

Acknowledgements: We are grateful for the help during<br />

the field work by Barish Yerli and Claus Fleischer. Part of<br />

the work was financially supported by Deutsche Forschungsgemeinschaft<br />

(DFG HI660/2-1).<br />

G<br />

F<br />

E<br />

D<br />

C<br />

B<br />

A<br />

Z<br />

Y<br />

X<br />

W<br />

log 10 Displacement (m)<br />

-1<br />

-2<br />

-3<br />

-4<br />

-5<br />

0 4 8 12 16 20<br />

T m e s<br />

X Z B D F<br />

Layer<br />

Fig. 6: Same layout as Fig. 3 for deformations of a<br />

synthetic earthquake with the ground motions shown<br />

in the insert; upper trace EW component lower trace<br />

NS component of ground motion.<br />

further quantify the ground motion parameters which<br />

caused the deformations.<br />

CONCLUSIONS<br />

The mausoleum located above the Roman forum of<br />

the ancient city of Pnara is heavily damaged;<br />

however, most of its simple block structure is still<br />

standing. Deformations indicate that the building suffered<br />

from dynamic loading and the location suggests<br />

rockfall and earthquake ground motion as possible<br />

d ( m )<br />

0.12<br />

0.08<br />

0.04<br />

-0.04<br />

0<br />

References<br />

Fleischer, C., K.-G. Hinzen & S. Schreiber (2010). Laser<br />

scanning eines roemischen Brunnens in der<br />

Archaeologischen Zone Koeln. Allgemeine Vermessungs-Nachrichten<br />

5/2010, 176-180.<br />

Goupillaud, P., A. Grossman & J. Morlet (1984). Cycle and<br />

related transforms in seismic signal analysis.<br />

Geoexploration 23, 85-102.<br />

Hinzen, K.-G., S. Schreiber & B.Yerli, (2010). The Lycian<br />

Sarcophagus of Arttumpara, Pnara (Turkey) -Testing<br />

Seismogenic and Anthropogenic Damage Scenarios.<br />

Bulletin of the Seismological Society of America 100 (6),<br />

3148-3164.<br />

Schreiber, S., K. G. Hinzen & C. Fleischer (2010). Closerange<br />

Laserscanning in the Archaeological Zone<br />

Cologne, Germany, Workshop on Remote Sensing<br />

Methods for Change Detection and Process Modelling,<br />

Cologne, November 18-19 2010.<br />

Schreiber, S., K. G. Hinzen & C. Fleischer (2009). An<br />

application of 3D laser scanning in archaeology and<br />

archaeoseismology: The medieval cesspit in the<br />

archaeological zone Cologne, Germany, 1st INQUA-<br />

IGCP-567 International Workshop on Earthquake<br />

Archaeology and Palaeoseismology, Baelo Claudia,<br />

Cadiz, Spain, 7-13 September 2009, 136-138.<br />

causes. Sintubin, M., P. Muchez, D. Similox-Tohon, G. Verhaert, E.<br />

Paulissen & M. Waelkens (2003). Seismic catastrophes<br />

Numeric modelling of rock fall impact indicates a<br />

different damage pattern from that one observed in<br />

the field. Impacting rocks tend to heavily deform the<br />

western wall facing the cliff; however, the rest of the<br />

building, especially block layers below the impact,<br />

are only insignificantly affected. Rocks with high<br />

impact velocities push blocks of the building to the<br />

inside; however, all fallen blocks from the top of the<br />

building were found outside of the structure.<br />

Tests of the dynamic reaction of the mausoleum with<br />

analytic ground motion signals show a high vulnerability<br />

of the colonnade to horizontal movements<br />

with frequencies around 1 Hz. Ground motions with<br />

frequencies of 2 Hz and above initiate rocking of<br />

blocks with an increasing tendency from bottom to<br />

top. Morlet wavelets with main frequencies of 2 Hz<br />

dislocate the blocks in the same pattern as it is<br />

observed today. Synthetic ground motions of local<br />

earthquakes with M W 5.6 to 6.0 also produce<br />

displacements similar to the observations.<br />

at the ancient city of Sagalassos (SW Turkey) and their<br />

implications for seismotectonics in the Burdur-Isparta<br />

area. Geological Journal 38 (3-4), 359-374.Vitruvius<br />

Pollio, M. & A. Rode (1796). Baukunst. Translated by<br />

August Rode, G. J. Goeschen (eds.), Leipzig, 10 Books.<br />

Wang, R. (1999). A simple orthonormalization method for<br />

stable and efficient computation of Green's functions.<br />

Bulletin of the Seismological Society of America 89, 733-<br />

741.<br />

Yerli, B., S. Schreiber & K.-G. Hinzen (2011). Seismotectonic<br />

background for archaeoseismologic studies in<br />

the Een Basin, SW Turkey. Natural Hazards, submitted.<br />

Yerli, B., J. Ten Veen, M. Sintubin, C. Karabacak, C.<br />

Yaliciner & E. Altunel (2010). Assessment of seismically<br />

induced damage using LIDAR: the ancient city of Pinara<br />

(SW Turkey) as a case study. In: Ancient Earthquakes.<br />

Geological Society of America, Special Paper 471, M.<br />

Sintubin, I. Stewart, T. M. Niemi and E. Altunel (eds.),<br />

157-170.<br />

Yerli, B., S. Schreiber, K. G. Hinzen, J. H. t. Veen, M.<br />

Sintubin & M. Sintubin (2009). Testing the Hypothesis of<br />

earthquake-related damage in structures in the Lycian<br />

ancient City of Pinara, SW Turkey, 1st INQUA-IGCP-567<br />

International Workshop on Earthquake Archaeology and<br />

Palaeoseismology, Baelo Claudia, Cadiz, Spain, 7-13<br />

September 2009, 173-176.<br />

83


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

NEOTECTONICS OF GRACIOSA ISLAND (AZORES) – UNCERTAINTY IN SEISMIC<br />

HAZARD ASSESSMENT IN A VOLCANIC AREA WITH VARIABLE SLIP-RATES<br />

Hipólito, Ana (1, José Madeira (2), Rita Carmo (1), João Luís Gaspar (1)<br />

(1) Centro de Vulcanologia e Avaliação de Riscos Geológicos, Universidade dos Açores, Complexo Científico, 3º Piso – Ala Sul,<br />

Rua da Mãe de Deus, 9500-321 Ponta Delgada, Açores, PORTUGAL. Email: ana.rc.hipolito@azores.gov.pt<br />

(2) Universidade de Lisboa, GeoFCUL, LATTEX/IDL-Laboratório Associado, Faculdade de Ciências das Universidade de<br />

Lisboa, Edifico C6, Campo Grande, 1749-016 Lisboa, PORTUGAL.Email: jmadeira@fc.ul.pt<br />

Abstract (Neotectonics of Graciosa Island (Azores) – uncertainty in seismic hazard assessment in a<br />

volcanic area with variable slip-rates): Graciosa is a mid-Pleistocene to Holocene volcanic island that lies on a<br />

complex plate boundary between the North American, Eurasian and Nubian plates. Large fault scarps displace the<br />

oldest volcanic units, but in the younger areas recent volcanism hides the surface expression of faulting, limiting<br />

neotectonic observations. Slip-rates deduced from neotectonic surveys are higher than those provided by kinematic<br />

plate motion models. This suggests a variability of deformation rates, alternating between high tectonic deformation<br />

periods, decreasing the recurrence interval of surface rupturing earthquakes, and phases of low slip-rate.<br />

Nevertheless, in historical time a few destructive earthquakes affected the island attesting for its seismic hazard.<br />

Key words: Neotectonics, Azores, Graciosa Island<br />

GEODYNAMIC AND VOLCANIC SETTING<br />

The Azores archipelago lies on a complex<br />

geodynamic setting: the Eurasian (Eu), North<br />

American (NA) and Nubian (Nu) triple junction<br />

(Azores Triple Junction - ATJ) (Fig.1).<br />

Graciosa Island is located on the west segment of<br />

the Eu-Nu boundary, a diffuse and complex<br />

deformation zone (Fig.1) sheared by a dextral<br />

transtensile regime (e.g. Madeira & Brum da Silveira,<br />

2003; Carmo, 2004; Hipólito, 2009).<br />

Fig. 1: Location of the Azores and main<br />

morphotectonic features of the region. The shaded<br />

area represents the sheared western segment of the<br />

Eu-Nu plate boundary; Plates: NA – North American;<br />

Eu – Eurasian; Nu – Nubian; Tectonic structures:<br />

MAR – Mid-Atlantic Ridge; EAFZ – East Azores<br />

Fracture Zone; GF – Gloria Fault; Islands: G –<br />

Graciosa Island. World topography and bathymetry<br />

from ESRI ® (2008), Azores bathymetry adapted from<br />

Lourenço et al. (1997). Datum: WGS 1984 (modified<br />

from Hipólito, 2009).<br />

84<br />

Fig. 2: Seismicity in and around Graciosa island from<br />

2003 to 2009, defining three alignments in West<br />

Graciosa Basin and one at the western border of the<br />

East Graciosa Basin. Red dots mark earthquake<br />

epicenters and magnitudes (M D ). Data from CIVISA<br />

(Centro de Informação e Vigilância Sismovulcânica<br />

dos Açores), 2009.<br />

This boundary acts as an ultra-slow oblique<br />

expansive center (Vogt and Jung, 2004) and a<br />

transfer zone accommodating the differential motion<br />

between Eu and Nu plates, due to the higher


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

spreading rates north of the Azores (e.g. DeMets,<br />

1994; Fernandes et al., 2003).<br />

Due to its geodynamic setting, the Azores<br />

archipelago presents frequent seismicity. Graciosa<br />

Island was affected by some significant earthquakes<br />

since its settlement in mid 15 th century. One of them<br />

(1837) had probably its epicenter on land (Madeira,<br />

1998; Silva, 2005). However, instrumental seismicity<br />

does not show significance seismic activity within the<br />

island (Fig.2). The epicenter distribution of the<br />

current activity reveals four NNW-SSE trending<br />

offshore alignments, one on the east flank of West<br />

Graciosa Basin, two crossing this basin floor to the<br />

west, and another alignment just east of the island<br />

(Fig. 2).<br />

STRUCTURAL DATA<br />

Geometric and kinematic fault analysis<br />

In Graciosa Island, neotectonic studies are limited by<br />

the absence of outcrops with well exposed fault<br />

planes. Tectonic features, corresponding to important<br />

fault scarps, several tens to hundreds of meters high,<br />

occur in the central region of the island. Unfortunately<br />

the faults producing these large features do not crop<br />

out and the small size of the island limits the<br />

observation of the full length of the faults. The thick<br />

and non-cohesive young volcanic fall deposits and<br />

the lava flows that mantle the topography were not<br />

tectonically displaced yet, hiding the trace of the<br />

faults in areas covered by recent volcanic units.<br />

Therefore, paleoseismological studies were not made<br />

due the sheer size of the fault scarps and the<br />

absence of recent surface faulting with favorable<br />

conditions for trenching. The observed faults crop out<br />

either in quarry walls exploring cinder cones or in<br />

inaccessible sea cliffs. In the first case the nature of<br />

the deposits (homogeneous, coarse size and low<br />

cohesion) does not allow the generation of<br />

slickensides, hindering kinematic analyses.<br />

Fig. 3: Digital Elevation Model of Graciosa and the main<br />

morphologic regions. Based on Carta Militar de Portugal,<br />

Sheet 21 – Stª Cruz da Graciosa, Instituto Geográfico do<br />

Exército (2001); UTM Projection; Datum: WGS 1984.<br />

Graciosa Island comprises one quiescent trachytic<br />

polygenetic volcano (the Caldera Volcano), an older<br />

volcanic complex (the Central-Southern Complex),<br />

resulting from the dismantling of an important central<br />

volcano, and several monogenetic eruptive centers<br />

included in the NW basaltic Platform that mantle the<br />

older units (Gaspar, 1996; Fig.3). The oldest<br />

volcanic-stratigraphic unit is around 620 ± 120 ka old<br />

(Féraud et al., 1980). The most recent volcanic event<br />

was a pre-settlement basaltic hawaiian-strombolian<br />

eruption at about 2 ka B.P. (Walker, unpublished<br />

data, in Gaspar, 1996). Currently the volcanic activity<br />

is just expressed by secondary manifestations,<br />

namely by thermal springs, fumarolic fields and<br />

diffuse degassing (e.g. Ferreira et al., 1993; Gaspar,<br />

1996).<br />

The island presents in general a smooth relief with<br />

maximum altitude of 402m (Fig.3). The central part is<br />

crossed by several NW-SE trending fault scarps<br />

parallel to the shape of the island (Fig.4). Those<br />

faults define a graben structure that is crossed in the<br />

SE by a NNE-SSW fault scarp which separates the<br />

older from the most recent volcanic units to the<br />

south.<br />

Fig. 4: Morphotectonic map of Graciosa Island. Based on<br />

Carta Militar de Portugal, Sheet 21 – Stª Cruz da Graciosa,<br />

Instituto Geográfico do Exército (2001); UTM Projection;<br />

Datum: WGS 1984.<br />

Generally, the mapped structures are normal faults or<br />

present normal component. Although, in most cases<br />

it was difficult to recognize a strike-slip component,<br />

these structures may also have strike-slip component<br />

(dextral or sinistral) typical of a tectonic transtensile<br />

regime.<br />

85


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The tectonic map (Fig.4; Gaspar and Queiroz, 1995;<br />

Hipólito, 2009), the geometric (Fig.5) and kinematic<br />

(Fig.6) fault analysis (Hipólito, 2009) show the<br />

presence of two main fault systems: system A –<br />

composed of two sets of conjugated faults, one<br />

trending NW-SE, dipping to SW, presenting normaldextral<br />

or dextral-normal oblique slip, and another<br />

striking NNE-SSW, dipping to ESE, with oblique<br />

normal-left lateral or left lateral-normal slip (Fig. 6a);<br />

system B – includes NNE-SSW to NE-SW trending<br />

faults, dipping to WNW or NW, presenting normaldextral<br />

or dextral-normal oblique slip (Fig. 6b). A<br />

family of faults conjugated with these structures was<br />

not found. The strong fault dips (the 80º to 90º range<br />

dominate) also suggest oblique (normal/strike-slip)<br />

faulting (Fig.5).<br />

Slip-rates<br />

Only three tectonic structures allowed slip-rate<br />

estimation: the Saúde-Serra das Fontes-Hortelã, the<br />

South Serra das Fontes and the East Serra das<br />

Fontes faults (Fig. 4). However, in two of them only<br />

the normal slip component was determined, while in<br />

the other only the strike-slip could be measured<br />

(Table I). The Saúde-Serra das Fontes-Hortelã Fault<br />

trace is marked by a 5750m-long south–facing scarp<br />

trending N306º to N328º producing 49m of dextral<br />

strike-slip displacement. The South Serra das Fontes<br />

Fault has a 200m-high south-facing scarp 3500m<br />

long, trending N282º to N327º. The East Serra das<br />

Fontes Fault trace is marked by a 4750m-long and<br />

185m- maximum high east-southeast facing scarp,<br />

trending N20º. The total length of these faults is<br />

unknown because on-shore they are fossilized by<br />

younger units and there is no off-shore data allowing<br />

to trace them.<br />

The estimated values indicate slip-rates somewhat<br />

higher than those provided by plate motion kinematic<br />

models (Table II) and contrasting with the present<br />

low seismicity and neotectonic deformation in the<br />

island (Fig. 2). These slip-rates could be overestimated<br />

because the age of the used marker may<br />

not represent the whole deformation period. On the<br />

other hand, the real displacements may be greater<br />

than the measured ones due to infilling of tectonically<br />

depressed areas by younger deposits.<br />

Table I: Slip-rates of three faults in Graciosa Island.<br />

Fig. 5: Geometry of all analyzed faults: a) stereographic<br />

plot of fault planes (lower hemisphere; Schmidt net) – <br />

diagram; b) stereogram of fault poles density – <br />

diagram; c) Circular histogram of unweighted frequencies<br />

of fault plane directions and dip angles. TectonicsFP<br />

software (Ortner et al., 2002).<br />

Paleostress analysis suggests that the region is<br />

affected by two different stress fields that can<br />

alternate in time and/or in space; variations of the<br />

local stress field may occur, which are responsible for<br />

the generation of new faults or reactivation of preexisting<br />

structures (Hipólito, 2009).<br />

Table II: Relative velocities and azimuth directions of<br />

slip vector for Eu and Nu plates and for those plates<br />

relatively to NA plate in ATJ zone, according to the<br />

NUVEL-1A (DeMets et al., 1994), REVEL (Sella et al.,<br />

2002) e DEOS2K (Fernandes et al., 2003) global<br />

kinematic models.<br />

Fig. 6: Stereographic plot of two main fault systems (lower<br />

hemisphere; Schmidt net – diagram). a: system A - NW-<br />

SE to NNW-SSE faults, with normal-dextral oblique slip,<br />

conjugate of NNE-SSW to NE-SW normal-left lateral<br />

structures; b: system B - NNE-SSW to NE-SW faults with<br />

normal-dextral oblique slip. Black arrows: strike-slip<br />

sense. TectonicsFP software (Ortner et al., 2002).<br />

DISCUSSION AND FINAL REMARKS<br />

As in other zones in the Azores (Hipólito et al., 2011),<br />

several limitations prevent a more detailed<br />

neotectonic survey of Graciosa Island. Recent<br />

volcanic deposits, that mantle the topography, hide<br />

86


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

the surface expression of major tectonic structures;<br />

basaltic lapilli deposits do not allow the generation of<br />

linear kinematic markers; there are no structures with<br />

scarp heights allowing paleoseismological studies,<br />

thus contributing to the assessment of seismic<br />

hazard. Nevertheless, the tectonic information from<br />

Graciosa Island is in agreement with the stress<br />

pattern proposed by other authors for the Azores<br />

region (e.g. Madeira, 1998, Lourenço et al., 1988;<br />

Carmo, 2004).<br />

The absence of seismic events producing surface<br />

rupture since settlement and the current low seismic<br />

activity in Graciosa Island, contrast with the youthful<br />

aspect of its tectonic morphology. The calculated<br />

slip-rates and the evident loss of geomorphic<br />

expression of fault scarps into the areas covered with<br />

the younger units (NW Platform and Caldeira<br />

Volcano) suggest that the occurrence of a period of<br />

important tectonic activity before 31 ka, with higher<br />

slip-rates than those observed in present times,<br />

responsible for the deformation of the central part of<br />

the island. That period was followed by a magmatic<br />

dominated phase, with the build-up of Caldeira<br />

Volcano and the installation of the fissural basaltic<br />

volcanism responsible for the formation of the NW<br />

Platform. During that period there was a reduction of<br />

fault slip-rates and consequent increase of the<br />

recurrence period of surface rupturing event. This is<br />

consistent with the occurrence of variations in<br />

deformation rates in the archipelago, with periods<br />

with slip-rates higher or lower than the average, as<br />

proposed by Madeira (1998).<br />

Thus, the current tectonics in Graciosa Island is not<br />

particularly active. Nevertheless, as it is located<br />

between two important seismogenic sources (West<br />

and East Graciosa Basins) and on a complex<br />

geodynamic setting, seismic hazard cannot be<br />

disregarded.<br />

Acknowledgements: Ana Hipólito and fieldwork were<br />

supported by CARIGE Project – Carta de Riscos<br />

Geológicos, Governo Regional da Região Autónoma dos<br />

Açores, Secretaria Regional da Habitação e Equipamentos.<br />

References<br />

Carmo, R. (2004). Geologia estrutural da região Povoação -<br />

Nordeste (ilha de S. Miguel, Açores). MSc thesis, Azores<br />

Univ., 133 p.<br />

DeMets, C., R. Gordon, D. Argus, S. Stein (1994). Effect of<br />

recent revisions to the geomagnetic reversal time scale<br />

on estimates of current plate motions. Geophys. Res.<br />

Lett. 21(20), pp. 2191-2194.<br />

Féraud, G., I. Kaneoka, C.J. Allegre (1980). K-Ar Ages and<br />

Stress Pattern in the Azores - Geodynamic Implications.<br />

Earth and Planet. Sci Lett. 46(2), pp. 275-286.<br />

Fernandes, R. M. S., B.A.C. Ambrosius, R. Noomen, L.<br />

Bastos, M.J.R. Wortel, W. Spakman, R. Govers (2003).<br />

The relative motion between Africa and Eurasia as<br />

derived from ITRF2000 and GPS data, Geophys. Res.<br />

Lett. 30 (16), 1828, pp. 1-1 - 1-5. doi:<br />

10.1029/2003GL017089.<br />

Ferreira, T., J. L. Gaspar, G. Queiroz (1993).<br />

Considerações sobre as emanações gasosas da Furna<br />

do Enxofre (Ilha Graciosa, Açores). Açoreana VII (4), pp.<br />

603-612.<br />

Gaspar, J.L.; G. Queiroz (1995). Carta Vulcanológica dos<br />

Açores, ilha Graciosa. Sheets A and B, scale 1:10000.<br />

Ed. Azores Univ. and Câmara Municipal de Santa Cruz<br />

da Graciosa.<br />

Gaspar, J.L. (1996). Ilha Graciosa (Açores): História<br />

Vulcanológica e Avaliação do Hazard. PhD thesis,<br />

Azores Univ., 361p.<br />

Hipólito, A. (2009). Geologia Estrutural da ilha Graciosa –<br />

Enquadramento no âmbito da Junção Tripla dos Açores.<br />

MSc thesis, Azores Univ., 155 p.<br />

Hipólito, A., F. Viveiros, R. Carmo, C. Silva, N. Cabral, J.<br />

Cabral, V. Alfama, T. Ferreira, J.L. Gaspar (2011). Soil<br />

degassing surveys as a tool to identify hidden faults in<br />

volcanic areas: preliminary results at the Ribeira Grande<br />

graben (Fogo Volcano, S. Miguel Island, Azores).<br />

Geophysical Research abstract. 13, EGU2011-8992.<br />

Lourenço, N., J. Miranda, J. Luís, A. Ribeiro, L. Mendes-<br />

Victor, J. Madeira, H. Needham (1998). Morpho-tectonic<br />

analysis of the Azores Volcanic Plateau from a new<br />

bathymetric compilation of the area. Marine Geophys.<br />

Res. 20, pp. 141-156.<br />

Luís, J.F., J. M. Miranda (2008). Reevaluation of magnetic<br />

chrons in the North Atlantic between 35ºN and 47ºN:<br />

Implications for the formation of yhe Azores Triple<br />

Junction and associated plateau. J. of Geophys. Res.<br />

113, B1o1o5. doi:10.1029/2007Jb005573.<br />

Madeira, J. (1998). Estudos de neotectónica nas ilhas do<br />

Faial, Pico e S. Jorge: uma contribuição para o<br />

conhecimento geodinâmico da junção tripla dos Açores.<br />

PhD thesis, Lisbon Univ., 428 p.<br />

Madeira, J., A. Brum da Silveira (2003). Active tectonics<br />

and first paleoseismological results in Faial, Pico and S.<br />

Jorge Islands (Azores, Portugal). Annals of Geophysics<br />

46 (5), pp. 733-761.<br />

Ortner, H., F. Reiter, P. Acs (2002). Easy handling of<br />

tectonic data: the programs TectonicVB for Mac and<br />

TectonicsFP for Windows. Computers and Geosciences<br />

28, 1193-1200.<br />

Sella, G., T. H. Dixon, A. Mao (2002). REVEL: a model for<br />

recent plate velocities from space geodesy. J. Geophys.<br />

Res. 107 (B4), 2081, pp. 11-1 – 11-12. doi:<br />

10.1029/2000JB000033.<br />

Silva, M. (2005). Caracterização da sismiciade histórica dos<br />

Açores com base na reinterpretação de dados de<br />

macrossísmica: contribuição para a avaliação do risco<br />

sísmico nas ilhas do Grupo Central. MSc thesis. Azores<br />

Univ., 163 p.<br />

Vogt, P.R., W.Y. Jung (2004). The Terceira Rift as hyperslow,<br />

hotspot dominated oblique spreading axis: A<br />

comparison with other slow-spreading plate boundaries.<br />

Earth Planet. Sci. Lett. 218, pp. 77–90.<br />

doi:10.10116/S0012-821X(03)00627-7.<br />

87


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EVIDENCE FOR HOLOCENE TSUNAMI-IMPACT ALONG THE SHORELINE OF OMAN<br />

Gösta Hoffmann (1, Klaus Reicherter (2), Thomas Wiatr (2), Christoph Grützner (2)<br />

(1) Department of Applied Geosciences, German University of Technology in Oman, PO Box 1816, Athaibah, Muscat, PC<br />

130, Sultanate of Oman, email: goesta.hoffmann@gutech.edu.om<br />

(2) RWTH Aachen University, Inst. für Neotektonik und Georisiken, Aachen, 52056, Germany<br />

Abstract (Evidence for Holocene tsunami-impact along the shoreline of Oman): Three independent sets of evidence of past<br />

tsunami along the coastline of Oman are reported. The rocky coastline of the Sultanate of Oman between Fins and Sur is<br />

decorated by a number of large boulders and boulder accumulations forming ramparts. The boulders occur as individual blocks of<br />

almost 50 tons weight, as imbricated sets and “boulder trains”. The coast is made up of folded Tertiary limestones and beach rock<br />

of Quaternary age. The transport distance from the fractured cliff front of 6-10 m height above mean sea level varies between<br />

several meters up to 70 m inland. We found individual blocks of recent corals and overturned blocks with oysters and pools. T-<br />

LIDAR was used to analyse geomorphologic features and for volumetric estimates of boulder weight. Tropical cyclones such as<br />

Gonu in 2007 or Phet in 2010 as well as historical tsunamis are known to have affected Oman’s coastline in the past. Coastal<br />

changes by cyclones are known to have been negligible; therefore, we interpret the boulder ridges as tsunamigenic deposits.<br />

Additionally, fine grained lagoonal sediments were analyzed. A distinct shell layer with allochthonous species is documented. A<br />

tsunamigenic origin is most likely. Although no dating evidence of the observed boulder and lagoon deposits is available at the<br />

moment we conclude that the 1945 Makran tsunami affected Oman’s coastline. This conclusion is based on interviews with local<br />

people.<br />

Key words: Oman, tsunami, boulder deposits, T-LiDAR<br />

INTRODUCTION<br />

Recent tsunami events like the Indian Ocean tsunami<br />

on 26th December 2004 and the Tôhoku earthquake<br />

and tsunami on 11th March 2011 resulted in large<br />

numbers of casualties and immense damage to<br />

infrastructure. These events underline the need for<br />

tsunami hazard research and assessment for any<br />

potentially vulnerable region. In most cases this can<br />

only be done by studying past tsunami records. The<br />

coastlines of the Sultanate of Oman are prone to<br />

various natural hazards such as tropical cyclones,<br />

landslides and tsunamis. The devastating effects of<br />

the cyclone Gonu, caused by flash floods and<br />

landslides in June 2007 illustrated the need to<br />

investigate the recurrence intervals of such events in<br />

order to assess the vulnerability and to mitigate<br />

damages. So far no scientific research concerning<br />

recurrence intervals of natural hazards has been<br />

carried out. However, studies published by<br />

Heidarzadeh et al. (2008a, 2008b, 2009) and Jordan<br />

(2008) reveal past tsunami events in the Indian<br />

Ocean with possible effects on the coastline of Oman<br />

(Fig. 1). As the population of Oman and the<br />

neighboring countries is concentrated along the<br />

coastline and large infrastructure projects are<br />

planned or already completed a holistic scientific<br />

approach to decipher the geological record of past<br />

extreme events is overdue. On 27 th November 1945<br />

an earthquake occurred in the Makran Subduction<br />

Zone offshore Pakistan and triggered a tsunami<br />

(Jaiswal et al., 2009). Up to 4000 casualties were<br />

reported along the coastlines of NW India and<br />

Pakistan, including 5 m run-up along the coastlines of<br />

the Sultanate of Oman. Donato et al. (2008, 2009)<br />

analysed shallow sediment cores from the lagoon in<br />

Sur and recorded a 5-25 cm thick shell bed close to<br />

the surface. Based on the taphonomy and<br />

fragmentation a tsunamigenic origin is discussed as<br />

the most likely form of deposition related to the 1945<br />

tsunami. However, there are almost no historical<br />

documents available for Oman for this period as the<br />

country was isolated with no international contacts<br />

until the 1970s, living conditions were poor and no<br />

modern technology was in use. We report geological<br />

and historical evidence for the tsunami along Oman's<br />

coastline. These evidence are: (a) fine grained<br />

lagoon sediments, which show distinct layers with<br />

allochthonous, offshore species (mollusks and<br />

foraminifera); (b) boulder deposits encountered along<br />

cliff-coastlines and (c) eyewitness-reports of old<br />

people we interviewed.<br />

OBSERVATIONS<br />

The coastal area under investigation is situated in the<br />

eastern part of Oman between the cities Quariat and<br />

Sur. The area is sparsely populated as most of the<br />

country; small fishing villages are scattered along the<br />

coast. Only since 2008 there is a paved road<br />

connecting the cities Quariat in the north and Sur in<br />

the south.<br />

The geology of the area is dominated by Paleogene<br />

to Neogene limestone formations, which rise from the<br />

coast up to 1500 m to form the Selma Plateau.<br />

Geomorphologic evidence of Quaternary land-uplift is<br />

obvious along the entire coastline: coast-parallel,<br />

wave-cut terraces are encountered up to elevations<br />

of 300 m. Within the study area these terraces are<br />

88


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 1: Historical tsunamis in the Indian Ocean and working area in the Sultanate of Oman.<br />

cut into Paleocene-Early Eocene limestone<br />

formations. Quaternary deposits are either of fluvial<br />

origin or ancient to subrecent littoral deposits, usually<br />

preserved as beachrock. In most cases only<br />

erosional remnants of the beachrock are found and<br />

the underlying older strata dominate along the cliff<br />

coast. Several intertidal lagoons exist in the vicinity of<br />

Sur and Ras al Hadd. These lagoons serve as<br />

geological archives with a preservation potential for<br />

palaeo-tsunami and were investigated during several<br />

field campaigns in 2010.<br />

We collected seven sediment cores at various<br />

locations within Sur lagoon. The deepest core<br />

reaches 10 m below the present surface. The<br />

sequence is characterized by silty fine sand in the<br />

lower part (10-6 m) and fine-sand in the upper part (6<br />

– 0 m). Within the uppermost meter several distinct<br />

shell beds were identified. The shell and foraminifera<br />

assemblage contains allochthonous species living in<br />

the subtidal zone and offshore. Additionally, we<br />

collected 4 sediment cores in the lagoon of Ras al<br />

Hadd. The longest core is 3 m long. The base of the<br />

sequence is made up of sandy gravel partly<br />

Fig. 2: Study area along the east coast of the Sultanate of Oman. Inset shows bathymetric sections.<br />

89


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 3: Study area along the east coast of the Sultanate of Oman with the observed ramparts.<br />

cemented as beachrock, overlain by fine to medium<br />

sand. The mollusk and foraminifera assemblages in<br />

the upper 100 cm show a variety of species.<br />

Especially some bivalve species within this layer are<br />

allochthonous as their habitat is characterized as<br />

subtidal and offshore.<br />

In our 2011 survey we found boulder deposits south<br />

of the village of Fins (Fig. 2) and more distal finergrained<br />

deposits yielding shells and coral blocks on<br />

the cliff top and to approx. 60-70 m inland (Fig. 3).<br />

The blocks are partly imbricated and reach volumes<br />

of more than 20 m 3 (determined with terrestrial LiDAR<br />

scanning), corresponding to a weight of almost 50<br />

tons. Also, we found so called tsunami boulder trains,<br />

where blocks are aligned in a row (Fig. 3). Some<br />

blocks are toppled or upright with hit marks on the<br />

surface, erosional pot holes, Lithophaga borings and<br />

attached oysters (which provide dating material,<br />

dating is in progress). The boulders form ramparts<br />

and have a wavy, lobe-like pattern. Most blocks have<br />

a platy shape, which origins in layer thickness of<br />

reworked material, mainly beach rock and Tertiary<br />

limestones. We measured the long axis (a-axis) of 60<br />

boulders, a vast majority is oriented N30, possibly<br />

pointing towards the wave direction. The cliff tops are<br />

“cleaned”, however, drift wood of the tropical cyclone<br />

Phet in 2010 is found in height of approximately 6 m<br />

above mean sea level. Inland, boulders are found in<br />

a gravel/sand matrix with various fossil remains like<br />

shell and corals (Fig. 4). The finer-grained layers<br />

show fining-up cycles. We also interviewed old<br />

people living in the towns of Sur and Tiwi. An old<br />

man in Sur recalled an event that happened most<br />

probably during the 1940s: first the sea retreated,<br />

then, two waves washed onshore. The event took<br />

place at 02:00 am.<br />

Another old man of the village of Tiwi reported from<br />

hearsay, as he was born in 1946. He heard about an<br />

event that destroyed the local graveyard in the<br />

1940s. He described that the graveyard was located<br />

much further inland. Furthermore, he gave an<br />

account of fish (sardines) and mollusks (oysters)<br />

being washed into the Wadi Shab. The women who<br />

used to get freshwater from the wadi could not walk<br />

there anymore, but boats had to be used. The marine<br />

fishes lived in the wadi after the event.<br />

DISCUSSION<br />

The uppermost ~1m in the lagoon of Sur as well as in<br />

the lagoon of Ras al Hadd clearly indicates an eventlayer<br />

which can be either storm- or tsunamigenerated<br />

(see Kortekaas and Dawson 2007). As the<br />

lagoons are intertidal, reworking and bioturbation of<br />

the sediments is a common phenomenon that<br />

hampers a clear stratification. Boulders deposits<br />

along the east coast of the Sultanate of Oman form<br />

ramparts between Fins and Sur. Inland, boulder<br />

deposits are incorporated in finer-grained sediments.<br />

Observations of the coastline changes and<br />

modifications of the last two very strong tropical<br />

cyclones Gonu and Phet rather exclude tropical<br />

cyclones a “moving agent” for the large 50 tons<br />

boulders. This was also proven by comparing time<br />

series of Google Earth, where blocks are detectable,<br />

but remained in the position (before and after).<br />

90


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The information gained from the interview in Sur is<br />

very helpful. The timing of an event resulting in seawater<br />

entering the house was given with 02:00 am<br />

which is 22.00 GMT (local time zone is GMT +4).<br />

This fits quite well with the time of the Makran<br />

earthquake on 27.11.1945 which is reported by<br />

Pendse (1948) as 21:57 GMT. The modeled travel<br />

time of the tsunami wave is 20-30 minutes<br />

(Heidarzadeh and Kijko. 2011). The description of a<br />

retreating sea fits general tsunami descriptions. The<br />

arrival of two separate waves fits observations in<br />

India (see Rajendran et al. 2008) where also two<br />

waves with disparity in arrival time are reported. A<br />

submarine landslide is assumed for the second<br />

wave. The descriptions given in the interview in Tiwi<br />

cannot unambiguously be related to any know event.<br />

However, the accounts are more likely to be the<br />

effect of a tsunami wave rather than wadi-flooding.<br />

Acknowledgements: This study was financially supported<br />

by the German Research Foundation (DFG-project<br />

Re 1361/14-1) and GUTech in Mascat. Our students Sina<br />

Rausch, Kathrin Wagner and Tobias Rausch are thanked<br />

for field and lab support, and Tobias also for graphic design.<br />

References<br />

Fig. 4: Sketch of the possible tsunamigenic<br />

deposits.<br />

Hence, we propose a tsunamigenic event being<br />

responsible for the rampart formation and the boulder<br />

deposits along this part of the coast. Dating of<br />

oysters, which have grown on the blocks and died<br />

during the deposition, is in progress.<br />

Donato, S. V., Reinhardt, E.G., Boyce, J.I., Rothaus, R.,<br />

Vosmer, T. (2008). Identifying tsunami deposits using<br />

bivalve shell taphonomy, Geology, 36, 199-202.<br />

Donato, S. V., Reinhardt, E. G., Boyce, J. I., Pilarczyk, J. E.,<br />

Jupp, B. P. (2009) Particle-size distribution of inferred<br />

tsunami deposits in Sur lagoon, Sultanate of Oman, Mar.<br />

Geol., 257, 54-64.<br />

Heidarzadeh, M., Pirooz, M., Zaker, N. H., Yalciner, A. C.,<br />

Mokhtari, M., Esmaeily, A. (2008a) Historical tsunami in<br />

the Makran subduction zone off the southern coasts of<br />

Iran and Pakistan and results of numerical modeling,<br />

Ocean Eng., 35, 774-786.<br />

Heidarzadeh, M., Pirooz, M., Zaker, N. H., Synolakis, C.<br />

(2008b). Evaluating tsunami hazard in the northwestern<br />

Indian Ocean, Pure Appl. Geophys., 165, 2045-2058,<br />

Doi:10.1007/s00024-008-0415-8.<br />

Heidarzadeh, M., Pirooz, M., Zaker, N. H., Yalciner, A. C.<br />

(2009). Preliminary estimation of the tsunami hazards<br />

associated with the Makran subduction zone at the<br />

northwestern Indian Ocean, Nat. Hazards, 48, 229-243,<br />

Doi:10.1007/s11069-008-9259-x.<br />

Heidarzadeh, M., Kijko, A. (2011). A probabilistic tsunami<br />

hazard assessment for the Makran subduction zone at the<br />

northwestern Indian Ocean, Nat. Hazards 56, 577-593.<br />

Jaiswal, R., Singh, A., Rastogi, B. (2009). Simulation of the<br />

Arabian Sea tsunami propagation generated due to 1945<br />

Makran earthquake and its effect on western parts of<br />

Gujarat (India), Nat. Hazards, 48, 245-258,<br />

Doi:10.1007/s11069-008-9261-3.<br />

Jordan, B. R. (2008). Tsunamis of the Arabian Peninsula - a<br />

guide of historic events, Science of Tsunami Hazards, 27,<br />

31-46.<br />

Kortekaas, S., Dawson, A. G. (2007). Distinguishing<br />

tsunami and storm deposits: An example from Martinhal,<br />

SW Portugal, Sed. Geol. 200, 208-221.<br />

Pendse, C. G. (1948). The Makran earthquake of the 28th<br />

of November, 1945, Scientific Notes, Indian<br />

Meteorological Dept. 10, 141-145.<br />

Rajendran, C. P., Ramanamurthy, M.V., Reddy, N.T.,<br />

Rajendran, K. (2008). Hazard implications of the late<br />

arrival of the 1945 Makran tsunami, Current Science 95,<br />

1739-1743.<br />

91


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

MONITORING COASTAL CHANGES ON THE IONIAN ISLANDS (NW-GREECE)<br />

BY MULTI-TEMPORAL TERRESTRIAL LASER SCANNING<br />

Hoffmeister, Dirk (1, Ntageretzis, Konstantin (2), Tilly, Nora (1), Curdt, Constanze (1), Bareth, Georg (1), Brückner, Helmut (1),<br />

Andreas Vött (2)<br />

(1) Institute for Geography, University of Cologne, Albertus-Magnus-Platz, 50923 Cologne. GERMANY. Email:<br />

dirk.hoffmeister@uni-koeln.de<br />

(2) Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong> <strong>Mainz</strong>. Johann-Joachim-Becher-Weg 21, 55099 <strong>Mainz</strong>.<br />

GERMANY. Email: konstantin.ntageretzis@uni-mainz.de<br />

Abstract: This study deals with the application of terrestrial laser scanning (TLS) for monitoring both gradual and abrupt coastal<br />

changes, the latter for example associated to seismic events. TLS is a widely-used method for accurate measurements and<br />

monitoring of surface changes over time. Several sites in northwestern Greece were surveyed during TLS campaigns in 2009 and<br />

2010. Our main objectives were (i) to measure the present configuration and geometry of a mushroom rock at Poros (Cefalonia<br />

Island), which repeatedly experienced co-seismic uplift and (ii) to monitor annual coastal changes of selected shoreline sections<br />

on Lefkada Island. At the Poros mushroom rock, up-to-date 3D data of different generations of notches, which were uplifted by<br />

palaeoseismic events offer a solid base for evaluating future event-related uplifts. At Lefkada, TLS datasets of different years were<br />

compared with each other, resulting in a net balance of elevation changes. Results are checked by photographs taken by a<br />

camera on top of the TLS. Comparing the years 2009 and 2010, clear differences in the grain size distribution on beach sediments<br />

can be observed as well as slight changes in the coastline configuration. However, shadowing effects of the complex surface and<br />

noise caused by sea water represent lead to considerable problems analysing the data.<br />

Key words: terrestrial laser scanning, coseismic uplift, coastal change, Ionian Islands<br />

INTRODUCTION<br />

Western Greece, especially the Ionian Islands,<br />

belong to the most active seismic regions in the<br />

Mediterranean Sea as they are directly exposed to<br />

the Hellenic Trench system and the Cefalonia<br />

transform fault (Cocard et al. 1999, Hollenstein et al.<br />

2008). Transform faulting, collision, and subduction<br />

can be found within less than 100 km distance<br />

(Sachpazi et al. 2000).<br />

ideal approach, easy to realize by multi-temporal<br />

surveys. Laser scanning is an active remote sensing<br />

technique, also known as Light Detection and<br />

Ranging (LIDAR).<br />

The tectonic setting of Cefalonia Island, located at<br />

the northwestern edge of the Hellenic Trench,<br />

consists of E dipping and NW/NNW-SE/SSE striking<br />

thrust sheets (Stiros et al. 1994). The vertical<br />

movement of Cefalonia Island is dominated by<br />

gradual subsidence interrupted by co-seismic uplifts<br />

(Hollenstein et al. 2008).<br />

In the southeastern part of Cefalonia, a mushroom<br />

rock (Fig. 1) in the harbour of Poros is well known for<br />

two uplifted Holocene notches at +0.6 m and +1.2 m<br />

above present sea level (a.s.l.), respectively<br />

(Pirazzoli et al. 1994, Stiros et al. 1994). The lower<br />

notch was co-seismically uplifted during the 1953<br />

earthquake, and the upper notch during a seismic<br />

event around 1.500 yr BP (Stiros et al. 1994,<br />

Pirazzoli et al. 1994). Due to the high seismic activity<br />

of the region, further uplift due to future earthquakes<br />

can be expected.<br />

For measuring the dimensions and for monitoring<br />

geomorphological features which are moved, due to<br />

co-seismic crustal movements, laser scanning is an<br />

Fig. 1: Mushroom rock with two elevated notches in the<br />

harbour of Poros (southeastern Cefalonia) (Photo: K.<br />

Ntageretzis 2009).<br />

Direct measurement of distances and angles<br />

between the sensor and reflecting targets provides<br />

highly accurate 3D point clouds. LIDAR can be<br />

applied from the ground surface as terrestrial laser<br />

scanning (TLS) (Heritage & Large 2009). The<br />

interpretation of 3D point clouds is used within the<br />

framework of various applications (Vosselmann &<br />

Maas 2010). For example, TLS is used to study<br />

erosion and denudation processes along cliffs (Lim et<br />

92


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

al. 2010) and hillslopes, such as landslides and rock<br />

fall (Abellan et al. 2011, Nguyen et al. 2011). Richter<br />

et al. (2011) quantified the erosion of a dune cliff and<br />

the change of beach width by multi-temporal airborne<br />

laser scanning. Rosser et al. (2005) used terrestrial<br />

laser scanning for dune cliff erosion monitoring. Also,<br />

the rapidly changing geomorphology of fluvial<br />

systems can be monitored by laser scanning<br />

approaches (Heritage & Milan 2009).<br />

In this study, we used TLS to retrieve high-resolution<br />

data of the introduced mushroom rock of Poros<br />

(Cefalonia Island), which provides a reliable base for<br />

long-term monitoring of uplifting movements by<br />

seismic events. Furthermore, a recent notch at<br />

Kaminia Beach (Lefkada Island) was scanned to<br />

monitor annual changes in the coastal sedimentary<br />

system by gradual sediment transport and the<br />

influence of storms.<br />

LOCATIONS<br />

Several sites were surveyed by TLS along the shores<br />

of the eastern Ionian Sea within the framework of an<br />

interdisciplinary research project on palaeo-tsunami<br />

impacts (Vött et al. 2010). TLS field campaigns were<br />

carried out in 2009 and 2010. In this paper, we focus<br />

on the sites of Poros (Cefalonia Island) and Kaminia<br />

Beach (Lefkada Island) (see Fig. 2).<br />

For the mushroom rock at Poros four scan positions<br />

were chosen to retrieve a good coverage. For the<br />

area of Kaminia Beach, two scan positions were<br />

selected, which cover the whole site. At one of these<br />

two scan positions, the scanner was tilted to obtain a<br />

view of the entire study area. The resolution for a<br />

detailed scan was around 0.8 cm at a distance of<br />

10m.<br />

A Topcon HiPer Pro DGPS instrument was used to<br />

measure the different scan positions with a relative<br />

accuracy of 1 cm. Positions were recorded in the<br />

WGS84 system, UTM Zone 34 N. Furthermore,<br />

positions of cylindrical reflectors on ranging poles<br />

were recorded.<br />

For the annual measurements, the base point of the<br />

local DGPS net was marked by a metal mark and<br />

measured 500 times to achieve a mean, enhanced<br />

position. All measurements in relation to this base<br />

point were within the stated accuracy. Additionally, in<br />

each year the similar scan positions were chosen.<br />

Point clouds were subsequently georeferenced by<br />

the DGPS points of the scan positions and the<br />

reflectors. Afterwards, the registration was enhanced<br />

by the ICP-algorithm. The mushroom rock at Poros<br />

was reconstructed using the Geomagic Studio 12<br />

software and textured according to the photographs.<br />

The results of the two campaigns at Kaminia Beach<br />

were clipped and high resolution digital elevation<br />

models (HRDEMs) for each year were established. In<br />

a first step, the results of the HRDEM comparison<br />

were checked visually by the pictures of mounted<br />

camera.<br />

RESULTS AND DISCUSSION<br />

The mushroom rock at Poros with the two<br />

generations of uplifted notches were reconstructed<br />

as an 3D-object and textured by photographs for a<br />

near-realistic model (Fig. 3). The data set of Fig. 3 is<br />

a valuable tool to precisely measure the present<br />

situation as well as the amount of uplift by<br />

palaeoseismic events. It also provides a reliable base<br />

for monitoring after future events.<br />

Fig. 2: Map of the study areas in the Ionian Sea (NW<br />

Greece). Poros lies on Cefalonia Island and Kaminia Beach<br />

on the Island of Lefkada. Map based on Modis and ASTER<br />

GDEM.<br />

METHODS<br />

We used a TLS LMS-Z420i Riegl instrument for this<br />

survey. The time-of-flight range measurements have<br />

an accuracy of 0.6 cm with a range between 2 m and<br />

1,000 m. A high-resolution digital Nikon D200 camera<br />

mounted on the head of the laser scanner took RGBphotos<br />

which were used to colourize the TLS point<br />

clouds and to control the results. For data acquisition<br />

and first post-processing steps, the RiSCAN PRO<br />

Riegl-software was applied.<br />

Fig. 3: Perspective view of the mushroom rock model at<br />

Poros (Cefalonia Island).<br />

93


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 4: Morphological net elevation changes of the shoreline at Kaminia Beach between 2009 and 2010 calculated from TLS data<br />

(center). Elevation changes go hand in hand with changes in the sedimentary structure along the shoreline (top and bottom<br />

photos); exemplary areas are marked by ellipses and discussed in the text.<br />

Comparing the point clouds from the years 2009 and<br />

2010 regarding the notch at Kaminia Beach no<br />

changes were observed. However, comparing the<br />

beach morphology of 2009 with the one of 2010<br />

considerable changes in the littoral zone can be<br />

observed.<br />

Results of the net elevation differences calculated<br />

from multi-temporal TLS data with according<br />

photographs are illustrated in Fig. 4. The sea weed,<br />

which in 2009 covered a large part of the beach has<br />

mostly disappeared in 2010 (yellow and blue<br />

ellipses). At the same time, a deposit of sandy<br />

sediments (orange ellipse) was considerably reduced<br />

up to 1 m in size. Another sand cover close to a wall<br />

(white ellipse) was replaced by coarse-grained<br />

material in 2010. The yellow ellipse marks an area<br />

close to the Kaminia notch where gravel was<br />

accumulated inland and the coast line was<br />

smoothened. This proves active littoral abrasion and<br />

documents that the present notch is formed by wave<br />

erosion and not by bio-erosion. The red ellipse shows<br />

an area where a sand cover was partly erroded and<br />

underlying gravel exhumed. However, some pieces<br />

of gravel with 30-50 cm diameter seem to have also<br />

been displaced. We assume that the observed<br />

changes in the sedimentary budget are mainly due to<br />

winter storm events.<br />

Post-processing of point clouds turned out to be<br />

difficult due to noise caused by water and shadow<br />

effects at places where dense gravel occurs (Fig. 5).<br />

A major problem, especially at Kaminia beach, is that<br />

measurements from the seaside are not possible.<br />

CONCLUSIONS AND OUTLOOK<br />

Our studies show that the application of TLS in<br />

different littoral settings is an appropriate tool for<br />

monitoring both abrupt and gradual coastal changes.<br />

94


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

With regard to palaeoseismic research, TLS based<br />

high-resolution 3D models of geomorphologic<br />

features moved by co-seismical uplift yield detailed<br />

and precise data, for instance at Poros (Cefalonia<br />

Island). Moreover, the 3D model may be used to<br />

quantify effects from future earthquakes.<br />

Furthermore, multi-temporal TLS datasets allows to<br />

detect and to monitor gradual coastal changes.<br />

Fig. 5: Two major methodological problems arose in<br />

coastal monitoring using TLS in littoral zones: noise caused<br />

by water and shadowing effects by coarse material in the<br />

upper littoral zone.<br />

The results in this paper represent an essential<br />

progress in coastal monitoring by calculating a net<br />

balance of sediment budget achieved by a<br />

comparison of HRDEMs from different years,<br />

generated by TLS data .<br />

Acknowledgements: We gratefully acknowledge financial<br />

support by the German Research Foundation (DFG, VO<br />

938/3-1).<br />

References<br />

Abellan, A., J. M. Vilaplana, J. Calvet, D. Garcia-Selles & E.<br />

Asensio (2011). Rockfall monitoring by Terrestrial Laser<br />

Scanning a case study of the basaltic rock face at<br />

Castellfollit de la Roca (Catalonia, Spain). Nat. Hazards<br />

Earth Syst. Sci. 11, 829-841.<br />

Cocard, M., H.-G. Kahle, Y. Peter, A. Geiger, G. Veis, S.<br />

Felekis, D. Paradissis & H. Billiris (1999). New<br />

constraints on the rapid crustal motion of the Aegean<br />

region: Recent results inferred from GPS measurements<br />

(1993-1998) across the West Hellenic Arc, Greece. Earth<br />

Planetary Science Letters 172, 39–47.<br />

Heritage, G. L. & A.R.G. Large (Eds.) (2009). Laser<br />

Scanning for the Environmental Sciences. 1. ed., Wiley-<br />

Blackwell, Chichester, UK. 278 p.<br />

Heritage, G. L. & D.J. Milan (2009). Terrestrial Laser<br />

Scanning of grain roughness in a gravel-bed river.<br />

Geomorphology 113 (1-2), 4-11. doi:<br />

10.1016/j.geomorph.2009.03.021<br />

Hollenstein, C., M.D. Müller, A. Geiger & H.G. KAHLE<br />

(2008). Crustal motion and deformation in Greece from a<br />

decade of GPS measurements, 1993-2003.<br />

Tectonophysics 449, 17-40.<br />

Lim, M., N.J. Rosser, D.N. Petley & M. Keen (2011).<br />

Quantifying the Controls and Influence of Tide and Wave<br />

Impacts on Coastal Rock Cliff Erosion. Journal of Coastal<br />

Research 27 (1), 46-56. doi: 10.2112/jcoastres-d-09-<br />

00061.1<br />

Nguyen, H. T., T.M. Fernandez-Steeger, T. Wiatr, D.<br />

Rodrigues, D. & R. Azzam (2011). Use of terrestrial laser<br />

scanning for engineering geological applications on<br />

volcanic rock slopes - an example from Madeira island<br />

(Portugal). Nat. Hazards Earth Syst. Sci. 11 (3), 807-817.<br />

doi: 10.5194/nhess-11-807-2011<br />

Pirazzoli, P.A., S.C. Stiros, J. Laborel, F. Laborel-Deguen,<br />

M. Arnold, S. Papageorgiou & C. Morhange (1994). Late-<br />

Holocene shoreline changes related to paleoseismic<br />

events in the Ionian Islands, Greece. The Holocene 4 (4),<br />

397-405.<br />

Richter, A., D. Faust & H.G. Maas (2011, in press). Dune<br />

cliff erosion and beach width change at the northern and<br />

southern spits of Sylt detected with multi-temporal Lidar.<br />

Catena.<br />

Rosser, N. J., D. N. Petley, M. Lim, S. A. Dunning & R. J.<br />

Allison (2005). Terrestrial laser scanning for monitoring<br />

the process of hard rock coastal cliff erosion. Quarterly<br />

Journal of Engineering Geology and Hydrogeology 38(4),<br />

363-375. doi: 10.1144/1470-9236/05-008<br />

Sachpazi, A., C. Hirn, M. Clement, F. Laigle, E. Haslinger,<br />

P. Kissling, Y. Charvis, J.C. Hello, M. Lepine, M. Sapin, &<br />

J. Ansorge (2000). Western Hellenic subduction and<br />

Cephalonia transform: local earthquakes and plate<br />

transport and strain. Tectonophysics 319 (4), 301-319.<br />

Stiros, S.C., P. A. Pirazzoli, J. Laborel & F. Laborel-Deguen<br />

(1994): The 1953 earthquake in Cephalonia (Western<br />

Hellenic Arc): coastal uplift and halotectonic faulting.<br />

Geophysical Journal International 117, 834-849.<br />

Vött, A., G. Bareth, H. Brückner, C. Curdt, I. Fountoulis, R.<br />

Grapmayer, H. Hadler, D. Hoffmeister, N. Klasen, F.<br />

Lang, P. Masberg, S. M. May, K. Ntageretzis, D.<br />

Sakellariou & T. Willershäuser (2010). Beachrock-type<br />

calcarenitic tsunamites along the shores of the eastern<br />

Ionian Sea (western Greece) – case studies from<br />

Akarnania the Ionian Islands and the western<br />

Peloponnese. Zeitschrift für Geomorphologie N.F., Suppl.<br />

Issue 54 (3), 1-50.<br />

Vosselmann, G. & H.-G. Maas (Eds.) (2010). Airborne and<br />

terrestrial laser scanning. 1st ed., Whittles Publishing,<br />

Dunbeath, UK. 311 p.<br />

95


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PRELIMINARY REPORT ON THE VODICE FAULT ACTIVITY AND ITS POTENTIAL FOR<br />

SEISMIC HAZARD IN THE LJUBLJANA BASIN, SLOVENIA<br />

Jamšek, Petra (1), Lucilla Benedetti (2), Miloš Bavec (1), Jure Atanackov (1), Marko Vrabec (3), Andrej Gosar (4,3)<br />

(1) Geological Survey of Slovenia. Dimieva 14. SI-1000 Ljubljana. SLOVENIA. Email: petra.jamsek@geo-zs.si<br />

(2) Centre Européen de Recherche et d’Enseignement des Géosciences de l’Environnement. BP 80, Europôle Méditerranéen de<br />

l’Arbois. 13545 Aix-en-Provence Cedex 4. FRANCE.<br />

(3) University of Ljubljana, Faculty of Natural Sciences and Engineering, Department of Geology. Aškereva 12. SI-1000<br />

Ljubljana. SLOVENIA.<br />

(4) Slovenian Environment Agency, Seismology and Geology Office. Dunajska 47. SI-1000 Ljubljana. SLOVENIA.<br />

Abstract (Preliminary report on the Vodice fault activity and its potential for seismic hazard in the Ljubljana Basin,<br />

Slovenia): In Ljubljana Basin (Slovenia) the Vodice fault was investigated to decipher its recent activity and seismic hazard<br />

potential for the densely populated central part of Slovenia. Preliminary geomorphological analysis and field observations suggest<br />

it may have been recently reversely active. Moreover, using preliminary optical stimulation ages we estimated a slip-rate along the<br />

Vodice fault at about 0.2 – 0.4 mm/yr over the last 115 ± 32 ka. An earthquake of magnitude 6.2 – 6.3 could be expected on this<br />

fault which may have been the source of the Ljubljana 1895 magnitude 6.1 earthquake. Further analysis will follow to better<br />

constrain characteristics of this fault and seismic hazard of the area.<br />

Key words: active fault, seismic hazard, Ljubljana Basin, Slovenia<br />

INTRODUCTION<br />

The Ljubljana Basin (Fig. 1), the most densely<br />

populated and urbanized area of Slovenia,<br />

experiences constant seismic activity and has been<br />

the site of strong historical seismic events with<br />

magnitude as high as 6.1 (intensity VIII–IX EMS-98,<br />

Ljubljana earthquake 1895 (Ribari, 1982)). The<br />

basin is filled with Quaternary sediments reaching a<br />

thickness of up to 280 m in some parts, which may<br />

significantly enhance site effects and therefore<br />

increase the earthquake hazard in the area (Gosar et<br />

al., 2010). However, active faults capable of<br />

producing strong magnitude earthquakes in this area<br />

are poorly known.<br />

The Ljubljana Basin probably results from<br />

transpression with dextral strike-slip movement along<br />

NE-SW faults and thrusting along smaller scale E-W<br />

structures (Vrabec, 2001; Bavec et al., 2003;<br />

Benedetti et al., 2000). In this report we focus on the<br />

Vodice fault escarpment, located 15 km north of<br />

Ljubljana, the capital city located within the<br />

Quaternary basin infill. This escarpment is offsetting<br />

Quaternary surfaces for 5 to 25 m along a length of<br />

10 – 11 km. This feature was previously described as<br />

a terrace riser of the Sava river (Žlebnik, 1971), and<br />

later as a reverse fault (Bavec et al., 2004; Verbi,<br />

2006). To decipher whether the Vodice fault has<br />

been recently active and to asses its potential for<br />

seismic hazard we performed a detailed study of its<br />

morphology and topography. Herein, we present our<br />

preliminary interpretations.<br />

METHODS<br />

Fig. 1: The Ljubljana Basin map with main active<br />

structures (modified according to Buser, 2009),<br />

historical earthquake epicentres with magnitude above<br />

3.9 (according to Živi, 2009; note that earthquakes<br />

magnitudes are obtained from macroseismic data and<br />

that locations of historical earthquake epicentres are<br />

not well constrained), and isoseisms of the Ljubljana<br />

earthquake 1895 (according to Lapajne, 1989).<br />

We investigated its surface expression trough<br />

geomorphological analysis of topographical maps (1 :<br />

5.000 and 1 : 10.000 scale), digital elevation model<br />

(resolution 5 m), aerial images, and 2.5 m resolution<br />

SPOT images in stereo pairs. Each alluvial surface<br />

was carefully mapped and a series of topographical<br />

profiles were extracted across, and parallel to the<br />

scarp from 5 m resolution digital elevation model.<br />

96


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Based on analysis of topographical profiles, detailed<br />

mapping and field observations we were able to<br />

evidence recent movement along the fault. Based on<br />

preliminary optical stimulation luminescence ages<br />

(Bavec et al., 2005) we evaluated the slip-rate along<br />

the Vodice fault, possible earthquake magnitude and<br />

average displacement per event.<br />

VODICE FAULT SURFACE EXPRESSION<br />

The flat Ljubljana Basin surface resulting from<br />

Quaternary deposition is perturbed in the area of<br />

Vodice by an unusual linear feature, ENE-WSW<br />

oriented, 10-11 km long scarp between Pšata river<br />

on the east and Sava river on the west (Fig. 2). East<br />

of Vodice, the scarp splays into two branches. A<br />

series of detailed topographic profiles levelled<br />

across, and parallel to the scarp show that the height<br />

of the scarp varies from 25 to 5 m for the southern<br />

branch and from 18 to 3 m for the northern branch<br />

(Figs. 2 and 3). Evidence of ongoing uplift is attested<br />

by the presence of abandoned streams across the<br />

structure and perched valleys (Fig. 3). Active streams<br />

strongly incise the northern, hanging wall<br />

compartment, this upper surface being clearly older<br />

as shown by its degradation mostly due to dolines.<br />

These observations lead us to interpret the Vodice<br />

scarp as the surface expression of an active reverse<br />

fault. The latter suggests that the scarp can not be<br />

interpreted as a terrace riser of a former course of<br />

the Sava river.<br />

Fig. 2: The Vodice fault surface expression.<br />

SUBSURFACE DEFORMATIONS<br />

Fig. 3: The series of topographic profiles across, and<br />

parallel to the Vodice fault scarp (red arrows = perched<br />

valeys). Note that the profile in the image is showing<br />

two fault branches because it is shalow, but in the<br />

depth the two branches are most probably joined into<br />

one fault.<br />

Further evidences of the fault activity are provided<br />

from subsurface deformations. Folding was reported<br />

in a clay pit at the eastern end of the scarp (Fig. 4,<br />

Drobne et al., 1960; Šifrer, 1961). Westward, where<br />

the fault cuts the N-S running Sava River, Quaternary<br />

conglomerates are also folded and offset (Fig. 5). In<br />

both cases the deformations are consistent with a<br />

reverse fault interpretation.<br />

FIRST EVALUATION OF SEISMIC HAZARD<br />

POTENTIAL<br />

Preliminary optical stimulation luminescence ages of<br />

deformed sediments, located at the eastern tip of the<br />

fault (Bavec et al., 2005), suggest an age of 115 ± 32<br />

ka for the upper alluvial surface (Fig. 2). Using this<br />

age and assuming a northward dip of 35 – 45°, we<br />

estimated a minimum slip-rate along the Vodice fault<br />

at about 0.2 – 0.4 mm/yr over the last 115 ± 32 ka.<br />

Fig. 4: Folded Quaternary sediments at the eastern end<br />

of the scarp reported by Šifrer (1961).<br />

On a 10 – 11 km long reverse fault, an earthquake of<br />

magnitude 6.2 – 6.3 could be expected. Such<br />

earthquake could trigger an average displacement of<br />

~ 0.5 m with an average recurrence time of 1200 –<br />

3000 yr (Wells & Coppersmith, 1994).<br />

97


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Acknowledgements: This work is funded by the Slovenian<br />

Research Agency, through 1) The project L1–2383<br />

Seismotectonic model of the Ljubljana Basin (cofounded by<br />

the Slovenian Environment Agency), 2) The programme P1<br />

– 0011 Regional geology, and 3) The support for young<br />

researcher (contract 1000-09-310068).<br />

References<br />

Fig. 5: One of the Vodice fault outcrops. Fault plane (red<br />

line) dips towards NNW for 35°. Features that may also be<br />

interpreted as near surface expression of faulting are<br />

marked in black.<br />

DISCUSSION<br />

This preliminary geomorphological analysis and field<br />

observations suggest recent activity along E-W<br />

trending 10-11 km long Vodice reverse fault and<br />

correlates well with results of previous studies (Bavec<br />

et al., 2004; Verbi, 2006). However, to further prove<br />

Vodice scarp as the surface expression of a reverse<br />

fault the geophysical investigations are planned.<br />

Also, to better constrain the Vodice fault<br />

displacements and its slip-rate, further accurate<br />

quantitative analysis of the morphology are<br />

warranted (levelling survey using theodolite or<br />

diferential GPS) and Quaternary geochronology to<br />

date the offset surfaces.<br />

A first approximation suggests the Vodice fault could<br />

trigger magnitude 6.2 – 6.3 earthquake events.<br />

Considering its location and proximity to Ljubljana,<br />

the Vodice fault is a candidate to be the source of the<br />

Ljubljana earthquake 1895. To better assess the<br />

seismic hazard of this area further analysis will<br />

follow, such as geophysical investigations of the fault<br />

geometry (high-resolution seismic reflection and<br />

ground penetrating radar profiling) and<br />

paleoseismological investigations along the fault to<br />

define its seismic behaviour and decipher its seismic<br />

history.<br />

Recognition of seismic hazard is crucial in areas<br />

such as Ljubljana Basin, where destructive<br />

earthquakes can represent a huge danger for<br />

population and infrastructure. Investigations in active<br />

tectonics are the first step towards seismic hazard<br />

assessment, leading to protection of human lives as<br />

well as to a decrease of economical damage in case<br />

of a destructive seismic event. To determine the<br />

seismic hazard in the Ljubljana Basin we will also<br />

extend our investigations to other presumably active<br />

faults bounding the basin.<br />

Bavec, M., B. Celarc, M. Demšar, M. Poljak, D. Rajver, I.<br />

Rižnar, F. Preusser, M. Toman, J. Žibrat, M. Novak & K.<br />

Hribernik, (2005). Izdelava geoloških kart: letno poroilo<br />

za leto 2005. Geological Survey of Slovenia. Ljubljana.<br />

Bavec, M., B. Celarc, M. Poljak, M. Demšar, D. Rajver, D.<br />

Skaberne, T. Verbi, I. Rižnar, F. Preusser & M. Culiberg,<br />

(2004). Izdelava geoloških kart. Karta aktivnih prelomov v<br />

Sloveniji: letno poroilo za leto 2004. Geological Survey<br />

of Slovenia. Ljubljana.<br />

Bavec, M., M. Poljak, M. Demšar, D. Rajver, M. Komac, M.<br />

Toman, S. Stojanova, B. Muši, M. Vrabec, T. Verbi & I.<br />

Rižnar, (2003). Izdelava geoloških kart. Karta aktivnih<br />

prelomov v Sloveniji : letno poroilo za leto 2003. 1. del :<br />

Raziskave na obmoju Ljubljansko-kranjske kotline.<br />

Geological Survey of Slovenia. Ljubljana. 56 f.<br />

Benedetti, L., P. Tapponnier, G. C. P. King, B. Meyer & I.<br />

Manighetti, (2000). Growth folding and active thrusting in<br />

the Montello region, Veneto, northern Italy. Journal of<br />

Geophysical Research – Solid Earth 105, 739–766.<br />

Buser, S. (2009). Geological map of Slovenia 1:250.000.<br />

Geological Survey of Slovenia. Ljubljana.<br />

Drobne, F., R. Pavlovec & A. Šercelj, (1960). Some<br />

analyses and problems of Pleistocene sediments at<br />

Lokarji near Vodice. Kamniški zbornik 6, 163–194.<br />

Gosar, A., J. Rošer, B. Šket Motnikar, P. Zupani, (2010).<br />

Microtremor study of site effects and soil-structure<br />

resonance in the city of Ljubljana (central Slovenia).<br />

Bulletin of earthquake engineering 8 (3), 571–592.<br />

Lapajne, J. (1989). Veliki potresi na Slovenskem – III.:<br />

potres v Ljubljani leta 1895. Ujma 3, 55–61.<br />

Ribari, V. (1982). Seismicity of Slovenia. Catalogue of<br />

earthquakes (792 A.D.–1981). Seismological Survey of<br />

Slovenia. Ljubljana. 649 p.<br />

Šifrer, M. (1961). The basin of Kamniška Bistrica during the<br />

Pleistocene period. Slovenian Academy of Sciences and<br />

Arts. Ljubljana. 211 p.<br />

Verbi, T. (2006). Quaternary-active reverse faults between<br />

Ljubljana and Kranj, central Slovenia. Razprave IV.<br />

razreda SAZU 47 (2), 101–142.<br />

Vrabec, M. (2001). Structural analysis of the Sava Fault<br />

zone between Trstenik and Stahovica. PhD thesis.<br />

University of Ljubljana. Faculty of Natural Sciences and<br />

Engineering. Department of Geology. 94 f.<br />

Wells, D. L. & K. J. Coppersmith, (1994). New Empirical<br />

Relationships among Magnitude, Rupture Length,<br />

Rupture Width, Rupture Area, and Surface Displacement.<br />

Bulletin of the Seismological Society of America 84 (4),<br />

974–1002.<br />

Živi, M. (2009). Catalogue of earthquakes in Slovenia.<br />

Internal documentation. ARSO – Slovenian Environment<br />

Agency. Ljubljana.<br />

Žlebnik, L. (1971). Pleistocene Deposits of the Kranj, Sora<br />

and Ljubljana Fields. Geologija 14, 5–51.<br />

98


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

DISTRIBUTION AND SEDIMENTARY CHARACTERISTICS OF TSUNAMI DEPOSITS ON<br />

PHRA THONG ISLAND, THAILAND<br />

Jankaew, Kruawun (1, Dominik Brill (2), Maria E. Martin (3), Yuki Sawai (4)<br />

(1) Department of Geology, Faculty of Science, Chulalongkorn University, Bangkok, 10330 THAILAND<br />

Email:kruawun.j@chula.ac.th<br />

(2) Department of Geography, University of Cologne, GERMANY<br />

(3) Department of Earth and Space Sciences, University of Washington, Seattle, USA<br />

(4) Geological Survey of Japan, National Institute of Advanced Industrial Science and Technology (AIST), Tsukuba, JAPAN<br />

Abstract (Paleotsunami): Phra Thong Island is located along the western coast of Thailand directly facing the source area of the<br />

2004 Indian Ocean tsunami. Here we report a distribution of paleotsunami sand layers on Phra Thong Island, collected during four<br />

years of field work (2007-2011). Sedimentary characteristics of paleotsunami deposits are compared with those of the 2004<br />

tsunami. Thicknesses of the 2004 and paleotsunamis sand layers are controlled by micro-topography and subjected to postdepositional<br />

alteration. Sand layers of 2004 and paleotsunami layers contain fining upward sequence(s), which suggest deposition<br />

from suspension, with similar grain size distribution. The sedimentary structures of the 2004 tsunami, where present, are already<br />

obliterated six years after the event, making it difficult to interpret the flow direction and to differentiate between deposits from<br />

inflow and outflow of tsunami. This will affect accuracy of future work in estimating flow characteristics of past tsunamis based on<br />

the depositional record in the vegetated area such as Thailand.<br />

Key words: Paleotsunami, tsunami deposits, Indian Ocean, Thailand<br />

INTRODUCTION<br />

Phra Thong Island is located about 580 km<br />

southwest of Bangkok and about 120 km north of<br />

Phuket (Fig.1). It is about 15 km in length and 9 km in<br />

width, with the long axis in an approximate N-S<br />

direction and the western side directly facing the<br />

Andaman Sea. The island composes of beach-ridge<br />

plain on the western side and the mangrove-fringed<br />

tidal inlet to the east. The 2004 tsunami severely<br />

struck the island and completely wiped out one of the<br />

only three villages on the island. Paleotsunami<br />

deposits on Phra Thong Island were reported by<br />

Jankaew et al. (2008) and Fujino et al. (2009).<br />

o<br />

9 00’ N<br />

8<br />

o<br />

00’ N<br />

o<br />

98 00’ E<br />

20 km<br />

ANDAMAN<br />

SEA<br />

PT<br />

o<br />

98 30’ E<br />

PHANG NGA<br />

PHUKET<br />

o<br />

99 00’ E<br />

Fig. 1: Map showing location of Phra Thong Island<br />

(PT).<br />

Apart from the frequency, a magnitude of large<br />

tsunamis is important information needed to be<br />

established in certain coastal area in order have<br />

better mitigation measure in place for possible future<br />

catastrophic events. The magnitudes of tsunamis<br />

responsible for the sand layers at Phra Thong, can<br />

be assessed using numerical models with help from<br />

tsunami deposits to validate the model results.<br />

Because of its relatively flat terrain, Phra Thong<br />

offers a suitable site to study velocity and tsunami<br />

flow depth from characteristics of the deposits.<br />

Thickness and grain size distribution of tsunami<br />

deposits can be used in sediment transport models<br />

(Jaffe & Gelfenbaum, 2002; Jaffe & Gelfenbaum,<br />

2007) to derive estimates of flow velocity and water<br />

depth (Peters et al., 2007), which are important in<br />

building design and in planning evacuation route.<br />

We present distribution of paleotsunami deposits on<br />

Phra Thong Island and sedimentary characteristics of<br />

2004 and paleotsunamis.<br />

METHOD<br />

We dug shallow pits and trenches to observe<br />

sedimentary characteristics of tsunami sediments<br />

(both 2004 tsunami and paleotsunamis) and to<br />

collect samples for grain size analyses. In order to<br />

better observe and to be able to make a meaningful<br />

conclusion of the deposit thickness and grain size<br />

variation, observed in the 2004 and paleotsunami<br />

deposits, we collected samples along two transects<br />

parallel to the tsunami flow (Fig. 2). Sediment<br />

samples were collected to the end of the 2004<br />

99


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

tsunami sedimentation limit. Total length of the two<br />

transects are 1.5 km and 2 km, respectively. Along<br />

these transects wherever the paleotsunami sand<br />

layer is present, we also collected samples for grain<br />

size analyses. The samples were collected at 0.5 cm<br />

interval, and analysed using settling tube, measuring<br />

size range from 0-5 with 0.1 intervals.<br />

landward location and sediments remaining in<br />

suspension. The 2004 tsunami deposits contain up to<br />

o<br />

98 15’ E<br />

o<br />

98 21’ E<br />

9<br />

o<br />

09’ N<br />

Transect 1<br />

Transect 2<br />

Fig. 4: Thick 2004 tsunami deposit (almost 30 cm<br />

thick) in a swale. Sedimentary structures indicating<br />

flow direction are clearly visible in this picture. The<br />

sea is to the right of the picture.<br />

o<br />

9 03’ N<br />

RESULT<br />

5KM<br />

Fig. 2: Sediment samples were collected from two<br />

transect from the northern part of the island.<br />

The thickness of deposits of the 2004 tsunami and its<br />

predecessors varies across the beach-ridge plain<br />

and is primarily controlled by local topography.<br />

Thickness tends to be greatest in a low-lying swale<br />

(as much as 30 cm). On ridges and higher ground it<br />

tends to be thinner (few cm, 10 cm at most). Fig. 3<br />

shows a variation in thickness of the 2004 tsunami<br />

deposit along transect 2 (in Fig. 2). A thicker, and<br />

sometimes coarser, deposit in topographic lows is a<br />

response of the flow to the deeper water depth,<br />

which slows the flow and drive deposition (Apotsos et<br />

al., 2009). In the case of Phra Thong, if the swale is<br />

located on a path of back flow, which is often<br />

channelized, thickness is much higher. Back flow<br />

deposits contain both sediments eroded from<br />

4 fining-upward sequences. Fig. 4 is an example of<br />

2004 tsunami deposit in the swale located on a path<br />

of the back flow. In Fig. 4, the lowest fining-upward<br />

sequence was probably deposited out of the first<br />

tsunami wave. Above this sequence, a continuous<br />

dark gray layer of fine silt-clay which was probably<br />

deposited when the tsunami flow was still and<br />

between the two main waves. Above this silty clay<br />

layer there are 2-3 fining upward sequences, which<br />

were possibly deposited from back flow of the first or<br />

second wave. Sets of cross bedding in these upper<br />

sequences indicate a flow direction towards the sea.<br />

Revisiting the sites allow us to observe that dense<br />

vegetation in the tropics, especially in the low lying<br />

fresh water swales, already destroy most of these<br />

cross bedding sets. This will complicate the<br />

interpretation of flow depths and velocities based on<br />

the deposit thickness.<br />

On high ground of Phra Thong Island (about 4 m<br />

above msl.), with less vegetation cover, the 2004<br />

tsunami deposit lies above the sand ridge sediments<br />

with faint soil in between them. It is still possible to<br />

identify the 2004 tsunami deposits in the high ground<br />

6 years after the event, but allowing more time<br />

differentiating it from the sand ridge below will be<br />

problematic. At many locations, distinguishing the<br />

2004 tsunami laid sand and the underlying ridge<br />

sand was already difficult due to the lack of organic<br />

soil formation on top of the ridge sand. On the high<br />

ground immediately next to the swale, after the<br />

tsunami dumped a majority of sediments in the<br />

swale, the 2004 tsunami deposit composes mainly of<br />

silt which fell out of suspension (Fig. 5). The tsunami<br />

flow then picked up more sediment from the ridge<br />

area along its flow path. As a result, grain size of<br />

2004 tsunami on high ground is composed mainly of<br />

silt in contrast to sand and silt in the swale deposit.<br />

The thicknesses of the 2004 deposits at different<br />

locations are also shown.<br />

Fig. 3: Thicknesses of 2004 tsunami deposit along<br />

transect 2 (blue vertical line: VE x10).<br />

100


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

appear as a massive bed or faintly normally graded,<br />

more so than the 2004 deposits, but mean grain size<br />

at 0.5 cm thickness intervals show that they contain<br />

up to two fining-upward sequences<br />

DISCUSSION<br />

Fig. 5: Thin 2004 tsunami deposit (about 5 cm thick)<br />

on high ground. The deposit contains mainly silt and<br />

very fine sand with no sedimentary structure. The sea<br />

is to the left of the picture.<br />

Apart from variation in the thickness of the deposit,<br />

internal characteristics of 2004 deposit on Phra<br />

Thong also varies greatly from one place to another.<br />

Often the deposits show no internal structure or<br />

layering. Typically the deposits appear as a massive<br />

bed or faintly normally graded. In locations close to<br />

the sea, some deposits are coarsened upward.<br />

The 2004 tsunami sediments range in grain size from<br />

coarse sand to coarse silt, with distinctive bimodal<br />

distribution with the first mode in fine-sized range (2<br />

), and the second and highest mode in very finesize<br />

range (3.3 ). The coarser sediments are<br />

composed of coarse sands, derived from beach berm<br />

and possibly from offshore areas, and broken shell<br />

fragments, large foraminiferas and other<br />

macrofossils, whereas the finer sediments were<br />

possibly derived from the subtidal zone. The 2004<br />

sediment composes about 85% of clear to white<br />

quartz grains, 8% of shell fragments and microfossils,<br />

4% of muscovite and 3% of heavy minerals - mostly<br />

small grains of tin. Although typically 2004 tsunami<br />

deposits at Phra Thong appear as a massive bed or<br />

faintly normally graded, detailed grain size analyses<br />

show that they can contain 2-3 fining-upward<br />

sequences. The mean grain size of the 2004 tsunami<br />

is slightly bigger than that of the paleotsunami sand.<br />

The 2004 tsunami layer contains 2-3 fining-upward<br />

sequences while paleotsunami deposits contain 2<br />

fining-upward sequences.<br />

Paleotsunami deposits at Phra Thong Island are<br />

composed of grains ranging in size from coarse sand<br />

to coarse silt, with bimodal distribution. A small first<br />

mode in fine-sized range (2.3 phi), and second and<br />

the highest mode in very fine-size range (3.2 phi).<br />

Grains compose of about 91% of clear to white<br />

quartz grains, 7% of muscovite and 2% of heavy<br />

minerals - mostly small grains of tin. The shell<br />

fragments and microfossils are absent in the paleosand<br />

layer. Phra Thong paleotsunami deposits<br />

Deposition of tsunami sediment, although generally<br />

has a sheet-like geometry, is greatly controlled by the<br />

local topography with thicker deposit in the low-lying<br />

area. The thickness of tsunami deposit is further<br />

altered by post-depositional processes such as<br />

human and animal disturbances, erosion and<br />

bioturbation. However, the thick deposits in the lowlying<br />

area have a greater tendency to be preserved<br />

and remained as a geological record. Dense<br />

vegetation in a study area quickly destroy most of the<br />

sedimentary structures contain in the deposit which<br />

may complicate differentiation between inflow and<br />

outflow deposits, leading to over or under estimation<br />

of the thicknesses of tsunami deposited by each<br />

wave. Sedimentation limit of the paleo-event is also<br />

difficult to define, especially in the case of Phra<br />

Thong Island where deposits are mainly preserved in<br />

the swales and are subjected to bioturbation.<br />

CONCLUSION<br />

Future work in estimating hydrodynamic properties of<br />

past tsunamis from sediment deposits in the tropics<br />

should not be dependent on thickness of the<br />

deposits.<br />

Acknowledgements: This work was supported by Japan<br />

Society for the Promotion of Science (to YS and KJ), DFG<br />

grant numbers BR 877/27 and KE 190/26 (to DB and KJ)<br />

and NRCT grant number GE3/2554 (to KJ).<br />

References<br />

Apotsos, A.; B. Jaffe, G. Gelfenbaum, and E. Elias, (2009).<br />

Modeling time-varying tsunami sediment deposition.<br />

<strong>Proceedings</strong> of Coastal Dynamics 2009, 1-15.<br />

Fujino, S., H. Naruse, D. Matsumoto, T. Jarupongsakul, A.<br />

Sphawajruksakul, and N. Sakakura, (2009). Stratigraphic<br />

evidence for pre-2004 tsunamis in southwestern<br />

Thailand. Marine Geology 262, 25-28.<br />

Jaffe, B. E. and G. Gelfenbaum, (2002). Using tsunami<br />

deposits to improve assessment of tsunami risk.<br />

Solutions to Coastal Disasters’02, Conference<br />

<strong>Proceedings</strong>, ASCE, 836-847.<br />

Jaffe, B. E. and G. Gelfenbaum, (2007). A simple model for<br />

calculating tsunami flow speed from tsunami deposits,<br />

Sedimentary Geology 200, 347-361.<br />

Jankaew, K., B.F. Atwater, Y. Sawai, M. Choowong, T.<br />

Charoentitirat, M.E. Martin, and A. Prendergast, (2008).<br />

Medieval forewarning of the 2004 Indian Ocean tsunami<br />

in Thailand. Nature 455, 1228–1231.<br />

Peters, R.; Jaffe, B. & Gelfenbaum, G. (2007). Distribution<br />

and sedimentary characteristics of tsunami deposits<br />

along the Cascadia margin of western North America,<br />

Sedimentary Geology 200, 372-386.<br />

101


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ARCHAEOSEISMOLOGY OF THE AD 1545 EARTHQUAKE<br />

IN CHIANG MAI, NORTHERN THAILAND<br />

Kázmér, M. (1, Kamol Sanittham (2), Punya Charusiri (3), Santi Pailoplee (3)<br />

(1) Department of Palaeontology, Eötvös University, Pázmány sétány 1/c, Hungary. Email: mkazmer@gmail.com<br />

(2). Department of Mathematics and Statistics, Chiangmai Rajabhat University, Chiang Mai, Thailand<br />

(3) Department of Geology, Chulalongkorn University, Bangkok, Thailand<br />

Abstract (Archaeoseismology of the A.D. 1545 earthquake in Chiang Mai, northern Thailand): The A.D. 1545 Chiang Mai<br />

earthquake in northern Thailand was studied by historical and archaeological sources.The temple Wat Chedi Luang has lost about<br />

half of the original 80-metres height due to southward-directed collapse. Twenty-one temple sites – out of 74 visited – has tilted<br />

pagodas, up to 5° in various directions, dominated by a SE trend. All damaged temples were built before the 1545 earthquake. We<br />

suggest that a city-wide liquefaction event caused tilting. The responsible earthquake possibly occurred along the Doi Suthep<br />

Fault within city limits. Possible activity of distant faults is assessed.<br />

Key words: palaeoseismology, Thailand, liquefaction, Wat Chedi Luang<br />

INTRODUCTION<br />

An important obstacle to the assessment of<br />

earthquake hazard at present is the lack of<br />

information about old earthquakes (Ambraseys,<br />

2009: xii). The locations of larger historical<br />

earthquakes have been found to be known well<br />

enough to guide field studies for further in situ<br />

investigations. Properly run field studies provide<br />

reliable observations for the assessment of damage,<br />

intensity, and its distribution, ground effects and<br />

surface faulting. Field studies of old earthquakes are<br />

time-consuming and often present subtle problems<br />

but they are essential (Ambraseys 2009: 16). Here<br />

we provide a brief description of traces of a<br />

significant earthquake in Northern Thailand, and<br />

provide assessment of seismic parameters of the<br />

event.<br />

Fig. 2: Archaeological reconstruction of the pre-earthquake<br />

dimensions of Wat Chedi Luang as seen from the east.<br />

Total height was approx. 80 m. The portion above the heavy<br />

line is the art historian’s vision about its looks.<br />

Fig. 1: Wat Chedi Luang in Chiang Mai, Thailand, seen from<br />

the southeast. Damaged during the AD 1546 earthquake,<br />

the upper half of the stupa fell to the south.<br />

Both historical documents and archaeological data<br />

are available describing the A.D. 1545 earthquake in<br />

Northern Thailand. We studied the Buddhist temples<br />

in and around the old city of Chiang Mai (Kázmér &<br />

102


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Sanittham, 2011) to identify possible earthquakeinduced<br />

damages preserved in the buildings’<br />

structure and orientation.<br />

Currenty earthquake activity in northern Thailand is<br />

interpreted within the framework of Thoen Fault<br />

(Chiang Saen, May 2007, M L = 6.3), Mae Tha and<br />

Pha Youv fault zones tens of kilometres away<br />

(Pailoplee et al., 2010). Since recurrence time of<br />

major earthquakes seems to be longer than the<br />

instrumental period of 50 years, archaeoseismology<br />

is a necessary tool to extend the observation period<br />

to centuries.<br />

high was built in 1479-1481. The base was enlarged<br />

and strengthened in 1512 (Podjarawaraporn, 2547).<br />

On 28 July 1545 there was a huge rainstorm and an<br />

earthquake, which caused the chedi to topple,<br />

leaving only half if its structure to stand (Fig. 2). The<br />

power and richness of the Medieval Lanna Empire<br />

already in the decline, no funds have ever been<br />

available to restore the damaged building to its<br />

former glory. The chedi was left in this damaged<br />

condition for more than four centuries. A cosmetic<br />

restoration in 1992 completed the strengthening by a<br />

60 x 60 m base, .<br />

METHODS<br />

Historical, archaeological, and geologicalgeophysical<br />

data are combined to understand the<br />

Chiang Mai earthquake of AD 1545. The published<br />

historical description was cross-checked with<br />

archaeological data of the site of Wat Chedi Luang<br />

and elsewhere. We visited 74 temples of Chiang Mai<br />

city. While recording earthquake archaeological<br />

effects (Rodriguez-Pascua et al. 2011), we measured<br />

the angle and direction of tilt of the chedi (stupa) by a<br />

stonemason’s tiltmeter and a compass, respectively.<br />

Coordinates of chedi location were taken from the<br />

digital map of Northern Thailland (ThinkNet 2010).<br />

Construction ages were drawn from Thai-language<br />

publications. When no printed source was available,<br />

we accepted the dating of tourist information tablets<br />

in the monasteries.<br />

TILTED BUILDINGS CITYWIDE<br />

In addition to the famous damaged chedi, numerous<br />

religious and secular monuments in and around the<br />

old city bear evidence for some kind of earthquake<br />

damage. The most obvious evidence is tilting of<br />

chedis: the pointed top part of the monument clearly<br />

deviates from the vertical by a few degrees (Fig. 3).<br />

(The lightweight metal decoration at the very top is<br />

almost always heavily tilted; we did not take these<br />

into account, only the brick portion below.) Historical<br />

data on construction time of the chedis indicate that<br />

all of them were built in the 14-15 th century AD,<br />

before the A.D. 1545 earthquake (Fig. 4). Locations<br />

and tilt directions are mapped on Fig. 5.<br />

Tilt directions are dominated by a conspicuous SE<br />

trend (Fig. 6).<br />

There is no official English transliteration system for<br />

the Thai language. English spelling of Thai names is<br />

inconsistent to the extent that one’s own name is<br />

written differently on subsequent occasions. In this<br />

paper we use names as found on the electronic map<br />

of ThinkNet (2010), which is neither official, nor better<br />

than any other spelling.<br />

HISTORICAL DATA<br />

There was a damaging earthquake in Chiang Mai city<br />

(Northern Thailand) on 28 July 1545 in the afternoon<br />

hours between 4.30 and 6.00 pm. „The earth<br />

trembled and shook, groaned and moaned, very<br />

intensely. The finials, (top parts) yòt, of the Jedi<br />

Luang and of the jedi in Wat Phra Sing broke off and<br />

fell down, and also the finials of many other jedis”,<br />

recorded the contemporary Chiang Mai Chronicle in<br />

Lanna language (translated by Penth, 2006).<br />

WAT CHEDI LUANG IN CHIANG MAI<br />

The largest chedi (stupa, pagoda) ever built in what<br />

is Thailand today is the Wat Chedi Luang, standing in<br />

the monastery of the same name in the centre of old<br />

Chiang Mai city (Fig. 1). Built in 1391, it has been<br />

reconstructed and enlarged several times, A huge<br />

chedi, 56 x 56 m rectangular basement, approx. 80 m<br />

Fig. 3: Tilted buildings citywide<br />

103


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The Doi Suthep fault, the master fault of the halfgraben<br />

of the Chiang Mai Basin is 3 km away. It is<br />

not known to be active: a minimum M 5.5 seismic<br />

event here could cause liquefaction.<br />

Fig. 4: All tilted chedis (stupas) were built before the AD<br />

1545 earthquake. 4 – Wat Chiang Yeun, 8 – Wat Hua<br />

Khuang, 12 – Wat Lok Mo Li, 17 – Wat Phra SIngh, 23 –<br />

Wat Chai Prakiat, 28 – Wat Phuak Hong, 34 – Wat Chet<br />

Rim, 42 – Wat Umong, 46 – Wat Chiang Man, 50 – Wat<br />

Srisupan, 51 – Wat Nantharam, 56 – Wat Daowadueng (no<br />

tilting was observed), 67 – Wat Bupharam, 72 – Wat<br />

Chomphu.<br />

The SW-NE trending, left-lateral Mae Kuang Fault 32<br />

km to the NE is possibly inactive since the Tertiary,<br />

althought the fault trace is particularly conspicuous in<br />

the landscape (Rhodes et al. 2004). A minimum M<br />

6.3 seismic even would have been sufficient to cause<br />

liquefaction in Chiang Mai.<br />

SUBSOIL<br />

The tilted chedis are all on the alluvial plain of the<br />

Ping River, extending over at least 4 km 2 .<br />

Groundwater lever was high during our survey in<br />

August 2010, about 70 to 100 cm below ground, as<br />

seen in several wells within the temple compounds.<br />

Historical data indicate a rainy summer season for<br />

1545, too.<br />

We suggest that a city-wide liquefaction event,<br />

caused uneven settlement and subsidence of the<br />

buildings in the saturated soil. The dominant SE-ward<br />

tilt direction possibly reflects strong motion<br />

directionality.<br />

INTENSITY<br />

While modified Mercalli intensity VII is the damage<br />

threshold for many archaeological sites (Kovach and<br />

Nur, 2006), we assume that damages to Wat Chedi<br />

Luang related to the 1545 earthquake require a<br />

larger intensity due to the especially compact<br />

construction of the building. The pagoda, built like a<br />

pyramid, is certainly a more earthquake-resistant<br />

structure than any ordinary city house, even palace.<br />

Intensity IX or higher (good masonry damaged<br />

seriously, in areas of loose sediment, sand, mud, and<br />

water ejected – Rapp, 1986) seems more probable.<br />

Fig. 5: Tilted chedis (Fig. 4) (black dots) on the alluvial plain<br />

of Ping River (Margane & Tatong, 1999). Ticks towards<br />

direction of tilting. Untilted chedis are marked with empty<br />

circles. Rectangle indicates walled city of old Chiang Mai.<br />

The Lampang-Thoen fault zone 120 km to the SE is<br />

active (Chiang Saen, May 2007, M L = 6.3). The<br />

segments are long enough to produce M 7<br />

earthquakes (Pailoplee et al., 2009). A minimum M 7<br />

seismic event is needed to cause liquefaction in<br />

Chiang Mai city.<br />

Intensity VIII to IX (heavily damaging to destructive)<br />

is assumed on the ESI 2007 environmental intensity<br />

scale (Michetti et al., 2007): liquefaction with<br />

settlement up to 30 cm or more. The total affected<br />

area was in the order of 1000 to 5000 km 2<br />

(Reicherter et al., 2009), i.e. all of the Chiang Mai-<br />

Lamphun Basin.<br />

EPICENTER AND MAGNITUDE<br />

Known and possibly active faults were assessed for<br />

source of the earthquake, and the minimum<br />

magnitude for liquefaction calculated after Obermeier<br />

(1996, Fig. 42).<br />

Fig. 6:Tilt directions plotted in polar bar chart. Horizontal<br />

axis – number of pagodas tilted in a certain direction. Note<br />

the prominent southeastward tilting of several chedis.<br />

104


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The Sagaing Fault in Myanmar, forming the boundary<br />

between the Sunda and Burma plates is 200 km to<br />

the W. It regularly produces M > 7 earthquakes (M<br />

7.0-7.4) (Hurukawa & Maung, 2011). However, the M<br />

7.5 earthquake on December 3, 1930, did not cause<br />

any liquefaction event in Chiang Mai we are aware<br />

of. A lack of proper attenuation model for Thailand<br />

(Chintanapakdee et al., 2008) prevents formulating a<br />

suitable explanation. There is local model developed<br />

for Chiang Mai (Kannika & Takada, 2009), although<br />

for rock sites, not for alluvium. We suggest that<br />

earthquake intensities display a strong directionality<br />

along the right-lateral Sagaing Fault: higher<br />

intensities occurring parallel and lower intensities<br />

perpendicular to the fault, thus protecting Chiang Mai<br />

from major plate-boundary events.<br />

Whether any of the above or another fault is<br />

resonsible for the AD 1545 earthquake is an open<br />

question as yet. Studies on strong motion direction<br />

causing the major damages (see, for example.<br />

Korjenkov & Mazor, 1999, 2003; Kázmér & Major,<br />

2010; Hinzen, 2008, 2009) may help tor resolve<br />

some of the open questions.<br />

Acknowledgements: Our thanks are due to the staff of<br />

Mahamakat Buddhist University, Lanna Campus library for<br />

their help in accessing relevant literature in the Thai<br />

language. Financial help from Hungarian National Science<br />

Foundation (OTKA) grant K 67.583 and and an IGCP 567<br />

travel grant are sincerely acknowledged here. Raúl Pérez<br />

López is thanked for his comments, which improved this<br />

manuscript.<br />

References<br />

Publication dates of Thai-published books in Thai only are<br />

given as of the Thai calendar. Deduce 543 for conversion to<br />

western, Gregorian years.<br />

Ambraseys, N.N. (2009). Earthquakes in the Mediterranean<br />

and Middle East. A Multidisciplinary Study of Seismicity<br />

up to 1900. Cambridge, Cambridge University Press. 968<br />

p.<br />

Chintanapakdee, C., Naguit, M.E. & Charoenyuth, M.<br />

(2008). Suitable attenuation model for Thailand. – 14th<br />

World Conference on Earthquake Engineering:<br />

Innovation Practice Safety, Beijing, 2008. [8 p.]<br />

Download: http://www.14wcee.org/<strong>Proceedings</strong>/files/02-<br />

0088.PDF (accessed 25 June 2011)<br />

Hinzen, K.-G. (2008). Can ruins indicate a back azimuth?:<br />

Seismological Research Letters, 79 (2), 290.<br />

Hinzen, K.-G. (2009). Simulation of toppling columns in<br />

archaeoseismology. Bulletin of the Seismological Society<br />

of America 99 (5), 2855–2875.<br />

Hurukawa, N. & Maung, P.M. (2011). Two seismic gaps on<br />

the Sagaing Fault, Myanmar, derived from relocation of<br />

historical earthquakes since 1918. – Geophysical<br />

Research Letters 38, L01310, 5 PP.,<br />

doi:10.1029/2010GL046099<br />

Kannika, P. & Takada, T. (2009). Updating framework for<br />

site-specific attenuation relation of seismic ground motion<br />

in Thailand. MSc thesis, Building Research Institute,<br />

Tsukuba, Japan, 70 p.<br />

Kázmér, M., & Major, B. (2010). Distinguishing damages of<br />

two earthquakes – archeoseismology of a Crusader<br />

castle (Al-Marqab citadel, Syria). In: Ancient<br />

Earthquakes. (Stewart, I., Sintubin, M., Niemi, T., Altunel,<br />

E. eds). Geological Society of America Special Paper,<br />

471, 186 –199.<br />

Kázmér, M. & Sanittham, K. (2011). Archeoseismology of<br />

the A.D. 1545 earthquake in Chiang Mai, northern<br />

Thailand. In: International Symposium on the 2001 Bhuj<br />

Earthquake and Advances in Earthquake Science, AES<br />

2011. 22-24 January 2011, Institute of Seismological<br />

Research, Raisan, Gandhinagar, India. Abstract Volume<br />

AES 2011, 117-118.<br />

Korjenkov, A.M. & Mazor, E. (1999). Seismogenic origin of<br />

the ancient Avdat Ruins, Negev Desert, Israel. Natural<br />

Hazards 18, 193–226.<br />

Korjenkov, A.M. c Mazor, E. (2003). Archaeoseismology in<br />

Mamshit (southern Israel): Cracking a millennia-old code<br />

of earthquakes preserved in ancient ruins:<br />

Archäologischer Anzeiger, 2003 (2), 51–82.<br />

Kovach, R.L & Nur, A. (2006). Earthquakes and archeology:<br />

Neocatastrophism or science?: Eos, Transactions,<br />

American Geophysical Union 87 (32), 317.<br />

Margane, A. & Tatong, T. (1999): Aspects of the<br />

hydrogeology of the Chiang Mai-Lamphun Basin,<br />

Thailand that are important for the groundwater<br />

management. – Zeitschrift für angewandte Geologie 45,<br />

188-197.<br />

Michetti, A.M., Esposito, E. et al. (2007). Environmental<br />

Seismic Intensity Scale 2007 – ESI 2007. In: Vittori, E:,<br />

Guerreri, L. eds. Memorie Descrittive della Carta<br />

Geologica d’Italia. LXXIV. Servizio Geologico d’Italia,<br />

Roma, 7–54.<br />

Obermeier, S.F. (1996): Use of liquefaction-induced<br />

features for paleoseismic analysis. – Engineering<br />

Geology 44, 1-76.<br />

Pailoplee, S., Takashima, I., Kosuwan, S. & Charusiri, P.<br />

(2009). Earthquake activities along the Lampang-Theon<br />

fault zone, northern Thailand: evidence from<br />

paleoseismological and seismicity data. – Journal of<br />

Applied Science Research 5 (2), 168-180.<br />

Pailoplee, S., Sugiyama, Y. & Charusiri, P. (2010).<br />

Probabilistic seismic hazard analysis in Thailand and<br />

adjacent areas by using regional seismic source zones. –<br />

Terestrial, Atmospheric and Ocean Sciences 21, 757–<br />

766.<br />

Penth, H. (2006). Earthquakes in Old Lan Na: part of natural<br />

catastrophes. Chiang Mai University CMU Journal 5 (2),<br />

255-265.<br />

Podjarawaraporn (2547). Pra wad Wat Chedi Luang wor ra<br />

wi harn. [History of Wat Chedi Luang.] Text by Pra Bhuda<br />

Podjanawaraporn (Chan Kusato), pictures by Pra Kro<br />

Soponkaweewat (Thanachan Suramanee). Published by<br />

Wat Chedi Luang, Chiang Mai, Thailand, 160 p. ISBN<br />

974-92159-8-2 (In Thai)<br />

Rapp, G., Jr. (1986). Assessing archaeological evidence for<br />

seismic catastrophes. Geoarchaeology 1, 365–379.<br />

Reicherter, K., Michetti, A.M. & Silva Barroso P.G. (2009).<br />

Palaeoseismology: historical and prehistorical records of<br />

earthquake ground effects for seismic hazard<br />

assessment. Geological Society, London, Special<br />

Publication 316, 1-10.<br />

Rhodes, B.P., Perez, R., Lamjuan, A. & S. Kosuwan, S.<br />

(2004). Kinematics and tectonic implications of the Mae<br />

Kuang Fault, northern Thailand. Journal of Asian Earth<br />

Sciences 24 (1): 79-89.<br />

Rodríguez-Pascua M.A., R. Pérez-López, J.L. Giner-<br />

Robles, P.G. Silva, V.H. Garduño-Monroy and K.<br />

Reicherter (2011). A Comprehensive Classification of<br />

Earthquake Archaeological Effects (EAE) in<br />

Archaeoseismology: application to ancient remains of<br />

Roman and Mexican cultures. Quaternary International.<br />

DOI: 10.1016/j.quaint.2011.04.044.<br />

ThinkNet (2010). Map of 14 northern provinces of Thailand<br />

+ CD. Bilingual Mapping Software. ThinkNet Co., Ltd,<br />

Bangkok, Thailand. (In Thai and English)<br />

105


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

OUTLINE OF THE 3.11 TOHOKU EARTHQUAKE IN JAPAN<br />

Yoshihiro Kinugasa (1<br />

(1) Association for the Development of Earthquake Prediction, Tokyo, Japan. Email: king@8f.adep.or.jp<br />

Abstract (Outline of the 3.11 Tohoku earthquake in Japan): Author will discuss the following subjects on the 3.11 Tohoku<br />

Earthquake (Mw=9.0), based on the published data by the time of the Corinth2011 workshop.<br />

1. Seismological Characteristics of the Earthquake; Location, Magnitudes, Focal Mechanism, Intensity Distribution, etc.<br />

2. Tsunami<br />

3. Ground Deformations<br />

4. Re-activation of Normal Fault<br />

5. Paleoseismology along the Tohoku Coast<br />

6. Impacts on the Fukushima NPPs.<br />

Key words: Tohoku Earthquake, Tsunami, Paleoseismology, Fukushima NPPs<br />

106


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE EVIDENCE OF TSUNAMIGENIC DEPOSITS IN THE GULF OF CORINTH (GREECE)<br />

WITH GEOPHYSICAL METHODS FOR SPATIAL DISTRIBUTION<br />

Koster, Benjamin (1), Klaus Reicherter (1), Andreas Vött (2), Christoph Grützner (1)<br />

(1) Neotectonics and Natural Hazards Group, RWTH Aachen University, Lochnerstr. 4 - 20, 52056 Aachen, Germany. Email:<br />

b.koster@nug.rwth-aachen.de<br />

(2) Natural Hazard Research and Geoarchaeology, Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong> <strong>Mainz</strong>, Johann-<br />

Joachim-Becher-Weg 21, 55099 <strong>Mainz</strong>, Germany<br />

The evidence of tsunamigenic deposits in the Gulf of Corinth (Greece) with geophysical methods for spatial distribution:<br />

Drill core sampling in coastal areas in the Mediterranean proved evidence for tsunamis. Sedimentary analyses were conducted to<br />

identify tsunamigenic deposits, but did not reveal larger scale sedimentary structures or spatial distribution of tsunamites in a<br />

regional scale. We used ground penetrating radar (GPR) in combination with electrical resistivity tomography (ERT)<br />

measurements and sedimentological research methods in different areas. The combination of these three methods allows us to<br />

generate 3D visualizations, which give clues for tsunamite distribution and sediment architecture. GPR data indicate<br />

unconformable thicknesses of tsunamigenic layers, channel-like structures of backwash deposits, in some extent non-planar<br />

erosion basement, as well as abrasion-scours in various places, and boulder accumulation inside the deposits.<br />

Key words: tsunami, GPR, ERT, Greece<br />

INTRODUCTION<br />

Former studies of various authors on tsunamis<br />

mainly focused on a hydromechanical analysis of<br />

specific tsunami events (e.g., Bondevik et al., 2005)<br />

or sedimentary analyses of drill cores (e.g.,<br />

Reicherter et al., 2010; Shiki et al., 2008; Vött et al.,<br />

2009). The latter method encompasses sieve curves<br />

as well as magnetic susceptibility measurements and<br />

micropaleontology to prove tsunamigenic features.<br />

Characteristics of the sediments, such as finingupward<br />

sequences, coarse shell debris, upward<br />

rising magnetic susceptibility and marine foraminifera<br />

in sandy sediments, give amongst others evidence<br />

for tsunami events. X-ray fluorescence spectroscopy<br />

measurements (XRF) were performed in some<br />

cases. OSL and 14 C-dating can be used for dating.<br />

All of these studies defined characteristics of the<br />

deposits (e.g., Bryant et al., 2005; Shiki et al., 2008),<br />

but do not show the spatial distribution of an event or<br />

the larger scale sediment structures of tsunami<br />

deposits. With the knowledge of spatial distribution<br />

and extent of erosion due to tsunamis it would be<br />

easier to understand processes during a tsunami<br />

event and to estimate the possible damage by a<br />

future tsunami. Since drilling is time-intensive and<br />

expensive (depending on extend), this method can<br />

by far not cover an entire coastal area. As the<br />

distribution and preservation of tsunamigenic<br />

deposits is highly variable according to several<br />

studies (e.g., Dawson & Stewart, 2007), there is a<br />

strong interest in a low-cost and easy to use imaging<br />

technique.<br />

Only one published study dealt with GPR for<br />

detecting tsunami deposits. Switzer et al. (2006)<br />

investigated a wash-over fan, but did not validate the<br />

data by other geophysical methods. Therefore, it is<br />

still not clear whether or not the detected sediments<br />

were deposited by a tsunami. Furthermore, a limiting<br />

factor of GPR measurements is a wet environment. A<br />

shallow ground water table or even sea water<br />

intrusions close to the ocean can reduce data quality<br />

significantly. Relative dielectric permittivity r and the<br />

conductivity of tsunamigenic deposits are<br />

unknown. However, we can show that GPR has the<br />

ability to distinguish between tsunami deposits made<br />

up of marine sands, boulders and shells and clayey<br />

background sediments although this is an ambitious<br />

challenge.<br />

STUDY AREA<br />

Our study area is located near Lechaion, one of the<br />

ancient harbors of Corinth (Fig. 1). It was probably<br />

the most important harbor of this type in antiquity,<br />

and one of the most important harbors in Greece for<br />

more than one millennium (Rothaus, 1995). Today it<br />

is partially buried by up to 2 meters of sediment.<br />

Fig. 1: Study area in Greece, Lechaion close to Corinth,<br />

brown areas illustrate topographic elevation; red box<br />

indicates area of GPR measurements (see Fig. 2 for<br />

details); green arrow displays possible tsunami propagation<br />

Lechaion may very well have been affected by a<br />

series of seismic events and possible tsunami in late<br />

fourth century after Christ. Reconstruction of the<br />

107


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

harbor in AD 353 - 358 (Stiros et al., 1996)<br />

contingently supports this idea.<br />

We collect GPR data in combination with drill cores<br />

and electrical resistivity tomography (ERT) in order to<br />

test our method in an extraordinary environment, an<br />

ancient harbor which could have been affected by a<br />

tsunami (Soloviev, 1990).<br />

400 MHz antenna promised the best compromise.<br />

Data processing included static correction,<br />

background removal, gain adjustment and velocity<br />

adaption for depth calculations based on a hyperbola<br />

analysis. Boulders were detected due to hyperbolic<br />

features in the data. Results of sedimentary drill<br />

cores (Fig. 4) and ERT profiles in the study area give<br />

evidence for three tsunami events (Hadler et al., this<br />

volume).<br />

VISUALIZATION & RESULTS<br />

Three GPR profiles were taken parallel to the coast,<br />

one profile was recorded perpendicular to the<br />

shoreline (Fig. 2). Three drill cores (Fig. 4) were<br />

taken between 50 and 150 meters away from GPR<br />

profiles in the ancient inner harbor. All GPR<br />

measurements took place on the top of the ancient<br />

harbor facility, which is buried under a possible third<br />

tsunami event layer. The base and inner structures of<br />

the possible tsunami deposits could be imaged in all<br />

the profiles.<br />

Fig. 2: Map of the study area with locations of drill points<br />

and GPR measurements<br />

METHODS<br />

GPR measurements were performed in patterns<br />

directly adjacent to drilling locations and ERT<br />

profiles. We used the GSSI 400 MHz antenna with a<br />

survey wheel, the SIR-3000 unit, and a handheld<br />

GPS (Fig. 3).<br />

Fig. 3: GPR with 400 MHz antenna, survey wheel, SIR-3000<br />

unit and GPS<br />

Trace increment was set to 0.02 m for detailed<br />

investigation, the range was set to 120 ns TWT and<br />

the sample rate to 512. From drillings and field<br />

observations a target depths up to 3.50 m could be<br />

assumed. The thickness of the assumed tsunami<br />

sediment layer reaches up to 2.00 m, so the<br />

Fig. 4: Correlation of drill cores in the study area of<br />

Lechaion; two possible tsunami layers were detected (red<br />

boxes with red dashed lines for correlation); these layers<br />

include fining-upward sequences as well as erosive bases<br />

108


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 5: GPR profile 276 taken parallel to the coastline at the ancient harbor Lechaion (Greece) in combination with ERT profile<br />

LEC_ERT_3: A) processed GPR data, B) processed GPR and ERT data; C) processed and combined analyzed data; yellow<br />

colored areas show tsunamigenic deposits; the base is illustrated by orange line; dashed orange lines refers to inner structures<br />

Fig. 6: GPR profiles 275, 276 & 277 taken parallel to the coastline at the ancient harbor Lechaion (Greece); direction of profiles is<br />

plotted in the circle on top left (red arrow is north; both black linings show x- and y-direction of ground surface); the yellow plain is<br />

a visualized inner structure of the possible upper tsunamigenic deposit; orange lines on the GPR data underline the reflected<br />

boundaries of layers or inner structures (dashed orange line); big channel structures within the possible tsunamite are clearly<br />

visible with accumulation of boulders in depressions; as well some smaller channels could be construed as abrasion-scours<br />

The combination of ERT and GPR measurements in<br />

the study area suggests that there are bigger channel<br />

structures with erosive bases and boulder deposits<br />

inside these channels (Fig. 5 & 6). They point toward<br />

the ocean and are not part of the buried harbor. The<br />

channels can be part of a flow-system during the<br />

backwash processes after a tsunami<br />

(Dawson & Stewart, 2007). In some cases, channel-<br />

109


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

like structures could also be interpreted as abrasionscours,<br />

which can originate by backwash processes<br />

with high backflow velocities. Some kind of crossbedding<br />

is visible in the GPR data as well (Fig. 5).<br />

The tsunamigenic deposits reaches depths up to<br />

2.00 m. The sedimentary evidence from the drill<br />

cores could not be verified due to high attenuation in<br />

lower depths (>2.50 m). The inner layer-structures of<br />

the tsunamites (maybe due to multiple waves) show<br />

an unconformable thickness (Fig. 5). Boulders in the<br />

sediments appear as hyperbolas (v = 0.12 m/ns).<br />

Boulders with diameters larger than the resolution<br />

limit of the 400 MHz GPR antenna are located inside<br />

the deposits and could be detected by GPR. ERT<br />

profiles show as well electrical resistivity contrasts at<br />

the boundary between the tsunamigenic deposits and<br />

underlying harbor sediments. For the other GPR<br />

profiles the correlation with ERT data has been done<br />

similarly, if ERT profiles were available.<br />

CONCLUSIONS<br />

Due to highly variable sedimentation processes and<br />

materials (gravels, sand or silt/clay to some extend<br />

with boulders) in the Mediterranean and worldwide in<br />

the context of a tsunami event, deposited sediments<br />

differ extremely. Therefore, no specific values for<br />

relative dielectric permittivity r or the conductivity <br />

can be declared for these variable sediments. Drill<br />

cores or outcrops are always necessary to prove<br />

tsunami characteristics and to correlate these results<br />

with the GPR data. Distinctive contrast changes in<br />

the GPR data help proving the spatial distribution of<br />

tsunami deposition interfaces. Only the combination<br />

of the presented methods is the key for conclusions<br />

on detailed spatial distribution of tsunamites.<br />

The main result is the visualization of channelized<br />

structures in the tsunamigenic horizon. The<br />

structures most likely originate from backwash<br />

processes. Our data lead us to conclude that the<br />

topography of an effected area plays an important<br />

role for the expansion of the channels, since we<br />

observed different channel types as well e. g., in<br />

Spain. It is possible to detect the upper and lower<br />

boundary of the tsunamite in some cases, depending<br />

on the grain size of the tsunamigenic material in<br />

contrast to the background sedimentation.<br />

With the 400 MHz GPR antenna it is also possible to<br />

detect bigger structures like abrasion-scours,<br />

unconformable thicknesses of tsunamigenic<br />

beddings, in some extent non-planar erosion<br />

basement and boulder accumulations inside the<br />

deposits. Typical thinning-inland structures (Dawson,<br />

1994) could not be detected in this case within the<br />

GPR profile perpendicular to the coast. A GPR with<br />

higher resolution should be useful to detect further<br />

sedimentary structures in tsunamigenic deposits in<br />

the future.<br />

OUTLOOK<br />

In the future, GPR and other shallow geophysical<br />

methods will be used to detect run-up distances and<br />

for creating large-scale models considering<br />

topography to detect sediment thickness and volume.<br />

With these data it would be possible to calculate the<br />

physical power and the possible damage of the<br />

tsunami wave and typical sediment structures. Due to<br />

account on spatial distribution information it could be<br />

possible the reconstruct the topography of the<br />

landscape before and after a paleo-tsunami.<br />

Another aim is to get detailed information of the<br />

deposits by trenching and using methods like LiDAR,<br />

multispectrometry, magnetic susceptibility and the<br />

documentation commonly used in archeological<br />

excavations.3-dimensional data or block-plots can be<br />

generated based on these methods to evaluate new<br />

features and characteristics of tsunamigenic<br />

deposits.<br />

References<br />

Bondevik, S., F. Løvholt, C. Harbitz, J. Mangerud, A.<br />

Dawson & J.I. Svendsen, (2005). The Storegga Slide<br />

tsunami – comparing field observations with numerical<br />

simulations. Marine and Petroleum Geology 22, 195-208.<br />

Bryant E., R.L. Wiegel, F.P. Shepard & D. Myles, (2005).<br />

Natural hazards. Cambridge University Press, 2nd<br />

edition, 214-225.<br />

Dawson, A.G. & I. Stewart, (2007). Tsunami deposits in the<br />

geological record. Sedimentary Geology (200), 166-183.<br />

Dawson, A. G. (1994). Geomorphological effects of tsunami<br />

run-up and backwash. Geomorphology 10, 83-94.<br />

Hadler, H., A. Vött, B. Koster, M. Mathes-Schmidt, T.<br />

Mattern, K. Ntageretzis, K. Reicherter, D. Sakellariou, T.<br />

Willershäuser, (2011). Lechaion, the ancient harbour of<br />

Corinth (Peloponnese, Greece) destroyed by<br />

tsunamigenic impact. 2nd INQUA-IGCP-567 International<br />

Workshop on Active Tectonics, Earthquake Geology,<br />

Archaeology and Engineering, Corinth, Greece. This<br />

volume.<br />

Reicherter, K., D. Vonberg, B. Koster, T. Fernández-<br />

Steeger, C. Grützner & M. Mathes-Schmidt, (2010). The<br />

sedimentary inventory of the 1755 Lisbon tsunami along<br />

the southern Gulf of Cádiz (southwestern Spain).<br />

Zeitschrift für Geomorphologie N.F., Suppl. Issue 54 (3),<br />

147-173.<br />

Rothaus, R., (1995). Lechaion, western port of Corinth: a<br />

preliminary archaeology and history. Oxford Journal of<br />

Archaeology 14 (3), 293-306.<br />

Shiki, T., T. Tachibana, O. Fujiwara, K. Goto, F. Nanayama<br />

& T. Yamazaki, (2008). Characteristic features of<br />

tsunamiites. In: Tsunamiites – Features and Implications<br />

(T. Shiki, Y. Tsuji, T. Yamakazi & K. Minoura eds).<br />

Elsevier Scientific Publishing Company, 319-336.<br />

Soloviev, S.L., (1990). Tsunamigenic Zones in the<br />

Mediterranean Sea. Natural Hazards (3), 183-202.<br />

Stiros, S., P. Pirazzoli, R. Rothaus, S. Papageorgiou, J.<br />

Laborel & M. Arnold, (1996). On the Date of Construction<br />

of Lechaion, Western Harbor of Ancient Corinth, Greece.<br />

Geoarchaeology 11 (3), 251-263.<br />

Switzer A.D., C.S. Bristow & B.G. Jones, (2006).<br />

Investigation of large-scale washover of a small barrier<br />

system on the southeast Australian coast using ground<br />

penetrating radar. Sedimentary Geology 183, 145-156.<br />

Vött, A., H. Brückner, S.M. May, D. Sakellariou, O. Nelle, F.<br />

Lang, V. Kapsimalis, S. Jahns, R. Herd, M. Handl & I.<br />

Fountoulis, (2009). The Lake Voulkaria (Akarnania, NW<br />

Greece) palaeoenvironmental archive - a sediment trap<br />

for multiple tsunami impact since the mid-Holocene.<br />

Zeitschrift für Geomorphologie N.F., Suppl. Issue 53 (1),<br />

1-3.<br />

110


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

COSEISMIC SURFACE RUPTURING IN THE EPICENTRAL AREA OF GERMANY’S<br />

STRONGEST HISTORICAL EARTHQUAKE<br />

S. Kübler(1), A. M. Friedrich(1), M. R. Strecker(2)<br />

(1) Ludwig-Maximilians-University Munich, Department of Earth and Environmental Sciences, Luisenstr. 37, 80333 Munich,<br />

Germany, Email: kuebler@iaag.geo.uni-muenchen.de<br />

(2) University of Potsdam, Institute of Geosciences, Karl-Liebknecht-Str. 24, 14476 Potsdam, Germany<br />

Abstract (Coseismic surface rupturing in the epicentral area of Germany’s strongest historical earthquake): The Lower<br />

Rhine Embayment is one of the most seismically active regions in intraplate Europe. A trenching investigation carried out south of<br />

the city of Düren near the estimated epicenter of Germany's largest historical event (1756 AD, M L 6.2 ± 0.2) revealed evidence of<br />

significant coseismic deformation of the earth's surface. Deformation is expressed by co-planar sets of fractured and rotated<br />

pebbles as well as liquefaction of fine-sand deposits. Two event horizons indicate an older rupturing event that occurred<br />

presumably in the Early to Mid Holocene and a younger rupturing event that occurred in the Latest Holocene, and may correspond<br />

to the 1756 Düren earthquake.<br />

Key words: paleoseismology, historical earthquake, coseismic rupture, Germany<br />

DISCUSSION<br />

Based on historical documents and instrumental data<br />

the Lower Rhine Embayment is currently one of the<br />

most seismically active regions in intraplate Europe.<br />

At least 21 M > 5 instrumental and historical<br />

earthquake events have been recorded and<br />

documented, respectively, for western Germany,<br />

eastern Belgium and southern Netherlands<br />

(Leydecker, 2002). One of the largest of these events<br />

was the earthquake of February 18 th 1756 near the<br />

city of Düren at the western border of the LRE.<br />

Damage related to this event included triggered<br />

landslides and destroyed buildings and castles, but<br />

the occurrence of surface ruptures has not been<br />

reported for this or any other historic event in this<br />

region.<br />

The LRE is characterized by NW-SE striking normal<br />

faults. Fault plane solutions indicate an extensional<br />

normal faulting regime (Hinzen, 2003). Several<br />

paleoseismic studies in the LRE indicate that strong,<br />

surface-rupturing events may have repeatedly<br />

occurred in this region since the Late Pleistocene<br />

(e.g., Camelbeeck & Meghraoui, 1998; Vanneste et<br />

al., 2001; Vanneste & Verbeeck, 2001), but there has<br />

been (Ahorner, 1996) and continues to be (Houtgast<br />

et al., 2003, 2005) controversy about whether faults<br />

in the LRE mainly move by creep, or whether<br />

earthquakes that are large enough to break the<br />

surface commonly occur in this region. The 1756<br />

Düren earthquake had a local magnitude of 6.2 ± 0.2<br />

based on empirical studies (Meidow, 1994), the total<br />

moment for this event was estimated at 1.6 x 10 17<br />

Nm (M W 6.1; Hinzen & Reamer, 2007). Events of this<br />

magnitude are commonly associated with surface<br />

ruptures, but there is still great uncertainty in the<br />

reliability of historic documents for estimating the<br />

correct magnitude of historic events. Therefore, the<br />

identification of a surface rupture related to the Düren<br />

event would be important to better understand the<br />

mechanical behaviour of faulting in this tectonically<br />

active region.<br />

In contrast to paleoseismic and geomorphic studies<br />

in arid regions, where fault scarps are exposed for<br />

many kilometers along strike and their preservation<br />

potential is excellent, the recognition and<br />

characterization of potentially active faults is much<br />

more difficult in the moderately humid Lower Rhine<br />

area. The dense vegetation cover and intense<br />

agricultural landuse hamper the recognition of<br />

seismogenic surface ruptures in the densely<br />

populated region. Low displacement rates on<br />

individual faults, and hence a sparse earthquake<br />

record due to long recurrence intervals of large<br />

seismic events, additionally aggravate seismic<br />

hazard evaluation. Thus, multidisciplinary<br />

paleoseismic studies are important to identify<br />

seismically active fault segments in the LRE.<br />

In order to search for a potential surface rupture<br />

related to the 1756 Düren event, we carried out<br />

multidisciplinary reconnaissance studies including<br />

geomorphic mapping, shallow drilling and shallow<br />

geophysical prospecting in the vicinity of Düren,<br />

wherever we could find late Holocene fluvial deposits<br />

covering the projection of potentially active faults. At<br />

one location, we identified two gentle 0.5 and 0.6 m<br />

high surface scarp in late Holocene deposits in<br />

prolongation of the E-dipping, 16-km long Schafberg<br />

fault. The trends of the two scarps differ: the south-<br />

111


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

western step strikes 125° and is slightly curved,<br />

whereas the north-eastern step is perfectly straight<br />

and strikes 155°. The study site is situated a few<br />

kilometres W of Untermaubach, where the Schafberg<br />

fault crosses the Rur river valley. Geophysical<br />

prospecting identified two anomalous zones<br />

characterized by resistivity minima, high near-surface<br />

conductivity, and offset radar reflectors that coincide<br />

with the two 0.5 and 0.6 m high surface scarps<br />

(Streich, 2003).<br />

We excavated the fault along an up to 5 m deep and<br />

85 m long trench approximately 200 m west of the<br />

current Rur river course. At the trench site, the<br />

Schafberg fault is covered by < 5 m thick, scarcely to<br />

moderately well layered Holocene sandy gravel<br />

deposits and fine-grained flood deposits overlaying<br />

Lower Devonian shale and sandstone. The trench<br />

exposes two asymmetric channels that coincide with<br />

the position of the observed surface scarps. The<br />

channel fill consists of sandy silt and clay-rich layers.<br />

At the base of both channels, we detected a high<br />

concentration of organic material including trees and<br />

branches. We mapped various types of soft-sediment<br />

deformation in the gravel deposits including tilted and<br />

rotated clasts as well as fractured clasts that show<br />

offsets on a millimeter to centimeter scale. We further<br />

detected liquefaction features in sand and silt<br />

deposits.<br />

Our mapping revealed a narrow zone of localized<br />

deformation below the north-eastern scarp expressed<br />

by abundant fractures with aligned and broken clasts<br />

extending vertically throughout the entire gravel unit.<br />

We carried out detailed gravel analysis including<br />

mapping of 237 fractured clasts and the long-axis<br />

orientation of ~ 10.000 clasts. Results define a ~ 10<br />

m wide fault zone that coincides with the surface<br />

offset and an almost 40 m wide deformation zone<br />

where gravel deformation is still prominent. In<br />

contrast, gravel below the south-western scarp<br />

exhibits no indicators of coseismic deformation and is<br />

therefore most likely a fluvial channel. Liquefaction<br />

occurs 20 m east and 50 m west of the suspected<br />

rupture zone. As the overlaying flood deposits do not<br />

show signs for seismogenic deformation, the contact<br />

between the gravel and overlaying silt and clay<br />

appears to mark the most recent event horizon. An<br />

older event horizon – a paleosoil with high amounts<br />

of organic material – is preserved at a depth of 2.5 m<br />

within the gravel deposits. Here, underlying gravel<br />

deposits show clear deformation features while<br />

gravel above this marker horizon is undisturbed. The<br />

maximum vertical displacement across the fault<br />

appears to range between 0.8 and 1.2 m based on<br />

two offset marker horizons. This is in accordance<br />

with estimates derived from borehole data,<br />

geophysical prospecting and morphometric analysis.<br />

We interpret the observed deformation features,<br />

especially the co-planar sets of rotated and fractured<br />

clasts, as the result of coseismic deformation at the<br />

near-surface end of the rupture, and we rule out slow<br />

deformation due to aseismic creep as governing<br />

process to cause rupturing of pebbles this close to<br />

the surface. Preliminary radiocarbon data bracket the<br />

younger event horizon to Latest Holocene age, which<br />

rules out periglacial processes as cause for the<br />

observed sediment deformation. Age results also<br />

imply that this fault may be a possible source for the<br />

1756 earthquake. Further analyses are in progress.<br />

We identified coseismic deformation at the trench<br />

site, because special conditions produced a number<br />

of features not normally observed in other fault<br />

exposures. The thin sedimentary cover (< 5 m)<br />

above basement rocks and the high water table may<br />

have played an important role in producing this<br />

unusual deformation pattern. However, this newly<br />

investigated trench site yields the first evidence for<br />

Late Holocene seismogenic surface rupturing in the<br />

German part of the LRE, and thus confirms the<br />

importance of paleoseismic studies for seismic<br />

hazard analysis in humid low-strain intraplate<br />

regions.<br />

Acknowledgements: Trench excavation and geological<br />

field work was financed by the DFG, research grant<br />

awarded to Friedrich and Strecker (AF 6513).<br />

References<br />

Ahorner, L., 1996, How reliable are speculations about large<br />

paleo-earthquakes at the western border fault of the Roer<br />

Valley Graben near Bree, In: Bonatz, M., ed., Comptes-<br />

Rendus des 81ièmes Journées Luxembourgeoises de<br />

Géodynamique, Walferdange, Grand Duchy of<br />

Luxemburg, p. 39-57.<br />

Camelbeeck, T. and M. Meghraoui (1998). Geological and<br />

geophysical evidence for large paleoearthquakes with<br />

surface faulting in the Roer Graben (northwest Europe).<br />

Geophys. J. Int., 132, 347 - 362.<br />

Hinzen, K. G. (2003). Stress field in the Northern Rhine<br />

area, Central Europe, from earthquake fault plane<br />

solutions, Tectonophysics, 377, 325 - 356.<br />

Hinzen, K. G., and S. K. Reamer (2007). Seismicity,<br />

seismotectonics, and seismic hazard in the Northern<br />

Rhine Area, In: Intraplate Seismicity (Stein, S. and<br />

Mazzotti, S. eds). GSA Books, 225-243.<br />

Houtgast, R.F., R. T. van Balen, K. Kasse, and J.<br />

Vandenberghe (2003). Late Quaternary tectonic evolution<br />

and postseismic near surface fault displacements along<br />

the Geleen fault (Feldbiss fault zone-Roer valley rift<br />

system, the Netherlands), based on trenching, Geologie<br />

en Mijnbouw, 82(2), 177-196.<br />

Houtgast, R.F., R. T. van Balen, and C. Kasse (2005). Late<br />

Quaternary evolution of the Feldbiss fault (Roer valley rift<br />

system, the Netherlands) based on trenching, and its<br />

potential relation to glacial unloading, Quaternary<br />

Science Reviews, 24(3-4), 489-508.<br />

Leydecker, G. (2002). Earthquake Catalogue for the<br />

Federal Republic of Germany and Adjacent Areas for the<br />

Years 800 - 2001. Datafile, Federal Institute for<br />

Geosciences and Natural Resources, Hannover,<br />

Germany.<br />

Meidow, H. (1994). Comparison of the macroseismic field of<br />

the 1992 Roermond earthquake, the Netherlands, with<br />

those of large historical earthquakes in the Lower Rhine<br />

Embayment and its vicinity, Geologie en Mijnbouw, 73(2-<br />

4), 282-289.<br />

Streich, R (2003). Geophysical prospecting of suspected<br />

Holocene fault activity in the Lower Rhine Embayment,<br />

Germany, MSc Thesis. University of Potsdam, Germany.<br />

Vanneste, K. and Verbeeck, K. (2001). Paleoseismological<br />

analysis of the Rurrand fault near Jülich, Roer Valley<br />

graben, Germany: coseismic or aseismic faulting history?<br />

Geologie enMijnbouw, 80, 155-169.<br />

112


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Vanneste, K., Verbeeck, K., Camelbeeck, T., Paulissen, E.,<br />

Meghraoui, M., Renardy, F., Jongmans, D. and Frechen,<br />

M. (2001). Surface-rupturing history of the Bree fault<br />

scarp, Roer Valley graben: evidence for six events since<br />

late Pleistocene. Jour. Seism., 5, 329-359.<br />

113


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SEISMOTECTONIC AND SEISMIC HAZARD MAPS OF LITHUANIA (BALTIC REGION) –<br />

RECENT IMPLICATIONS OF INTRACRATONIC SEISMICITY<br />

Lazauskiene, Jurga (1 and Pacesa, Andrius (2)<br />

(1) Lithuanian Geological Survey under the Ministry of Environment. S. Konarskio 35; Vilnius University, M.K. iurlionio 21/27,<br />

(1) LT-03123 Vilnius, Lithuania. Email: jurga.lazauskiene@lgt.lt,<br />

(2) Lithuanian Geological Survey under the Ministry of Environment. S. Konarskio 35, LT-03123 Vilnius, Lithuania. Institute of<br />

(2) Geology and Geography, T.Ševenkos, 13, LT-03123 Vilnius, Lithuania. Email: andrius.pacesa@lgt.lt,<br />

Abstract (Seismotectonic and seismic hazard maps of Lithuania (Baltic region) – recent implications of intracratonic<br />

seismicity): Lithuania, situated in the western part of the East European Craton, is regarded as an intracratonic area of low<br />

seismicity. Several dozens of earthquakes of intensity up to VII (MSK-64) were recorded since 1616 implying the possible<br />

occurrence of stronger earthquakes. The northern part of in the Baltic region is seismically more active than the southern, but the<br />

Kaliningrad earthquakes of 2004 showed the necessity to re-assess the seismicity of the region. The identification of seismogenic<br />

faults in the Baltic region is rather complicated due to the small scale of tectonic structures and significant errors of location of<br />

seismic events and even faults. Still, recently the seismic hazard and seismotectonic maps of Lithuania were compiled implying the<br />

highest seismic hazard of 32,6 cm/s 2 PGA in eastern and 25-30 cm/s 2 in northern Lithuania. The majority of the territory is<br />

described by PGAs of 10-20 cm/s 2 .<br />

Key words: intracratonic seismicity, Baltic region, seismic hazard, seismotectonic framework,<br />

INTRODUCTION<br />

The territory of Lithuania comprises a part of the<br />

Baltic sedimentary basin situated in the western part<br />

of the East European Craton that is characterised by<br />

low seismic activity - the historical sources since<br />

1616 to 1964 record only a few tens of weak or<br />

moderate earthquakes (Pasa et al., 2005). The<br />

historical seismic activity in the eastern Baltic region<br />

is significantly lower comparing with seismicity of the<br />

Fennoscandian shield. High seismic activity of the<br />

Fennoscandian shield and adjacent Baltic Sea<br />

territories during the Late Glacial and Holocene (last<br />

13 000 years) is also well documented by numerous<br />

paleoseismic investigations and corresponding<br />

publications. The earthquakes caused landslides in<br />

glacial till, seismically-induced soft sediment<br />

deformation structures, “seismites”, are common in<br />

trench exposures in the vicinity of the faults in<br />

northern Sweden and even with tsunami events<br />

reported in the Baltic Sea (Mörner, 2005, 2008).<br />

Still, the eastern Baltic region is more seismically<br />

active comparing with the more “inland” aseismic<br />

territories of the craton. Several tens (~ 40) of<br />

earthquakes of intensities of VI-VII (MSK-64 scale)<br />

and local magnitudes up to M L = 5 are recorded in the<br />

Baltic region and neighbouring since 1616 (fig. 1)<br />

The strongest instrumentally registered earthquakes<br />

are Ossmussare (Estonia) earthquake of 1976 (with<br />

maximal magnitude up to M L = 4,75) and the<br />

Kaliningrad (Russia) earthquakes of 2004 of<br />

magnitudes, respectively, M L = 4.75 and M L = 5.0 (M w<br />

= 5.2, table 1). The other more significant<br />

earthquakes in the region are:<br />

Fig. 1: Main tectonic faults and seismic events in the Baltic<br />

region (after Sharov (ed.) et al., 2007; Pasa et al., 2005 ).<br />

1–3 faults: 1 – superregional, 2 – regional, 3 – subregional;<br />

5–8 – epicentres of earthquakes with local magnitudes:<br />

5 – M L = 1-2, 6 – M L = 2-3, 7 –M L = 3-4, 8 – M L = 4-5.<br />

February 22, 1821, Kokneses (Estonia), M L = 4.5;<br />

December 28, 1908, Gudogai (Belarus), M L = 4.5;<br />

114


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

December 29, 1908, Madona (Latvia); M L = 4.5<br />

(Boborikin et al., 1993). All the mentioned<br />

earthquakes within the same Baltic sedimentary<br />

basin in the same tectonic setting show the recent<br />

seismic activity of the Baltic region.<br />

INTRACRATONIC SEISMICITY OF THE BALTIC<br />

REGION<br />

The seismic activity in the Baltic region has an<br />

irregular distribution - the northern part of the region<br />

is more seismically active than the southern one (fig.<br />

1). The maximum activity is recorded in Latvia that is<br />

characterized by most intense faulting of the<br />

sediment layers. Seismic activity is slightly lower in<br />

Estonia, while the territory of Lithuania seems to be<br />

the most quite. The boundary between two areas of<br />

different seismicity approximately coincides with the<br />

northern state border of Lithuania. A quite similar<br />

boundary between northern and southern parts of the<br />

region was established in previous global seismic<br />

hazard assessment studies e.g. WSHAP (Giardini,<br />

1999) and European–Mediterranean Seismic Hazard<br />

Map (Jimenez et al., 2003).<br />

No earthquakes are registered instrumentally or<br />

reliably recorded historically in the territory of<br />

Lithuania. Peter of Duisburg, a 14th Century<br />

chronicler of the Teutonian Knights (Chronicon<br />

Terrae Prussiae) reported “a ground shaking which<br />

was felt in Skirsnemune castle” (south-western part<br />

of Lithuania) in 1328; after the event the castle was<br />

abandoned. However, this historical record causes<br />

some doubts and is considered controversially:<br />

A. Nikonov (personal communication) includes this<br />

seismic event of 1328 (and one of 1303 in Prussia;<br />

(Chronicon Terrae Prussiae)) into revised<br />

seismological catalogue of the South-East Baltic<br />

region, while Grunthal and Riedel (2007) strongly<br />

deny the existence of these events. Thus, this event<br />

was not included to the complete catalogue of<br />

seismic events of Eastern Baltic Region. There are<br />

also three records about the ground shaking during<br />

the years 1908-1909 in western and central part of<br />

Lithuania, but the primary sources rises some doubts<br />

and could not be considered as reliable. The first<br />

local seismic event with magnitude of M = 2.1<br />

presumably of tectonic origin was instrumentally<br />

registered 4-th of September, 2001 by one of short<br />

period seismic station, located in the NE part of<br />

Lithuania. The seismic signal of this seismic event<br />

was the typical one for tectonic events (Pasa et al.,<br />

2002). Still, as it was registered only in one station,<br />

no accurate location of this event was possible and<br />

the epicentre of the earthquake could have been<br />

located in a distance of ~80 km from the station. The<br />

passive seismic experiment PASSEQ was carried out<br />

in the eastern part of Europe, including territory of<br />

Lithuania in 2006–2007. The preliminary data<br />

received from the project indicated some possibility<br />

of tectonic seismic event in the middle Lithuania 4-th<br />

July, 2007 (coordinates - 55,053; 24,264) and one<br />

more event in Kaliningrad offshore of the Baltic Sea,<br />

14-th June, 2007 (coordinates - 54,796; 19,232)<br />

(Kozlovskaya et al., 2010). The coordinates of the<br />

epicentre of the onshore Lithuania event were<br />

identified near Kaunas town, located within the<br />

Middle Lithuanian Suture zone (the zone between<br />

two different Precambrian domains) with possibly<br />

increased tectonic activity. Both onshore and the<br />

Baltic Sea offshore events have occurred at the<br />

night-time. The seismic signals have the<br />

characteristics of tectonic events, but the quality of<br />

data is too poor to define magnitudes and depths of<br />

epicentres of these seismic events and, respectively,<br />

too poor for the reliable conclusions. Also, it must be<br />

noted, that the network of local seismic stations in the<br />

eastern Baltic region, especially Lithuania, is rather<br />

sparse and the instrumental seismological<br />

information is quite poor.<br />

No doubts, the seismic events in Kaliningrad area<br />

showed the necessity to review the understanding of<br />

the seismicity of the Baltic region - earlier it was<br />

considered that the maximum magnitude of the<br />

earthquakes in this part of region might be M L = 4.8;<br />

but the magnitude of the Kaliningrad earthquake was<br />

M L = 5.0 (M w = 5.2; table 1). Thus, taking into<br />

consideration the accepted margin of 0.5 (based on<br />

common agreement), the maximum magnitude could<br />

be assumed to be as high as M w = 5.7 that implies<br />

the possibility of occurrence of some stronger<br />

earthquakes in the Baltic region.<br />

Date Time Lat Long Magn Depth<br />

10/25/1976 8:39:00 59.2 23.58 4.7 M L 15 - 18<br />

9/21/2004 11:05:04 54.908 20.029 5.0 M w 16 ± 9<br />

9/21/2004 13:32:31 54.849 20.088 5.2 M w 20 ± 10<br />

Table 1. The strongest seismic events of Eastern Baltic<br />

region with magnitude M>4 (compiled by A. Pasa: the first<br />

one - Ossmussare (Estonia) earthquake, the other two –<br />

Kaliningrad (Russia) earthquakes. Magnitude types: M L –<br />

local magnitude, M w – moment magnitude. Coordinates of<br />

epicentres are provided in geographical system of<br />

coordinates: Lat - latitude (North), Long – longitude (East);<br />

time – in GMT; Magn – magnitude; Depth - in km. After:<br />

Gregersen et al., 2007.<br />

The seismotectonic framework of the study areas has<br />

been outlined several times during the last decades<br />

(Sharov et al., 2007; Suveizdis, 2003). A number of<br />

faults and fault zones have been distinguished in the<br />

Baltic region and adjacent territories based on<br />

geological and geophysical data. Still, different<br />

authors provide quite different tectonic and<br />

seismotectonic maps of the eastern Baltic region and<br />

there is no single commonly accepted tectonic map of<br />

this region currently; location, orientation, length or the<br />

other parameters of the same fault might be interpreted<br />

differently. No doubts, summarizing different maps one<br />

can infer some dominating fault zones, directions and<br />

spacing (fig. 1).<br />

DISCUSSIONS<br />

As it was mentioned, majority of seismic events in the<br />

eastern Baltic region are historical ones. The primary<br />

sources of information do not provide any evaluations<br />

of errors of epicentre locations for historical events,<br />

but, most likely, the errors vary from ten to several<br />

115


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

tens of kilometres. Even the epicentres for the two<br />

strongest Kaliningrad earthquakes were scattered in<br />

the area with a diameter of about 30 km (Pasa et<br />

al., 2005). It is very hard to associate single<br />

earthquakes with some certain faults unambiguously<br />

due to significant errors of location of seismic events<br />

and even location of faults. Additionally, the<br />

identification of the seismogenic faults in the region is<br />

rather complicated due to the small scale of tectonic<br />

structures located within an intracratonic area. It must<br />

also be pointed out that not all the earthquakes in the<br />

Baltic region are related to fault zones. Moreover, the<br />

majority of the previous global seismogenic and<br />

seismic hazard studies (Grünthal et al., 1999;<br />

Jimenez et al., 2003) that included also the territory<br />

of the Baltic region, strongly implied the local<br />

seismogenic sources with diffused seismicity. Thus, it<br />

was rather complicated to understand the<br />

geodynamic control on the seismicity of the relatively<br />

seismotectonically non-active cratonic area.<br />

Accordingly, until recently, no any seismic hazard<br />

map has been compiled neither for the territory of<br />

Lithuania, nor for the wider Baltic region. Finally, in<br />

year 2011, on request of the Lithuanian Geological<br />

Survey, the seismic hazard and seismotectonic maps<br />

(Šliaupa, 2011) have been compiled for the territory<br />

of Lithuania. The complex analysis of the geological,<br />

geophysical, geodetic, structural, seismic and<br />

geodynamic data allowed to distinguish 5 active<br />

seismogenic zones and 5 potentially active<br />

seismogenic zones in the territory of Lithuania. The<br />

maximal seismic potential has been implied for the<br />

W-E trending Silute-Polock and Northern Prieglius–<br />

Birstonas fault zones (Northern Prieglius has<br />

„hosted“ Kaliningrad earthquakes), transecting the<br />

central and southern part of Lithuania, with a<br />

maximum magnitude up to M = 5,5 of the possible<br />

earthquake predicted. It is implied that possible<br />

seismic activity of the faults is related to the regional<br />

stress field that affects the lithosphere of the Baltic<br />

region. No uniform stress pattern can be found for<br />

Baltic countries. Two stress provinces are suggested<br />

in Lithuania: NW-SE horizontal compression in the<br />

western part, whereas the main horizontal stress in<br />

the eastern part is NE-SW oriented. The western<br />

zone is attributed to the North Atlantic stress<br />

province, while the eastern is possibly a part of the<br />

Mediterranean stress province (Šliaupa, 2011; fig. 2).<br />

Therefore, as the seismic stresses are rather variable<br />

within the territory of Lithuania this might influence<br />

the character of the fault activation.<br />

The newly compiled seismic hazard map of Lithuania<br />

(Šliaupa, 2011) shows that the highest seismic<br />

hazard with peak ground acceleration equal to 32,6<br />

cm/s 2 (0,025 g) with 10% probability of exceedance<br />

within 50 years could be expected in the eastern part<br />

of Lithuania; also, the increased seismic hazard<br />

(peak ground acceleration being equal to 25-30<br />

cm/s 2 (0,030 g) with 10% probability of exceedance<br />

within 50 years) also is estimated in the northern part<br />

of Lithuania. Therefore, the majority of the territory of<br />

Lithuania is described by rather law seismic hazard -<br />

peak ground acceleration values varies in a range of<br />

10-20 cm/s 2 with 10% probability of exceedance<br />

within 50 years.<br />

Fig. 2. Directions of maximum dilatation in the territory of<br />

Lithuania (Sliaupa, 2011). Grey areas roughly indicate N-S<br />

maximum dilatation, blank – NW-SE maximum dilatation;<br />

areas subjected to bi-axial compression are bounded by<br />

dotted countures.Triangules indicates GPS points of the<br />

zero class; squares – GPS points if the first class; numbers<br />

-the numbers of GPS stations; dotted lines – maximal<br />

dilatation.<br />

The results of the new seismic hazard assessment<br />

show good coincidence with the results of previous<br />

global seismic hazard assessment - according to the<br />

results of Global Seismic Hazard Assessment<br />

Program (Grünthal et al., 1999) seismic level in the<br />

territory of Lithuania was estimated in the range of<br />

10–30 cm/s 2 for the standard seismic level of civil<br />

engineering (10% probability of exceedance within 50<br />

years or for a return period of 475 years) and<br />

according to the European map of seismic hazard<br />

(ESC-SESAME project, (Jimenez et al., 2003) the<br />

standard seismic level of civil engineering (return<br />

period of 475 years) was estimated as high as<br />

20cm/s 2 .<br />

References<br />

Boborikin, A..M., Avotinia, I.Y., Yemelianov, A. P., Sildvee,<br />

N. N., Suveizdis, P. (1993). Catalogue of historical<br />

earthquakes of Belarus and the Baltic Region.<br />

Seismological report of seismic stations of Minsk–<br />

Pleshchenitsi and Naroch for 1988. Minsk. 126–137.<br />

Kozlovskaya, E., Budraitis, M., Janutyte, I., Motuza, G.,<br />

Lazauskiene, J. and PASSEQ-Working Group, (2010). 3-<br />

D crustal velocity model for Lithuania and its application<br />

to local event studies. EGU General Assembly 2010,<br />

Vienna, Austria, 02-07 May, 2010, Abstract and<br />

programme book, Abstract No. EGU2010-10625.<br />

Giardini, D. (Ed.). 1999. The Global Seismic Hazard<br />

Assessment Program 1992–1999. Annali Geofis. 42 (6)<br />

Special Issue.<br />

Gregersen S., P. Wiejacz, W. Debski, B. Domanski, B.<br />

Assinovskaya, B. Gutterch, P. Mantiniemi, V.G. Nikulin,<br />

A. Pacesa, V. Puura, A.G. Aronov, T.I.Aronova, G.<br />

Grunthal, E.S. Husebye, S. Sliaupa. (2007). The<br />

exceptional earthquakes in Kaliningrad district, Russia on<br />

September 21, 2004. Physics of the Earth and planetatry<br />

interiors 164, 63-74.<br />

Grünthal, G. and GSHAP Region 3 Working Group. (1999).<br />

Seismic Hazard Assessment for Central, North and<br />

Northwest Europe: GSHAP Region 3. Annali di Geofisica<br />

42 (6), 999–1011.<br />

116


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Grünthal, G., Riedel, P. (2007). Zwei angebliche Erdbeben<br />

in den Jahren 1303 und 1328 im heutigen Raum<br />

Kaliningrad. Zeitschrift für Geologische Wissenschaften.<br />

Berlin, 35 (3). 157–163 (in Germany with English<br />

summary).<br />

Jiménez, M. J., Giardini, D. and Grünthal, G. (2003). The<br />

ESC-SESAME Unified Hazard Model for the European–<br />

Mediterranean Region. EMSC/CSEM Newsletter. 19, 2–<br />

4.<br />

Mörner, N.A. (2005). An interpretation and catalogue of<br />

Paleoseismicity in Sweden. Tectonophysics 408, 265-<br />

307.<br />

Mörner, N.A. (2008). Tsunami events within Baltic. Polish<br />

Geological Institute Special Papers, 23. 71-76.<br />

Pasa, A., Šliaupa, S., Satknas J. (2005). Recent<br />

earthquake activity in the Baltic region and seismological<br />

monitoring in Lithuania. Geologija 50, 8-18. (in<br />

Lithuanian).<br />

Pasa, A., Narbuntas, J., Nemcovas, G. (2002). The<br />

annual bulletin of the seismological monitoring in<br />

Lithuania. 2001. Vilnius. 40 p. (internal report in<br />

Lithuanian with English summary).<br />

Sharov N. V., Malovichko A. A., Schukin Ju. K. (eds.).<br />

(2007). Earthquakes and microseismicity of Eats<br />

European Platform – recent geodynamic approach. Book<br />

1 Earthquakes. Petrozavodsk: KarNC RAN, 381 p. (in<br />

Russian).<br />

Suveizdis (ed.). (2003). Tectonic structure of Lithuania.<br />

Institute of Geology and Geography, Vilnius. 160 p. (in<br />

Lithuanian).<br />

Šliaupa, S. (2011). Assessment of the seismic observations<br />

in the territory of Lithuania. Book 2 - Seismic hazard map<br />

of Lithuania. 136 p. (internal report, in Lithuanian).<br />

117


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PRELIMINARY STUDY ON DAMAGED STONE MONUMENTS IN GYEONGJU, SE KOREA<br />

Minjung Lee (1), Young-Seog Kim (1,*)<br />

(1) GSGR, Dept. of Earth Environmental Sciences, Pukyong National University, Daeyeon3-dong, Busan 608-737, Korea.<br />

*Email: ysk7909@pknu.ac.kr<br />

Abstract (Preliminary study on damaged stone monuments in Gyeongju, SE Korea): Concerns over the possibility of future<br />

earthquakes are high in Eastern Asia, since the catastrophic earthquake in north-eastern Japan on the 11 th of March, 2011. The<br />

Korean peninsula is regarded as a relatively safe area from earthquakes because it is located within the Eurasian intracontinental<br />

region. According to historical records, the Gyeongju area was struck by many large earthquakes. Recently a long crack was<br />

discovered on the east facing exterior surface of Seokgatap, a famous pagoda in Gyeongju. The crack length is 132 cm and its<br />

width is 5 mm. During the investigation into the cause of the damage, we performed kinematic analysis of the cracks on the<br />

pagoda. We cannot find any consistency between the damage of the pagoda and earthquakes, recently cracks were found within<br />

the pagoda, which was most probably developed by a prolonged continuous force. Other probable earthquake damage can,<br />

however, be seen in other historical heritage sites within this city and its surrounding regions. Some of the damage at this and<br />

other sites are also recorded in historical references. Therefore, more careful studies are necessary to identify and distinguish the<br />

origins of the damage.<br />

Key words: stone heritage, physical damage, Gyeongju, paleoseismology.<br />

INTRODUCTION<br />

The Korean peninsula is located within the relatively<br />

safe Eurasian intracontinental region. In some<br />

neighbouring countries around Korea such as Japan<br />

and Taiwan, big earthquakes occur frequently. Over<br />

forty Quaternary faults were recently discovered<br />

along the Yangsan and Ulsan faults which are major<br />

tectonic features in SE Korea (Fig. 1). According to<br />

Korean historical records (e.g. Samguksagi,<br />

Mukseojipyeon), several relatively strong seismic<br />

events, greater than intensity VII, have affected the<br />

Korean peninsula in the past (Lee and Yang, 2006).<br />

Gyeongju is a good place to study damage of stone<br />

heritages, because it has many stone artefacts at its<br />

heritage sites and it was the capital city of the Silla<br />

Dynasty for almost 1000 years from 57 BC to 935 AD.<br />

During this time, many large seismic events (over<br />

intensity VII) affected the Gyeongju and Ulsan areas<br />

destroying many heritage sites. In AD 779, a huge<br />

seismic event was recorded and resulted in about<br />

100 casualties. A number of buildings were<br />

destroyed due to this earthquake. It is inferred to be a<br />

destructive earthquake with a magnitude of 6.7 (Lee<br />

& Yang, 1983, 2006). The purpose of this study is to<br />

examine stone heritage sites and to distinguish and<br />

interpret the affects of recorded paleoseismological<br />

events at cultural heritage sites in the Gyeongju area.<br />

GENERAL GEOLOGY<br />

The basement of the study area consists of<br />

Cretaceous sedimentary rocks, the Taegu Formation,<br />

which forms part of the Gyeongsang Basin. This<br />

formation is discordantly intruded by Cretaceous and<br />

Tertiary igneous rocks (Fig. 1). In recent years, more<br />

than forty Quaternary faults have been reported near<br />

the Yangsan and Ulsan fault systems, which are the<br />

major fault systems in and around the Gyeongsang<br />

Basin (Lee and Jin, 1991; Kyung & Okada, 1995;<br />

Chang, 2001; Kim and Jin, 2006). The geometry of<br />

the intersection zone between the Yangsan and<br />

Ulsan faults is similar to simulated -faults (a low<br />

angle merging fault system; Du & Aydin, 1995, Kim et<br />

al., 2000).<br />

Fig. 1: Location map and geological map of the study area<br />

(modified from Lee et al., 1995).<br />

EARTHQUAKE RECORDS<br />

Instrumental recording of earthquakes in Korea<br />

began in 1905, and about one thousand earthquakes<br />

of mostly small magnitude (


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

between 2 to 1904 AD (Jin et al., in press, Lee &<br />

Yang, 2006).<br />

Based on the compiled earthquake catalogue (Fig. 2),<br />

more reliable records were reported after the Koryeo<br />

Dynasty. We have identified 3 clusters and their<br />

respective recurrence intervals. We observed that<br />

after the first cluster there was a recurrence interval<br />

of 368 years. After this recurrence interval,<br />

earthquakes occurred frequently over a period of 193<br />

years. This was followed by another recurrence<br />

interval of 53 years which was in turn followed by a<br />

110 year period of repeated earthquakes. We are<br />

now 200 years into a period without earthquake<br />

events.<br />

DESCRIPTION OF THE SEOKGATAP<br />

Bulguksa temple was built in AD 751 and is one of<br />

the most famous temples in Korea. There are two<br />

main pagodas in the temple site. The three-story<br />

Seokgatap which stands at 8.2 m is a traditional<br />

Korean-style granite pagoda with simple lines and<br />

minimal detailing. Seokgatap is over 13 centuries old.<br />

The other one - Dabotap - is 10.4 m tall. In contrast<br />

to Seokgatap, Dabotap is famous for its highly ornate<br />

structure.<br />

The temple was renovated during the Koryeo<br />

Dynasty (AD 918 – 1392) and the early Joseon<br />

Dynasty (AD 1392 – 1910). Historical records say<br />

that the Seokgatap was destroyed twice, in 1024 and<br />

in 1036 by big earthquakes (National Museum of<br />

Korea, 1997). After 1604, reconstruction and the<br />

expansion of Bulguksa started, this was followed by<br />

about 40 renovations leading up to 1805. During the<br />

Colonization of Korea between 1910 and 1945, the<br />

Japanese conducted restoration, but there were no<br />

records of the work done, and many known treasures<br />

disappeared during this time. After World War and<br />

the Korean War, partial restoration was conducted<br />

and completed in 1966. Upon an expansive<br />

archeological investigation, major restoration was<br />

conducted between 1969 and 1973, bringing<br />

Bulguksa to its current form. These stone pagodas<br />

are now preserved in their original Silla style.<br />

To interpret the factors related to fracturing and the<br />

instability of the pagoda, we performed a kinematic<br />

analysis on the developed cracks. Cracks on the<br />

structure demonstrate the factors that control crack<br />

creation and spreading. We can use crosscutting<br />

relationship, kinematic indicators, inferring stress and<br />

damage zone theory to interpret the conditions that<br />

the pagoda has been subjected to.<br />

According to an engineering survey for the<br />

Seokgatap, the following problems of structural<br />

stability are observed. The pagoda is tilled 0.9 to the<br />

Northwest. Foundation stones are dislocated with<br />

respect to each other by about 4 cm (Fig. 3). The<br />

center of the pagoda has subsided by about 3 cm.<br />

Also, we observed that there are cracks in the flat<br />

stones that are laid on other stones ranging in length<br />

between 1 and 6.5 cm. Some blocks have a gap<br />

between each other ranging from 3 to 4.5 cm. Also, a<br />

long crack of 132 cm in length and width of 5 mm<br />

have recently been detected on the eastern side of<br />

the pagoda. By comparing these results with<br />

previous state, the structural stability of the upper<br />

part of the pagoda is not safe and damage currently<br />

occurring.<br />

Fig. 3: (a) Photograph of western part of Seokgatap (b)<br />

sketch of western part of Seokgatap.<br />

Fig. 2: Histogram of historic earthquake catalogue from 2 AD to 1810 AD for >ML = 4. It shows three earthquake clusters from<br />

1013 AD and various recurrence intervals in the study area (modified from Jin et al., in press).<br />

119


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

DISCUSSION<br />

Quaternary faults have mainly been focused on the<br />

subject of paleoseismological studies in Korea.<br />

Recently, an archaeoseismological approach was used<br />

to infer the causes of the falling of a Buddha statue in<br />

Gyeongju area (Jin et al., in press).<br />

Archaeoseismological research in Gyeongju is<br />

important because deformed man-made structures<br />

with known age and original state offer supplementary<br />

information on past seismic events. Studying damaged<br />

monuments such as Seokgatap, Cheomseongdae and<br />

Sukbinggo suggest different major factors contributed<br />

to the damage at each site. The Cheomseongdae was<br />

built as an observatory and leans to the north by about<br />

4 and it shows a horizontal shift in its large ashlars<br />

(Fig. 4), which may have been caused by episodic<br />

forces. If we carry out a similar engineering survey and<br />

a kinematic analysis (slip sense, opening and tilting<br />

etc.) in order to know how the cracks and gaps are<br />

developed in the pagoda at Cheomseongdae as we did<br />

at Seokgatap, we could be able to estimate the reason<br />

of the damage.<br />

Fig. 4: (a) Overview of the damaged Cheomseongdae<br />

observatory in Gyeongju. (b) Sketch of southern part of<br />

Cheomseongdae.<br />

Acknowledgements: This work was funded by the Korea<br />

Meteorological Administration Research and Development<br />

Program under Grant CATER 2008-5502<br />

References<br />

Chang, T.W. (2001). Quaternary tectonic activity at the<br />

Eastern Block of the Ulsan Fault. Journal of the Geological<br />

Society of Korea, 37, 431-444 (in Korean with English<br />

abstract).<br />

Du, Y and Aydin, A., (1995). Shear fracture patterns and<br />

connectivity at geometric complexities along strike-slip<br />

faults. Journal of Geophysical Research, 100, 18093-18102.<br />

Jin, K., Lee, M., Kim, Y.-S., Choi, J.-H. (2011).<br />

Archaeoseismological studies on historical heritage sites in<br />

the Gyeongju area, SE Korea. Quaternary International,<br />

doi:10.1016/j.quaint.2011.03.055.<br />

Kim, Y.-S., Andrews, J.R., Sanderson, D.J. (2000). Damage<br />

zones around strike-slip fault systems and strike-slip fault<br />

evolution, Cracking Haven, southwest England.<br />

Geoscience Journal, 4, 53-72.<br />

Kim, Y.-S., Jin, K. (2006). Estimated earthquake magnitude<br />

from the Yugye Fault displacement on a trench section in<br />

Pohang, SE Korea. Journal of the Geological Society of<br />

Korea, 42, 79-94 (in Korean with English abstract).<br />

Kyung, J.G., Okada, A. (1995). Liquefaction phenomena due<br />

to the occurrences of great earthquakes; some cases in<br />

central Japan and Korea. Journal of the Geological Society<br />

of Korea, 31, 237-250.<br />

Lee, K. and Na, S. H. (1983). A study of microearthquake<br />

activity of the Yangsan fault. Journal of the Geological<br />

Society of Korea, 19, 127-135.<br />

Lee, K., Jin, Y.G. (1991). Segmentation of the Yangsan fault<br />

system: geophysical studies on major faults in the<br />

Kyeongsang basin. Journal of the Geological Society of<br />

Korea, 27, 434-449.<br />

Lee, K. (1998). Historical earthquake data of Korean, Journal<br />

of the Korea Geophysical Society, 1, 3-22.<br />

Lee, K., Yang, W.-S. (2006). Historical seismicity of Korea.<br />

Bulletin of the Seismo-logical Society of America 96,<br />

846e855.Stiros, S. 1988b. Archaeology, a tool to study<br />

active tectonics – The Aegean as a case study. Eos,<br />

Transactions, American Geophysical Union 13, 1636-1639.<br />

National Museum of Korea (1997). News.<br />

Kim, B.-S. (1145) Samguksagi.<br />

CONCLUSIONS<br />

Gyeongju is located around the junction between the<br />

Yangsan and Ulsan faults. Recently many Quaternary<br />

faults have been reported along the Yangsan and<br />

Ulsan faults. This has encouraged research on fault<br />

activities such as Quaternary faults and related<br />

paleoseismicities. According to historical records, the<br />

study area was significantly affected by earthquakes.<br />

Large seismic events produced heavy casualties and<br />

destroyed historical property. To identify the reason for<br />

the damage, we have performed kinematic analysis on<br />

the Seokgatap, in Gyeongju. Comparing previous<br />

report of structural stability with its present state, it is<br />

being destroyed by anonymous forces and movements<br />

such as foundation subsidence. Further detailed<br />

research on other destroyed stone monuments in<br />

Gyeongju would be helpful to clarifying the relationship<br />

between damage and historical earthquakes. So this<br />

study can give some valuable information to heritage<br />

preservation and earthquake hazard study in Korea.<br />

120


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKE GEOLOGY AND RELATED HAZARD IN KACHCHH, GUJARAT,<br />

WESTERN INDIA<br />

Malik, Javed N.(1), Michio Morino (2), Mahendra S. Gadhavi (3,4), Khalid Ansari (1), Chiranjeeb Banerjee (1),<br />

B. K. Rastogi (3),Fumio Kaneko (2), Falguni Bhattacharjee (3), Ashok. K. Singhvi (5)<br />

(1) Department of Civil Engineering, IIT Kanpur. Kanpur 208016. UP. INDIA. Email: javed@iitk.ac.in<br />

(2) OYO International Corporation Yushima 1-Chome Bldg. 4F, 1-6-3 Yushima, Bunkyo-ku, Tokyo 113-0034, JAPAN.<br />

Email: morino@oyointer.com; kaneko@oyointer.com<br />

(3) Institute of Seismological Research, Gandhinagar 382018, Gujarat, INDIA. Email: mahendrasinh@gmail.com;<br />

brastogi@yahoo.com; falguni.geo@gmail.com<br />

(4) L. D. College of Engineering, Ahmedabad, Gujarat. INDIA<br />

(5) Physical Research Laboratory, Ahmedabad. INDIA. Email: aksprl@yahoo.com<br />

Abstract: The Kachchh falls under seismic zone V outside the Himalaya. It is marked by several E-W striking longitudinal faults<br />

viz. the Allah Bund Fault (ABF), Island Belt Fault (IBF), South Wagdh Fault (SWF), Kachchh Mainland Fault (KMF), and Katrol Hill<br />

Fault (KHF). Several large to moderate magnitude earthquakes have struck this area during last 300 years, viz. 893 AD, 1668<br />

Indus Delta (M7); 1819 Allah Bund (M7.8), 1956 Anjar earthquake (Ms6.1), and the recent 2001 Bhuj earthquake (Mw7.6). The<br />

rupture of 2001 remained concealed, suggestive of occurrence on blind fault. We carried out active fault identification, mapping<br />

and paleoseismic studies along ABF, KMF and KHF. CORONA satellite photos were used for identification of active fault traces.<br />

Ground Penetrating Radar (GPR) profiling helped us in locating appropriate site for trenching across KMF and KHF. Our study<br />

suggests that all three fault are active and were ruptured during recent historic past.<br />

Key words: Kachchh, active faults, paleoseismic studies, Western India.<br />

INTRODUCTION<br />

The state of Gujarat has two fold hazard posed by<br />

earthquakes, one on-land along the active faults in<br />

Kachchh region and its neighbourhood, and another<br />

offshore along the Makran Subduction Zone (MSZ)<br />

located in the west (Figure 1). The entire coastline is<br />

highly vulnerable from the tsunami triggered by an<br />

earthquake occurring along the MSZ. Two such<br />

tsunami events have been recorded during historic<br />

and recent times by earthquakes along MSZ, i.e.,<br />

during 325 BC and 1945 AD. Apart from this the<br />

Kachchh region has been struck by several large to<br />

moderate magnitude earthquakes during last 300<br />

years, viz. 893 AD, 1668 Indus Delta (M7); 1819<br />

Allah Bund (M7.8), 1956 Anjar earthquake (Ms6.1),<br />

and the recent 2001 Bhuj earthquake (Mw7.6) (Malik<br />

et al., 1999a; Bilham, 1999). From these events, only<br />

1819 Allah Bund earthquake has been reported to<br />

have accompanied with 80-90 km long surface<br />

rupture and uplift resulting into formation of about 4-6<br />

m high scarp (Quittmeyer and Jacob, 1979; Johnstan<br />

and Kanter, 1990). The recent 2001 Bhuj earthqukae<br />

with magnitude Mw7.6, the rupture remained<br />

concealed below the ground at a depth of 7-10 km,<br />

suggestive of occurrence on blind fault (Mandal and<br />

Horton, 2007). It is a matter of concern that if<br />

movements on a blind fault are capable of producing<br />

large magnitude earthquakes, than having<br />

earthquakes of similar magnitude or larger on active<br />

faults with surface rupture cannot be ruled out (Malik<br />

et al., 2008; Morino et al., 2008). Active faults are<br />

considered to be the source for large magnitude<br />

earthquakes in seismically active regions. Their<br />

proper identification and distribution significantly help<br />

in knowing the seismic potential and associated<br />

hazard in the region. The landscape of Kachchh is<br />

marked by several E-W striking longitudinal faults viz.<br />

the Allah Bund Fault (ABF), Island Belt Fault (IBF),<br />

South Wagdh Fault (SWF), Kachchh Mainland Fault<br />

(KMF), and Katrol Hill Fault (KHF). Under the project<br />

sponsored by Gujarat State Disaster Management<br />

Authority (GSDMA) on Seismic Microzonation of<br />

Gandhidham, Kachchh, we carried out active fault<br />

mapping and paleoseismic investigations along ABF,<br />

KMF and KHF. In this paper we highlight in brief our<br />

findings based on detailed satellite data interpretation<br />

for identification of active faults and related<br />

geomorphic features as well as paleo-earthquake<br />

signatures preserved in sediment succession<br />

(Figures 1).<br />

Active fault and paleoseismic investigations:<br />

Katrol Hill Fault:<br />

Several new active fault traces were identified along<br />

Katrol Hill Fault (KHF) (Figure 1). These fault traces<br />

were identified based on satellite photo interpretation<br />

and field survey. Trenches were excavated to identify<br />

the paleoseismic events, pattern of faulting and the<br />

nature of deformation. Active fault traces were<br />

recognized about 1km north of the topographic<br />

boundary between the Katrol Hill and the plain area.<br />

121


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Figure 1. Generalized structural map of Kachchh region (after Biswas and Deshpande, 1970). Inset at the left top show DEM of<br />

India highlight the location of Kachchh peninsula and Makran Subduction Zone (MSZ). Inset at the lower right shows major<br />

geomorphic zones of Kachchh. Yellow box marks the study area along Kachchh Mainland Fault and Katrol Hill Fault, and along<br />

Allah Bund Fault. NHR- Northern Hill Range, KHR- Katrol Hill Range, KMF Kachchh Mainland Fault and KHF- Katrol Hill Fault.<br />

The fault exposure along the left bank of Khari River<br />

with 10m wide shear zone in the Mesozoic rocks and<br />

showing displacement of the overlying Quaternary<br />

deposits is indicative of continued tectonic activity<br />

along the ancient fault. The E-W trending active fault<br />

traces along the KHF in the western part changes to<br />

NE-SW or ENE-WSW near Wandhay village.<br />

Trenching survey across a low scarp near Wandhay<br />

village reveals three major fault strands F1, F2, and<br />

F3 (Figures 2a & b). These fault strands displaced<br />

the older terrace deposits comprising Sand, Silt and<br />

Gravel units along with overlying younger deposits<br />

from units 1 to 5 made of gravel, sand and silt.<br />

Stratigraphic relationship indicates at least three<br />

large magnitude earthquakes along KHF during Late<br />

Holocene or recent historic past.<br />

Kachchh Mainland Fault:<br />

Two trenches were dug along the KMF, one near<br />

Jhura and another near Lodai villages. We reported<br />

first identified active fault exposure from Kachchh<br />

region along the Kachchh Mainland Fault (KMF)<br />

other than the 1819 Allah Bund earthquake. The<br />

active fault scarps striking E-W were identified near<br />

Lodai village along KMF. North facing scarps with<br />

height from 10-15 m are the manifestation of the<br />

displaced alluvial fan surface along this fault.<br />

Occurrence of discontinuous linear mound ranging in<br />

height from 3-5 m aligned along the strike about 100<br />

m north of the main scarp are suggestive of younger<br />

tectonic movement and progressive shift of tectonic<br />

activity towards north along new imbricated fault<br />

(Figure 3). Three low to high angle reverse fault<br />

strands (F1, F2 and F3) displacing young Quaternary<br />

deposits (late Pleistocene-Holocene?) classified as A<br />

to F units comprising gravel and sand-silt facies were<br />

identified in a trench excavated at the base of the<br />

linear mound along KMF (Figure 3). Our preliminary<br />

observations revealed occurrence of at least two<br />

large magnitude earthquakes along the F3 fault, and<br />

may be older events along the F1 and F2. Latest<br />

event (Event-I) occurred along F3 after the deposition<br />

of unit B registering the displacement of ~33 cm,<br />

penultimate event (Event-II) occurred after the<br />

deposition of unit C with ~40 cm of displacement.<br />

The maximum displacement of about 73 cm along F3<br />

indicates cumulative displacement accommodated<br />

during more than one event. The total displacement<br />

of ~98 cm along F2 strand displacing the E and F<br />

units is the result of more than one event, and since<br />

the F2 probably displaced the unit C suggests that<br />

the movements occurred during penultimate (Event<br />

II) and during the Event III, older than penultimate.<br />

Displacement of Mesozoic succession during older<br />

events and unit B during the latest Event I along F1<br />

suggests repetitive movement along this fault. The<br />

fragile nature of ~3-4 m wide shear zone formed in<br />

Mesozoic rocks (shale+sandstone) also point<br />

towards repetitive tectonic movement along KMF.<br />

122


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Figure 2a. Photo-mosaic of the eastern wall of Wandhay trench exposed across the Katrol Hill Fault near Wandhay Dam. The<br />

trench was excavated across a very low fault scarp (refer Figure 2, for location). Three strands of thrust faults (F1, F2 and F3)<br />

dipping towards south were identified registering the latest event along F1, the penultimate as well as latest event along F2 and<br />

the older event along F3. 2b.Trench log of eastern wall of Wandhay trench. Terrace deposits are composed of Sand, Gravel-1,<br />

Silt, and Gravel- 2. These terrace deposits are covered by units 1 to 5 (after Morino et al., 2008).<br />

Figure 3. East wall view of trench excavated across active fault scarp near Lodai village along Kachchh Mainland Fault (KMF).<br />

Three main fault strands F1, F2 and F3 shows reverse faulting with variable dip ranging from 10°-55° towards south.<br />

To confirm further active faulting along the KMF,<br />

paleoseismic investigation near Jhura Village about<br />

30 km west of Lodai revealed an active fault<br />

displacing overbank deposits of Kaila River (Figure 2<br />

and 6). Two fault strands F1 and F2 were identified in<br />

the trench. The northern F1 shows a low-angle<br />

reverse fault with inclination of 15° towards the south.<br />

At least two faulting events were inferred on the basis<br />

of upward fault termination with clear angular<br />

unconformity. The net-slip during a single faulting<br />

event considering deformation on the hanging wall of<br />

F1 fault is over 5 m, suggestive of a large magnitude<br />

event during late Holocene period.<br />

Allah Bund Fault in Great Rann of Kachchh:<br />

The Kachchh region is not only well known for the<br />

occurrence of large magnitude earthquakes, but also<br />

for having a major Harappan (4000-4500 year) and<br />

historical sites. One of such major sites was<br />

Dholavira located on Khadir Island (Figure 1). Few<br />

sites in Great Rann of Kachchh (GRK), probably<br />

flourished until 1819 Allah Bund earthquake (?). Till<br />

date it is not fully understood as whether these sites<br />

were affected by the major seismic events in the past<br />

and also the presently evolved landscape was<br />

influenced by tectonic movements. The geologists,<br />

archaeologists, and scholars of ancient Indian history<br />

have mentioned the existence of numerous mighty<br />

southwest flowing rivers viz. the Sindhu (Indus),<br />

Shatadru or Nara (Sutlej) and Sarasvati, during Pre-<br />

Vedic and Vedic times (~4000 yr). These rivers<br />

flowed into then existing Arabian Sea, presently the<br />

GRK.<br />

We excavated 6-8 trenches in Allah Bund region<br />

GRK. Study reveals occurrence of at least 3 major<br />

events during recent past, which were probably<br />

responsible for the disruption of major channels (?),<br />

changing the landscape and destruction of the<br />

settlements (Figure 4). Trenches on the hanging wall<br />

of ABF shows thick massive yellowish medium-fine<br />

sand overlain by 1-1.5 m thick laminated sequence of<br />

123


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

silty-sand and clay. This suggests change in<br />

depositional environment from fluvial to fluvial-marine<br />

or tidal environment (high sea-level during 4000-6000<br />

yr?). Trench at Vigukot revealed prominent sandsheets<br />

at three levels indicative of 3 major<br />

liquefaction events, triggered by near source<br />

earthquakes (Figure 4), the latest event probably be<br />

the 1819 Allah Bund. Preliminary OSL ages of the<br />

sediments dated from the sand blow, soft sediment<br />

deformational structures, faulted sedimentary units<br />

from the trenches excavated on the hanging wall and<br />

across the Allah Bund Fault suggests occurrence of<br />

at least 2-3 events during 2.0-3.0 ka, with the most<br />

recent event during 2.0 ka.<br />

(a)<br />

(b)<br />

(c)<br />

Figure 4. (a) Vigukot Fort in GRK, fortification area is marked by broken while line. VT1, VT2, VT3 and VT4 are the locations of<br />

trenches excavated in and around Vigukot Fort, (b) field-photo showing location of trench VT1 and VT3. Location of VT3 is marked<br />

by yellow-box. Oriented bricks on the surface represents remnant of old foundation and (c) east wall view of 1.2 m deep and 3.5 m<br />

long trench excavated across a 80 cm thick wall (for location refer figure 4b). Exposed succession exhibits massive medium to fine<br />

sand at the base – suggestive of fluvial environment, capped by 1 m thick laminated succession – indicative of tidal environment.<br />

Preliminary interpretations suggests occurrence of at least three major event marked by prominent liquefied sand embedded in<br />

form of sand-sheet. CL1, CL2 and CL3 mark three cultural or occupation levels, the foundation in the centre of the trench belongs<br />

to the latest cultural level (CL3), which got severely affected during latest event, most likely the 1819 Allah Bund event(?).<br />

Acknowledgements: The authors are thankful to GSDMA,<br />

Gujarat, for funding this project. We are grateful to Prof. S.<br />

K. Jain, IIT Kanpur, now at IIT Gandhinagar and Dr. Alpa<br />

Sheth, VMS Consultants Private Limited, for their constant<br />

encouragement, support and discussion during this work.<br />

References<br />

Bilham, R. (1999). Slip parameters for the Rann of<br />

Kachchh, India, 16 June 1819 earthquake quantified<br />

from contemporary ccounts. Coastal Tectonics.<br />

Stewart, I. S., and Vita-Finzi, C. (eds.). Geological<br />

Society of London 146, 295-318.<br />

Biswas, S. K. (1980): Structure of Kutch-Kathiawar Region,<br />

W. India.- Proc. 3rd Ind. Geol. Congr., Pune, 255-272.<br />

Biswas, S. K., and Deshpande, S. V. (1970): Geological<br />

and tectonic maps of Kutch. - In: , 7, 115-116<br />

Johnston, A. C. and Kanter, L. R. (1990). Earthquakes in<br />

stable continental crust. Scientific American, 68-75.<br />

Malik, J. N., Sohoni, P. S., Karanth, R. V. and Merh, S.<br />

S. (1999a). Modern and Historic seismicity of Kachchh<br />

Malik, J. N., Morino, M., Mishra, P., Bhuiyan, C., and<br />

Kaneko, F. (2008). First active fault exposure identified<br />

along Kachchh Mainland Fault: Evidence from trench<br />

excavation near Lodai village, Gujarat, Western India.<br />

Journal Geological Society of India, 71, 201-208.<br />

Morino, M., Malik, J. N., Gadhavi, M. S., Ansari, K., Mishra,<br />

P., Bhuiyan, C., and Kaneko, F. (2008). Active Low-<br />

Angle Reverse Fault and Wide Quaternary Deformation<br />

Identified in Jhura Trench across Kachchh Mainland<br />

Fault, Kachchh, Gujarat, India. Journal of Active Fault<br />

Research, Japan, 29, 71-79.<br />

Morino, M., Malik, J. N., Mishra, P., Bhuiyan, C., and<br />

Kaneko, F. (2008). Active fault traces along Bhuj Fault<br />

and Katrol Hill Fault, and trenching survey at Wandhay,<br />

Kachchh, Gujarat, India. Journal of Earth System<br />

Sciences, 117(3),181–188.<br />

Quittmeyer, R. C. and Jacob, K. H. (1979). Historical and<br />

modern seismicity of Pakistan, Afghanistan,<br />

Northwestern India and South-eastern Iran. Bull. Seism.<br />

Soc. of America, 69, 773-823.<br />

124


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SEISMOGENIC SLUMPS IN PALAEO-DEAD SEA SEDIMENTS<br />

Marco, Shmuel (1) and G. Ian Alsop (2)<br />

(1) Department of Geophysics and Planetary Sciences, Tel Aviv Univesity. Ramat-Aviv, Tel-Aviv 69978, ISRAEL<br />

(shmulikm@tau.ac.il)<br />

(2) Department of Geology and Petroleum Geology, School of Geosciences, University of Aberdeen, Aberdeen AB24 3UE, UK<br />

(ian.alsop@abdn.ac.uk)<br />

Abstract (Seismogenic slumps in palaeo-Dead Sea sediments): We analyze a series of slumps in lake sediments overlying the<br />

Dead Sea Fault. The slumps are interpreted as seismites that have been triggered by earthquakes, thus providing a palaeoseismic<br />

record for the DSF. The direction of slumping inferred from the geometry and orientations of folds and thrusts varies systematically<br />

along the entire ~100 km length of the western Dead Sea Basin. They are interpreted to form part of a large-scale radial slump<br />

system directed towards the depocentre of the precursor to the Dead Sea. The recognition that slumps may be reworked by<br />

younger seismically-triggered events suggests that in some cases the seismic recurrence interval may be shorter than previously<br />

anticipated.<br />

Key words: paleoseismology; seismites; slumps; Dead Sea Fault<br />

DISCUSSION<br />

Seismites found in sediments of lakes that straddled<br />

the Dead Sea Fault have provided a 70-kyr-long<br />

palaeoseismic record, one of the longest on Earth<br />

(e.g., Ferry et al., 2011; Kagan et al., 2011; Ken-Tor<br />

et al., 2001; Marco et al., 1996; Migowski et al.,<br />

2004). Still, questions concerning the significance of<br />

their detailed shapes and the physical process that<br />

governed the formation of these seismites remain<br />

open. Understanding the seismites evolution and the<br />

underlying physics are tools for reconstructing the<br />

properties of past earthquakes on the basis of the<br />

observed deformations. These could also be useful in<br />

detecting and understanding off-fault earthquake<br />

indicators in other submarine environments.<br />

The Rayleigh-Taylor instability can explain<br />

mushroom-like symmetric deformation of inverseddensity<br />

stratification, where heavy strata overlay<br />

lighter strata (Figure 1a). However, stable<br />

stratification and asymmetric folds predominate the<br />

seismites in the Lisan Formation, which was<br />

deposited during the last glacial period in Lake Lisan<br />

(the precursor to the Dead Sea). Seismites in stably<br />

stratified lacustrine marls favor a mechanism of<br />

earthquake-triggered shear known as the “Kelvin-<br />

Helmholtz Instability” (Heifetz et al., 2005). Field<br />

observations and numerical simulations show that<br />

the deformation begins as moderate wave-like folds<br />

due to shear at the water-sediment interface (Wetzler<br />

et al., 2010). It evolves into asymmetric folds, then<br />

reclining folds, and in cases of backwash and/or<br />

vertical relaxation we observe backward-oriented<br />

folds (Fig. 1). If the flow becomes turbulent the layers<br />

are fragmented, re-suspended, and ultimately redeposited<br />

as breccia. During past earthquakes this<br />

process stopped at different stages, depending on<br />

the strength and duration of the shaking.<br />

Our detailed analysis of the seismites (Alsop and<br />

Marco, 2011b) reveals that some of them were<br />

affected by later events. Failure to recognize<br />

reworked seismites would reduce the number of<br />

earthquakes inferred from simple counting. In cases<br />

of incomplete reworking careful scrutiny can<br />

distinguish two earthquakes represented in one<br />

seismite. For example, we observe breccia layers<br />

that are folded as part of slumps, which we interpret<br />

as multiple event seismites (Fig. 2).<br />

We recognize that the direction of slumping inferred<br />

from the folds and thrusts varies systematically along<br />

the entire ~100 km length of the western Dead Sea<br />

Basin (Fig. 3). Hence the asymmetry of the folds<br />

represents the sense of shear, which was determined<br />

by very subtle slope of less than 1°. In the case of<br />

Lake Lisan coherent pattern of the fold vergence<br />

indicates that the location of the depocenter dictated<br />

the flow (Alsop and Marco, 2011a).<br />

Acknowledgments: We thank Mr. John Levy, together with<br />

the Carnegie Trust and the Royal Society of Edinburgh for<br />

travel grants to IA, and the Israel Science Foundation for<br />

grant 1539/08 to SM.<br />

References<br />

Alsop, G. I., and Marco, S., 2011a, A large-scale radial<br />

pattern of seismogenic slumping towards the Dead Sea<br />

Basin: Journal of the Geological Society. Accepted.<br />

Alsop, G. I., and Marco, S., 2011b, Soft-sediment<br />

deformation within seismogenic slumps of the Dead Sea<br />

Basin: Journal of Structural Geology, v. 33, no. 4, p. 433-<br />

457.<br />

El-Isa, Z. H., and Mustafa, H., 1986, Earthquake<br />

deformations in the Lisan deposits and seismotectonic<br />

implications: Geophys. J. R. Astron. Soc., v. 86, p. 413-<br />

424.<br />

Ferry, M., Meghraoui, M., Abou Karaki, N., Al-Taj, M., and<br />

Khalil, L., 2011, Episodic Behavior of the Jordan Valley<br />

125


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Section of the Dead Sea Fault Inferred from a 14-ka-Long<br />

Integrated Catalog of Large Earthquakes: Bulletin of the<br />

Seismological Society of America, v. 101, no. 1, p. 39-67.<br />

Hall, J. K., 1996, Digital topography and bathymetry of the<br />

area of the Dead Sea Depression: Teconophysics, v.<br />

266, p. 177-185.<br />

Heifetz, E., Agnon, A., and Marco, S., 2005, Soft sediment<br />

deformation by Kelvin Helmholtz Instability: A case from<br />

Dead Sea earthquakes: Earth Planet. Sci. Lett., v. 236, p.<br />

497-504.<br />

Kagan, E., Stein, M., Agnon, A., and Neumann, F., 2011,<br />

Intrabasin paleoearthquake and quiescence correlation of<br />

the late Holocene Dead Sea: J. Geophys. Res., v. 116,<br />

no. B4, p. B04311.<br />

Ken-Tor, R., Agnon, A., Enzel, Y., Marco, S., Negendank, J.<br />

F. W., and Stein, M., 2001, High-resolution geological<br />

record of historic earthquakes in the Dead Sea basin: J.<br />

Geophys. Res., v. 106, no. B2, p. 2221-2234.<br />

Marco, S., Stein, M., Agnon, A., and Ron, H., 1996, Long<br />

term earthquake clustering: a 50,000 year paleoseismic<br />

record in the Dead Sea Graben: J. Geophys. Res., v.<br />

101, no. B3, p. 6179-6192.<br />

Migowski, C., Agnon, A., Bookman, R., Negendank, J. F.<br />

W., and Stein, M., 2004, Recurrence pattern of Holocene<br />

earthquakes along the Dead Sea transform revealed by<br />

varve-counting and radiocarbon dating of lacustrine<br />

sediments: Earth and Planetary Science Letters, v. 222,<br />

no. 1, p. 301-314.<br />

Wetzler, N., Marco, S., and Heifetz, E., 2010, Quantitative<br />

analysis of seismogenic shear-induced turbulence in lake<br />

sediments: Geology, v. 38, no. 4, p. 303-306.<br />

Figure 1. Schematic line<br />

drawings and photographic<br />

examples of structures<br />

generated during slump<br />

sheet initiation (a),<br />

translation (b, c), cessation<br />

(d), relaxation (e) and<br />

compaction (f). Folded beds<br />

are shown in yellow, while<br />

axial planes (blue) and<br />

thrusts (red) are also<br />

highlighted together with<br />

possible deformation styles.<br />

Not all slump sheets will<br />

show the full range and<br />

evolution of structures<br />

depicted here. Northeast is<br />

on the right of photographs<br />

(Figure from Alsop and<br />

Marco, 2011b).<br />

126


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Figure 2. Seismites 1 and 2 that are<br />

recognized as breccia layers are<br />

included in a seismogenic slump. Note<br />

that the upper part of the slump is also<br />

brecciated. We therefore interpret the<br />

sequence as containing evidence for<br />

three seismic events. Coin diameter is<br />

20 mm.<br />

Figure 3. Schematic illustrations of slump structures showing radial vergence directions. We show the mean vergence directions of<br />

over 350 measured folds on the west and slump directions from El-Isa and Mustafa (1986) on the east, suggesting the control of<br />

the depocenter on the slump transport (Alsop and Marco, 2011a). Map from Hall (1996).<br />

127


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

MAPPING AND MEASURING HOLOCENE FAULT SCARPS IN DENSE FORESTS WITH<br />

LIDAR<br />

McCalpin, James P.<br />

GEO-HAZ Consulting, Inc., P.O. Box 837, Crestone, Colorado 81131 USA. Email: mccalpin@geohaz.com<br />

Abstract (Mapping and Measuring Holocene Fault Scarps in Dense Forests with Lidar): In the past decade new fault scarps<br />

have been discovered in the forests of North America by the use of Lidar DEMs. Most of the surveyed regions had been examined<br />

previously with aerial photographs, but no scarps were seen through the tree canopy. Lidar in those same areas shows obvious<br />

fault scarps. The scarp heights and slope angles can be accurately measured directly from the DEM, and compare favourably to<br />

field measurements at the same sites. We describe several case histories in the USA where Lidar-detected scarps have been<br />

trenched and analyzed for their seismic source characteristics.<br />

Key words: fault scarps, Holocene, Lidar<br />

INTRODUCTION: THE PROBLEM OF FINDING<br />

FAULT SCARPS IN DENSE FORESTS<br />

The identification and mapping of Quaternary fault<br />

scarps is well advanced in arid and semi-arid regions<br />

that contain little vegetation. Sub-humid and humid<br />

regions, in contrast, are often heavily forested and<br />

the ground surface cannot be seen in aerial<br />

photographs. Until recently the only way to locate<br />

and map fault scarps in forested areas was to enter<br />

the forest on foot and begin searching, but without a<br />

previously-identified target, a search would be<br />

random and extremely time-consuming. As a result,<br />

the density of fault scarps mapped to date in forested<br />

areas is a small fraction of that mapped in open<br />

areas. The scarcity of mapped Quaternary faults in<br />

forested areas has, in turn, translated into lower<br />

predicted seismic hazards, particularly in areas with<br />

low historic seismicity. This low predicted seismic<br />

hazard may be incorrect for many regions, resulting<br />

mainly from the lack of a reliable technique for finding<br />

Quaternary faults in forests.<br />

The advent of light detection and ranging (LiDAR) in<br />

the 1990s has changed the situation dramatically. In<br />

this paper I describe four case histories where new<br />

fault scarps have been discovered in dense forests of<br />

the western USA and Alaska.<br />

LiDAR Basics<br />

LiDAR is also known by the more descriptive phrase<br />

“Laser altimeter terrain mapping.” An airborne<br />

scanning laser rangefinder collects millions of<br />

distance measurements between the airplane and<br />

the ground, as many as 30,000 points per second at<br />

~15 cm accuracy. The plane’s position is measured<br />

by differential GPS and an inertial navigation system.<br />

Each laser pulse measures multiple returns (distance<br />

measurements) along a single beam, with the first<br />

return from the top of local vegetation, and the last<br />

return from the ground surface. The last returns are<br />

amalgamated into a “bare earth” digital elevation<br />

model (DEM) which shows the shape of the ground<br />

surface beneath the forest canopy (virtual<br />

deforestation). The current cost of a LiDAR survey<br />

and bare-earth DEM is ca. $150–$400/km2.<br />

Fig. 1: Schematic diagram of LiDAR data acquisition<br />

The Puget Sound area in Washington State: The<br />

Seattle Fault and its Backthrusts<br />

The first LiDAR survey of the Puget Sound area in<br />

2000 revealed two previously unsuspected late<br />

Quaternary faults, the Toe Jam and Waterman Point<br />

faults. The Toe Jam fault scarp on Bainbridge Island<br />

was trenched by US Geological Survey in 1998 and<br />

1999 (Nelson et al., 2003). Trenching confirmed that<br />

the scarp records at least one, and probably more,<br />

large earthquakes since the latest glacial maximum<br />

(LGM, ca. 18 ka). A fossil beach terrace that<br />

surrounds this part of Bainbridge Island was uplifted<br />

about 7 meters in a single large earthquake about<br />

1,100 years ago. Likewise, trenching in August 2001<br />

confirmed that the Waterman Point scarp, like the<br />

Toe Jam scarp, follows a south-verging thrust fault<br />

that has moved since the Latest Glacial Maximum.<br />

128


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

blanketed by a dense coastal rain forest, which has<br />

defeated past attempts to unravel its neotectonics.<br />

Fig. 2. LiDAR image that revealed the scarp of the<br />

Toe Jam fault on Bainbridge Island, west of<br />

downtown Seattle, Washington, USA. The fault is an<br />

east-west-trending, north-dipping reverse fault.<br />

Western USA; The Upper Rio Grande Rift, Colorado<br />

The 1000 km-long Rio Grande rift traverses the<br />

desert state of New Mexico and much of the state of<br />

Colorado. In the desert Quaternary fault scarps are<br />

unobscured by vegetation, but as the rift floor<br />

gradually rises northward to 2500 m and ultimately<br />

3000 m in central Colorado, the rift margins become<br />

densely forested. Fault scarps higher than about 10<br />

m can be seen on aerial photographs (i.e., about half<br />

as high as the average tree height of 20-25 m).<br />

Smaller scarps cannot be seen on airphotos, but can<br />

be seen on LiDAR.<br />

The Williams Fork normal fault was discovered in<br />

2002 in a dense pine forest at the foot of the Williams<br />

Fork Mountains in central Colorado. Even though the<br />

aerial photographs did not show any fault scarps in<br />

the dense forest, Kirkham (2004) walked to the range<br />

front and discovered multiple-event Quaternary fault<br />

scarps there. That fault is now the northernmost<br />

known Quaternary fault associated with the Rio<br />

Grande rift zone, where scarps are clearly late<br />

Quaternary in age, and trenches show displacement<br />

of late Quaternary strata.<br />

This discovery encouraged the collection of 2400<br />

km2 of additional LiDAR of the rift margins in 2010 by<br />

the Colorado Geological Survey and USGS. The new<br />

LiDAR DEMs show numerous previously unknown<br />

fault scarps on the forested rift margins. These fault<br />

scarps fill in large gaps in the maps of the Quaternary<br />

Fault and Fold Database of the USA.<br />

Alaska: The Yakutat Microplate, Reverse Faults and<br />

Sackungen<br />

The Yakutat microplate lies in south coastal Alaska,<br />

comprising a small plate fragment caught between<br />

the North American plate and the Pacific Plate (Fig.<br />

3). The northward plate convergence of 55 mm/yr<br />

has created a spectacular barrier of coastal hills and<br />

mountains. These mountains intercept Pacific storms<br />

and lead to an annual precipitation of about 2000<br />

mm/yr. Due to the high precipitation the terrain is<br />

Fig. 3. Location map of the Yakutat Terrane<br />

(microplate) in south coastal Alaska.<br />

Re-folding of foreland fold and thrust belt structures<br />

has created Holocene fault scarps in mountain blocks<br />

throughout the western part of the microplate. The<br />

morphology of the fold belt is atypical of classical<br />

foreland fold belts, where valleys and ridges are<br />

typically elongated parallel to fold axes. The Yakutat<br />

landscape is composed by irregular to elongated,<br />

north to northeast trending mountain blocks<br />

separated by flat-floored valleys filled with alluvial<br />

and glacial deposits. This morphology reflects<br />

Quaternary re-folding of foreland folds about plunging<br />

hinge-lines and vigorous erosion by glaciers and<br />

rivers. The mountain blocks are covered with swarms<br />

of discontinuous scarps (Fig. 4), mainly formed by<br />

displacement along bedding during second-phase<br />

folding of Tertiary strata.<br />

Fig. 4. Photograph of flexural-slip scarps on a typical<br />

mountain ridge in the Yakutat microplate.<br />

Flexural-slip scarps occur in a variety of orientations<br />

with respect to hill slopes and ridge crests; parallel to<br />

ridge crests, cutting across ridge crests, and<br />

extending across alpine valleys. Scarps forming by<br />

flexural-slip folding are tens to hundreds of meters in<br />

length, have vertical offsets up to several meters,<br />

most face 'up-slope', and some have a dominant<br />

lateral slip component (Li et al., 2010). Ratios of fault<br />

scarp displacement to length (D/L) range between<br />

0.1 to 0.001, overlapping with, but generally larger<br />

than, D/L for discrete tectonic faults. Distances<br />

between scarps measured normal to bedding are<br />

tens of meters. The fundamental controls on<br />

129


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

localization of bedding parallel scarps are the<br />

mechanical competence of bedding and mechanical<br />

anisotropy imparted by the first-phase structures (Li<br />

et al., 2010). Other factors may include complex<br />

stress distributions within steep-sided mountain<br />

blocks, and stress transients generated by ground<br />

motion during large to great magnitude earthquakes.<br />

Fig. 5. Faults (red) and gravitational scarps (green) shown by LiDAR around the Martin Lake detailed study site, in the<br />

western Yakutat microplate. Black dashed line with hachures shows the LGM glacial trimline. From McCalpin et al., in<br />

press.<br />

Alaska: The Foothills Fault Zone<br />

The Foothills fault zone is a zone of north-verging<br />

thrusts and interconnecting strike-slip faults on the<br />

north side of the Alaska Range in central Alaska.<br />

Some alternative routes for the proposed Alaska-<br />

Canada Gas Pipeline cross the fault zone. Recent<br />

neotectonic studies in the crossing areas (Carver et<br />

al., 2008; 2010) have discovered fault scarps along<br />

the densely-forested range front, visible on LiDAR<br />

imagery but not detectable on aerial photographs.<br />

DISCUSSION<br />

LiDAR DEMs promise to change the way<br />

neotectonic-paleoseismic investigations are<br />

performed in forested regions. LiDAR should be the<br />

standard tool used for the critical step of identifying<br />

and mapping Quaternary fault scarps (Fig. 6). The<br />

success of this step determines the success of the<br />

remaining steps in the investigation.<br />

Acknowledgements: I thank the St. Elias Erosion-<br />

Tectonics Project (STEEP) and the National Science<br />

Foundation for funding the neotectonic work in the Yakutat<br />

microplate.<br />

References<br />

Carver, G.A., Bemis, S.P., Solie, D.N., and Obermiller, K.E.<br />

(2008). Active and potentially active faults in or near the<br />

Alaska Highway corridor, Delta Junction to Dot Lake,<br />

Alaska: Preliminary Interpretive Report 2008-3d, Alaska<br />

Division of Geological and Geophysical Surveys,<br />

Fairbanks, AK, 32 p<br />

Carver, G.A., Bemis, S.P., Solie, D.N., Castonguay, S. and<br />

Obermiller, K.E. (2010). Active and potentially active<br />

faults in or near the Alaska Highway corridor, Dot Lake to<br />

Tetlin Junction, Alaska: Preliminary Interpretive Report<br />

2010-1, Alaska Division of Geological and Geophysical<br />

Surveys, Fairbanks, AK, 41 p.<br />

Kirkham, R.K. (2004) Quaternary faulting in the Williams<br />

Fork Valley graben, north-central Colorado, and<br />

comparison with late Quaternary deformation near<br />

Spinney Mountain, central Colorado. Unpublished Final<br />

Technical Report submitted to the U.S. Geological<br />

Survey, NEHRP Program, Grant 02HQGR0102, revised<br />

09-JAN-2004, 50 p.<br />

Li, Z., Bruhn, R.L., Pavlis, T.L., Vorkink, M. and Zeng, Z.<br />

(2010). Origin of sackung uphill-facing scarps in the Saint<br />

Elias orogen, Alaska: LIDAR data visualization and stress<br />

modeling. Geol. Soc. Amer. Bull., 122 (9-10), 1585-1599.<br />

McCalpin, J.P. (ed). (2009). Paleosesmology, 2 nd<br />

Edition:<br />

Academic Press-Elsevier, 848 p.<br />

McCalpin, J.P., Bruhn, R.L., Pavlis, T.L., Gutierrez, F.,<br />

Guerrero, J and Lucha, P., in press, Antislope scarps,<br />

gravitational spreading, and tectonic faulting in the<br />

western Yakutat microplate, south coastal Alaska.<br />

Geosphere., 2011.<br />

Nelson, A. R., Johnson, S. Y., Kelsey, H. M., Wells, R. E.,<br />

Sherrod, B. L., Pezzopane, S. K., Bradley, L. A., Koehler,<br />

R. D., and Bucknam, R. C. (2003). Late Holocene<br />

earthquakes on the Toe Jam fault, Seattle fault zone,<br />

Bainbridge Island, Washington. Geol. Soc. Am. Bull.<br />

115(11), 1388–1403.<br />

Pavlis, T. L. and Bruhn, R. L. (2011). Application of LIDAR<br />

to resolving bedrock structure in areas of poor exposure:<br />

An example from the STEEP study area, southern<br />

Alaska, Geological Society of America Bulletin, v. 123, p.<br />

206-217, doi: 10.1130/B30132.1.<br />

130


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 6. Flow chart of a typical paleoseismic investigation that includes fault trenching, showing the early need for<br />

LiDAR surveys in forested areas. You cannot trench a fault scarp if you never find it. Adapted from McCalpin (2009).<br />

131


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

CHANGES ON THE GEOMORPHIC SETTINGS OF<br />

SAND-POOR ENVIRONMENT COAST OF BANDA ACEH, INDONESIA<br />

SUBJECT TO TECTONIC AND TSUNAMI EVENTS<br />

Meilianda, Ella (1), Ben Maathuis (2), Marjolein Dohmen-Janssen (3)<br />

(1) INTERCOAST, University of Bremen, MARUM - Leobener Strasse D-28359 Bremen, GERMANY. Email:<br />

emeilianda@marum.de<br />

(2) Water Resource Department, Faculty of GeoInformation Science and Earth Observation (ITC), The Netherlands.+31 53<br />

487 4391; Email: maathuis@itc.nl<br />

(3) Water Engineering & Management (WEM), University of Twente, PO Box 217, 7500 AE Enschede, NETHERLANDS, email:<br />

c.m.dohmen-janssen@utwente.nl<br />

Abstract: Geomorphic settings before and after the December 2004 tsunami of the Banda Aceh coast, Sumatra, Indonesia were<br />

investigated in this study. The coast of Banda Aceh is a sand-poor environment contains thin layer of mobile sand perched on top<br />

of a consolidated Holocene prograding delta. Changes on the coastal shoreface morphology as the response to the tsunami<br />

waves were varying subject to the difference of geomorphic settings. The seawater inundated to the coastal plain as far inland as<br />

the shoreline position of 0.6 ky BP, during which a comparable magnitude of tsunami confirmed to have occurred in this region.<br />

This study demonstrates that such huge tsunami event occurred abruptly, but led to changes of the geomorphology developed in<br />

the Holocene.<br />

Key words: tsunami, tectonic, geomorphic setting, Holocene<br />

INTRODUCTION<br />

Despite the growth of tsunami-related studies since<br />

December 2004’s tsunami event, there was little<br />

discussion on the geomorphological adjustment and<br />

development of the affected coast. Given the fact that<br />

such a powerful earthquake leading to an<br />

extraordinary magnitude of tsunami is a rare event in<br />

human-life time history, there are ample knowledge<br />

gaps that hinder thorough investigations towards this<br />

level. In addition, even before the tsunami of<br />

December 2004 occurred, that the topic of<br />

paleotsunami requires inevitable sedimentological<br />

and geomorphological research, since such extreme<br />

event effects on sedimentary transport or<br />

considerable alterations of the coastal configuration.<br />

In spite of this, only 5% of the existing tsunami<br />

literature is related to such issues (Scheffers &<br />

Kelletat, 2003).<br />

The geomorphological state and development of the<br />

coast in pre-tsunami time is essential to investigate<br />

the extent of the environmental damage of the coast<br />

after being hit by such a huge catastrophe. In the<br />

absence of this knowledge for the coastal<br />

management practice, it is impossible, to forecast or<br />

to set-up future scenarios of the geomorphological<br />

development of the affected coasts.<br />

In terms of sedimentary transport and<br />

geomorphological research, our main challenge is<br />

the lack of readily available data of the pre-existing<br />

geomorphology of the affected coast, particularly in<br />

the less-investigated study location of Banda Aceh,<br />

Sumatra Island, Indonesia. Pragmatically, the<br />

investigation on the coastal geomorphological<br />

development using various types of available<br />

information such as literature, historical maps,<br />

satellite images, bathymetric charts and in-situ<br />

sample data are tailored to construct the<br />

geomorphological interpretation. The objective of the<br />

present study is therefore to qualitatively measure<br />

the impact of the December 2004 tsunami on the<br />

coastal geomorphology based on the reconstruction<br />

of geomorphology of the Banda Aceh coast in the<br />

past.<br />

The work in this study rests on the notion that each<br />

older geomorphological unit provides a boundary for<br />

the more recent units and therefore co-determines<br />

more recent geomorphological developments. Some<br />

preliminary studies after the tsunami event revealed<br />

that the extent of the tsunami inundation on the<br />

Banda Aceh coastal plain reached as far as about 5<br />

km away from the coastline (Dohmen-Janssen et al.,<br />

2006).<br />

Given the lack of geological data to support the<br />

quantification of the coastal destruction, we use data<br />

on the Holocene sea-level fluctuation of the region as<br />

a proxy and combined it with a Digital Elevation<br />

Model (DEM) of the coastal plain to interpret the<br />

geomorphological development of the coastal area.<br />

Additionally, this study also uses the generic<br />

knowledge of geomorphology and sediment<br />

stratification from bore hole samples to support the<br />

geomorphological interpretation. The geomorphic<br />

settings of the coastal area are also given by<br />

synthesizing the various but very limited geological<br />

and geomorphological studies of the coastal region.<br />

Overall, the study was aimed to answer the following<br />

research questions: 1) Which geomorphological units<br />

132


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

can we distinguish at the Banda Aceh coast and how<br />

do they respond to the earthquake and tsunami?; 2)<br />

To what extent have the earthquake and tsunami of<br />

26 December 2004 affected this coastal system?<br />

STUDY SITE, DATA SETS AND METHODS<br />

Banda Aceh is located at the northwestern tip of<br />

Sumatra Island, Indonesia (Fig. 1). The coastal plain<br />

has elevations from -0.5 m to +11 m relative to<br />

present sea level and occupies 125 square km of the<br />

northwest valley of the Barisan mountain range,<br />

which is the backbone of Sumatra Island. Krueng<br />

Aceh River is the main river crossing this low-lying<br />

coastal plain. In the 1990s, the middle reach of the<br />

river was bifurcated by a floodway channel, as a<br />

means to divert the surplus discharge of the river<br />

which caused annual flooding of the city. The<br />

coastline is dissected by parallel lagoon inlets in<br />

between some of the beach ridges sections.<br />

This study interpreted the geomorphology of the<br />

Banda Aceh coast from various types of maps and<br />

satellite images covering the past century and from<br />

literature covering the geological and<br />

geomorphological evolution of the surrounding<br />

region. All maps, images and charts were georeferenced<br />

to a master map (i.e. ortho-rectified aerial<br />

photo of 2005) which was processed in ArcGIS TM .<br />

The nautical charts, bathymetric data and a posttsunami’s<br />

topographic map of the coastal plain were<br />

transformed into Digital Elevation Models (DEMs)<br />

using Triangulation Irregular Network (TIN).<br />

topographic maps, bathymetric charts and bore hole<br />

samples as the proxy to the non-existing data (e.g.<br />

from core sampling, etc.) to reconstruct the<br />

development of the geomorphic settings of Banda<br />

Aceh.<br />

GEOMORPHIC SETTINGS OF BANDA ACEH<br />

Similar to the low latitude regions in the world, Banda<br />

Aceh (5° N) experienced a sea-level fall during the<br />

late Holocene (e.g. since 6.0 ky BP). Sea-level raised<br />

in Malacca Strait from -13 m in 8.0 ky BP to about +5<br />

m around 5.5 ky BP and then declined towards the<br />

present level (Geyh et al., 1979). Similar phenomena<br />

were also observed in Australia (Woodroffe and<br />

Horton, 2005), Singapore (Hesp et al. 1998) and<br />

Thailand (Sinsakul, 1990) for the same period.<br />

Tjia & Fujii (1989) suggested that at the west coast of<br />

Malaysian Peninsula the sea level reached its<br />

maximum of about 5 metres above present sea-level<br />

at approximately 5.0 ky BP. They used the carbon<br />

dating technique to investigate Holocene shoreline of<br />

the coasts of Malaysian Peninsula using abrasional<br />

and biogenic indicators. The sampled location at the<br />

west coast of Langkawi Island which is situated in the<br />

same tectonic region as to Banda Aceh and both<br />

locations are situated in the solitary Andaman Sea<br />

environment. In the present study, the sea-level<br />

fluctuation indicated by Tjia and Fujii (1989) was<br />

compared with the elevations as well as the units of<br />

morphology identified from the DEM of Banda Aceh<br />

coastal plain, which was derived from the 0.5-mcontour-interval<br />

topographic map. In this way, the<br />

shorelines associated with different phase of<br />

transgression and regression of sea-level during the<br />

Holocene were extracted. At the same time, the predominant<br />

process leads to the coastal morphology<br />

can be interpreted.<br />

Geomorphic settings of Banda Aceh coast<br />

Fig. 1: Topographic map of Banda Aceh coastal plain,<br />

Sumatra, Indonesia derived from DEM (RSTM TM<br />

2003 for<br />

the land part and nautical chart 1978 by BAKOSURTANAL,<br />

Indonesia).<br />

The Holocene sea-level fluctuation curves produced<br />

by Tjia & Fujii (1989) were used in our study to<br />

reconstruct the Holocene shoreline changes and the<br />

development of the Banda Aceh coastal plain<br />

towards the modern shoreline position. Overall, we<br />

use the results of these various studies in<br />

combination with the analysis of satellite images,<br />

Banda Aceh coastal plain and shoreface region<br />

consists of two distinctive geomorphic settings. The<br />

northeastern part is a tilted coastal zone (during the<br />

Pleistocene) which developments during the<br />

Holocene was influenced by the marine regression<br />

process. During this period, broad parallel coastal<br />

ridges and swales were developed over the modern<br />

coastal plain, and the shoreface has a concave -<br />

shape profile with sandy surface, indicating the predominant<br />

influence of the marine regression. On the<br />

other hand, the southwestern part consists of a broad<br />

alluvial flat-plain over a depression zone associated<br />

with the Sumatran Fault zone. The shoreface<br />

stratigraphic layers mainly consisted of silty-clay<br />

structures which profile is convex-shape, indicating<br />

an alluvial progradation which was developed during<br />

the Holocene (Meilianda, 2009).<br />

From the shoreline reconstruction using this method<br />

we analyse that the shoreline response to sea level<br />

fluctuations at Banda Aceh during the Holocene was<br />

highly influenced by the mechanisms of repeated<br />

changes of river mouths position and intermittent<br />

beach-ridge formation caused by periods of above<br />

133


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

average onshore winds; a similar climate condition of<br />

Jakarta Bay in Java Island according to Verstappen<br />

(1973).<br />

Modern coastal geomorphic settings and changes<br />

after December 2004’s tsunami<br />

Parallel to the completion of this study, several new<br />

findings about the recurrence of tsunami events in<br />

the Indian Ocean region have emerged. Monecke, et<br />

al. (2008) suggested that the three sediment coring<br />

samples at Meulaboh (west coast of Aceh), Simeulue<br />

(offshore west coast of Aceh) and Phra Thong (west<br />

coast of Thailand) show sediment deposit layers<br />

unconformity which age range are correlated with the<br />

historic tsunami occurrences in this region.<br />

Two sample units in Meulaboh and Phra Trong<br />

suggested the ranges of age of the deposit of AD<br />

1290 to 1400 (or 660 to 550 y BP) and AD 1300 to<br />

1450 (or 650 to 500 y BP), respectively (Monecke et<br />

al., 2008). In their report, they used the age<br />

estimates of coastal terraces in the Andaman Islands<br />

region by Rajendran, et al. (2008) to confirm their<br />

interpretation as evidence for subduction<br />

earthquakes. The problem with estimates made by<br />

Rajendran, et al. (2008) was that the range of age<br />

they provide was considerably large. The marine<br />

terrace of Andaman Islands was suggested to age of<br />

AD 1170-1600 and AD 550-1330 (or equal to 750-<br />

350 y BP and 1400-620 y BP). In this study, we<br />

produced an alternative proxy to that provides<br />

comparable confirmation to the analysis of great<br />

tsunami recurrence studied by Monecke, et al.<br />

(2008).<br />

In the 2004 tsunami, the inundation by seawater<br />

reached as far as 5 km inland, crossing the entire<br />

coastal plain of Banda Aceh that was already<br />

established since 3.5 ky BP (compare the tsunami<br />

inundation with the shoreline position of 2.8 ky BP in<br />

Fig. 1). It left deposits of various thickness and<br />

texture among the ruins and debris. The outer belt of<br />

beach ridge was mostly breached and even<br />

completely disappeared. After four days, the<br />

inundation reduced to about 50% from its initial<br />

extent (see also Dohmen-Janssen et al., 2006),<br />

which changed the low-lying housing areas, wetland<br />

and lagoon system behind the barrier islands into<br />

submerged areas. These areas are somehow<br />

associated with the one that geomorphologically<br />

modified by the last marine transgression (i.e.<br />

tsunami) in 0.6 ky BP. This shows that a huge<br />

tsunami event occurs only in a very short time-scale,<br />

but it leads to changes in geomorphology that has<br />

developed in a long-term scale, i.e. in centuries to<br />

millennia.<br />

The Banda Aceh coast in the recent pre-tsunami time<br />

consisted of a narrow sandy beach perched on top of<br />

the toe of the beach ridges from the previous coastal<br />

development (Pleistocene and Holocene<br />

morphological units). The preceding development<br />

has made the modern coast of Banda Aceh to be<br />

highly dissected within a 25-km coastal stretch.<br />

Figure 2 shows the five representative cross-shore<br />

shoreface profiles at Banda Aceh (P1, P2, P3, P4<br />

and P5 in Fig. 2). Each profile shows the elevation<br />

changes of three points in time within the recent<br />

century, i.e., before the tsunami (1893 and 1924),<br />

and after the tsunami (2006).<br />

elevation (m)<br />

(a)<br />

elevation (m)<br />

-50<br />

(b)<br />

elevation (m)<br />

(c)<br />

(d)<br />

elevation (m)<br />

(e)<br />

elevation (m)<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

0<br />

-10<br />

-20<br />

-30<br />

-40<br />

-50<br />

1893 1924 2006<br />

1893 1924 2006<br />

1893 1924 2006<br />

1893 1924 2006<br />

1893 1924 2006<br />

cross profile P1<br />

0 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 3600 3900 4200 4500 4800<br />

distance from shoreline (m)<br />

cross profile P2<br />

0 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 3600 3900 4200 4500 4800<br />

distance from shoreline (m)<br />

cross profile P3<br />

0 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 3600 3900 4200 4500 4800<br />

distance from shoreline (m)<br />

cross profile P4<br />

0 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 3600 3900 4200 4500 4800<br />

distance from shoreline (m)<br />

cross profile P5<br />

0 300 600 900 1200 1500 1800 2100 2400 2700 3000 3300 3600 3900 4200 4500 4800<br />

distance from shoreline (m)<br />

silty clay (soft)<br />

Sandy silty clay<br />

(medium stiff)<br />

Poorly graded sand<br />

(very loose)<br />

Sandy clayey silt<br />

(stiff)<br />

Sandy clayey silt<br />

(stiff)<br />

Silty sand (loose)<br />

Sandy clayey silt<br />

(very soft)<br />

Silty sand (loose)<br />

Silty clay (very soft)<br />

Sandy silt (very<br />

soft)<br />

Poorly graded sand<br />

(loose)<br />

Sandy silty clay<br />

(very soft)<br />

Thickness scale : = 1 meter deep<br />

Fig. 2: Shoreface profile development in the past century<br />

and the associated bore hole samples. The downward<br />

arrows point at the sampling location at a depth ca. 10 m<br />

from mean sea level using rotational machine bore drilling<br />

log technique. The right panels of each profile display the<br />

stratification of the samples. Profile locations are in Fig. 1.<br />

From the borehole samples shown in Fig. 2 it can be<br />

observed that most of the intercalated layers<br />

underneath the top layers of each profile on the<br />

southwestern part of the coast (P1 and P2) consist of<br />

stiff deposits. These layers were associated with the<br />

alluvial delta progradation during the last Holocene.<br />

Furthermore, we suppose that the intercalating layers<br />

found near the Sumatran Fault zone (e.g. P2) were<br />

influenced by some tectonic-related event (e.g. land<br />

subsidence or tsunami). Mobile fine sediments were<br />

subsequently perched on top of these Holocene<br />

deposits. The deposits associated with the<br />

prograding delta process consist of very fine<br />

sediments (e.g. silty clay or clayey silt) which<br />

eventually become stiff structure of deposit layers in<br />

134


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

a long time and becomes the basement of the<br />

subsequent development (marine regression).<br />

The shoreface profile of Banda Aceh in response to<br />

the tsunami can be seen in cross shore profiles in<br />

1924 and 2006 shown in Fig. 2. The amount of<br />

eroded shoreface sediments varied from one profile<br />

to another. Interestingly, each shoreface profile<br />

maintained its convexity or concavity upon being<br />

eroded by the tsunami. The upper layer of the<br />

shoreface consists of young loose sediments<br />

underlain by the more consolidated morphological<br />

units (see bore hole data in Fig. 2) which indicate the<br />

combination between alluvial and regression<br />

processes during the Holocene. As a result, the<br />

upper layer were easily eroded during the tsunami<br />

(e.g. soft clay or silty layer deposit at the upper<br />

shoreface of profile P1 and the entire profile P3),<br />

while the shoreface composed by older deposits<br />

were less eroded (e.g. profiles center part of profile<br />

P1, profile P2 and P4). The shoreface erosion has<br />

modified the shoreface profiles of the entire coast<br />

and therefore has eventually modified the<br />

hydrodynamics of the coastal system.<br />

DISCUSSION<br />

The shoreface profile changes in Figure 2 show that<br />

the convex shoreface was influenced by the predominant<br />

alluvial progradation during sea-level<br />

regression, whilst the concave shoreface was<br />

influenced by sea-level regression process during the<br />

Holocene. In addition, the resistance of the shoreface<br />

against severe erosion by the tsunami waves may<br />

depend on the degree of consolidation or the rate of<br />

weathering of the shoreface deposits. On the other<br />

hand tectonic intensity must be influential in the<br />

consolidation processes (e.g. liquefaction or<br />

compaction) on a specific coastal section preceding<br />

the December 2004 tsunami. In particular, tectonics<br />

and extreme waves such as tsunami have left their<br />

signature in the present geomorphology.<br />

Regarding the complexity of the processes involved<br />

and the resulting geomorphology, proper attention<br />

should therefore be given to the boundary conditions<br />

which are imposed by the development of older<br />

geomorphological units. The complex<br />

geomorphology underlying the Banda Aceh coastal<br />

plain resulted from various processes in history and<br />

ultimately determines the shoreface geomorphic<br />

settings and the morphological development,<br />

composition and texture of beach sediments, and the<br />

rates of shoreline position changes. This case<br />

demonstrated that it is essential to understand the<br />

geomorphic settings of a coast before attempting to<br />

model the large-scale behaviour of these types of<br />

coastal systems.<br />

This study was aimed to establish the general<br />

framework for future research on this specific study<br />

location which deserves attention to gain knowledge<br />

on the tectono-tsunami process and its impact on the<br />

coastal system. Borehole samples and reviews from<br />

other studies about this study area incorporated in<br />

the analysis in the present study fairly ascertain our<br />

interpretations on the geomorphic settings inherited<br />

from various morphological developments in history.<br />

Further study on the stratification of sediment deposit<br />

layers, carbon-dating, and side-looking radar may be<br />

utilized in future to support the overall interpretations<br />

of the present study.<br />

Acknowledgements: This study was conducted in<br />

the Water Engineering and Management Group of<br />

University of Twente, Enschede, The Netherlands, in<br />

collaboration with the International Institute for Geo-<br />

Information Science and Earth Observation (ITC),<br />

Enschede, The Netherlands. The project is co-funded<br />

by the Syiah Kuala University of Banda Aceh,<br />

Indonesia and the University of Twente, Enschede,<br />

The Netherlands. Data sets was provided by BRR-<br />

Nias/NAD Banda Aceh.<br />

References<br />

Dohmen-Janssen, C.M., Meilianda, E., Maathuis, B.H.P. &<br />

Wong, P.P. (2006). State of Banda Aceh beach before<br />

and after the tsunami. Proceeding of 30 th International<br />

Conference on Coastal Engineering, ASCE, Sand<br />

Diego, California.<br />

Geyh, M.A., Kudrass, H.R. & Streif, H., (1979). Sea-level<br />

changes during the late Pleistocene and Holocene in<br />

the Strait of Malacca. Nature 278(5703), 441 – 443.<br />

Hesp, P.A., Hung, C.C., Hilton, M. , Ming, C.L., Turner, I.M.,<br />

1998. A first tentative Holocene sea-level curve for<br />

Singapore. Journal of Coastal Research 14(1), 308-314.<br />

Meilianda, E., (2009). Past, present and future<br />

morphological development of a tsunami-affected coast:<br />

A case study of Banda Aceh. Ph D Thesis, University of<br />

Twente, Enschede, The Netherlands.<br />

Monecke, K., Finger, W., Klarer, D., Kongko, W., McAdoo,<br />

B.G., Moore, A.L. & Sudrajat, S.U., (2008). A 1000-year<br />

sediment record of tsunami recurrence in northern<br />

Sumatra. Nature (Letter) 455, 1232-1234.<br />

Rajendran, K., Rajendran, C.P., Earnest, A., Ravi Prasad,<br />

G.V., Dutta, K., Ray, D.K. & Anu, R., (2008). Age<br />

estimates of coastal terraces in the Andaman and<br />

Nicobar islands and their tectonic implications.<br />

Tectonophysics 455, 53-60.<br />

Scheffers, A. & Kelletat, D., (2003). Sedimentologic and<br />

geomorphologic tsunami imprints worldwide – a review.<br />

Earth-Science Reviews 63, 83-92.<br />

Sinsakul, S., 1990. Evidence of sea-level changes in the<br />

coastal area of Thailand: A review. In: <strong>Proceedings</strong><br />

IEM/ICE Joint Conference on Coastal Engineering in<br />

National Development (1991), pp. B1-B30.<br />

Tjia, H.D., (1992). Holocene sea-level changes in the<br />

Malay-Thai Peninsula, a tectonically stable<br />

environment. Geology Society of Malaysia 31, 157-176.<br />

Tjia, H. D. & Fujii, S., (1989). Late Quaternary shorelines in<br />

Peninsular Malaysia. In: Spafa Final Report Seminar in<br />

Prehistory of Southeast Asia 1987, 239–257.<br />

Verstappen, H. Th., (1973). A geomorphological<br />

reconnaissance of Sumatra and adjacent islands.<br />

Wolters-Noordhof Publication, Groningen.<br />

Woodroffe, S.A. & Horton, B.P., (2005). Holocene sea-level<br />

changes in the Indo-Pacific. Journal of Asian Earth<br />

Science 25, 29-43.<br />

135


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SEISMIC ANALYSIS OF LIQUID STORAGE TANKS<br />

Konstantin Meskouris, Britta Holtschoppen, Christoph Butenweg, Julia Rosin (1)<br />

(1) Chair of Structural Statics and Dynamics, Mies-van-der-Rohe Strasse 1, 52074 Aachen, GERMANY.<br />

Email: meskouris@rwth-aachen.de<br />

Abstract (Tank design for seismic loading): In the event of strong earthquakes, it is important that the structural integrity of<br />

tanks containing liquids is maintained in order to not jeopardize the population's supply with essential goods. The continuous<br />

operation of tanks after strong earthquakes requires safe and effective design rules. The overall seismic behaviour of tanks is,<br />

however, quite complex, since the dynamic interaction effects between tank wall and liquid must be considered. The interaction<br />

can be simplified with the concept of generalized single-degree-of-systems representing the convective, rigid impulsive and flexible<br />

impulsive vibration modes of tank and liquid. This concept is well accepted for anchored tanks with a fix connection to a rigid<br />

foundation. This paper presents the state of the art of tank design with special focus on the practicability of the available design<br />

rules. Analytical and numerical calculation approaches are compared on the example of a typical tank geometry, taken the<br />

relevant interaction effects into account.<br />

Key words: liquid filled tank, seismic design, impulsive vibration mode, critical facilities<br />

INTRODUCTION<br />

Recent earthquake events showed that heavy<br />

seismic damages of tanks may lead to environmental<br />

hazards, fire following earthquakes and temporary<br />

loss of essential facilities. The vibration of liquid filled<br />

tanks subject to seismic loading depends on the<br />

inertia of the liquid and on the interaction effects<br />

between the liquid and the tank shell. Different<br />

calculation methods are available for describing the<br />

vibration behaviour and the earthquake loads. These<br />

methods are either quite simple (Housner, 1963) or<br />

very accurate but complex (Fischer et al., 1991).<br />

Therefore a well comprehensive and feasible method<br />

is needed, that provides realistic results for the<br />

seismic behaviour of tanks with an acceptable<br />

computing time. The following considerations apply<br />

to cylindrical, anchored tanks with a fix connection to<br />

a rigid foundation.<br />

SEISMICALLY INDUCED LOAD COMPONENTS<br />

OF LIQUID FILLED TANKS<br />

The seismic loads acting on wall and bottom of<br />

cylindrical tanks (Figure 1) can be divided into the<br />

following components (Meskouris et al., 2010):<br />

- the convective load component; the fluid<br />

vibration in the rigid tank (sloshing),<br />

- the impulsive rigid load component; caused by<br />

the inertia of the liquid, if the rigid tank moves<br />

together with the foundation,<br />

- the impulsive flexible load component;<br />

representing the combined vibration of the<br />

flexibile tank shell (e.g. steel tanks) with the<br />

liquid.<br />

Fig. 1: Cylindrical tank<br />

Convective pressure<br />

Figure 2 shows the mode of vibration and the<br />

pressure distribution corresponding to the convective<br />

pressure component.<br />

The pressure distribution is defined as:<br />

∞<br />

p k<br />

(ξ, ζ, θ, t) = 2∙R∙ρ L<br />

(λ 2 J 1(λ n ∙ ξ)<br />

n -1) J 1 (λ n ) cosh(λ n∙ γ ∙ ζ)<br />

<br />

cosh(λ n ∙ γ)<br />

n=1 (1)<br />

⋅ [cos(θ)] [a kn (t) ∙ Γ kn ]<br />

with<br />

p k convective pressure component due to<br />

horizontal excitation<br />

n summation index; number of considered<br />

sloshing modes (here: n = 1)<br />

R inner tank radius<br />

ρ L liquid density<br />

136


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

J 1<br />

first order Bessel function:<br />

∞<br />

J 1 (λ n ∙ ξ) = <br />

(-1)k<br />

2k+1<br />

k! ∙ Γ(1 + k + 1) ∙ λ n ∙ ξ<br />

2 <br />

k=0<br />

λ n null derivation of Bessel function:<br />

λ 1 = 1,841, λ 2 = 5,331, λ 3 = 8,536<br />

ξ dimensionless radius: ξ = r/R<br />

ζ dimensionless height: ζ = z/H<br />

θ angle of circumference<br />

γ tank slenderness: γ = H/R<br />

a kn(t) horizontal acceleration-time history as a result<br />

of an equivalent single-degree-of-freedom<br />

system with a period T kn for the n th -eigenmode<br />

of the sloshing wave. By using the response<br />

spectrum analysis the spectral accelerations<br />

corresponding to the natural periods T kn<br />

should be calculated based on the elastic<br />

response spectrum.<br />

Γ kn participation factor for the convective pressure<br />

component for the n th -eigenmode.<br />

Fig. 2: Convective pressure - mode of vibration and<br />

pressure distribution<br />

Taking into account the first sloshing eigenmode<br />

(n = 1) and the pressure distribution of the tank shell<br />

(ξ = 1), equation (1) can be simplified to:<br />

cosh(1,841∙ γ ∙ ζ)<br />

p k<br />

(ξ = 1, ζ, θ, t) = R ∙ ρ L<br />

0,837 ∙ <br />

cosh(1,841∙ γ) (2)<br />

⋅ [cos(θ)][a k1 (t) ∙ Γ k1 ]<br />

ν n ν n = 2n + 1<br />

π<br />

2<br />

modified first order Bessel function:<br />

I 1<br />

I 1<br />

’<br />

I 1 ν n<br />

γ ∙ ξ = J 1 i∙ ν n<br />

γ<br />

∙ ξ<br />

i n<br />

∞<br />

ν<br />

2k+1<br />

nγ<br />

1<br />

∙ ξ<br />

= <br />

∙ <br />

k! ∙ Γ(1 + k + 1) 2 <br />

k=0<br />

Derivation of the modified Bessel function<br />

regarding to DIN EN 1998-4 (2007)<br />

I ν n<br />

γ ∙ξ = I 0 ν n<br />

γ ∙ξ - I 1 ν n<br />

γ<br />

∙ξ<br />

ν n<br />

γ<br />

∙ ξ<br />

ν<br />

2k+0<br />

∞<br />

n<br />

I 1<br />

∙ξ γ<br />

= <br />

k! ∙ Γ(0 + k + 1) ∙ 2 <br />

k=0<br />

∑<br />

∞<br />

k=0<br />

ν<br />

2k+1<br />

nγ<br />

1<br />

∙ ξ<br />

∙ <br />

k! ∙ Γ(1+k+1) 2 <br />

-<br />

ν n<br />

∙ ξ γ<br />

a is,h(t) horizontal acceleration-time history. By using<br />

the response spectrum analysis a is,h(t) should<br />

be replaced by the spectral acceleration S a<br />

corresponding to T = 0 s.<br />

a gR reference peak ground acceleration on type A<br />

ground<br />

S soil factor<br />

γ I importance factor according to DIN EN 1998-<br />

1 (2010) or DIN EN 1998-4 (2007)<br />

Γ is,h participation factor for the rigid impulsive<br />

pressure component: Γ is,h = 1,0, because the<br />

rigid tank is moving together with the<br />

foundation.<br />

The natural period T kn for the n th -eigenmode of the<br />

sloshing wave is calculated with:<br />

2π<br />

T kn =<br />

g ∙ λ n ∙ tanh (λ n∙ γ)<br />

(3)<br />

R<br />

Rigid impulsive pressure<br />

Figure 3 shows the mode of vibration and the<br />

pressure distribution corresponding to the rigid<br />

impulsive pressure component.<br />

The pressure distribution is given by the expression:<br />

∞<br />

p is,h<br />

(ξ, ζ, θ, t) = 2 ∙ R∙ γ∙ ρ L ∙(-1)n<br />

with:<br />

p is,h<br />

I 1 ν n<br />

∙ ξ γ<br />

<br />

<br />

γ<br />

ν2<br />

n I ν n<br />

n=0 (4)<br />

⋅ [cos(ν n ∙ ζ)][cos(θ)]a is,h (t) ∙ Γ is,h <br />

rigid impulsive pressure component due to<br />

horizontal excitation<br />

Fig. 3: Rigid impulsive pressure - Mode of vibration<br />

and pressure distribution<br />

Taking into account the pressure distribution of the<br />

tank shell (ξ = 1), equation (4) is simplified to:<br />

p is,h<br />

(ξ = 1, ζ, θ, t)= R∙ρ L<br />

2 ∙ γ ∙ (-1)n<br />

∞<br />

n=0<br />

ν n<br />

2<br />

I 1 ν n<br />

γ<br />

<br />

I ν n<br />

γ<br />

cos(ν n∙ζ)<br />

⋅ [cos(θ)]a is,h (t)∙Γ is,h (5)<br />

The corresponding natural period is T = 0.<br />

137


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Flexible impulsive pressure<br />

Figure 4 shows the mode of vibration and the<br />

pressure distribution corresponding to the flexible<br />

impulsive pressure component.<br />

The flexible impulsive pressure component is<br />

calculated in an iterative procedure using the addedmass-model<br />

according to DIN EN 1998-4 (2007),<br />

Annex A. Within the framework of the procedure the<br />

tank wall is loaded with iterative calculated additional<br />

mass portions of the activated fluid. The pressure<br />

distribution is given by the expression:<br />

p if,h<br />

(ξ, ζ, θ, t) =<br />

∞<br />

I 1 ν n<br />

∙ ξ γ<br />

2R ρ L<br />

ν n<br />

∙ I γ ν n cos(ν n∙ ζ) f(ζ)∙cos(ν n ∙ ζ) dζ<br />

n=0<br />

γ 0<br />

with:<br />

p if,h<br />

f(ζ)<br />

1<br />

⋅ [cos(θ)] a if,h (t) ∙ Γ if,h <br />

flexible impulsive pressure component due to<br />

horizontal excitation<br />

deflection curve of the first (anti-symmetric)<br />

mode of oscillation of the tank-fluid<br />

combination<br />

a if,h(t) horizontal acceleration-time history as a result<br />

of an equivalent single-degree-of-freedom<br />

system. By using the response spectrum<br />

analysis the spectral accelerations<br />

corresponding to the first natural period T if,h,1<br />

should be used.<br />

Γ if,h<br />

(6)<br />

participation factor for the flexible impulsive<br />

pressure component due to horizontal<br />

excitation<br />

F(y) correction factor: F(γ) = 0,157 ∙ γ 2 + γ + 1,49<br />

s(ζ = 1/3) wall thickness of the tank at 1/3 filling<br />

height<br />

Fig. 4: Flexible impulsive pressure - Mode of vibration<br />

and pressure distribution.<br />

The function curve of f(ζ) in (8) is generally not<br />

known. It depends on the impact of the liquid onto<br />

the tank, in other words the aforementioned pressure<br />

function p if.h. Thus the joint bending form must be<br />

correctly determined iteratively. In DIN EN 1998-4<br />

(2007), Annex A the "added-mass concept" is<br />

proposed. According to this the resonating fluid,<br />

activated with the first bending shape, is added to the<br />

tank wall density. Then with the new "dry" tank<br />

model, the more accurate bending form is<br />

determined (Fischer et. al., 1991).<br />

The participation factor Γ if,h for the flexible impulsive<br />

pressure component is calculated as follows:<br />

Γ if,h =<br />

∫ 1<br />

0<br />

p if,h (ζ)dζ<br />

1<br />

∫ f(ζ) ∙ p if,h<br />

(ζ)dζ<br />

0<br />

, s(ζ) = const. (7)<br />

with:<br />

p if,h (ζ) pressure function of the flexible impulsive<br />

pressure component as a function of the filling<br />

height<br />

s(ζ) wall thickness of the tank<br />

Taking into account the pressure distribution of the<br />

tank shell (ξ = 1), equation (6) can be simplified to:<br />

p if,h<br />

(ξ = 1, ζ, θ, t) =<br />

∞<br />

I 1 ν n<br />

γ<br />

= R∙ρ L<br />

2∙ ν nγ I ν n ∙cos(ν n∙ζ) f(ζ)∙cos(ν n ∙ζ) dζ<br />

n=0 γ 0<br />

1<br />

[cos(θ)]a if,h (t) ∙ Γ if,h <br />

(8)<br />

Fig. 5: Iteration process (Holtschoppen et. al., 2011)<br />

The first natural period T if,h,1 is calculated as follows:<br />

W L<br />

T if,h,1 = 2 ∙ F(γ)<br />

π ∙ g ∙ E ∙ s(ζ = 1/3)<br />

with:<br />

W L<br />

H ∙ρ<br />

= 2 ∙ R ∙ F(γ)<br />

L<br />

E ∙ s(ζ = 1/3)<br />

fluid weight: W L = π ∙ R 2 ∙ H ∙ ρ L<br />

∙ g<br />

(9)<br />

However as Figure 5 shows, even this iterative<br />

process is impractical because the calculation is<br />

based on a complex mathematical pressure function<br />

(6) which requires a coupling of a mathematical<br />

software tool with a finite element program.<br />

Though, with comprehensive parameter studies it is<br />

shown that the bending form of any tank can be<br />

described by using a parameterized sine wave which<br />

138


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

can be mapped correctly to the natural frequencies<br />

for all common geometric and material configurations<br />

(Cornelissen, 2010):<br />

f(ζ)=a⋅ sin π ⋅(ζ-b)⋅c+d (10)<br />

2<br />

The bending form of the combined vibration – and<br />

with this the parameters a, b, c, d – depends on<br />

- the tank slenderness (γ = H/R),<br />

- the Poisson’s ratio ν,<br />

- the ratio fluid mass to tank mass and<br />

- a changing wall thickness along the tank heigh.<br />

However, the last named influences are<br />

comparatively small. By using the sine function (10) it<br />

is possible to determine the flexible impulsive<br />

pressure without iteration. For practical use and to<br />

represent the pressure the equation<br />

p if,h<br />

(ξ,ζ,θ,t)=R⋅ρ L<br />

⋅ cos(θ) ⋅Γ if,h ⋅a if,h (t)⋅C if,h (ξ,ζ) (11)<br />

is suitable. C if,h(ξ,ζ) corresponds to the normalized<br />

pressure at θ=0. Figure 6 shows the variation of the<br />

factor C if,h(ξ=1,ζ) for the tank shell (ξ = 1) for different<br />

tank slendernesses.<br />

For simplicity, the formulations of all pressure<br />

functions are also provided in a standardized,<br />

tabulated form (Meskouris et. al., 2011).<br />

Fig. 6: Normalized flexible impulsive pressure C if,h.<br />

The participation factor for the flexible impulsive<br />

pressure Γ if,h, which is specified in equation (7), can<br />

be tabulated as well, assuming that the tank wall is<br />

constant and the tank mass is insignificant.<br />

CONCLUSION<br />

This paper provides guidance for the implementation<br />

of normative demands for the seismic design of liquid<br />

filled tanks. With the tabulation of different factors the<br />

cumbersome mathematical formulas for calculating<br />

the load components are avoidable, which allows an<br />

easy load generation for finite element analysis.<br />

References<br />

Cornelissen, P., (2010). Erarbeitung eines vereinfachten<br />

impulsiv-flexiblen Lastansatzes für die Berechnung von<br />

Tankbauwerken unter Erdbebenlast. Diplomarbeit,<br />

Fakultät für Bauingenieurwesen, RWTH Aachen<br />

DIN EN 1998-1, (2010). Auslegung von Bauwerken gegen<br />

Erdbeben – Teil 1: Grundlagen, Erdbebeneinwirkungen<br />

und Regeln für Hochbauten, Deutsche Fassung EN<br />

1998-1:2004+AC:2009, Deutsches Institut für Normung<br />

(DIN), Berlin<br />

DIN EN 1998-4, (2007). Auslegung von Bauwerken gegen<br />

Erdbeben – Teil 4: Silos, Tanks und Pipelines, Deutsche<br />

Fassung EN 1998-4:2006, Deutsches Institut für<br />

Normung (DIN), Berlin<br />

Fischer, F.D., Rammerstorfer, F.G., Scharf, K., (1991).<br />

Earthquake Resistant Design of Anchored and<br />

Unanchored, Liquid Storage Tanks under Three-<br />

Dimensional Earthquake Excitation. In: Structural<br />

Dynamics (G.I.). Springer Verlag, 317-371, ISBN 3-540-<br />

53593-4<br />

Holtschoppen, B., Cornelissen P., Butenweg, C., Meskouris,<br />

K., (2011). Vereinfachtes Berechnungsverfahren<br />

zur Berücksichtigung der Interaktionsschwingung bei<br />

flüssigkeitsgefüllten Tankbauwerken unter seismischer<br />

Belastung. Baustatik Baupraxis 11, Institut für Baustatik,<br />

Technische <strong>Universität</strong> Graz, ISBN 978-3-85125-115-9<br />

Housner, G.W., (1963). The dynamic behaviour of water<br />

tanks, Bulletin of the Seismological Society of America,<br />

Vol. 53, 381-387<br />

Meskouris, K., Holtschoppen B., Butenweg, C., Park, J.,<br />

(2010). Seismische Auslegung von Silo- und<br />

Tankbauwerken. Festschrift für Prof. Dr.-Ing. Mangering,<br />

<strong>Universität</strong> der Bundeswehr München<br />

Meskouris, K., Hintzen, K.-G., Butenweg, C., Mistler, M.,<br />

(2011). Bauwerke und Erdbeben. Vieweg und Teubner<br />

Verlag<br />

139


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

GEOLOGICAL CRITERIA FOR EVALUATING SEISMICITY: LESSONS LEARNED FROM<br />

THE PO PLAIN, NORTHERN ITALY<br />

Michetti, Alessandro M. (1, Leonello Serva (1), Andrea Berlusconi (1, Livio Bonadeo (1, Fabio Brunamonte (1, Francesca<br />

Ferrario (1Gianfranco Fioraso (2), Franz Livio (1, Giancanio Sileo (1), Eutizio Vittori (3)<br />

(1) Dipartimento di Scienze Chimiche e Ambientali, Università dell’Insubria, Via Valleggio 11, 22100, Como. ITALY. Email:<br />

alessandro.michetti@uninsubria.it<br />

(2). CNR, Istituto di Geoscienze e Georisorse, Via Valperga Caluso, 35,10125, Torino. ITALY. E-mail: g.fioraso@csg.to.cnr.it<br />

(3). Dipartimento Difesa del Suolo, Servizio Geologico d’Italia, ISPRA, Via V. Brancati 48, 00144 Roma, ITALY. Email:<br />

eutizio.vittori@isprambiente.it<br />

Abstract (Geological criteria for evaluating seismicity: lessons learned from the Po Plain, N Italy): We compare available<br />

literature and new geological and paleoseismological data, in particular from the Monte Netto - Brescia site, to maintain that A)<br />

strong earthquakes similar to the intensity IX to X MCS - M6.5 to 7 - historical events occurred in 1117 and 1222, though very rare,<br />

must be considered unlikely but credible events along virtually all the Quaternary faults throughout the whole Po Plain Foredeep,<br />

at least until we learn from geological and geophysical studies how/if the Quaternary tectonic structures near Lake Garda are truly<br />

different from those in other areas such as, for example, the Turin Hills, Monferrato, Insubria or around Piacenza, B) the late<br />

Quaternary (and especially Holocene) history of deformation and the local seismic landscape are far more valuable tools in<br />

estimating seismic hazard than has generally been appreciated until now in Italy – as well as in most parts of the world. The lack of<br />

significant earthquakes for 8 centuries in the Garda region and for longer elsewhere must prompt an effort to much deeper<br />

understanding of the actual seismic potential in one of the most populated and economically developed areas of Europe.<br />

Key words: Paleoseismology, late Quaternary tectonics, Po Plain, Seismic Hazard<br />

INTRODUCTION<br />

In the field of paleoseismology and earthquake<br />

hazard assessment, the year 2011 has been<br />

characterized by two “unexpected” destructive<br />

events, the Christchurch eq. in New Zealand and the<br />

Tohoku eq. in Japan. In particular, the devastating<br />

impact from the March 11, 2011, M9 Tohoku<br />

coseismic environmental effects – and of course in<br />

particular the massive tsunami – is definitely bound<br />

to deeply influence the standards for seismic hazard<br />

assessment worldwide for decades.<br />

In fact, these events have clearly shown that “the<br />

most important contribution to the understanding of<br />

long term seismicity, which is critical to the siting and<br />

design of safe structures and to the establishment of<br />

realistic building codes, is to learn more – region by<br />

region – of the late Quaternary history of<br />

deformation”. This of course includes the evidence<br />

for paleoseismicity. This is a quote from Allen (1975),<br />

a paper that has inspired also the title of this note. A<br />

retrospective look to the origin of the science of<br />

Paleoseismology is surprisingly instructive today,<br />

while we are still shocked by the effects of the 2011<br />

catastrophic seismic crises.<br />

The purpose of the present note is to argue that the<br />

approach discussed by Allen (1975), and further<br />

developed – among others – by Serva (1996) and<br />

Michetti et al. (2005) with the introduction of the<br />

notion of seismic landscape, is definitely validated<br />

not only by the investigations conducted before and<br />

after the 2011 events in New Zealand and Japan,<br />

which are seismically very active countries, but also<br />

by the lessons learned during the 4 decades of<br />

seismotectonic studies in the Po Plain, in Northern<br />

Italy, one of the most populated and developed areas<br />

of Europe, hosting a substantial portion of the Italian<br />

industrial production, many infrastructures and a<br />

number of high risk plants.<br />

From historical seismicity studies we know that the<br />

Po Plain, near Lake Garda, has been hit by two<br />

strong earthquakes in the XII – XIII centuries, the<br />

January 13, 1117, Verona, and Christmas 1222,<br />

Brescia, events, both in the range of intensity IX-X<br />

MCS, equivalent to M6.5 to 7 (Fig. 1). No other<br />

comparable earthquakes have occurred since than in<br />

the Central-Western Po Plain region.<br />

We use available literature and new geological and<br />

paleoseismological data to maintain that A) strong<br />

earthquakes similar to the Verona and Brescia must<br />

be considered credible events – though unlikely –<br />

along virtually all the Quaternary faults throughout<br />

the Po Plain Foredeep, at least until we understand<br />

from geological and geophysical studies if/how the<br />

Quaternary tectonic structures near Lake Garda are<br />

truly different from those in other areas such as in the<br />

Turin Hills, Monferrato, Insubria or around Piacenza<br />

(Figs. 1 and 2), B) the late Quaternary (and<br />

especially Holocene) history of deformation and the<br />

local seismic landscape are far more valuable tools in<br />

estimating seismicity and associated seismic hazard<br />

than has generally been appreciated until now in Italy<br />

- like in most parts of the world, and C) the “absence<br />

of evidence” for historical or instrumental strong/large<br />

140


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 1: Historical and instrumental earthquakes catalogue (modified from CPTI, 2004) and map of Quaternary capable faults in the<br />

Po Plain (after Livio et al., 2009). SP, Superga; MF, Monferrato; SV, Spina Verde; SL, San Colombano al Lambro; MN, Monte<br />

Netto; CV, Ciliverghe; CS, Castenedolo; MR, Mirandola; SO, Soncino.<br />

seismic events within a region should never be<br />

construed as “evidence of absence” of seismic<br />

hazard, unless systematic and specific paleoseismic<br />

analyses have been conducted in that region.<br />

Indeed, in the past decade our knowledge about the<br />

magnitude and rates of the key geomorphic agents<br />

which control the late Quaternary landscape<br />

evolution of the Po Plain and surrounding piedmont<br />

belts (along the margins of the Apennines to the S,<br />

and the Southern Alps to the N), has greatly<br />

improved. In particular, due to the large amount of<br />

novel information on the active tectonics and seismic<br />

potential of this region, it has become increasingly<br />

clear the role played by strong seismic events in this<br />

process.<br />

As already pointed out in the literature (e.g., Desio,<br />

1965; Carraro et al., 1995), both the Apennines and<br />

the Southern Alps piedmont belts are characterized<br />

by Quaternary tectonic features, i.e. drainage<br />

anomalies, isolated hills (e.g., Superga, Spina Verde,<br />

Fig. 2: Digital Elevation Model of the Superga and Monferrato Hills (vertical exaggeration 5x).<br />

141


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

San Colombano, Figs. 1 and 2), and buried structural<br />

highs (e.g., Soncino, Mirandola), evidence of the<br />

Quaternary growth of the two mountain belts beneath<br />

the Po Plain. However, only in the past decade,<br />

thanks to the integrated research conducted by the<br />

University of Insubria and partners (ISPRA-<br />

Geological Survey of Italy, ENI E&P, UCL London,<br />

Colorado University-Boulder, Innsbruck University,<br />

Università Statale di Milano, Università di Brescia,<br />

INGV, CNR Torino, Regione Lombardia) it has been<br />

possible to study the first paleoseismic site showing<br />

evidence for repeated latest Pleistocene to Holocene<br />

compressional surface faulting earthquakes along the<br />

Monte Netto Backthrust, Brescia (Livio et al., 2009;<br />

Figs. 3 and 4). At this site, 3 paleoseismic events<br />

have been recorded in the past ca. 40 kyr B.P. Based<br />

on trench data and geomorphic investigations, the<br />

magnitude estimated for these events is consistent<br />

with the earthquake size of the Christmas 1222<br />

Brescia event (M6.2 to 6.8; e.g., Serva, 1990;<br />

Guidoboni & Comastri, 2005).<br />

It is important to note that this result has been<br />

possible only due to the availability of extensive oil<br />

exploration subsurface data (seismic reflection and<br />

deep stratigraphic boreholes, courtesy of ENI E&P);<br />

and large quarry excavations at Monte Netto site,<br />

which provided outstanding exposures of late<br />

Quaternary growth anticlines, bending moment<br />

surface faulting, and coseismic liquefaction (data and<br />

site access courtesy of Fornaci Laterizi Danesi SpA).<br />

In fact, the Monte Netto structure is a unique case<br />

study in the Po Plain. Which is logic in a sense, due<br />

to i) the moderate seismicity level of this tectonic<br />

province, ii) the local high erosional and depositional<br />

rates compared to slip rates of capable faults, iii) the<br />

need for expensive investigations (seismic reflection<br />

profiles, boreholes, trenching, dating) due to the<br />

reverse style of faulting, and a geomorphic and<br />

human environment unfavorable to paleoseismic<br />

analyses.<br />

Literature data and our ongoing research strongly<br />

suggest that several Quaternary structures in the Po<br />

Plain share the same tectonic and geomorphic<br />

features of the Monte Netto Backthrust. However,<br />

virtually no comparable paleoseismic information and<br />

detailed structural characterization is available for<br />

these structures.<br />

DISCUSSION<br />

We are systematically investigating the structural<br />

features and evidence for capability along the<br />

Quaternary faults in Figure 1; to be scientifically<br />

prudent, at present we cannot rule out that many of<br />

these faults are able to generate earthquakes similar<br />

to the Verona and Brescia Middle Age events. So, in<br />

view of the difficulties in interpreting the historic<br />

record from Middle Age events, and the lack of<br />

detailed paleoseismic information for most of the<br />

structures showing some evidence for capability, we<br />

stress the need to be exceedingly conservative in<br />

estimating the probability of major damaging<br />

earthquakes in the Po Plain. “We have been<br />

surprised too often in the past, and we cannot afford<br />

to be surprised too many times in the future” (Allen,<br />

1975); this must be the case for the Po Plain, point<br />

made even more clear after the Japan and New<br />

Zealand experience in 2011, where, it must be said,<br />

the basic knowledge was seemingly much better.<br />

The Po Plain hosts one of the most industrially<br />

developed areas in the world, and like in many other<br />

“megacities” located in seismically active regions,<br />

every year the seismic risk keeps growing up, with<br />

more and more serious societal effects to be<br />

expected. The Po Plain has not been the subject of<br />

an intensive seismotectonic research effort since the<br />

end of the Italian Nuclear Program more than 20<br />

years ago (e.g., Serva, 1990). The state of the art in<br />

paleoseismology, and the seismic events of 2011,<br />

demonstrate that our conservatism is scientifically<br />

sound. The lack of events in the last 8 centuries in<br />

Fig. 3: Decametric Late Quaternary growth anticline at the Monte Netto site. Note the gravity graben due to due to bending<br />

moment normal faulting on the fold crest; new exploratory trenching across this graben revealed 3 paleoseismic surface faulting<br />

events in the past 40 kyr B.P. The coseismic liquefaction shown in Figure 4 is located near the core of the anticline, indicated by<br />

the person for scale.<br />

142


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

the Garda region and longer (how much?) elsewhere<br />

must not indulge us to optimism, but, on the contrary,<br />

warn us to get prepared, how unlikely a M6-7 seismic<br />

event can be. As a matter of fact, siting and design of<br />

safe structures, and realistic building codes for the<br />

Po Plain should be based on a much better<br />

understanding of the long term seismicity, that is on<br />

the detailed knowledge of the late Quaternary history<br />

of deformation and Holocene paleoseismicity. This is<br />

the kind of knowledge revealed by the new data<br />

collected at the Monte Netto site, as of yet basically<br />

the only paleoseismic site available in the region.<br />

Fig. 4: Monet Netto site, liquefaction feature near the<br />

core of the anticline described in Figure 3.<br />

Acknowledgements: This work has been partly funded by<br />

Project S1 INGV-DPC 2007-2009; and by grants from the<br />

Operational Programme Cross Border Cooperation IT / CH<br />

2007-2013 – project “SITINET: census, networking and<br />

development of geological and archaeological sites” ID<br />

7621984.<br />

References<br />

Allen, C., (1975), Geological criteria for evaluating<br />

seismicity. Geological Society of America Bulletin; 86 (8),<br />

1041-1057; DOI: 10.1130/0016-7606<br />

Desio, A., (1965), I rilievi isolati della Pianura Lombarda ed i<br />

movimenti tettonici del Quaternario. Rendiconti Istituto<br />

Lombardo Accademia Scienze e Lettere, Sez. A, 99, 881-<br />

894.<br />

Carraro, F., G. Collo, M.G. Forno, M. Giardino, F. Maraga,<br />

A. Perotto & D. Tropeano, (1995), L’evoluzione del<br />

reticolato idrografico del Piemonte centrale in relazione<br />

alla mobilità quaternaria. In: R. Polino & R. Sacchi, Eds.,<br />

Atti del Convegno Rapporti Alpi-Appennino e guide alle<br />

escursioni, Peveragno (CN), 31 maggio - 1 giugno 1994,<br />

Accademia Nazionale delle Scienze detta dei 40, 1995 -<br />

XII, 441-465.<br />

Guidoboni, E. & A. Comastri, (2005), Catalogue of<br />

earthquakes and tsunamis in the Mediterrean area from<br />

the 11th to the 15th century. 1037 pp., SGA Storia<br />

Geofisica Ambiente srl, Bologna.<br />

Livio, F., A. Berlusconi, A.M. Michetti, G. Sileo, A. Zerboni,<br />

L. Trombino, M. Cremaschi, K. Mueller, E. Vittori, C.<br />

Carcano, & S. Rogledi, (2009), Active fault-related folding<br />

in the epicentral area of the December 25, 1222 (Io = IX<br />

MCS) Brescia earthquake (Northern Italy):<br />

seismotectonic implications. Tectonophysics, 476 (1-2),<br />

320-335, doi:10.1016/j.tecto.2009.03.019<br />

Michetti A.M., F. Audemard & S. Marco, (2005), Future<br />

trends in paleoseismology: Integrated study of the<br />

seismic landscape as a vital tool in seismic hazard<br />

analyses. Tectonophysics, 408 (1-4), 3-21.<br />

Serva, L., (1990). Il ruolo delle Scienze della Terra nelle<br />

analisi di sicurezza di un sito per alcune tipologie di<br />

impianti ad alto rischio. Il terremoto di riferimento per il<br />

sito di Viadana (MN). Bollettino Società Geologica<br />

Italiana, 109 (3), 375-411.<br />

Serva, L., (1996), Criteri geologici per la valutazione della<br />

sismicità: considerazioni e proposte. Atti dei Convegni<br />

Lincei 122, “Terremoti in Italia”, Accademia Nazionale dei<br />

Lincei, 1-2 Dicembre 1994, Roma, pp. 103-116.<br />

143


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ANCIENT SEISMITES AS GEODYNAMICAL INDICATOR: APPROACH TO CONSTRUCT A<br />

REACTIVATION EVENT ON THE MAIN BOUNDARY THRUST IN THE HIMALAYAN<br />

REGION<br />

Mishra, Anurag (1), D.C. Srivastava (2), Jyoti Shah (3)<br />

(1) Department of Earth Sciences, Indian Institute of Technology, Roorkee, Haridwar, Uttarakhand-247667.<br />

Email: anurag2009iitr@gmail.com<br />

Abstract: Diversified soft-sediment deformation structures exposed in the nearest vicinity of the Main Boundary Thrust in the<br />

South-eastern Kumaun Himalaya, show progressive increase in abundance and complexity towards the thrust. After establishing<br />

liquidization as dominant deformation mechanism, a whole to part study has been done to identify most possible triggering<br />

process, excluding all the other probable processes, integrating criteria based approaches in combination. The structures show<br />

their development due to seismic origin, hence called as seismites, define a clear relationship with the Main Boundary Thrust.<br />

Magnetostratigraphic dates of the associated sediments show 4-5 Ma as the age of development of the structures. Hence these<br />

structures, acting as geodynamical indicator, show paleoseismicity of the area, and dates back a reactivation event on the Main<br />

Boundary Thrust in the Kumaun Himalaya around 4-5 Ma, when established age for the formation of this thrust is around 11 Ma.<br />

Key words: Soft-sediment deformation structures, seismites, Main Boundary Thrust<br />

INTRODUCTION<br />

The Alpine-Himalayan orogeny originating in the late<br />

period of the Mesozoic era led to development of a<br />

series of orogen scale detachment boundaries,<br />

named as the South Tibet Detachment Zone, the<br />

Main Central Thrust, the Main Boundary Thrust and<br />

the Main Frontal Thrust (Fig. 1a) from North to South<br />

(Jackson and Bilham, 1994).<br />

The Main Boundary Thrust gains importance due to<br />

recent active seismicity of the thrust causing<br />

catastrophic earthquakes and accommodation of<br />

large scale crustal shortening along the thrust<br />

(Nakata, 1989). But the dating of the Main Boundary<br />

Thrust has not been given much consideration in<br />

previous works in comparison to all the other<br />

boundary thrusts, perhaps due to poor exposures<br />

and lack of cross-cutting relationships. The available<br />

dates for the formation of thrust do not converge<br />

perhaps due to ambiguous relation between the<br />

activity of the thrust and applied methods, such as<br />

one of the date suggest it’s development before 11<br />

Ma whereas another younger than 5 Ma (Yin, 2006).<br />

Most of the available dates for the activity of the Main<br />

Boundary Thrust are largely based on study of<br />

change in depositional pattern in the Himalayan<br />

Foreland and study of exhumation based on rate of<br />

subsidence of the Himalayan Foreland. It is<br />

presumed that exhumation is directly incorporated<br />

with the uplift in the hanging wall of the Main<br />

Boundary Thrust. But the Main Boundary Thrust is<br />

Southerly of the Thrusts cutting through the<br />

Himalaya, and exhumation can be related with any of<br />

the Northerly thrust cutting through the rocks.<br />

The soft-sediment deformation structures of seismic<br />

origin are called as seismites (Seilacher, 1969). They<br />

have been used for understanding geodynamic<br />

activity of causative seismogenic faults and<br />

paleoseismicity of an area.<br />

Fig. 1: (a) Litho-tectonic units and demarcating boundaries<br />

in the Himalayan terrain. (b) Satellite imagery showing<br />

Himalayan terrain, location of Kumaun Himalaya and our<br />

study area.<br />

We use the soft-sediment deformation structures<br />

exposed in the vicinity of the Main Boundary Thrust<br />

for constraining one of the movement events along<br />

the thrust, in case they are seismites and show an<br />

undisputed relationship with the Main Boundary<br />

144


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Thrust. We follow a 3 fold-process: To classify these<br />

structures, to understand their origin in terms of<br />

deformation mechanism, driving force system and<br />

triggering agent, and finally finding out possible<br />

relationship with the Main Boundary Thrust.<br />

OBSERVATIONS<br />

The convolute laminations developed in medium to<br />

fine grained sandstone lithology (Fig. 3). Folded<br />

laminations, thicker at the crest and thinner at the<br />

trough show that sediments became hydro-plastic in<br />

nature during development of these structures (Fig.<br />

4). At the crest, the laminations are blurred and show<br />

homogenization of sediments (Fig. 4).<br />

A large variety of the soft-sediment deformation<br />

structures are abundantly exposed in the vicinity of<br />

the Main Boundary Thrust in the southeastern<br />

Kumaun Himalaya (Fig. 1b). The structures identified<br />

in the area are the deformed cross-stratifications, the<br />

liquefaction pockets, the slump folds, the mushroom<br />

structures, the convolute laminations, the sand<br />

dykes, the synsedimentary faults, the fluid-escape<br />

structures and the flame structures.<br />

Fig.4: Convolute lamination in relatively medium grained<br />

sandstone. Arrows show blurring of laminations and<br />

homogenization of sediments.<br />

Fig. 2: Recumbent cross-stratification in sandstone. Arrows<br />

point to minor faults cutting through the liquefied<br />

laminations. Note the fluid-escape structure in the rightupper<br />

part originating from the liquefied bed below.<br />

Fig. 5. Liquefaction pocket in sandstone. Arrows mark the<br />

offset in laminations.<br />

Liquefaction pockets are developed due to complete<br />

homogenization of sediments and obliteration of<br />

primary structures (Fig. 5) by liquefaction within the<br />

local pockets of few centimeters (Lowe, 1975).<br />

Laminations are offset along these pockets (Fig. 5).<br />

Fig. 3: Typical convolute laminations in sandstone<br />

The recumbent cross-stratifications show well<br />

preserved laminations and thickness of individual<br />

laminae varies throughout the structure (Fig. 2). Most<br />

plausible explanation for their development is<br />

deformation of liquefied sediments due to current<br />

drag, as foresets are folded towards paleocurrent<br />

direction (Allen and Banks, 1972, Allen, 1982). The<br />

minor faults are developed due to cohesive nature of<br />

sediments, when grain to grain packing established<br />

after liquefaction (Owen, 1987, 1996).<br />

Sand dykes developed due to intrusion of coarsegrained<br />

sandstone upward, originated by drag force<br />

of fluids (Fig 6 and Fig. 7). Their origin is well<br />

established and related to fluidization.<br />

Mushroom structures developed due to fluidization in<br />

the upward regime, called as elutriation, by which a<br />

plume type (Fig. 8) of forceful injection took place,<br />

aligning clay minerals in the neck connected to the<br />

clay enriched dome shaped main body (Lowe, 1975,<br />

Allen, 1982).<br />

145


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

body, and these structures are confined within<br />

undeformed horizons.<br />

Fluid escape structures (Fig. 9) are developed due to<br />

expulsion of fluids to surrounding beds of same<br />

comparable lithology due to any of the processes that<br />

causes extraction of fluids from the beds such as<br />

liquefaction, fluidization or compaction of sediments.<br />

Fig. 6: Sand dyke in Sandstone. Note the alignment of<br />

primary laminations towards the direction of injection.<br />

Fig. 9. Fluid escape structures in sandstone<br />

Fig. 7. Multiple sand dyke system in medium-grained<br />

sandstone. Note the reorientation of parallel laminations<br />

due to drag of upward sand injection. A synsedimentary<br />

fault is cutting through the parallel laminated sandstone and<br />

sand dyke.<br />

Fig. 10. Syn-sedimentary faults in sandstone. Notice the<br />

geometry of faults.<br />

Several variably oriented planar or listric<br />

synsedimentary faults (Fig. 10) developed that cut<br />

through the soft-sedimentary deformation structures.<br />

Compaction may be one of the plausible<br />

mechanisms for the development of these structures,<br />

but observing the insufficient thickness of<br />

overburden, large scale offset of many faults and<br />

superposition on earlier phases of fluidized<br />

structures, seismic activity seems to be most<br />

possible cause.<br />

Fig. 8. Two classes of mushroom structures in medium<br />

grained sandstone. In left photograph, the upper domal part<br />

is detached from the main body, whereas in the right one it<br />

is connected through a narrow neck. Notice the darkening<br />

of structure in compared to surrounding lithology and<br />

warping and folding of surrounding laminations (shown by<br />

arrow)<br />

Their development indicates requirement of instant<br />

force, so that elutriation and sudden expulsion could<br />

take place. The best possible originating force for<br />

these structures is seismic shaking, as there is no<br />

difference in grain size between injecting and injected<br />

DISCUSSION<br />

Liquefaction and fluidization are the most dominant<br />

deformation mechanisms for development of these<br />

structures. Triggers such as groundwater movement,<br />

rapid sediment deposition, storm waves, meteoritic<br />

impact and earthquake shocks can be responsible for<br />

the development of the soft-sediment deformation<br />

structures (Van Loon and Brodzikowski, 1987). We<br />

follow the criteria given by Sims (1973), Owen and<br />

Moretti (2011) and Owen et al. (2011) for identifying<br />

the trigger of these structures. These structures are<br />

confined within rocks of varied lithology such as, in<br />

several varieties of sandstone and in mudstonemarlstone<br />

beds. This indicates the probability of out<br />

146


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

of system, allogenic trigger for the development of<br />

these structures. Rapid sedimentation cannot be a<br />

cause of development as magnetostratigraphic<br />

dating reveals same rate of sedimentation throughout<br />

the area (Kotalia et al., 2001). Storm waves can be<br />

neglected due to established fluviatile origin of these<br />

sediments (Tandon, 1976). We suggest seismic<br />

origin, after excluding all the probable triggers. Our<br />

inference gets more strength as it fits on criteria<br />

based approaches and gives several evidences for<br />

approval, given in the following lines:<br />

1) The structures confined within the same<br />

stratigraphic horizons, follow these beds for<br />

a sufficient lateral extent of about more than<br />

2 km in linear tract in the nearest vicinity of<br />

the Main Boundary Thrust.<br />

2) These are confined within cohesionless<br />

loose and friable sediments that are<br />

potentially liquefiable.<br />

3) These structures are comparable with the<br />

structures developed in seismic shaking<br />

experiments.<br />

4) The structures are exposed in the vicinity of<br />

the Main Boundary Thrust that is one of the<br />

most seismically active thrusts in the<br />

Himalayan Terrain.<br />

5) Several small scale structures such as<br />

minor faults in the recumbent-cross<br />

stratification are developed.<br />

6) Mushroom structures are very suggestive<br />

for seismic trigger.<br />

7) Frequent occurrence of faults with liquefied<br />

structures indicates need of instant shaking<br />

as soon as the grain to grain contact<br />

develops that suggest possible seismic<br />

origin.<br />

It is generally agreed that seismic activity of not less<br />

than 5.5 magnitude on Richter scale can possibly be<br />

recorded in sediments, as seismites (Ambraseys,<br />

1988). Hence this event shows an earthquake event<br />

of at least 5.5 magnitude on Richter scale. Two<br />

phases of seismic trigger has been established on<br />

the basis of overprinting of structures. The sand dyke<br />

representing the fluidization of one phase must has<br />

been developed earlier than synsedimentary fault of<br />

second phase cutting through it, that shows brittle<br />

nature of origin (Fig. 10).<br />

Confined within the 4-5 Ma horizon (Kotalia et al.,<br />

2001), the soft-sedimentary deformation structures<br />

increase in abundance and, towards the Main<br />

Boundary Thrust. As the development of the Main<br />

Boundary Thrust occurred c. 10-11 Ma. (Miegs et al.,<br />

1995). Hence, it is evident that the ancient seismites,<br />

discussed in this article, are related to the 4-5 Ma old<br />

reactivation event on the Main Boundary Thrust.<br />

Acknowledgements: I express my thanks to conference<br />

organizers for giving me chance to present my work.<br />

References<br />

Allen, J.R.L., Banks, N.L., 1972. An interpretation and<br />

analysis of recumbent-folded deformed cross-bedding.<br />

Sedimentology 19, 257-283.<br />

Allen, J.R.L., 1982. Sedimentary Structures: their character<br />

and physical basis. Development in Sedimentology. Vol.<br />

30. Elsevier, Amsterdem. 663 pp.<br />

Ambraseys, N.N., 1988. Engineering seismology.<br />

Earthquake Engineering Structural Dynamics 17, 1–105.<br />

Jackson, B., Bilham, R., 1994. Constraints on Himalayan<br />

Deformation Inferred from the vertical velocity fields in the<br />

Nepal and Tibet. Journal of Geophysical Research 99<br />

(B7), 13897-13,912.<br />

Kotalia, B.S., Nakayama, K., Bhalla, M.S., Phartiyal, B., T.<br />

Kosake, 2001. Lithology and magnetic stratigraphy of<br />

Lower-Middle Siwalik succession between Kathgodam<br />

and Ranibagh, Kumaun Himalaya. Journal Geological<br />

Society of India 58, 411-423.<br />

Lowe, D.R., 1975. Water escape structures in coarsegrained<br />

sediments. Sedimentology 22, 157–204.<br />

Miegs, A. J., Burbank, D. W., Beck, R. A., 1995. Middle late-<br />

Miocene (>10 ma) formation of the main boundary thrust<br />

in the western Himalaya. Geology. 23, 423-426.<br />

Nakata, T., 1989. Active faults of the Himalaya of India and<br />

Nepal. Geological Society of America. Special paper.<br />

232, 243-264.<br />

Owen, G., 1987. Deformation processes in unconsolidated<br />

sands. In: Jones, M.E., Preston, R.M.F. (Eds.),<br />

Deformation of Sediments and Sedimentary Rocks,<br />

Geological Society of (London) Special Publication. No.<br />

29, pp. 11–24.<br />

Owen, G., 1996. Experimental soft-sediment deformation<br />

structures formed by the liquefaction of unconsolidated<br />

sands and some ancient examples. Sedimentology 43,<br />

279–293.<br />

Owen, G., Moretti, M., 2011. Identifying triggers for<br />

liquefaction-induced soft-sediment deformation in sands.<br />

Sedimentary Geology 235, 141-147.<br />

Owen, G., Moretti, M., Alfaro, P., 2011. Recognising triggers<br />

for soft-sediment deformation: current understanding and<br />

future directions. Sedimentary Geology 235, 133-140.<br />

Seilacher, A., 1969. Fault-graded beds interpreted as<br />

seismites. Sedimentology 13, 155–159.<br />

Sims, J.D., 1973. Earthquake-induced structures in<br />

sediments in Van Norman Lake, San Fernando,<br />

California. Science 182, 161–163.<br />

Tandon, S.K., 1976. Siwalik sedimentation in a part of<br />

Kumaun Himalaya. Sedimentary Geology 16, 131–154.<br />

Van Loon, A.J., Brodzikowski, K., 1987. Problems and<br />

progress in the research on soft sediment deformations.<br />

Sedimentary Geology 50, 167–193.<br />

Yin, A., 2006. Cenozoic evolution of the Himalayan orogen<br />

as constrained by along-strike variation of structural<br />

geometry, exhumation history, foreland sedimentation.<br />

Earth Sciences Reviews 76, 1-131.<br />

147


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SAMPLING BIASES IN THE PALEOSEISMOLOGICAL DATA<br />

Mouslopoulou Vasiliki (1,2), Andrew Nicol (3), John J. Walsh (1),<br />

John G. Begg (3), Dougal B. Townsend (3), Dionissios T. Hristopulos (2)<br />

(1) Fault Analysis Group, University College Dublin, Dublin 4, Ireland. Email: vasiliki@mred.tuc.gr, john@fag.ucd.ie<br />

(2) Dpt. of Mineral Resources Engineering, Technical University of Crete, 73100, Greece. Email: dionisi@mred.tuc.gr<br />

(3) GNS Science, PO Box 30368, New Zealand. Email: a.nicol@gns.cri.nz ; j.begg@gns.cri.nz ; d.townsend@gns.cri.nz<br />

Abstract (Sampling biases in the paleoseismological data): The recent earthquakes in Christchurch, New Zealand, show that<br />

active faults, capable of generating large-magnitude earthquakes, can be hidden beneath the Earth’s surface. Here we combine<br />

near-surface paleoseismic data with deep (


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Pliocene (3-4 Ma) and has produced kilometre-scale<br />

cumulative displacements on normal faults (Nicol et<br />

al., 2007; Giba et al., 2010; Mouslopoulou et al., In<br />

review), active faulting at the ground surface<br />

(Townsend et al., 2010; Mouslopoulou et al., in<br />

review) and historical seismicity (Sherburn & White,<br />

2006).<br />

Crustal extension is also accompanied by<br />

widespread Late Quaternary ( <br />

on the Taranaki Peninsula. The most recent phase of<br />

volcanic activity started about 100 ka (last eruption<br />

ca. 250 years ago) and resulted in the formation of<br />

the impressive volcanic cone of Mt. Taranaki.<br />

Although Mt. Taranaki rises to 2518 m above sealevel<br />

(a.s.l.) and dominates the landscape, studies<br />

show no direct link between the timing of prehistoric<br />

large-magnitude earthquakes in the rift and volcanic<br />

eruptions or episodes of cone collapse (Townsend et<br />

al., 2010).<br />

Here we combine seismic-reflection and trench data<br />

for individual faults to chart their growth in<br />

displacement through time. Two-dimensional (2D)<br />

seismic-reflection lines occupying the southwest of<br />

Mt. Taranaki, extend across the entire onshore width<br />

of the rift and in a rift-parallel direction for about 20<br />

km (Fig. 1). The 2D seismic survey comprises a total<br />

of 21 fault-perpendicular and 17 fault-parallel lines<br />

spaced at 0.5 to 5 km. These seismic lines image all<br />

of the known active faults and provide estimates of<br />

<br />

sites. Seismic reflectors extending to depths of up to<br />

ca. 5 km have been tied to five wells (Figs 1 and 2a)<br />

in which the stratigraphy was dated using<br />

micropaleontology (King & Thrasher, 1996). The six<br />

youngest of these seismic reflectors (ca. 0.5, 2, 3,<br />

3.6, 5.5 and 10 Ma) were traced along and across<br />

the faults around the grid of seismic lines and provide<br />

information on the accumulation of fault<br />

displacements during the Plio-Pleistocene. For<br />

individual faults in the area of study fault vertical<br />

displacements are up to ca. 1.5 km and decrease<br />

with reduction in horizon age from 3.6 Ma (Fig. 3a).<br />

The paleoearthquake history of the six active faults<br />

with resolvable surface traces is defined by<br />

constraints from 10 excavated trenches. Associated<br />

trench data permit identification of a total of 23 large<br />

paleoearthquakes (M>5.5) that ruptured the ground<br />

surface since 27 ka (Fig. 2b). The earthquake record<br />

is complete since 5 ka on all six active faults and for<br />

27 kyr on two of these faults. Vertical displacements<br />

measured in the trenches range from 0.1 m to 4 m<br />

(e.g., Townsend et al., 2010; Fig. 2b) and accrued<br />

from earthquakes with single-event displacements of<br />

up to 1.2 m. These displacements, which represent<br />

point measurements on fault traces (accrued during<br />

one or more earthquakes), are supplemented by<br />

scarp heights recorded along the lengths of each<br />

active trace. Paleoearthquake ages are primarily<br />

constrained by 14 C dating of near-surface (< 6 m<br />

deep) stratigraphy displaced by active faults (Fig.<br />

2b). Further details of the trenching are documented<br />

by Townsend et al. (2010).<br />

SAMPLING BIASES<br />

Seismic-reflection lines record many more faults than<br />

can be inferred from field mapping of active traces.<br />

Thirteen faults displace the 0.5 Ma horizon and<br />

clearly extend upwards towards the ground surface<br />

(Fig. 2a). Field mapping and interpretation of aerial<br />

ortho-photographs suggest that six of these faults<br />

have surface scarps up to 4 m high and displace a<br />

diachronous landscape mainly ranging in age from<br />

~8 to 27 ka (Fig. 1) (Townsend et al., 2010;<br />

Mouslopoulou et al., in review). Trench data confirm<br />

that these faults are active and have ruptured the<br />

ground surface during large-magnitude prehistoric<br />

earthquakes (Figs. 2b) (Townsend et al., 2010). The<br />

remaining seven faults that displace the 0.5 Ma<br />

horizon generally have large cumulative<br />

displacements (>1000 m throw) and also displace<br />

near-surface horizons (


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

lengths are generally between 20 and 30 km.<br />

Analysis of individual faults suggests that the subsurface<br />

fault lengths are about 2-8 times larger than<br />

their equivalent surface traces. Some of this<br />

discrepancy arises because fault scarps


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

often through impressive fault scarps (Nicol et al.,<br />

2009). These faults are likely to be those which have<br />

recently accommodated successive earthquakes<br />

(e.g. during the Holocene). This view is supported by<br />

seismic reflection lines which show that the number<br />

of surface-rupturing faults in the rift is significantly<br />

less than the total number of faults in the rift (Figs 1<br />

and 2a). Examination of the buried faults indicates<br />

that, in addition to the six ground-rupturing faults,<br />

there are at least a further seven faults, including the<br />

larger faults in the system, that reach within ~50 m of<br />

the modern ground surface and are probably active<br />

(i.e. still accruing displacement and capable of<br />

producing future ground-rupturing earthquakes). For<br />

these additional seven active faults, long-term<br />

displacement rates are greater than Holocene rates<br />

by up to a factor of five (Fig. 3a). Thus, faults with<br />

increased and decreased displacement rates in the<br />

Holocene could be present in approximately equal<br />

proportions. As a result of these temporal rate<br />

changes, the sums of long-term rates on all faults in<br />

the system and Holocene displacement rates on the<br />

13 active faults identified are similar (i.e. long-term<br />

3.4±0.5 mm/yr and Holocene 3.1±0.5 mm/yr). This<br />

similarity in the sum of the rates over each time<br />

interval supports the view that regional rates of<br />

extension may not have changed, and that the<br />

temporal clustering of earthquakes reflects migration<br />

of the locus of fault activity from one fault to another.<br />

An important consequence of this migration is that<br />

faults which are currently in a relatively quiescent<br />

phase of earthquake activity may in the next 10 kyr,<br />

for example, become relatively active. Earthquake<br />

hazard analysis must take account of these relatively<br />

quiescent faults.<br />

CONCLUSIONS<br />

In the Taranaki Rift displacement rates vary<br />

temporally on individual faults by in excess of an<br />

order of magnitude over timescales of thousands to<br />

millions of years. These changes are attributed to<br />

fault interactions rather than to changes in regional<br />

strain rates. During the Holocene fault displacement<br />

rates were both faster (~50%) and slower (~50%)<br />

than their million-year averages. Faults moving faster<br />

than the long-term average can be identified in the<br />

landscape, while those moving slower cannot. The<br />

numbers of active faults and their active trace lengths<br />

are underestimated by at least 50% using<br />

geomorphic mapping of the scarps. Therefore, the<br />

number of potential earthquake sources may be<br />

significantly higher than it is represented in seismic<br />

hazard models. As a result, some future earthquakes<br />

will occur on faults that were not previously known to<br />

be active. Integration of seismic-reflection with<br />

paleoseismic data provides a basis for identifying<br />

active faults not observed at the ground surface,<br />

estimating maximum fault-rupture lengths and<br />

improving earthquake hazard assessment.<br />

References<br />

Coppersmith, K.J., (1989). On spatial and temporal<br />

clustering of paleoseismic events. Seismological<br />

Research Letters 59, 299-230.<br />

Giba, M., A. Nicol & J.J. Walsh, (2010). Evolution of faulting<br />

and volcanism in a back-arc basin and its implications for<br />

subduction processes. Tectonics 29,<br />

doi:10.1029/2009TC002634.<br />

King, P.R. & G.P. Thrasher, (1996). Cretaceous and<br />

Cenozoic geology and petroleum systems of the Taranaki<br />

Basin, New Zealand. Institute of Geological & Nuclear<br />

Sciences Monograph 13, 243 p.<br />

Marco, S., M. Stein & A. Agnon, (1996). Long-term<br />

earthquake clustering: a 50 000-year paleoseismic record<br />

in Dead Sea Graben. Journal of Geophysical Research<br />

101, 6179-6191.<br />

Mouslopoulou, V., J.J. Walsh & A. Nicol, (2009). Fault<br />

displacement rates on a range of timescales. Earth and<br />

Planetary Science Letters 278, 186-197.<br />

Mouslopoulou, V., A. Nicol, J.J. Walsh, J.G. Begg, D.B.<br />

Townsend & D.T. Hristopulos, (in review). Relations<br />

between paleoearthquakes and million year fault growth<br />

in an active rift. Journal of Structural Geology, 2011.<br />

Nicol, A., J.J. Walsh, K. Berryman, P. Villamor, (2006).<br />

Interdependence of fault displacement rates and<br />

paleoearthquakes in an active rift. Geology 34, 865-868.<br />

Nicol, A., C. Mazengarb, F. Chanier, G. Rait, C. Uruski & L.<br />

Wallace, (2007). Tectonic evolution of the Hikurangi<br />

subduction margin, New Zealand, since the Oligocene.<br />

Tectonics 26(4), TC4002. doi:10.1029/2006TC002090.<br />

Nicol, A., J.J. Walsh, V. Mouslopoulou, & P. Villamor,<br />

(2009). Earthquake histories and Holocene acceleration<br />

of fault displacement rates. Geology 37, 911–914.<br />

Palumbo, L., L. Benedetti, D. Bourles, A. Cinque & R.<br />

Finkel, (2004). Slip history of the magnolia fault<br />

(Apennines, Central Italy) from 36Cl surface exposure<br />

dating: evidence for strong earthquakes in the Holocene.<br />

Earth and Planetary Science Letters 225, 163–176.<br />

Schwartz, D.P. & K.J. Coppersmith, (1984). Fault behavior<br />

and characteristic earthquakes: examples from the<br />

Wasatch and San Andreas fault zones. Journal of<br />

Geophysical Research 89, 5681–5698.<br />

Sherburn, S. & R.S. White, (2006). Tectonics of the<br />

Taranaki region, New Zealand: earthquake focal<br />

mechanisms and stress axes. New Zealand Journal of<br />

Geology and Geophysics 49, 269–279.<br />

Sieh, K., M. Stuiver & D. Brillinger, (1989). A more precise<br />

chronology of earthquakes produced by the San Andreas<br />

Fault in Southern California. Journal of Geophysical<br />

Research 94, 603-623.<br />

Townsend, D., A. Nicol, V. Mouslopoulou, J.G. Begg, R.D.<br />

Beetham, D. Clark, M. Giba, D. Heron, B. Lukovic, A.<br />

McPherson, H. Seebeck & J.J. Walsh, (2010).<br />

Paleoearthquake histories across a normal fault system<br />

in the southwestern Taranaki Peninsula, New Zealand.<br />

New Zealand Journal of Geology and Geophysics 53, 1-<br />

20.<br />

Weldon, R., K. Scharer, T. Fumal, G. Biasi, (2004).<br />

Wrightwood and the earthquake cycle: what a long<br />

recurrence record tells us about how faults work. GSA<br />

Today 14, 9, 4-10.<br />

Wells, D.L. & K.J. Coppersmith, (1994). New empirical<br />

relationships among magnitude, rupture length, rupture<br />

width, rupture area and surface displacement. Bulletin of<br />

the Seismological Society of America 8, 974-1002.<br />

Acknowledgements: This research was funded by an IIF<br />

Marie Curie Fellowship of the European Community’s 7th<br />

Framework Program (contract no. PIIF-GA-2009-235931)<br />

and an IRCSET (Irish) Postdoctoral Fellowship.<br />

151


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKES IN AQABA, JORDAN OVER THE PAST 2,000 YEARS:<br />

EVIDENCE FROM HISTORICAL, GEOLOGICAL, AND ARCHAEOLOGICAL DATA<br />

Tina M. Niemi (1)<br />

(1) Department of Geosciences, University of Missouri-Kansas City, 5100 Rockhill Road, Flarsheim Hall 420, Kansas City, MO<br />

64110 U.S.A. Email: niemit@umkc.edu<br />

Abstract (Earthquakes in Aqaba, Jordan Over the Past 2,000 Years: Evidence from Historical, Geological,<br />

and Archaeological Data): Aqaba lies at the boundary between the Gulf of Aqaba and Wadi ‘Arabah segments of<br />

the Dead Sea Fault Zone (DSFZ) and is therefore situated to sustain earthquake damage from rupture of either<br />

segment. The rupture history of these fault segments remains rather enigmatic because of the low population<br />

density around them throughout history. This paper presents a synthesis of paleoseismic trenching, archaeological<br />

excavations, and historicl data and provides a model for the recurrence of earthquakes along the southern DSFZ.<br />

Significant periods of active seismicity in the 4 th , 7 th -8 th , 11 th -13 th , and 15-16 th centuries suggest a three- to fivecentury<br />

recurrence rate of faulting of the Gulf of Aqaba and Wadi ‘Arabah segments of the DSFZ. It is interesting to<br />

note that earthquakes have been coincident with major political transitions that have occurred in this region and thus<br />

are likely to have played a significant role in these cultural shifts.<br />

Key words: Archaeoseismology, Dead Sea Transform, Jordan, Earthquakes<br />

INTRODUCTION<br />

The Dead Sea fault zone (DSFZ) is an left-lateral,<br />

strike-slip transform plate boundary between the<br />

Arabian and Sinai plates (Fig. 1). Earthquakes in the<br />

southern ‘Arabah and Gulf of Aqaba of Jordan and<br />

Israel are created by motion along the DST. The Mw<br />

7.2 Nuweiba earthquake of November 22, 1995 that<br />

ruptured a submarine fault in the Gulf of Aqaba (Gulf<br />

of Eilat) was the largest earthquake in the modern<br />

instrumented era (Hofstetter et al., 2003) along the<br />

Dead Sea fault. Most of the significant damage was<br />

concentrated in cities in the Sinai Peninsula near the<br />

epicenter, but damage was also reported from the<br />

Saudi Arabian coastline and the cities of Aqaba,<br />

Jordan and Eilat, Israel; both about 70 km north of<br />

the epicenter.<br />

Historical earthquake data reported in recent<br />

catalogues (Guidoboni, 1994; Guidoboni and<br />

Comastri, 2005; Ambraseys, 2009) suggest that the<br />

seismic events in A.D. 110-114?, 363, 749, 1068,<br />

1212, 1458, and 1588 likely caused damage in the<br />

region of southern Jordan and Aqaba. It is unclear<br />

whether events recorded in the Dead Sea region or<br />

Jerusalem, such as the earthquakes of A.D. 419,<br />

597, 634, 659, 1293, and 1546, could have also<br />

caused damage in Aqaba. Data from earlier<br />

catalogues have largely been superseded by more<br />

recent catalogue compilations.<br />

The city of Aqaba is situated at the northern end of<br />

Gulf of Aqaba along the southern part of the Dead<br />

Sea Transform fault system that separates the Sinai<br />

and Arabian tectonic plates. Furthermore, Aqaba lies<br />

Fig. 1: Regional setting o the Dead Sea fault system showing<br />

recent location of recent earthquake foci. Map after Hartman<br />

(2011).<br />

152


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

along the transition from the marine to the continental<br />

Eilat/Aqaba sedimentary basins. Faults controlling<br />

the structurally dynamics of sedimentation as well as<br />

the seismic activity lie both onshore and offshore. In<br />

order to understand the history of earthquakes in the<br />

region, an understanding of the seismogenic faults in<br />

both the marine and continental environments is thus<br />

essential.<br />

1.3 m across the older Qf1 and Qf2 surfaces. These<br />

data indicate that scarp heights reflect cumulative slip<br />

events. The most recent scarp-forming event fault<br />

occurred after A.D. 1045-1278 based on a corrected,<br />

calibrated radiocarbon age from charcoal collected<br />

from a buried campfire at the base of the scarp in<br />

Trench T-1. This likely represents fault motion in one<br />

of the historical earthquakes affecting southern<br />

Jordan (e.g. 1068, 1212, 1458, or 1588).<br />

ACTIVE FAULTING<br />

Offshore geophysical surveys have recently identified<br />

four submarine fault zones in the northern Gulf of<br />

Aqaba (Tibor et al., 2010). Two fault zones flank the<br />

margins of the gulf (the Eilat fault and Aqaba fault)<br />

and continue onland as faults that truncate the distal<br />

portions of alluvial fan systems. Both the onshore<br />

and offshore Eilat marginal faults have normal fault<br />

displacement (e.g., Ben-Avraham 1985; Bowman &<br />

Gerson, 1986). Along the eastern margin of the gulf<br />

lie the Aqaba and west Aqaba fault zones. The very<br />

steep bathymetric escarpment with granitic bedrock<br />

truncated at the eastern shoreline indicates that a<br />

significant amount of vertical offset is accommodated<br />

on faults of the Aqaba fault zone. The zone is wide<br />

with three offshore strands. Hartman (2011) mapped<br />

strands of a “West Aqaba” fault that may also<br />

accommodate strike-slip motion. In the middle of the<br />

basin, the Ayla fault zone that appears to bound<br />

subsidence across a localized basin was apparently<br />

active in the early Holocene. The main offshore<br />

strike-slip fault continues northward as the Evrona<br />

fault in Israel. This fault extends into Jordan where it<br />

crosses the Wadi Muhtadi alluvial fan and continues<br />

northward to the Dead Sea. The fault in Jordan is<br />

called the Wadi ‘Arabah fault. Faulting in the 1068<br />

earthquake has been documented on this fault in the<br />

Evrona sabkha (Amit et al., 1999, 2002; Zilberman et<br />

al. 2005).<br />

In Aqaba a concerted effort has been made to map<br />

the onshore continuation of the Aqaba fault zone.<br />

Due to the dense urbanization, mapping the<br />

northward continuation of faults has been difficult. Air<br />

photo interpretation of the Aqaba regional surficial<br />

geology suggests that the Aqaba fault emerges from<br />

the gulf and that slip is transferred to northwesttrending<br />

cross faults (Niemi and Smith 1999; Slater<br />

and Niemi 2003; Mansoor et al. 2004). Because the<br />

cross faults are linear and not offset, this geometry<br />

constrains the location of the Aqaba fault to lie south<br />

and/or east of the cross faults and at the toe of the<br />

eastern alluvial fan surfaces.<br />

Geological trenches (T-1 through T-5) were<br />

excavated across four NW-trending cross-faults (Fig.<br />

2) that produce active tectonic subsidence at the<br />

head of the Gulf (Mansoor, 2002; Slater and Niemi,<br />

2003). Mapping of alluvial fan and buried soil<br />

horizons in the trenches reveal multiple fault ruptures<br />

on the highest scarps and fewer distinct ruptures on<br />

the lowest scarp (Mansoor, 2002). The scarp heights<br />

range from 25 cm across the youngest Qf3 surface to<br />

153<br />

Fig. 2: Map of the city of Aqaba showing the location of<br />

major archaeological sites. Active cross faults (CF) mapped<br />

from aerial photos and in the archaeological excavations of<br />

J-East are also shown (Thomas et al., 2007).<br />

ARCHAEOSEISMOLOGY<br />

Aqaba has a rich cultural history. The earliest<br />

evidence of sedentary occupation found are the Late<br />

Chalcolithic sites of Tell Magass, and Tell Hujayrat<br />

al-Ghuzlan located 3-4 km northeast of the gulf<br />

(Khalil & Schmidt, 2009). An Iron through Hellenistic<br />

period (8 th - 4 th centuries B.C.) site called Tell el-<br />

Kheleifeh (Pratico, 1993) is located along the Jordan-<br />

Israel border. Along the coast and extending<br />

underneath the modern city, the remains of the early<br />

Roman-Byzantine (1 st century B.C.-7 th century A.D.),<br />

Early Islamic (7 th -12 th century), and Mamluk through<br />

Ottoman (13 th -early 20 th century) cities of Aila (or al<br />

‘Aqabah) have been partially excavated (e.g.<br />

Whitcomb, 1994; Parker, 2007; De Meulemeester &<br />

al-Shqour, 2008).<br />

Excavation of the Aila ruins from the 1 st to the 8 th<br />

centuries in Area J-East of the Roman Aqaba Project<br />

exposed a monumental mudbrick structure heavily<br />

damaged by successive earthquakes. Nine faults<br />

were mapped across the site (Thomas et al., 2007).<br />

Based on subsidence across the fault locations,


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

changes in floor elevations, and layers of collapsed<br />

mudbrick, the archaeological data suggest that the<br />

site was ruptured in an early 2nd century earthquake,<br />

an early 4th century earthquake, and the 363 A.D.<br />

earthquake. The monumental use of the structure<br />

was converted to domestic use in the late 4th to early<br />

5th century.<br />

We also have evidence for primary ground rupture for<br />

at least four post-date 363 A.D. earthquakes that<br />

fault the ruins in the J-East area of Aila. Primary fault<br />

rupture is documented in stratigraphic sections and<br />

plan maps of walls of various construction ages. Two<br />

earthquakes occurred during the Byzantine to<br />

Umayyad period (sixth to eighth centuries). There is<br />

a hiatus of deposition at this location between the<br />

Umayyad and the modern age. The two most recent<br />

earthquakes, with 42 and 35 cm of dip slip, occurred<br />

some time after the 8 th century and likely correlate to<br />

the historical earthquakes after the 11th century. No<br />

stratified materials were found at this site that could<br />

be used to further refine the timing of these seismic<br />

events.<br />

The early Islamic (8 th -11 th centuries A.D) site of Ayla<br />

was excavated in 1986-1995. Whitcomb<br />

hypothesized that the drainage running through the<br />

ancient site originated in erosion along the structural<br />

weakness of a fault and placed such a fault on the<br />

site plan map (Whitcomb, 1994). However,<br />

excavations by Rucker and Niemi (2005) of the NE<br />

corner tower of the walled city in the wadi and<br />

interpretation of 1918, 1945, and 1953 air photos<br />

indicate the wadi is man-made. There is evidence at<br />

Islamic Ayla for damage as a result of the 749 (or<br />

746 or 757? See Ambraseys 2005; 2009) earthquake<br />

followed by extensive reconstruction at the beginning<br />

of the Abbasid period. Major damage occurred in the<br />

town of Ayla in the March 18, 1068 earthquake. One<br />

contemporary source living in Baghdad, Ibn al-<br />

Banna, wrote “As for Aila, its inhabitants all perished<br />

except for 12 persons who had gone fishing at sea,<br />

thus escaping death.” (Guidoboni and Comastri,<br />

2005, p. 53). The site of Islamic Ayla was apparently<br />

never reoccupied to any significant degree after the<br />

earthquake.<br />

The site of Early Islamic Ayla was not rebuilt, but a<br />

new castle or caravan station was built about 1 km to<br />

the southeast. Excavations in and around the Aqaba<br />

castle from 2000-2008 have revealed three different<br />

phases in the “khan” or castle from the late 12 th to<br />

16 th centuries (De Meulemeester and Al-Shqour,<br />

2008). The extant castle was built in 1515 and rebuilt<br />

in 1587/8, probably after the Gulf of Aqaba<br />

earthquake of January 4, 1588 which, based on<br />

historical accounts, was felt in NW Arabia, Aqaba,<br />

and Sinai (Guidoboni and Comastri, 2005;<br />

Ambraseys, 2009). The archaeological data from the<br />

Aqaba castle (De Meulemeester and Al-Shqour,<br />

2008) also appear to support rupture of the Gulf of<br />

Aqaba fault segment in the earthquake of 1212 and<br />

possibly of the Wadi ‘Arabah fault segment in 1458.<br />

Our data suggest significant periods of active<br />

seismicity in the 4 th , 7 th -8 th , 11 th -13 th , and 15-16 th<br />

Centuries. These data suggest a recurrence rate of<br />

faulting and damaging earthquakes either from the<br />

Gulf of Aqaba segment or the Wadi ‘Arabah segment<br />

of the Dead Sea fault system of about every three to<br />

four centuries. It is interesting to note that<br />

earthquakes have been coincident with major political<br />

transitions that occurred in the region. Seismic<br />

activity is thus likely to have played a significant role<br />

in these cultural shifts.<br />

Acknowledgements: The Wadi ‘Arabah Earthquake<br />

Project has been supported by the Roman Aqaba Project,<br />

the American School of Oriental Research, a UMKC Faculty<br />

Research grant, the University of Missouri Research Board,<br />

and two grants from the Committee on Research and<br />

Exploration of the National Geographic Society. I am<br />

grateful for the assistance of Dr. Fawwaz al Khraysheh,<br />

former General Director, and Dr. Sawsan Fakhri, Aqaba<br />

Region director of the Department of Antiquities of Jordan,<br />

who granted permission to work in Aqaba.<br />

References<br />

Ambraseys, N.N. (2005). The seismic activity in Syria and<br />

Palestine during the middle of the 8 th<br />

century; an<br />

amalgamation of historical earthquakes. Journal of<br />

Seismology 9, 115-125.<br />

Ambraseys, N. (2009). Earthquakes in the Mediterranean<br />

and Middle East: A Multidisciplinary Study of Seismicity<br />

up to 1900. Cambridge University Press. Cambridge,<br />

U.K. 947 p.<br />

Amit, R., E. Zilberman, N. Porat, N. & Y. Enzel, (1999).<br />

Relief inversion in the Avrona playa as evidence of largemagnitude<br />

historical earthquakes, southern Arava valley,<br />

Dead Sea rift. Quaternary Research 52, 76-91<br />

Amit, R., E. Zilberman, Y. Enzel & N. Porat, (2002).<br />

Paleoseismic evidence for time dependency of seismic<br />

response on a fault system in the southern Arava Valley,<br />

Dead Sea rift, Israel. Geological Society of America<br />

Bulletin 114 (2), 192-206.<br />

Ben-Avraham, Z. (1985). Structural framework of the Gulf of<br />

Eilat (Aqaba), Northern Red Sea. Journal of Geophysical<br />

Research 90, 703-726.<br />

Bowman, D. & R. Gerson, (1986). Morphology of the Latest<br />

Quaternary Surface Faulting in the Gulf of Elat Region,<br />

Eastern Sinai. Tectonophysics 128, 97-119.<br />

De Meulemeester, J. & Al-Shqour, R. (2008). The Islamic<br />

Aqaba Project 2008: Preliminary Unpublished Report.<br />

Ghent University, Belgium, 48 p.<br />

Guidoboni, E. (1994). Catalogue of Ancient Earthquakes<br />

and Tsunamis in the Mediterranean Area from the 11 th to<br />

the 15 th Century. Istituto Nazionale di Geofisica. Roma,<br />

Italy. 504 p.<br />

Guidoboni, E., & A. Comastri, (2005). Catalogue of Ancient<br />

Earthquakes in the Mediterranean Area up to the 10 th<br />

Century. Istituto Nazionale di Geofisica. Roma, Italy.<br />

1037 p.<br />

Hartman, G. (2011). Quaternary Evolution of a Transform<br />

Basin: The Northern Gulf of Eilat/Aqaba. Dissertation. Tel<br />

Aviv University, Israel. 100 p.<br />

Hofstetter, R., Y. Gitterman, V. Pinksy, N. Kraeva, & L.<br />

Feldman, (2008). Seismological observations of the<br />

northern Dead Sea basin earthquake on 11 February<br />

2004 and its associated activity. Israel Journal of Earth<br />

Sciences 57 (2), 101-124.<br />

Khalil, L., & Schmidt, K. (2009). Prehistoric ‘Aqaba I.<br />

Orient-Archäologie Band 23, Rahden/Westf., Germany,<br />

Verlag Marie Leidorf. 419 p.<br />

154


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SWATH BATHYMETRY AND MORPHOLOGICAL SLOPE ANALYSIS OF THE<br />

CORINTH GULF<br />

Nomikou, P. (1), Alexandri M. (2), Lykousis V. (2), Sakellariou D. (2), Ballas D. (2)<br />

(1) University of Athens, Department of Geology and Geoenvironment, Panepistimioupoli Zografou, 15784 Athens, Greece,<br />

evi@ath.hcmr.gr<br />

(2) Institute of Oceanography, Hellenic Centre for Marine Research, POBox 712, 19013 Anavyssos, Greece,<br />

matina@ath.hcmr.gr, vlikou@ath.hcmr.gr, sakell@ath.hcmr.gr, dballas@ath.hcmr.gr<br />

Abstract (Swath Bathymetry and Morphological Slope Analysis of Corinth Gulf): The swath bathymetric survey of the Gulf of<br />

Corinth has been conducted during multiple campaigns of HCMR’s vessel R/V "AEGAEO" between March 2001 and November<br />

2004, using the 20 kHz, SEABEAM 2120 system. The bathymetric map was produced using a 20 meter grid interval and plotted<br />

with a Mercator projection at 1:300.000 scale with 20 m contours. The major part of the central Gulf is a 800-870 m deep, 40km<br />

long, 9-12km wide, WNW-ESE elongate flat area. West of Aigio, the basin is shrinking towards its central run off axis, and finally<br />

diminishes at Rio-Antirrio straight. Eastwards the central basin terminates at the Alkyonides islets, east of which a sub-basin with<br />

maximum depth of 350 m is formed. The steep (30–40%) southern margin of the basin between Kiato to Aigio, is incised by<br />

numerous small canyons. The northern margin dips gently till the 400m isobaths and becomes steeper between 400-800m. The<br />

very detailed illustration of the Gulf’s bathymetry and morphology reflects the offshore active tectonics and faulting of the seafloor<br />

and the deformation during Quaternary.<br />

Key words: Swath Bathymetry, Sea Bed Morphology, Corinth Gulf<br />

INTRODUCTION<br />

The Corinth Rift, which separates the Peloponese from<br />

continental Greece, is a N100E oriented elongate<br />

graben, 105 km long, bounded by systems of very recent<br />

normal faults (less than 2 my old). The Corinth Gulf<br />

reflects the modification of the geomorphology and<br />

crustal structure formed during the alpine orogenesis by<br />

the extension tectonics and the resulted deformation of<br />

the Aegean region during Miocene-to-Quaternary. The rift<br />

transects obliquely the Pindos mountain chain resulted<br />

from the alpine nap tectonics.<br />

This structure is the most seismically active zone in<br />

Europe, and the fastest opening area of continental<br />

break-up, with up to 1.5 cm yr-1 of north-south extension<br />

rate (Clarke et al, 1998), and more than 1 mm yr-1 of<br />

uplift rate of the southern margin and 0.7-1 mm yr-1<br />

subsidence of the northern margin (Lykousis et al., 2009).<br />

The high rates of tectonic faulting and uplift lead to the<br />

outcropping of very recent fault planes with large offsets.<br />

SURVEY AND SYSTEM DEPLOYMENT<br />

The multibeam bathymetric surveys were carried<br />

out by the R/V AEGAEO of the Hellenic Centre<br />

for Marine Research, using a SEABEAM 2120<br />

swath system during 2001-2004 within the frame<br />

of various EU and National projects like ASSEM<br />

(Lykousis et. al., 2006) (Fig.1). The SEABEAM<br />

2120 is a hull-mounted swath system operating<br />

at 20 kHz in water depths not exceeding 6000 m.<br />

It has an angular coverage sector of 150 0 with<br />

149 beams, covering a swath width from 7.5 to<br />

11.5 times the water depth for depths from 20 m<br />

to 5 km. The maximum swath coverage can<br />

reach 9 km at maximum depth and gives<br />

satisfactory data quality at speeds up to 11<br />

knots. During some cruises we perform high<br />

resolution 3.5kHz and Air Gun profiling, side<br />

scan sonar imaging, gravity and box coring,<br />

deployment of a current-meter and CTD<br />

measurements.<br />

Offshore active faulting on the seafloor of the Gulf of<br />

Corinth has been studied in detail during dozens of<br />

offshore campaigns led by Greek and international<br />

research teams (Sakellariou et al, 2001; 2004; 2007;<br />

Stefatos et al, 2002; Moreti et al, 2004; Zelt et al, 2004;<br />

McNeill et al, 2005; Lykousis et al, 2006; 2008; 2007a,b;<br />

2009; Bell et al, 2008; Taylor et al, 2011).<br />

The tectonic graben of the Corinth Gulf has been filled up<br />

by gravitational deposits (turbidity flows, debris flows,<br />

mudflows etc.), while fun delta prograding deposits during<br />

Pleistocene sea-level changes, coupled by extensive<br />

slumping are the predominant sedimentary processes in<br />

the steep flanks of the gulf (Heezen et al. 1966;<br />

Ferentinos et al. 1988; Papatheodorou & Ferentinos<br />

1997; Lykousis et al. 1998; Perisoratis et al. 2000;<br />

Lykousis et al. 2007a).<br />

155<br />

Fig.1: Bathymetric Track lines during 2001-2004<br />

oceanographic cruises with R/V AEGAEO.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig.2: Multibeam Bathymetric Map of Corinth Gulf using 20m isobaths<br />

SWATH BATHYMETRY<br />

The multibeam data have been extensively processed by<br />

means of data editing, cleaning of erroneous beams,<br />

filtering of noise, processing of navigation data and<br />

interpolation of missing beams. The resulting bathymetric<br />

map of Corinth Gulf was originally compiled at 1/300.000<br />

scale, which was greatly reduced for publication with 9<br />

different colors corresponding to 100 m depth intervals<br />

and with additional isobaths of 20 m (Fig.2). This map<br />

permits the first detailed description of the overall<br />

topography of the sea floor as well as the mapping of the<br />

major morphotectonic structures within this area.<br />

shows that the distribution of slope values within<br />

the studied area can be subdivided into five<br />

categories: (1) areas of mean morphological<br />

slope 0-5%, (2) areas of 5-15%, (3) areas of 15-<br />

25%, (4) areas of 15-25% and (5) areas of 35-<br />

50%.<br />

Most of the central part of the Gulf from 800 up to 870 m<br />

depth is a very wide flat area forming an extensive WNW-<br />

ESE elongated basin of 40km length and a width of 9 km<br />

at the west up to 12 km at the east (Alexandri et al.,<br />

2003) (Fig. 2).<br />

West of Aigio, the basin is shrinking towards its central<br />

run off axis, and finally diminishes at Rio-Antirio straight.<br />

Eastwards the central basin is reaching the Alkyonides<br />

islets where a sub-basin with maximum depth of 350 m is<br />

formed. The steep southern margin, that reaches the<br />

800m isobath is scored, from Kiato to Aigio by numerous<br />

small canyons trending NE-SW transversal to the main<br />

direction of the gulf (Fig. 3). In contrast, the northern<br />

margin spans up to 400m depth where abruptly drops to<br />

800 m creating the northern border of the central basin.<br />

Fig. 3: 3D view of Corinth Gulf combined offshore and<br />

onshore data.<br />

MORPHOLOGICAL ANALYSIS<br />

The bathymetric map of the area was analyzed from the<br />

perspective of slope distribution and the results are<br />

presented as a slope distribution map (Fig.5). This map<br />

Fig. 4: 3D view of fan delta failures (debris flows) along<br />

the southern flanks of Western Gulf of Corinth (imaged<br />

from ENE) (Lykousis et al., 2009).<br />

156


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig.5: Slope distribution Map of Corinth Gulf<br />

Fig.6: Detailed swath bathymetry map of the Western Corinth Gulf using 10m isobaths<br />

This classification of the slope magnitudes clearly<br />

illustrates the zones where there is an abrupt change of<br />

slope, which usually reflects the position of active tectonic<br />

structures or steep slopes.<br />

The highest slope values (35-50%) are observed in the<br />

southern margin from Kiato to Aigio due to numerous<br />

small canyons. The northern margin of the gulf if<br />

characterized mostly by morphological slopes up to 25%<br />

showing a smooth transition to the basin. On the contrary,<br />

the central part of the gulf is characterized by a<br />

flat bottom of the basin with low morphological<br />

slopes from 0-5%.<br />

WEST CORINTH GULF<br />

The western Gulf of Corinth is the currently most<br />

active part of the Gulf and has been the site of<br />

the Aegion M:6 earthquake in 1995 and many<br />

other strong tremors during the last century.<br />

157


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The seabed morphology of the western Gulf of Corinth is<br />

dominated by slope canyon systems feeding an axial<br />

channel which is only present within this part of the gulf<br />

(Fig. 6). Canyons prevail the southern margin but are rare<br />

on the northern margin of the gulf. The basin floor<br />

reaches a maximum depth of 800m towards the east.<br />

Slope failures are abundant in the southern margin and<br />

are related to prodelta sediment instabilities and to a<br />

minor degree in the northern margin (Eratini area, etc.). <br />

couple of sites were related to historical tsunarnis and are<br />

expected to be the potential tsunarnigenic sites the W.<br />

Corinth Gulf in the future (Lykousis et al., 2007a;<br />

Lykousis et al., 2009). Infinite slope stability analysis<br />

indicated that slopes steeper than 2-3 degrees are<br />

capable to induce slope failures taken account the<br />

expected (seismic) ground accelerations for the next 50<br />

years (Lykousis et al., 2009). The northern margin is<br />

dominated by a basement horst uplifted by the North (N)<br />

and South (S) Eratini faults (Stefatos et al., 2002). The<br />

horst is most prominent near the Psaromita Peninsula<br />

and is progressively buried and reduced in topography to<br />

the east and west, with faults as long as 15 km (McNeil et<br />

al., 2005).<br />

DISCUSSION<br />

The swath bathymetric survey of the Gulf of Corinth<br />

provides insight into morphological expression of the<br />

active seafloor processes, like faulting and submarine<br />

sliding, erosions and sediment deposition and enables<br />

detailed mapping of the offshore structural elements.<br />

The major part of the central Gulf is a 800-870 m deep,<br />

40km long, 9-12km wide, WNW-ESE elongate flat area.<br />

West of Aigio, the basin is shrinking towards its central<br />

run off axis, and finally diminishes at Rio-Antirio straight.<br />

Eastwards, the central basin terminates at the Alkyonides<br />

islets, east of which a sub-basin with maximum depth of<br />

350 m is formed. Both, the steep (30–40%) southern<br />

margin and the gentler dipping north margin of the basin<br />

are fault controlled. The steep margins are incised by<br />

numerous small canyons.<br />

References<br />

Alexandri M., Nomikou P., Ballas D., Lykousis V. & Sakellariou<br />

D. (2003): Swath bathymetry map of Gulf of Corinth. Geoph.<br />

Res. Abstracts, Vol. 5, 14268, EGS 2003.<br />

Bell, R., McNeill, L.C., Bull, J.M., and Henstock, T.J., 2008.<br />

Evolution of the western Gulf of Corinth continental rift,<br />

Greece. Geological Society of America Bulletin, 120, 156-178.<br />

Clarke, P.J., Davies, R.R., England, P.C., Parson, B., Billiris, H.,<br />

Paradissis, D., Veis, G., Cross, P.A., Denys, P.H., Ashkenazi,<br />

V., Bingley, R., Kahle, H.-G., Muller, M.-V., and Briole, P.,<br />

1998, Crustal strain in central Greece from repeated GPS<br />

measurements in the interval 1989-1997: Geophysical Journal<br />

International, 135, 195-214.<br />

Ferentinos, G., Papatheodorou, G. & Collins,M.B., 1988.<br />

Sediment transport processes on an active submarine fault<br />

escarpment: Gulf of Corinth, Greece, Mar. Geol., 83, 43–61.<br />

Heezen, B.C., Ewing, M. & Johnson, G.L., 1966. The Gulf of<br />

Corinth floor, Deep-Sea Res., 13, 381–411.<br />

Lykousis V., Sakellariou D., Blandin J., Person R., Etiope G.,<br />

Alexandri M., Nomikou P., Rousakis G., 2006, The ASSEM<br />

sea bed observatory for long term multihazard monitoring: Gulf<br />

of Corinth experiment. 8 . . . & ,<br />

. 151-154.<br />

158<br />

Lykousis, V., Sakellariou, D., Papanikolaou, D., 1998.<br />

Sequence stratigraphy in the northern margin of the<br />

Gulf of Corinth: implications to upper Quaternary<br />

basin evolution, Bull. Geol. Soc. Greece, 32, 157–<br />

164.<br />

Lykousis V., Sakellariou D., Rousakis G., Alexandri S.,<br />

Kaberi H., Nomikou P., Georgiou P., Balas D.<br />

,2007a. Sediment failure processes in active<br />

grabens: the Western Gulf of Corinth (Greece). In:<br />

Lykousis V., Sakellariou D., Locat J. (eds.),<br />

Submarine Mass Movements and their<br />

Consequences, 297-305, Springer.<br />

Lykousis V., Sakellariou D., Moretti, I., Kaberi H.<br />

,2007b. Late Quaternary basin evolution of the Gulf<br />

of Corinth: Sequence stratigraphy, sedimentation,<br />

fault–slip and subsidence rates. Tectoniphysics, 440,<br />

29-51.<br />

Lykousis, V., G. Roussakis , D. Sakellariou, 2009.<br />

Slope failures and stability analysis of shallow water<br />

prodeltas in the active margins of Western Greece,<br />

northeastern Mediterranean Sea. J. Earth Sciences,<br />

98, 807–822.<br />

McNeill, L., Cotterill, C., Stefatos, A., Henstock, T.,<br />

Bull, J., Collier, R., Papatheoderou, G., Ferentinos<br />

G., and Hicks, S. 2005, Active faulting within the<br />

offshore western Gulf of Corinth, Greece:<br />

implications for models of continental rift deformation:<br />

Geology, 33, 241-244.<br />

Moretti, I., Lykousis, V., Sakellariou, D., Reynaud, J.-<br />

Y., Benziane, B., and Prinzhoffer, A., 2004,<br />

Sedimentation and subsidence rate in the Gulf of<br />

Corinth: what we learn from Marion Dufresne’s longpiston<br />

coring: Comptes Rendus Geoscience, 336,<br />

291-299.<br />

Papatheodorou, G. & Ferentinos, G., 1997. Submarine<br />

and coastal sediment failure triggered by the 1995,<br />

Ms = 6.1 Aegion earthquake, Gulf of Corinth,<br />

Greece, Mar. Geol., 137, 287–304.<br />

Perissoratis C., Piper D.J.W. and Lykousis V., 2000.<br />

Alternating marine and lacustrine sedimentation<br />

during late Quaternary in the Gulf of Corinth rift<br />

basin, central Greece. Mar. Geol. 167: 391-411.<br />

Sakellariou, D., Lykousis,V. & Papanikolaou, D. (2001)<br />

Active faulting in the Gulf of Corinth, Greece. In:<br />

36th CIESM Congress <strong>Proceedings</strong>, 36 pp. 43.<br />

Sakellariou, D.,Kaberi, H. & Lykousis,V. (2004)<br />

Infuence of active tectonics on the recent<br />

sedimentation of the Gulf of Corinth basin. In: 10th<br />

International Congress of Greek Geological Society,<br />

Abstracts 15-17 April, p. 232-233, Thessaloniki.<br />

Sakellariou D., Lykousis V., Alexandri S., Kaberi H.,<br />

Rousakis G., Nomikou P., Georgiou, P., Ballas D.,<br />

(2007): Faulting, seismic-stratigraphic architecture<br />

and Late Quaternary evolution of the Gulf of<br />

Alkyonides basin – East Gulf of Corinth, Central<br />

Greece. Basin Research, 19/2, p. 273-295.<br />

Stefatos, A., Papatheoderou, G., Ferentinos, G.,<br />

Leeder, M., and Collier, R., 2002, Active offshore<br />

faults in the Gulf of Corinth, Greece: Their<br />

seismotectonic significance: Basin Research, 14,<br />

487-502.<br />

Taylor, B., Weiss, J.R., Goodliffe, A. M., Sachpazi, M.,<br />

Laigle, M. & Hirn, A., 2011. The structures,<br />

stratigraphy and evolution of the Gulf of Corinth rift,<br />

Greece Geophys. J. Int. (2011) 185, 1189–1219<br />

Zelt, B.C., Taylor, B., Weiss, J.R., Goodliffe, A.M.,<br />

Sachpazi, M., and Hirn, A., 2004. Streamer<br />

tomography velocity models for the Gulf of Corinth<br />

and Gulf of Itea, Greece. Geophys. J. Int., 159, 333-<br />

346.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

IS THE DECADENCE OF LEPTIS MAGNA (LYBIA) THE CONSEQUENCE OF A<br />

DESTRUCTIVE EARTHQUAKE?<br />

Pantosti, Daniela (1), Stefano Pucci (1), Paolo Marco De Martini (1), Alessandra Smedile (1<br />

(1) Istituto Nazionale di Geofisica e Vulcanologia. Via di Vigna Murata, 605. 00143- Roma. ITALY.<br />

Email: daniela.pantosti@ingv.it, stefano.pucci@ingv.it, paolomarco.demartini@ingv.it, alessandra.smedile@ingv.it<br />

Abstract (Is the decadence of Leptis Magna (Lybia) the consequence of a destructive earthquake?): The relationships<br />

between human modification of the environment and natural events in the Roman city of Leptis Magna are analyzed. Historical<br />

and archaeological sources indicate extreme natural events as the cause of the town’s decline: earthquakes, flooding, and<br />

tsunamis. Stratigraphical and geomorphological surveys investigated history and dynamics of the depositional and erosional<br />

systems of the settlement area by integrating independent constraints as archaeological and absolute radiocarbon dating. The<br />

results highlighted that once the Romans society could no longer guarantee the maintenance of defensive structures, destructive<br />

floods affected the town. Conversely, large earthquakes or tsunami have been discarded as primary cause of the town decline.<br />

Key words: Natural Hazards, Flooding, Earthquakes, AD 365 Crete earthquake and tsunami<br />

The Roman city of Leptis Magna (UNESCO world<br />

heritage) in western Libya was a magnificent,<br />

flourishing, strategic town in Tripolitania reaching its<br />

maximum expansion during the Empire of Septimius<br />

Severus (193-211 A.D.). The town is built on a wide<br />

alluvial fan fed by wadi (creek in Arabic) Lebda<br />

whose outlet served as a harbor since Phoenician<br />

times (Figure 1). An ingenious hydraulic system<br />

composed by a dam and an artificial channel<br />

diverting the wadi water was built by the Romans to<br />

prevent the town from flooding. The decline of Leptis<br />

Magna irreversibly started in the 4th Century.<br />

Historical and archaeological sources suggest that<br />

the town decadence is related to the fact that the<br />

harbor became inefficient and was abandoned due to<br />

its complete sediment infill. Several causes are taken<br />

in consideration to explain this infilling, these are: i)<br />

violent flooding following the collapse of a dam built<br />

to regulate the course of the wadi because of the<br />

large 365 A.D. Crete earthquake (Salza Prina Ricotti,<br />

1995; Di Vita, 1990); ii) lack of maintenance due to<br />

the decline of the settlement induced by severe<br />

damage after the 365 A.D. earthquake (Di Vita, 1990<br />

and 1995), or other local seismic sources (Guidoboni<br />

et al., 1994; Stiros, 2001); iii) inundation of a tsunami<br />

wave caused by the 365 A.D. earthquake that left a<br />

huge amount of debris and modified the local coastal<br />

morphology (Guidoboni et al., 1994; Ambraseys et<br />

al., 1994; Lorito et al., 2007, Shaw et al., 2008); iv)<br />

bad orientation and geometry of the harbor structures<br />

with respect to the local marine currents that were<br />

bringing sediments inside (Salza Prina Ricotti, 1972-<br />

1973). Most of these hypotheses are connected to<br />

the M8.5, 365 A.D. Crete event that, being the largest<br />

historical earthquake and tsunami ever occurred in<br />

the Mediterranean, has certainly stimulated the<br />

formulation of many hypothesis that put in connection<br />

any critical situation recognized in the mid of the 4th<br />

Century. with the occurrence of this catastrophe.<br />

159<br />

We studied in detail the geomorphology and the<br />

stratigraphy of Leptis Magna to understand if the<br />

causes of the decline of the town can be deciphered in<br />

the natural events recorded in the local geology (Pucci<br />

et al., 2011). To build up a consistent chronological<br />

record of significant natural events and their interaction<br />

with human history, geomorphological analyses (the<br />

study of cores and natural/artificial exposures), and the<br />

reconstruction of the different depositional<br />

environments based on the palaeontological content,<br />

were integrated to radiometric dating and<br />

archaeological knowledge.<br />

Given the strategic role of the harbor in the<br />

flourishing/decline of the town, this is certainly a key<br />

site to study to understand the history and men-nature<br />

interactions but, to fully understand the functioning of<br />

the natural and human system and equilibrium besides<br />

the harbor, we investigated the wadi, along with its<br />

hydraulic system built to control flooding (dam and<br />

aegere), the town, and the surrounding coastal area.<br />

At the same time we have surveyed the buildings in<br />

town to check for evidence of seismically induced<br />

collapses.<br />

The integration of all these observations allowed us to<br />

reconstruct the following history for the site (Pucci et<br />

al., 2011 - Figure 2):<br />

1) Building of hydraulic protections (dam and a<br />

channel diverting the wadi along a defensive aegere)<br />

to prevent flooding and have at disposal a water<br />

reserve, 32-131 A.D. No deposition in the lower reach<br />

of wadi (north of the dam), in town and in the harbor (B<br />

in figure 2).<br />

2) Functioning of the hydraulic system, expansion of<br />

the town toward the lower reach into the dry wadi bed<br />

and on its lower terraces. Enlargement of the harbor<br />

into the Severian monumental structure allowed by the<br />

lack of sediment/water discharge at the wadi outlet.<br />

Efficiency at least up to the time of Septimius Severus


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

(193-211 A.D.). Deposition only in the upper reach of<br />

the wadi (south of the dam), siltation effects cleared<br />

by artificial sweeping of the dam basin (C in figure 2).<br />

3) Maintenance (sweeping) of the hydraulic<br />

protections start to fail during the 3rd Century; onset<br />

of the dam siltation and beginning of the decadence<br />

of the town. In about 100 yrs the reservoir is<br />

completely filled; the wadi starts to spill out from the<br />

dam with the first alluvial deposits passing again the<br />

lower wadi reach and harbor area and with beginning<br />

of erosion at the base of the right shoulder of the dam<br />

(D in figure 2).<br />

4) Collapse of the right shoulder of the dam,<br />

entrenchment of the siltation body, restoration of the<br />

natural equilibrium of the wadi longitudinal profile by<br />

remobilization of a large amount of deposits that were<br />

previously stored in the dam basin. At 320-440 A.D.<br />

the wadi floods again the alluvial plain, begins to bury<br />

the town with alluvial sediments and to discharge large<br />

amounts of sediments into the harbor (E in figure 2).<br />

Figure 1. Geological and geomorphological map of the area of Leptis Magna. Locations S1-S7 are those studied in detail through<br />

stratigraphic and laboratory analyses. Inset in the lower left locates the study area on the north African coast (modified from Pucci<br />

et al., 2011).<br />

160


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

5) Harbor is completely infilled by wadi sediments at<br />

the 6th Century (F in figure 2).<br />

6) No evidence for important seismic collapses or<br />

damage was found by inspecting the buildings, as<br />

they appear today and the dam. Unfortunately, there<br />

was no possibility to check the original documents<br />

related to the early archaeological excavations to<br />

confirm the lack of major shaking effects. A wall in a<br />

recent excavation (Prof. F. Tommasello, pers.<br />

comm.), collapsed on top of 1 m-thick alluvium<br />

burying already the town, may be evidence for an<br />

earthquake. If seismic, this collapse occurred well<br />

after that repeated flooding draped the town with a<br />

thick layer of alluvium and thus post-dates 320-440<br />

A.D.<br />

Figure 2. Conceptual sketch of the evolution of the wadi Lebda longitudinal profile, from the dam to the harbor, and its depositional<br />

history in relation with the different studied locations (upper wadi, lower wadi, harbor, town - modified from Pucci et al., 2011).<br />

161


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The data collected, although not definitive to<br />

characterize the true cause for the decline of Leptis<br />

Magna, allow evaluating the initial hypotheses (Pucci<br />

et al., 2011).<br />

1) Flooding due to the collapse of the dam because<br />

of the 365 A.D. earthquake is discarded. This is<br />

because no traces for earthquake effects were found<br />

in the existing structures (including the dam), but also<br />

because no water was present in the dam basin at<br />

the time of the dam shoulder collapse. In fact, as it is<br />

today, the basin should have been completely filled<br />

with sediments due to the siltation process incepted<br />

by lack of dam sweeping. Should the dam shoulder<br />

have collapsed when the basin was still containing<br />

water (to produce the violent flooding) no siltation<br />

would have ever occurred as the dam was no more<br />

efficient and the wadi would have re-gained its<br />

course and profile.<br />

2) Lack of maintenance due to severe damage to<br />

the town after the 365 A.D. This hypothesis is partly<br />

correct. In fact, lack of maintenance is the primary<br />

cause for siltation in the dam and for all the<br />

consequent effects (flooding in town and sediments<br />

infilling in the harbor) that made the initial<br />

(economic?) decline irreversible. However, the<br />

reason for this lack of maintenance does not seem<br />

related to the damage from an earthquake (and<br />

certainly not the 365 A.D. Crete event) as no serious<br />

evidence for seismically induced damage was found.<br />

3) 365 A.D. tsunami harbor inundation is discarded<br />

as well. Coring in the harbor we reconstructed the<br />

whole history of the sedimentation in that area, from<br />

marine to continental. No evidence for such a violent<br />

event was found. A possible layer representing the<br />

deposit from a paleotsunami was found in the upper<br />

part of harbor stratigraphy, when alluvium was<br />

already deposited in the harbor after the 6th Century.<br />

4) Bad orientation and geometry of the harbor as the<br />

cause of debris deposition from marine currents is<br />

clearly discarded because the reconstruction of the<br />

depositional environments in the harbor has shown<br />

initial marine conditions (up to the 3rd Century) that<br />

gradually had a continental input to become purely<br />

alluvial (5th Century). If the initial hypothesis was<br />

correct, the harbor infill would have only a marine<br />

origin.<br />

In conclusion, we can rule out the hypothesis that an<br />

earthquake, and particularly the Crete 365 A.D., was<br />

the cause for the decadence of Leptis Magna. On the<br />

contrary, from the analyses of the recent geological<br />

history of the site it is likely that social and economic<br />

criticalities (possibly related to the continuous<br />

robberies and attacks from the Austurians) are the<br />

primary cause for this decadence that made the<br />

Lepticians unable to cope with natural hazards and<br />

preserve their “risk management plans”. The fact that<br />

nature took up the place again, causing widespread<br />

flooding and huge sediments remobilization made it<br />

too difficult for the Lepticians to recover and the initial<br />

decline of the town became irreversible.<br />

Acknowledgements: This work was funded by Istituto<br />

Nazionale di Geofisica e Geofisica e Vulcanologia — INGV<br />

with contribution of the EU Transfer project. We are thankful<br />

to the Italian Archaeological Mission leaded by L. Musso for<br />

hosting us and for the fruitful discussions and collaboration.<br />

References<br />

Ambraseys, N.N., Melville, C.P., Adams, R.D., (1994). The<br />

Seismicity of Egypt, Arabia and the Red Sea: a<br />

Historical Review. Cambridge Univ. Press, Cambridge.<br />

Di Vita, A., (1990). Sismi, urbanistica e cronologia assoluta.<br />

Terremoti e urbanistica nella città di Tripolitania tra il I<br />

secolo a.C. ed il IV d.C. In: L’Afrique dans L’Occident<br />

Romani (Ier Siècle Av. J.-C.- IVe Siècle Ap. J.- C.),<br />

Actes du Colloque Organisé par l’Èrcole Français de<br />

Rome sous le Patronage de l’Institut National<br />

d’Archéologie et d’Art de Tunis (Rome, 3-5 décembre<br />

1987), Roma, pp. 425e494.<br />

Di Vita, A., (1995). Archaeologists and earthquakes: the<br />

case of 365 AD. Annali di Geofisica 38, 971-976.<br />

Guidoboni, E., Comastri, A., Traina, G., (1994). Catalogue<br />

of Ancient Earthquakes in the Mediterranean Area up to<br />

the 10th Century. ING SGA, Rome.<br />

Lorito, S., Tiberti, M.M., Basili, R., Piatanesi, A., Valensise,<br />

G., (2007). Earthquake generated tsunamis in the<br />

Mediterranean Sea: scenarios of potential threats to<br />

Southern Italy. Journal of Geophysical Research,<br />

doi:10.1029/2007JB004943.<br />

Pucci, S., D. Pantosti, P.M. De Martini, A. Smedile, M.<br />

Munzi, E. Cirelli, M. Pentiricci, L. Musso, (2011).<br />

Environment-man relationships in historical times: The<br />

balance between urban development and natural forces<br />

at Leptis Magna (Libya), Quaternary International,<br />

doi:10.1016/j.quaint.2011.03.050<br />

Salza Prina Ricotti, E., (1972-1973). I porti della zona di<br />

Leptis Magna. Rendiconti della Pontificia Accademia di<br />

Archeologia 45, 75-103.<br />

Salza Prina Ricotti, E., (1995). Leptis Magna la città delle<br />

ombre bianche. Archeo 9 (127), 50-91.<br />

Shaw, B., Ambraseys, N.N., England, P.C., Floyd, M.A.,<br />

Gorman, G.J., Higham, T.F.G., Jackson, J.A., Nocquet,<br />

J.M., Pain, C.C., Piggott, M.D., (2008). Eastern<br />

Mediterranean tectonics and tsunami hazard inferred<br />

from the AD 365 earthquake.Nature Geoscience 1, 268-<br />

276. doi:10.1038/ngeo151.<br />

Stiros, S.C., (2001). The AD 365 Crete earthquake and<br />

possible seismic clusteringduring the fourth to sixth<br />

centuries AD in the Eastern Mediterranean: a reviewof<br />

historical and archaeological data. Journal of Structural<br />

Geology 23, 545-562.<br />

162


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ON-SHORE PROLONGATION OF BATHYMETRICALLY RECOGNIZED FAULT ZONES<br />

BASED ON GEODETIC GPS OBSERVATIONS ALONG SANTORINI VOLCANO<br />

Papageorgiou, Elena (1, Nomikou, Paraskevi (2)<br />

(1) University of Athens, Department of Geophysics and Geothermy, Zografou 15784, Greece, Email: epapageo@geol.uoa.gr<br />

(2) University of Athens, Department of Dynamic, Tectonic and Applied Geology, Zografou 15784, Greece,<br />

Email: evi@ath.hcmr.gr<br />

Abstract (On-shore prolongation of bathymetrically recognized fault zones based on geodetic GPS observations along<br />

Santorini volcano): Ground deformation monitoring on Santorini Volcanic Complex (SVC) revealed different horizontal kinematics<br />

in several parts of the volcano. However, in the current reposed state of SVC, ground deformation is primarily the result of tectonic<br />

activity. This work aims to combine and interpret the results of both GPS measurements and bathymetric survey conducted at<br />

SVC. The GPS data were intermittently collected during numerous campaigns between 1994 and 2005, while bathymetric survey<br />

was completed by the Hellenic vessel R/V ‘AEGAEO’ in November 2001, using the 20 kHz, SEABEAM 2120 swath system and<br />

mapping a total area of 2480 km². In this attempt a correlation between the morpho-tectonic features exposed on the seabed by<br />

the bathymetric survey, and the surface displacements observed by the GPS measurements could be finally achieved. An<br />

interpretation will be held, on whether these submarine tectonic features continue and intersect the volcanic island, especially in<br />

the areas where the surface displacements enable the plausible existence of contemporary tectonics.<br />

Key words: Tectonics, Bathymetry data, GPS Displacements, Santorini Volcano<br />

INTRODUCTION<br />

Santorini Volcanic Complex (SVC) lies in an area of<br />

complex extensional and subduction-related<br />

tectonics in a continental environment (Le Pichon &<br />

Angelier, 1979; Fytikas et al., 1984; Papanikolaou,<br />

1993; Jackson, 1994). The development of the<br />

volcanic field has been strongly influenced by<br />

regional faults.The current geodynamic state of SVC<br />

is controlled by two major volcanoes, the Nea<br />

Kammeni Volcano built up in the central caldera and<br />

the submarine Columbo Volcano, NE of Cape<br />

Columbo. Two major faulting zones Kammeni and<br />

Columbo (KFZ, CFZ, respectively) cross the<br />

volcanoes in a NE-SW direction, for several km,<br />

forming vents through which the magma ascend to<br />

the surface (Fouqué 1879; Reck 1936). Generally,<br />

the major structural features in Santorini are striking<br />

in a NE direction (Heiken & McCoy 1984; Druitt et al.,<br />

1989, 1999; Jackson 1994; Mountrakis et al., 1998;<br />

Pe-Piper & Piper, 2005).<br />

Seismic data (Bohnhoff et al., 2006; Dimitriadis et al.<br />

2009) indicate that the main volcanic centers of the<br />

Santorini volcanic field (Christiana, Santorini,<br />

Columbo and minor volcanic cones) are aligned<br />

along the NE-SW trending Santorini-Columbo<br />

volcano-tectonic line, a deep-seated, strike-slip<br />

feature (Sakellariou et al., 2010; Nomikou et al.,<br />

2011a) which served as conduit for the rising<br />

magma.<br />

Herein, we present the joint data interpretation of<br />

GPS measurements and bathymetry survey of the<br />

surrounding area of Santorini, in an attempt to relate<br />

the submarine rupture system, with its onshore<br />

163<br />

continuation as deduced from surface displacement<br />

variations on either side of the possible prolonged<br />

faults.<br />

GPS OBSERVATIONS<br />

DGPS has proven to be an important tool for<br />

intermediate and long-term monitoring of active<br />

volcanoes, as it provides high resolution 3-D<br />

information about the deformation field (e.g. Dvorak<br />

& Dzurisin, 1997; Fernandez et al., 1999 and many<br />

others). At the SVC, DGPS measurements<br />

conducted during the period 1994-2005 have<br />

detected non-trivial amounts of ground displacement,<br />

in both the horizontal and vertical axes (e.g.<br />

Papageorgiou et al., 2007). This enabled the<br />

identification of several domains with different<br />

kinematic characteristics. The interested reader may<br />

find details about the field measurement and data<br />

reduction procedures in Papageorgiou et al. (2007,<br />

2010).<br />

In particular, the GPS analysis presents several<br />

distinct domains with different horizontal kinematics,<br />

which could be defined most possibly by the major<br />

tectonic fault zones of the area. Thus, areas with<br />

uniform GPS displacements both in rate and<br />

orientation that characterize homogeneous regions<br />

were separated among others with considerable<br />

changes in their kinematic pattern. The observed<br />

differential deformation between these deforming<br />

domains is thought primarily to be the result of<br />

differential motion across faulting structures located<br />

at the boundaries. For the purpose of this paper, a<br />

synthetic map was constructed showing the above<br />

mentioned homogeneous regions, based merely on


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

their kinematics, as well as the areas/zones which<br />

may correspond to the major regional faults, as they<br />

still affect the tectonic–volcanic evolution of the SVC,<br />

as has already been observed in the past (Fig. 1).<br />

Fig. 1: Horizontal displacements relative to MTPI (NE Thera) for the period 1994-2005 (Papageorgiou et al., 2010).<br />

In Northern Thera, in between the GPS stations 33<br />

and 43 a different kinematic pattern is observed,<br />

indicating the possible existence of an area/zone<br />

(case I) of kinematic change (Fig. 1). Similarly, in<br />

central caldera, at Nea Kammeni Volcano, a second<br />

kinematic differentiation is observed, between the<br />

northern GPS stations 15 and 22 and the central<br />

GPS station 45 (case II). In both cases (I and II), fault<br />

zones are implicated to consist the boundary of the<br />

different homogeneous deforming domains, where in<br />

either side of which the horizontal GPS<br />

displacements change both in rate and orientation<br />

(Fig. 1).<br />

This is a very complex pattern and certainly difficult<br />

to interpret. Nevertheless, the observed deformation<br />

pattern enables the drafting of a qualitative model of<br />

contemporary tectonics, which is also presented and<br />

discussed in detail in Papageorgiou et al. (2010).<br />

However, the most important differences of surface<br />

displacements which are observed for the time<br />

interval 1994-2005 are thought to be explained with<br />

the existence of tectonic features. As it appears,<br />

these features coincide with the major fault zones in<br />

the area, the Kammeni, and Columbo Fault Zones<br />

(Fig. 2).<br />

BATHYMETRIC SURVEY<br />

The multibeam bathymetric survey was completed by<br />

the Hellenic vessel R/V ‘AEGAEO’ in November<br />

2001, using the 20 kHz, SEABEAM 2120 swath<br />

system and mapping a total area of 2480 km². The<br />

164<br />

bathymetric map was produced using a 50 meter grid<br />

interval and plotted with a Mercator projection at a<br />

scale of 1:100.000 with 10m contours (Alexandri et<br />

al., 2003; Nomikou et al., 2011a). What comes out<br />

from the bathymetric map (Fig. 2) is the identification<br />

of several morphological discontinuities that finally<br />

correspond to fault zones as indicated from seismic<br />

profiling data (Fig. 3) (Sakellariou et al., 2010;<br />

Nomikou et al., 2011). The volcanic cones in the<br />

regional area of the SVC presented in Fig. 3 show<br />

linear distribution which is controlled by strike-slip<br />

faults that run parallel to the long axis of the<br />

Anydhros basin.<br />

Furthermore, the CFZ is an active, possibly rightlateral,<br />

40 km long, strike-slip fault-zone, which has<br />

enabled the upward migration of magmatic fluids, the<br />

creation of dikes and the development of submarine<br />

volcanic cones, with Columbo submarine volcano as<br />

the most productive among them (Sakellariou et al.,<br />

2010). The KFZ in addition, passing through Nea<br />

Kammeni volcano, in the centre of the caldera is<br />

directly related to CFZ and displays similar<br />

characteristics.<br />

A particularly noteworthy feature that should be<br />

mentioned is a hydrothermal vent field at the northern<br />

caldera that is located in line with the normal fault<br />

system of the Columbo rift, and also near the margin<br />

of a shallow intrusion that occurs within the<br />

sediments of the north caldera (Sigurdsoon et al.,<br />

2006; Nomikou et al., 2011).


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: Multibeam Bathymetric Map of Santorini Volcanic Field using 20m isobaths (Nomikou et al., 2011).<br />

Fig. 3: Tectonic map of Santorini-Columbo volcanic field (Sakellariou et al., 2010).<br />

165


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

DISCUSSION<br />

The results of GPS measurements agree with the<br />

swath bathymetry data, as the observed submarine<br />

tectonic fault zones seem to prolongate onshore<br />

crosscutting the whole of the volcanic edifice. This<br />

phenomenon is more obvious in Northern Thera with<br />

the continuation of the Columbus Fault Zone up to<br />

the northern part of the Santorini caldera. The GPS<br />

horizontal displacements vectors show evidently<br />

different kinematic pattern on both sides of the CFZ,<br />

agreeing for the distinct existence of the CFZ and its<br />

continuation on land. However, a consideration<br />

should be taken into account on whether the GPS<br />

data reflect the existence of two different tectonic<br />

blocks on both sides of the CFZ, or whether it may<br />

indicate an updoming due to ascending magma<br />

along the fault zone which at the same time is used<br />

by the volcanic dikes.<br />

This is also evident in other parts of the SVC. The<br />

displacement field in Nea Kammeni volcano, is<br />

differentiating in each side of the Kammeni Fault<br />

Zone (KFZ), indicating primarily two kinematic blocks<br />

bounded by the KFZ, and secondly its onshore<br />

prolongation at Nea Kammeni.<br />

However, the GPS results show a more complex<br />

kinematic model of the SVC and further numerical<br />

modelling should be considered in order to get a<br />

more representative tectonic regime. Yet, this study<br />

enabled the combination of two different methods, as<br />

is the GPS observations and bathymetric data, in<br />

order to successfully certify the existence of major<br />

fault zones on the surface.<br />

Acknowledgements: Part of the GPS raw data was given<br />

by Prof. E. Lagios. Part of the collection and analysis of<br />

GPS data was financed by (i) The European Union (75%),<br />

(ii) The General Secretariat for Research & Technology<br />

(25%), and (iii) Terramentor E.E.I.G. Funding for offshore<br />

work at Santorini Volcanic Field was provided by<br />

GEOWARN project. The European Commission is<br />

acknowledged for their financial contribution to the project.<br />

The officers and the crew of the R/V AEGAEO are gratefully<br />

acknowledged for their important and effective contribution<br />

to the field work.<br />

References<br />

Alexandri, M., D. Papanikolaou & P. Nomikou (2003).<br />

Santorini volcanic field - new insights based on swath<br />

bathymetry. Abstracts IUGG, 30 June-11July 2003,<br />

Sapporo, Japan.<br />

Bohnhoff, M., M. Rische, Th. Meier, D. Becker, G.<br />

Stavrakakis & H-P. Harjes (2006). Microseismic activity in<br />

the Hellenic Volcanic Arc, Greece, with emphasis on the<br />

seismotectonic setting of the Santorini–Amorgos zone.<br />

Tectonophysics, 423, 17–33.<br />

Dimitriadis, I., E. Karagianni, D. Panagiotopoulos, C.<br />

Papazachos, P. Hatzidimitriou, M. Bohnhoff, M. Rische, &<br />

T. Meier (2009). Seismicity and active tectonics at<br />

Coloumbo Reef (Aegean Sea, Greece): Monitoring an<br />

active volcano at Santorini Volcanic Center using a<br />

temporary seismic network. Tectonophysics, 465, 136-<br />

149.<br />

Druitt, T.H., R.A. Mellors, D.M. Pyle, & R.S.J. Sparks<br />

(1989). Explosive volcanism on Santorini, Greece. Geol.<br />

Mag., 126, 95-126.<br />

Druitt, T.H., L. Edwards, R.A. Mellors, D.M. Pyle, R.S.J.<br />

Sparks, M. Lanphere, M. Davies, & B. Barreirio (1999).<br />

Santorini Volcano. Geol. Soc. Mem., 19, 165.<br />

Dvorak, J.J., & D. Dzurisin (1997). Volcano geodesy; the<br />

search for magma reservoirs and the formation of<br />

eruptive event. Rev. Geophys., 35, 343–384.<br />

Fernández, J., J.M. Carrasco, J.B. Rundle, & V. Arána<br />

(1999). Geodetic methods for detecting volcanic unrest: A<br />

theoretical approach. Bull. Volcanol., 60, 534–544.<br />

Fouqué, F., 1879. Santorin et ses eruptions. Masson et<br />

Cie., Paris, 440p.<br />

Fytikas, M., F. Innocenti, P. Manetti, R. Mazuoli, A.<br />

Peccerilo & L. Vilari (1984). Tertiary to Quaternary<br />

evolution of the volcanism in the Aegean Region. In<br />

Dixon J.E. & Robertson A.H.F. (eds.). The Geological<br />

Evolution of the Eastern Mediterranean. Geol. Soc.<br />

London, Spec. Pub., 17, 687-699.<br />

Heiken, G., & F. McCoy (1984). Caldera development<br />

during the Minoan eruption, Thera, Cyclades, Greece. J.<br />

Geophys. Res., 89, 8441-8462.<br />

Jackson, J.A. (1994). Active tectonics of the Aegean region.<br />

Annual Reviews of Earth and Planetary Sciences, 22,<br />

239-271.<br />

Le Pichon, X., & J. Angelier (1979). The Hellenic Arc and<br />

trench system: a key to the Neotectonic evolution of the<br />

Eastern Mediterranean area. Tectonophysics, 60, 1-42.<br />

Mountrakis, D., S. Pavlides, A. Chatzipetros, S. Meletidis,<br />

M. Tranos, G. Vougioukalakis & A. Kilias (1998). Active<br />

deformation in Santorini. In: Casale R., Fytikas M.,<br />

Sigvaldarsson, G., & G. Vougioukalakis (eds): The<br />

European laboratory volcanoes. European Commission,<br />

EUR 18161, 13-22.<br />

Nomikou, P., S. N. Carey, D. Papanikolaou, K.L.C Bell, D.<br />

Sakellariou, M. Alexandri, & K. Bejelou (2011a).<br />

Exploration of the submarine cones in the Kolumbo<br />

Submarine Volcanic Zone of the Hellenic Arc (Aegean<br />

Sea, Greece). Journal of Global and Planetary Change,<br />

In review.<br />

Nomikou, P., D. Papanikolaou, M. Alexandri,& D.<br />

Sakellariou (2011b). Submarine Volcanoes along the<br />

Agean Volcanic Arc. Tectonophysics, In review.<br />

Papanikolaou, D. (1993). Geotectonic evolution of the<br />

Aegean. Bull. Geol. Soc. Greece, XXVII, 33-48.<br />

Papageorgiou, E., E. Lagios, S. Vassilopoulou & V. Sakkas,<br />

(2007). Vertical & Horizontal Ground Deformation of<br />

Santorini Island deduced by DGPS measurements.<br />

<strong>Proceedings</strong> of the 11th International Conference Geol.<br />

Soc. Greece, Athens, Greece. Bull. Geol. Soc. Greece,<br />

40 (3), 1219-1225.<br />

Papageorgiou E., A. Tzanis, P. Sotiropoulos, & E. Lagios<br />

(2010). DGPS and magnetotelluric constraints on the<br />

contemporary tectonics of Thera Island, Greece. Bull.<br />

Geol. Soc. Greece, XLIII, 1,344-356.<br />

Pe-Piper, G. & D.J.W. Piper (2005). The South Aegean<br />

active volcanic arc: relationships between magmatism<br />

and tectonics. Develop. in Volc., 7, 113-133.<br />

Sakellariou D., H. Sigurdsson, M. Alexandri, S. Carey, G.<br />

Rousakis, P. Nomikou, P. Georgiou & D. Ballas (2010).<br />

Active tectonics in the Hellenic volcanic arc: The<br />

Kolumbo submarine volcanic zone. Bull. Geol. Soc.<br />

Greece, XLIII, 2,1056-1063.<br />

Sigurdsson, H., S. Carey, M. Alexandri, G. Vougioukalakis,<br />

K. L. Croff, C. Roman, D. Sakellariou, C. Anagnostou, G.<br />

Rousakis, C. Ioakim, A. Gogou, D. Ballas, T. Misaridis,<br />

and P. Nomikou 2006. Marine investigations of Greece’s<br />

Santorini volcanic field. EOS, 87(34):337, 342.<br />

Reck, H. (1936). Santorini. –Der Werdergang eines<br />

Inselvulcans und sein Ausbruch 1925-1928. Dietrich<br />

Reimer, Berlin, 3 vols. MANCA.<br />

166


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE DYNAMIC ANALYSIS OF MULTI-DRUM ANCIENT STRUCTURES UNDER<br />

EARTHQUAKE EXCITATIONS<br />

Loizos Papaloizou (1), Petros Komodromos (2)<br />

(1) University of Cyprus, Department of Civil and Environmental Engineering, School of Engineering, University of Cyprus<br />

75 Kallipoleos Street, PO Box 20537, 1678 Nicosia, Cyprus. Email: loizop@ucy.ac.cy<br />

(2). University of Cyprus, Department of Civil and Environmental Engineering, School of Engineering, University of Cyprus<br />

75 Kallipoleos Street, PO Box 20537, 1678 Nicosia, Cyprus. Email: komodromos@ucy.ac.cy<br />

Abstract (The dynamic analysis of multi-drum ancient structures under earthquake excitations): The seismic behaviour of<br />

ancient monumental structures with monolithic or multi-drum classical columns and colonnade systems in two rows is investigated.<br />

In particular, the Discrete Element Method (DEM) is utilized in the study of ancient columns under strong ground excitations, by<br />

simulating the individual rock blocks as distinct bodies. A specialized software application is developed, using a modern objectoriented<br />

programming language, in order to enable the effective simulation of multi-drum columns and colonnades. A number of<br />

parametric studies is performed in order to investigate the effect of excitation characteristics on the behaviour of multi-drum<br />

columns under harmonic and earthquake excitations. The simulations reveal that the columns and colonnades have the capacity<br />

to successfully withstand earthquakes excitations that were selected from regions where these monuments are often built.<br />

Key words: Ancient Columns, Earthquake, DEM.<br />

INTRODUCTION<br />

Strong earthquakes are common causes of<br />

destruction of ancient monuments, such as classical<br />

columns and colonnades. Ancient classical columns<br />

of great archaeological significance can be<br />

abundantly found in high seismicity areas in the<br />

Eastern Mediterranean. Multi-drum columns are<br />

constructed of stone blocks that are placed on top of<br />

each other, with or without connecting material<br />

between the individual blocks. The seismic behaviour<br />

of these structures exhibits complicated rocking and<br />

sliding phenomena between the individual blocks of<br />

the structure that very rarely appear in modern<br />

structures.<br />

Entablature<br />

Column<br />

Cornice<br />

Frieze<br />

Architrave<br />

Capital<br />

Shaft<br />

Fig. 1: Architecture of a typical classical monument.<br />

In ancient Greece the temples formed the most<br />

important class of buildings erected during that era<br />

and can be classified into three "Orders of<br />

167<br />

Architecture", the Doric, Ionic and Corinthian order.<br />

An "order" in Greek architecture consists of the<br />

column, including the base and the capital, and the<br />

entablature (Fig. 1). The entablature is divided into<br />

the architrave (lower part), the frieze (middle part)<br />

and the cornice (upper part). The differences among<br />

these three orders refer on the dimensions,<br />

proportions, mouldings and decorations of the<br />

various parts.<br />

Today, the remains of most of these temples are<br />

often limited to series of columns with an entablature<br />

or only an architrave, and in some cases only<br />

standalone columns. The investigation of the seismic<br />

behaviour of such monuments is scientifically<br />

interesting, as it involves complicated rocking and<br />

sliding responses of the individual rock blocks. The<br />

understanding of the seismic behaviour of these<br />

structures contributes to the rational assessment of<br />

efforts for their structural rehabilitation and may also<br />

reveal some information about past earthquakes that<br />

had struck the respective region.<br />

Ancient monuments, compared to modern structures,<br />

have been exposed to large numbers of strong<br />

seismic events throughout the many centuries of their<br />

life spans. Those that survived have successfully<br />

withstood a natural seismic testing that lasted for<br />

several centuries. Thus, it is important to understand<br />

the mechanisms that enabled them to avoid<br />

structural collapse and destruction during strong<br />

earthquakes. Since analytical study of such multiblock<br />

structures under strong earthquake excitations<br />

is practically difficult, if not impossible for large<br />

numbers of blocks, while laboratory tests are very<br />

difficult and costly, numerical methods can be used<br />

to simulate their seismic response.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

very extensive review of the literature on the usage<br />

of numerical methods for the analysis of monuments<br />

until 1993 was published by Beskos (1993). The<br />

dynamic behaviour of infinitely rigid bodies during<br />

horizontal excitations was studied by Housner<br />

(1963), while, later on, other researchers (Psycharis<br />

et al., 2003, Pompei et al., 1998, Makris & Zhang,<br />

1998, Manos et al., 2001, Komodromos et al., 2008)<br />

investigated further, both analytically and<br />

experimentally, the required conditions to overturn<br />

rigid bodies. Such structures can be simulated<br />

utilizing the Discrete Element Methods (DEM), which<br />

have been specifically developed for systems with<br />

distinct bodies that can move freely in space and<br />

interact with each other with contact forces through<br />

an automatic and efficient recognition of contacts.<br />

Research efforts to use the DEM in the simulation of<br />

ancient structures have already shown promising<br />

results, motivating further utilization of this method.<br />

Recent research work based on commercial DEM<br />

software applications (Psycharis et al., 2003,<br />

Papantonopoulos, 2002), demonstrated that the DEM<br />

can be reliably used for the analysis of such<br />

structures, although they reported a sensitivity of the<br />

response to small perturbations of the characteristics<br />

of the structure or the excitation. However, similar<br />

sensitivity has also been observed in experiments<br />

with classical columns (Mouzakis et al., 2002).<br />

Hence, it is important to perform large numbers of<br />

simulations with varying earthquake characteristics<br />

and design parameters to properly assess and<br />

interpret the simulation results.<br />

Latest research studies in the fields of<br />

paleoseismology and archaeoseismology (Hinzen et<br />

al., 2010, Caputo et al., 2011) investigate the<br />

damage in ancient monument structures and propose<br />

various quantitative models to test the seismogenic<br />

hypothesis of observed damage. Papaloizou and<br />

Komodromos (2011) used the Discrete Element<br />

Method (DEM) as well as a modern object-oriented<br />

design and programming approach, in order to<br />

examine the simulation of multi-drum columns and<br />

colonnades under harmonic and earthquake<br />

excitations.<br />

A custom-made DEM software has been specifically<br />

designed (Papaloizou and Komodromos, 2011) and<br />

implemented to enable efficient performance of large<br />

numbers of numerical simulations with varying<br />

parameters, modelling these structures with<br />

independent distinct bodies, as they are constructed<br />

in practice. Such simulations allow us to assess the<br />

influence of different earthquake characteristics as<br />

well as the various mechanical and geometrical<br />

parameters of these structures on their seismic<br />

responses.<br />

NUMERICAL ANALYSIS<br />

A large number of parametric studies is performed in<br />

order to investigate the effect of excitation<br />

characteristics as well as the influence of geometrical<br />

and mechanical characteristics of multi-drum<br />

columns on their behaviour and response under<br />

168<br />

harmonic and earthquake excitations. Colonnade<br />

systems with two rows of columns, one over the<br />

other, are examined with earthquake ground motions.<br />

T=36.3 sec<br />

T=37.0 sec<br />

Fig. 2: Mexico City Earthquake (scaled 1.5 Times).<br />

T=4.4 sec<br />

T=12.2 sec<br />

Fig. 3: Athens Earthquake (scaled 9 Times).<br />

Parametric studies have been performed by varying<br />

the excitation frequency and acceleration, as well as<br />

the friction coefficient and the geometric


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

characteristics of the simulated columns and<br />

colonnades in order to assess the influence of these<br />

parameters in the seismic response of the structure.<br />

Fig. 2 and Fig. 3 show snapshots (in different time<br />

steps) of the response of such structural systems for<br />

the Mexico City and the Athens Earthquakes,<br />

respectively (Table 1). The results show that the<br />

frequency content of the ground motion affects<br />

significantly the response. The displacements of the<br />

upper level columns in respect to the displacements<br />

of the lower level column are affected by the<br />

frequency content of the excitation.<br />

Specifically, for the Mexico City earthquake, which<br />

has low predominant frequencies, the upper level<br />

columns overturn. For the Athens Earthquake, which<br />

has higher predominant frequencies, and for a much<br />

larger maximum ground acceleration the upper level<br />

columns are not affected by the seismic excitation.<br />

Date<br />

and<br />

Time<br />

9/19/1995<br />

(13:19CT)<br />

9/7/1999<br />

(11:56:50)<br />

Earthquake<br />

Component<br />

Mexico City<br />

(270)<br />

Athens<br />

(N46)<br />

PGA<br />

(m/sec2)<br />

Predominant<br />

Frequencies<br />

(Hz)<br />

0.98 0.45-0.53<br />

3.01 4.1-8.3<br />

Table 1. List of earthquake records that have been used in<br />

the analyses.<br />

Moreover, for low frequency harmonic excitations,<br />

the exhibited response is dominated by rocking, while<br />

sliding prevails in cases of excitations with very high<br />

frequencies. In between the two extremes, the<br />

response contains both rocking and sliding<br />

phenomena.<br />

Furthermore, the results indicate that the required<br />

acceleration to initiate rocking or sliding decreases as<br />

the excitation frequency increases. The acceleration<br />

that is needed to overturn the column also increases<br />

as the frequency increases.<br />

By examining the stability of multi-drum columns and<br />

colonnades for earthquakes that were selected from<br />

regions, where these monuments are often built,<br />

such as the Eastern Mediterranean regions, the<br />

simulations reveal that the columns have the capacity<br />

to successfully withstand strong earthquakes. The<br />

required acceleration to overturn a column decreases<br />

as the predominant frequency of the earthquake<br />

decreases.<br />

The investigation of the dynamic response of such<br />

monumental structures, combined with the research<br />

fields of paleoseismology and archaeoseismology,<br />

may reveal certain information from past strong<br />

earthquakes that have struck the respective regions.<br />

The investigation of the response of multi-drum<br />

structures under different ground motions can help in<br />

defining the frequency content of old destructive<br />

earthquakes.<br />

Acknowledgements: The authors would like to thank the<br />

Cyprus Research Promotion Foundation for funding the<br />

active research project entitled “Investigation for the<br />

protection of ancient multi-drum columns and colonnades<br />

from strong earthquakes (//<br />

0609()/23)”.<br />

References<br />

Beskos D. (1993), “Use of finite and boundary elements in<br />

the analysis of monuments and special structures”,<br />

Association of Civil Engineers of Greece, 217.<br />

Caputo, R., Hinzen, K.-G., Liberatore, D., Schreiber, S.,<br />

Helly, B., Tziafalias, A., (2011), “Quantitative<br />

archaeoseismological investigation of the Great Theatre<br />

of Larissa, Greece “Bulletin of Earthquake Engineering. 9<br />

(2), pp. 347-366.<br />

Hinzen, K.-G., Fleischer, C., Reamer, S.K., Schreiber, S.,<br />

Schütte, S., Yerli, B., (2010), “Quantitative methods in<br />

archaeoseismology” Quaternary International, DOI:<br />

10.1016/j.quaint.2010.11.006.<br />

Housner G.W (1963), ‘The behavior of inverted pendulum<br />

structures during earthquakes”. Bulletin of seismological<br />

Society of America, 53, 403-417.<br />

Komodromos P., Papaloizou L., Polycarpou P., (2008)<br />

“Simulation of the response of ancient columns under<br />

harmonic and earthquake excitations”, Engineering<br />

Structures 30(8), 2154-2164.<br />

Makris N. and Zhang J., (2001) “Rocking response of<br />

anchored blocks under pulse-type motions”, Engineering<br />

Mechanics 127, 484-493.<br />

Manos G.C., Demosthenous M., Kourtides V.,<br />

Hatzigeorgiou A., (2001) “Dynamic and earthquake<br />

behavior of models of ancient columns and colonnades<br />

with or without energy absorbtions systems”,<br />

<strong>Proceedings</strong> of Second Greek National Conference on<br />

Earthquake Engineering and Seismology, 1, 257-276.<br />

Mouzakis H., Psycharis I., Papastamatiou D., Carydis P.,<br />

Papantonopoulos C., and Zambas C., (2002).<br />

“Experimental investigation of the earthquake response<br />

of a model of a marble classical column”, Earthquake<br />

Engineering and Structural Dynamics 31, 1681-1698.<br />

Papaloizou, L., Komodromos, P., (2011) “Investigating the<br />

seismic response of ancient multi-drum colonnades with<br />

two rows of columns using an object-oriented designed<br />

software”, Advances in Engineering Software, DOI:<br />

10.1016/j.advengsoft.2011.05.030.<br />

Papantonopoulos C., Psycharis I., Papastamatiou D.,<br />

Lemos J. and Mouzakis H., (2002). “Numerical prediction<br />

of the earthquake response of classical columns using<br />

the distinct element method”, Earthquake Engineering<br />

and Structural Dynamics 31, 1699-1717.<br />

Pompei A., Scalia A. and Sumbatyan M.A. (1998),<br />

“Dynamics of Rigid Block due to Horizontal Ground<br />

Motion”, Engineering Mechanics 124, 713-717.<br />

Psycharis I., Lemos J., Papastamatiou D., Zambas C. and<br />

Papantonopoulos C. (2003), “Numerical study of the<br />

seismic behaviour of a part of the Parthenon Pronaos”,<br />

Earthquake Engineering & Structural Dynamics 32, 2063-<br />

2084.<br />

169


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

NEOTECTONIC AND ACTIVE DIVERGING RATES OF EXTENSION IN THE NORTHERN<br />

AND SOUTHERN HELLENIDES ACROSS THE CENTRAL HELLENIC SHEAR ZONE<br />

Papanikolau, Dimitrios (1, Royden, Leigh (2), Vassilakis, Emmanuel (1)<br />

(1) School of Geology and Geoenvironment, Department of Dynamics, Tectonics and Applied Geology, National and Kapodistrian<br />

University of Athens, 15784, Greece. Email: dpapan@geol.uoa.gr<br />

(2) Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, 54-826, Cambridge, MA<br />

02139, USA<br />

Abstract (Neotectonic and active diverging rates of extension in the Northern and Southern Hellenides across the Central<br />

Hellenic Shear Zone): Present day location and geometry of the Hellenic arc and trench system is only a small portion of the<br />

previously developed Hellenic arc that created the Hellenides orogenic system. The timing of differentiation is constrained in Late<br />

Miocene, when the arc was divided in a northern and a southern segment. This is based on: a) the dating of the last compressive<br />

structures observed all along the Hellenides during Oligocene to Middle-Late Miocene, b) on the time of initiation of the Kephalonia<br />

transform fault, c) on the time of opening of the North Aegean Basin and d) on the time of opening of new arc parallel basins in the<br />

south and new transverse basins in the central shear zone, separating the rapidly moving southwestward Hellenic subduction<br />

system from the slowly converging system of the Northern Hellenides. The driving mechanism of the arc differentiation is the<br />

heterogeneity produced by the different subducting slabs in the north (continental) and in the south (oceanic) and the resulted<br />

shear zone because of the retreating plate boundary producing a roll back mechanism in the present arc and trench system. The<br />

extension produced in the upper plate has resulted in the subsidence of the Aegean Sea and the creation of several neotectonic<br />

basins in southern continental Greece in contrast to the absence of new basins in the northern segment since Late Miocene.<br />

Crustal thickness, structural profiles, earthquake mechanisms and GPS rates are compared in the two segments, showing<br />

significant extension in the Southern Hellenides.<br />

Key words: Hellenides, neotectonic extension, slab roll-back<br />

INTRODUCTION<br />

The active portion of the Hellenic orogen is an<br />

arcuate belt reaching from the central Adriatic Sea<br />

southwards and eastwards to western Anatolia<br />

(McKenzie, 1978; Le Pichon and Angelier, 1979;<br />

Picha, 2002; Reilinger et. al., 2006). Within the<br />

northwest-trending portion of the Hellenic belt, two<br />

segments can be distinguished by subduction rate<br />

(slow in the northwest, fast in the southeast) and<br />

water depth of the foreland lithosphere (shallow in<br />

the northwest, deep in the southeast; e.g. Dewey and<br />

Sengor, 1979; Finetti, 1982; McKenzie, 1978; Le<br />

Pichon and Angelier, 1979, Fig. 1). These segments,<br />

dextrally offset from one another near the island of<br />

Kephalonia, have been referred as the Northern and<br />

Southern Hellenides (Papanikolaou and Royden,<br />

2007; Royden and Papanikolaou, 2011).<br />

THE SEPARATION OF THE NORTHERN AND<br />

SOUTHERN HELLENIDES BY THE CENTRAL<br />

HELLENIC SHEAR ZONE<br />

Several years of GPS measurements indicate a<br />

convergence rate of ~5-10 mm/yr across the<br />

Northern Hellenic subduction boundary, as measured<br />

between stations on the subducting plate (Apulia)<br />

and on the over-riding plate in northern Greece<br />

(Hollenstein et al., 2008; Bennett et al., 2008;<br />

Vassilakis et al., 2011) and seismic activity and focal<br />

solutions for local earthquakes attest to continuing<br />

convergence in this region (e.g. Louvari et al., 1999;<br />

Papazachos et al., 2000). Based on evidence from<br />

170<br />

seismology, morphology and industry seismic data,<br />

the active thrust front of the Northern Hellenides lies<br />

just west of the Ionian islands of Corfu and Paxos<br />

(Monopolis & Bruneton, 1982; Vassilakis et al., 2011)<br />

(Fig. 1).<br />

Even though gravity data indicate that the basement<br />

is flexed downward beneath the thrust front and the<br />

resulting depression filled with sedimentary foredeep<br />

deposits (Moretti and Royden, 1988), no trench is<br />

present in the bathymetry along the Northern<br />

Hellenides. For ease of reference we will refer to this<br />

zone of convergence as the Northern Hellenic trench,<br />

despite the fact that the trench has been entirely filled<br />

with sediments. The lithosphere entering the<br />

Northern Hellenic trench is continental or transitional<br />

in character, with a crustal thickness of ~25-30 km<br />

(Morelli et al, 1975; Marone et al, 2003, Cassinis et.<br />

al., 2003). Modern water depths near the thrust belt<br />

are generally ~1 km or less and overlie a shallow<br />

water sedimentary sequence of Triassic through<br />

Pliocene age (Jacobshagen et al., 1978).<br />

On the contrary the same GPS data indicate a<br />

convergence rate of ~35 mm/yr across the Southern<br />

Hellenides, as measured between Africa and points<br />

in the over-riding (Aegean) domain (McClusky et al.,<br />

2000; Reilinger et al., 2006). Behind the Southern<br />

Hellenides, a Benioff zone reaches to ~150 km depth<br />

(e.g. Papazachos et al., 2000) and an active volcanic<br />

arc is present ~200 km behind the trench (Fytikas et<br />

al., 1984). The zone along which basement is<br />

subducted beneath the Hellenides lies ~50 km west


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

of the southwestern coast of the Peloponnesus,<br />

passing beneath the deepest portions of the Hellenic<br />

trench (Fig. 1). Here, earthquake hypocenters and<br />

gravity data indicate that the depth to basement is<br />

~12-15 km depth (Royden 1993; Hirn et al., 1997;<br />

Clément et al., 2000; Sachpazi et al., 2000). Thrust<br />

faults and folds also occur within the accretionary<br />

prism outboard of the trench, over a width of several<br />

hundred kilometers to the Mediterranean ridge (Kopf<br />

et al., 2003), but thrusting here involves only<br />

sedimentary cover detached above the basement.<br />

The crust beneath the Ionian Sea is almost certainly<br />

oceanic, probably Triassic or Jurassic in age, and<br />

consists of approximately 8 km of crystalline crust<br />

overlain by 6-10 km of sedimentary cover (Makris,<br />

1985; Kopf et al., 2003). The water depth throughout<br />

much of the Ionian Sea region is 3-4 km, with the<br />

deepest depths of up to 5 km occurring along the<br />

Southern Hellenic Trench.<br />

the Northern Hellenides to ~3-4 km in front of the<br />

Southern Hellenides, particularly near the west coast<br />

of mainland Greece (Fig. 1). Here the Hellenic<br />

subduction boundary (trench) appears to be dextrally<br />

offset by ~100-150 km across the Kephalonia<br />

transform zone, although the precise offset is difficult<br />

to determine due to north-south variations in water<br />

depth and sediment thickness and due to the fact<br />

that some of the offset of the northern and southern<br />

segments is taken up by clockwise rotation of crustal<br />

units adjacent to the south side of the transform<br />

(Vassilakis et al., 2011). Northeastward, the<br />

Kephalonia transform zone extends into mainland<br />

Greece to merge with the broadly defined zone of<br />

dextral and extensional deformation in the Central<br />

Hellenic Shear Zone (Papanikolaou and Royden,<br />

2007; see also discussions by Goldsworthy et al.,<br />

2002; Armijo et al., 1996; Roberts and Jackson,<br />

1991).<br />

Global P-wave tomography indicates a northeastdipping<br />

zone of high P-wave speeds beneath the<br />

Southern and Northern Hellenides (Spakman et al.,<br />

1993; van der Hilst et al., 1997; Karason and van der<br />

Hilst, 2001; Wortel and Spakman, 2000; Suckale et<br />

al., 2009). Behind the Southern Hellenides, the<br />

subducted lithosphere appears to reach to the base<br />

of the upper mantle, perhaps deeper. A northeastdipping<br />

zone of high P-wave speed also exists north<br />

of the Kephalonia transform zone, but here the<br />

velocity contrast with the surrounding mantle is not<br />

as large as beneath the Southern Hellenides.<br />

COMPARISON OF TECTONIC STRUCTURES AND<br />

DEFORMATION RATES BETWEEN THE<br />

NORTHERN AND SOUTHERN HELLENIDES<br />

Fig. 1: Selected GPS velocities from McClusky et al. [2000]<br />

in a reference frame that minimizes velocities in the Aegean<br />

region. Some data were omitted in areas where high data<br />

density obscured the velocity pattern, mainly near the North<br />

Anatolian Fault zone. Shading indicates the zones of active<br />

oblique extension that bound the northwestern and<br />

northeastern margins of the largely undeforming Aegean<br />

domain.<br />

The northern and southern segments of the Hellenic<br />

subduction boundary are separated by the<br />

Kephalonia transform zone (e.g. Dewey and Sengor,<br />

1979; Finetti, 1982; Kahle and Muller, 1998; Kahle et<br />

al., 1995; Hollenstein et al., 2008), across which GPS<br />

data indicate ~25 mm/yr of dextral motion (Fig. 1).<br />

The Kephalonia transform separates the slowly<br />

subducting, continental foreland of the northern<br />

Hellenides from the rapidly moving upper plate of the<br />

southern Hellenides. Focal solutions from<br />

earthquakes located along the Kephalonia transform<br />

show right-slip on steep southwest-striking fault<br />

planes and also thrust faulting along northeaststriking<br />

fault planes.<br />

The Kephalonia transform coincides closely with a<br />

change in foreland water depth from ~1 km in front of<br />

171<br />

The difference of the tectonic structure and of the<br />

deformation rates in the Northern and Southern<br />

Hellenides can be studied in two time periods: the<br />

recent neotectonic period of the last 5 million years<br />

during the Pliocene-Pleistocene and the active<br />

deformation of the last 14000 years during the<br />

Holocene incorporating also the actual observations<br />

and measurements. The main points showing the<br />

diverging rates of extension in the two segments of<br />

the Hellenides are the following.<br />

1) The width of the Hellenic peninsula from the Ionian<br />

sea at the west to the Aegean Sea at the east, is<br />

longer approximately by 60 km across the southern<br />

profile which corresponds to the offset of the trench<br />

across the Kephalonia transform since the Aegean<br />

coast is rather rectilinear from the Olympus foothills<br />

to the Southern Evia and Andros-Tinos-Mykonos<br />

Cycladic islands.<br />

2) The crustal thickness in the Southern Hellenides is<br />

much thinner (less than 25 km) than in the Northern<br />

Hellenides (up to 45 km) (Makris, 1977; 1985).<br />

3) The tectonic structure of the Southern Hellenides<br />

is characterised by the occurrence of the External<br />

Tectonometamorphic Belt in Peloponnesus in the<br />

form of tectonic windows, which resulted from<br />

extensional detachments during Miocene


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

(Papanikolaou and Royden, 2007; Papanikolaou et<br />

al., 2009). Thus, the structure in the Northern<br />

Hellenides is rather simple with a regular nappe<br />

emplacement from the area west of the Olympus<br />

tectonic window up to the Ionian Sea in contrast to<br />

the more complex structure in the Southern<br />

Hellenides, where there are new metamorphic<br />

geotectonic units (such as the Mani and Arna units)<br />

appearing below the non metamorphosed nappes<br />

(such as the Tripolis and Pindos units) (Papanikolaou<br />

and Vassilakis, 2010).<br />

4) A number of arc parallel extensional neotectonic<br />

basins oriented NW-SE occur along the transverse<br />

profile of the Southern Hellenides (e.g. in Southern<br />

Peloponnesus) as described by Papanikolaou et al.<br />

1988 (Fig. 2). In contrast, no neotectonic basins are<br />

developed in the transverse profile of the Northern<br />

Hellenides with the exception of the Ioannina basin<br />

(the only extensive Plio-Quaternary structure west of<br />

the Oligo-Miocene Mesohellenic basin). The<br />

difference of extension in the two segments within<br />

the upper plate is more than 20%.<br />

5) The last onshore compressive structures in the<br />

Northern Hellenides are observed in Corfu and Parga<br />

areas, involving Late Miocene and early Pliocene<br />

sedimentary formations (including the Messinian<br />

evaporites) and in Kephalonia and Zakynthos islands<br />

in the Southern Hellenides with a dextral offset<br />

similar to the Kephalonia transform but somewhat<br />

smaller.<br />

6) The transition from compression to extension as<br />

determined by earthquake mechanisms occurs in<br />

central Epirus in the Northern Hellenides and in the<br />

shallow marine zone between the Ionian islands and<br />

mainland Greece in the Southern Hellenides.<br />

7) The rate of upper plate extension across the belt is<br />

~3-5mm/yr in the Northern Hellenides and more than<br />

30mm/yr in the Southern Hellenides.<br />

EPILOGUE<br />

All arguments displayed above point to the same<br />

conclusion of differentiation between the northern<br />

and southern Hellenides, regarding the amount of the<br />

extension both during the neotectonic period and the<br />

present day deformation. In many cases this<br />

differentiation can be quantified and the results are<br />

quite impressive showing that a major complex<br />

structure such as the Central Hellenic Shear Zone is<br />

acting as an oblique transform structure of the<br />

Hellenides by dividing the Hellenic peninsula and<br />

changing the present morphology.<br />

Acknowledgements: This research was supported by the<br />

Continental Dynamics Program at NSF, grant EAR-<br />

0409373.<br />

Fig. 2: In this vertically exaggerated (x10) SW-NE section across the Southern Hellenides the post-nappestacking extensional<br />

structures are indicated.<br />

172


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

References<br />

Armijo, R., et al. (1996), Quaternary evolution of the Corinth<br />

Rift and its implications for the late Cenozoic evolution of<br />

the Aegean, Geophys. J. Int, 126, 11-53.<br />

Bennett, R. A., et al. (2008), Eocene to present subduction<br />

of southern Adria mantle lithosphere beneath the<br />

Dinarides, Geology, 36(1), 3-6.<br />

Cassinis, R., et al. (2003), The deep crustal structure of Italy<br />

and surrounding areas from seismic refraction data: a<br />

new synthesis, Bollettino di Geofisica Teorica ed<br />

Applicata, 122, 365-375.<br />

Clément, C., et al. (2000), Seismic structure and the active<br />

Hellenic subduction in the Ionian islands,<br />

Tectonophysics, 329(1-4), 141-156.<br />

Dewey, J. F., and C. engör (1979), Aegean and<br />

surrounding regions: Complex multiplate and continuum<br />

tectonics in a convergent zone, Geological Society of<br />

America Bulletin, 90(1), 84-92.<br />

Finetti, I. (1982), Structure, stratigraphy and evolution of<br />

central Mediterranean (Pelagian Sea, Ionian Sea),<br />

Bollettino di Geofisica Teorica ed Applicata, 24(96), 247-<br />

312.<br />

Fytikas, M., et al. (1984), Tertiary to Quaternary evolution of<br />

volcanism in the Aegean region, in The geological<br />

evolution of the eastern Mediterranean, edited by J.<br />

Dixon and A. Robertson, pp. 687-701, Geological Society<br />

of London, London.<br />

Goldsworthy, M., et al. (2002), The continuity of active fault<br />

systems in Greece, Geophysical Journal International,<br />

148(3), 596-618.<br />

Hirn, A., et al. (1997), Increase in melt fraction along a<br />

south-north traverse below the Tibetan Plateau: evidence<br />

from seismology, Tectonophysics, 273(1-2), 17-30.<br />

Hollenstein, C., et al. (2008), Crustal motion and<br />

deformation in Greece from a decade of GPS<br />

measurements, 1993-2003, Tectonophysics, 449(1-4),<br />

17-40.<br />

Jacobshagen, V. (1986), Geologie von Griechenland, 363<br />

pp.<br />

Kahle, H. G., and S. Mueller (1998), Structure and<br />

dynamics of the Eurasian-African/Arabian plate boundary<br />

system: Objectives, tasks and resources of the<br />

WEGENER group, Journal of Geodynamics, 25(3-4),<br />

303-325.<br />

Kahle, H. G., et al. (1995), The strain field in northwestern<br />

Greece and the Ionian Islands: results inferred from GPS<br />

measurements, Tectonophysics, 249, 41-52.<br />

Kopf, A., et al. (2003), The Mediterranean Ridge: A mass<br />

balance across the fastest growing accretionary complex<br />

on Earth, J. Geophys. Res., 108(B8), 2372.<br />

Le Pichon, X., and J. Angelier (1979), The Hellenic arc and<br />

trench system: a key to the neotectonic evolution of the<br />

eastern Mediterranean area, Tectonophysics, 60, 1-42.<br />

Louvari, E., et al. (1999), The Cephalonia Transform Fault<br />

and its extension to western Lefkada Island (Greece),<br />

Tectonophysics, 308, 223-236.<br />

Makris, J. (1976), A dynamic model of the Hellenic Arc<br />

deduced from Geophysical data, Tectonophysics, 36,<br />

339-346.<br />

Makris, J. (1985), Geophysics and geodynamic implications<br />

for the evolution of the Hellenides, in Geological<br />

Evolution of the Mediterranean Basin, edited by D. J.<br />

Stanley and F. Wezel, pp. 231–248, Springer, Berlin<br />

Marone, F., et al. (2003), Joint inversion of local, regional<br />

and teleseismic data for crustal thickness in the Eurasia-<br />

Africa plate boundary region, Geophysical Journal<br />

International, 154(2), 499-514.<br />

McClusky, S., et al. (2000), Global Positioning System<br />

constraints on plate kinematics and dynamic in the<br />

eastern Mediterranean and Caucasus, Journal of<br />

Geophysical Research, 105(B3), 5695-5719.<br />

173<br />

McKenzie, D. (1978), Active tectonics of the Alpine-<br />

Himalayan belt: the Aegean Sea and surrounding<br />

regions, Geophysical Journal of the Royal Astronomical<br />

Society, 55(1), 217-254.<br />

Monopolis, D., and A. Bruneton (1982), Ionian sea (Western<br />

Greece): Its structural outline deduced from drilling and<br />

geophysical data, Tectonophysics, 83(3-4), 227-242.<br />

Morelli, C., et al. (1975), Bathymetry, gravity and magnetism<br />

in the Strait of Sicily and in the Ionian Sea, Bollettino di<br />

Geofisica Teorica ed Applicata, 17, 39-58.<br />

Moretti, I., and L. Royden (1988), Deflection, Gravity-<br />

Anomalies and Tectonics of Doubly Subducted<br />

Continental Lithosphere - Adriatic and Ionian Seas,<br />

Tectonics, 7(4), 875-893.<br />

Papanikolaou, D., (2010), Major palaeogeographic, tectonic<br />

and geodynamic changes from the last stage of the<br />

Hellenides to the actual Hellenic arc and trench system,<br />

Bulletin Geological Society of Greece, XLIII(1), 72-85.<br />

Papanikolaou, D., and E. Vassilakis (2010), Thrust faults<br />

and extensional detachment faults in Cretan tectonostratigraphy:<br />

Implications for Middle Miocene extension,<br />

Tectonophysics, 488(1-4), 233-247.<br />

Papanikolaou, D., and L. Royden (2007), Disruption of the<br />

Hellenic arc: Late Miocene extensional detachment faults<br />

and steep Pliocene-Quaternary normal faults - Or what<br />

happened at Corinth?, Tectonics, 26 (TC5003).<br />

Papanikolaou, D., et al. (1988), A comparative study of<br />

neotectonic basins across the Hellenic Arc: The<br />

Messiniakos, Argolikos, Saronikos and Southern Evoikos<br />

Gulfs, Basin Research, 1, 167-176.<br />

Papanikolaou, D., et al. (2009), The Itea-Amfissa<br />

detachment: a pre-Corinth rift Miocene extensional<br />

structure in central Greece, Geological Society, London,<br />

Special Publications, 311(1), 293-310.<br />

Papazachos, B. C., et al. (2000), The geometry of the<br />

Wadati-Benioff zone and lithospheric kinematics in the<br />

Hellenic arc, Tectonophysics, 319, 275-300.<br />

Picha, F. J. (2002), Late Orogenic Strike-Slip Faulting and<br />

Escape Tectonics in Frontal Dinarides-Hellenides,<br />

Croatia, Yugoslavia, Albania, and Greece, AAPG Bulletin,<br />

86(9), 1659-1671.<br />

Reilinger, R., et al. (2006), GPS constraints on continental<br />

deformation in the Africa-Arabia-Eurasia continental<br />

collision zone and implications for the dynamics of plate<br />

interactions, J. Geophys. Res., 111.<br />

Roberts, S., and J. Jackson (1991), Active normal faulting in<br />

central Greece: an overview, Geological Society, London,<br />

Special Publications, 56(1), 125-142.<br />

Royden, L. H. (1993), Evolution of retreating subduction<br />

boundaries formed during continental collision, Tectonics,<br />

12(3), 629-638.<br />

Royden, L., and D. Papanikolaou (2011), Slab<br />

segmentation and late Cenozoic disruption of the Hellenic<br />

arc, Geochem. Geophys. Geosyst., 12(3), Q03010.<br />

Sachpazi, M., et al. (2000), Western Hellenic subduction<br />

and Cephalonia Transform: local earthquakes and plate<br />

transport and strain, Tectonophysics, 319(4), 301-319.<br />

Spakman, W., et al. (1993), Travel-time tomography of the<br />

European-Mediterranean mantle down to 1400 km,<br />

Physics of The Earth and Planetary Interiors, 79(1-2), 3-<br />

74.<br />

Suckale, J., et al. (2009), High-resolution seismic imaging of<br />

the western Hellenic subduction zone using teleseismic<br />

scattered waves, Geophysical Journal International,<br />

178(2), 775-791.<br />

van der Hilst, R. D., et al. (1997), Evidence for deep mantle<br />

circulation from global tomography, Nature, 386(6625),<br />

578-584.<br />

Vassilakis, Emm., et al. (2011), Kinematic Links from the<br />

Hellenic Trench to the Western Gulf of Corinth. Earth &<br />

Planetary Science Letters, 303, 108-120.<br />

Wortel, M. J. R., and W. Spakman (2000), Subduction and<br />

Slab Detachment in the Mediterranean-Carpathian<br />

Region, Science, 290(5498), 1910-1917.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

CLUSTERING AND ANTICLUSTERING IN THE SOUTHERN APENNINES AS EVIDENCED<br />

FROM GEOLOGICAL FAULT SLIP-RATE SEISMIC HAZARD MAPS AND THE<br />

HISTORICAL RECORD<br />

Papanikolau Ioannis D. (1,2), Gerald Roberts (2)<br />

(1) Mineralogy-Geology Laboratory, Department of Earth and Atmoshperic Sciences, Agricultural University of Athens, Iera Odos<br />

75, 118-55, Athens, Greece<br />

(2) Department of Earth and Planetary Sciences, Birkbeck and University College London, Gower Street, WC1E 7HX. UK.<br />

i.papanikolaou@ucl.ac.uk, gerald.roberts@ucl.ac.uk<br />

Abstract (Clustering and anticlustering in the Southern Apennines as evidenced from geological fault slip-rate seismic<br />

hazard maps and the historical record): Geological fault slip rate seismic hazard maps were constructed in the area of the<br />

southern Apennines. By comparing short-term throw-rates (e.g. historical record) with longer-term throw-rates (e.g. offset<br />

postglacial geomorphic features), one might be able to differentiate areas that are currently in a cluster of earthquakes from areas<br />

that are currently in an anticluster, providing important insights into seismic hazards. In the southern Apennines, there seems to be<br />

an inconsistency between the historical record and the longer-term rates extracted from the geological fault slip-rate hazard maps,<br />

implying that the historical record is not representative of the longer-term throw-rates. This occurs because the northern part of the<br />

study area (Irpinia, northern Basilicata) appears to be ahead of its longer-term seismicity rate, thus perhaps in a temporal<br />

clustering period, whereas the southern part (Pollino, southern Basilicata) appears to be below its longer-term rate, implying that it<br />

is in a temporal anticlustering period. The latter agrees with the so-called Pollino seismic gap.<br />

Key words: Southern Italy, Pollino, active faults<br />

INTRODUCTION<br />

Gaps on seismicity maps are areas that constantly<br />

attract the interest of the scientific community. One<br />

fundamental question that puzzles scientists<br />

worldwide is whether a seismic gap on a seismicity<br />

map that is based on instrumental/historical data,<br />

could represent: a) an area of genuinely low longterm<br />

seismicity, leading to a correct low hazard<br />

interpretation; or b) a poor or incomplete historical<br />

record, suggesting that the region has experienced<br />

large earthquakes in the past, but they have not been<br />

recorded because the catalogues were either too<br />

short or incomplete; or c) a quiescent period in an<br />

area characterised by temporal clustering followed by<br />

a long recurrence interval (e.g. Marco et al. 1996).<br />

Scenario a suggests a low hazard whereas,<br />

scenarios b and c imply a high hazard potential. In<br />

particular, in scenario b the time elapsed since the<br />

last event, exceeds the duration of historic<br />

catalogues, implying that the region could probably<br />

have entered the last stage of the seismic cycle. In<br />

scenario c, if a perceived seismic gap coincides with<br />

a quiescent interval, then the region sooner or later<br />

will enter a period of earthquake spatial and temporal<br />

clustering, implying a very high potential hazard.<br />

However, without knowledge of the long-term<br />

seismicity record scenarios b and c are difficult to<br />

differentiate from scenario a. Geological data have<br />

the potential to extend the history of slip on a fault<br />

back many thousands of years, a time span that<br />

generally encompasses a large number of<br />

earthquake cycles (e.g. Yeats and Prentice, 1996),<br />

and thus elucidates the long-term pattern of fault-slip.<br />

As a result, multi-cycle seismic hazard maps based<br />

174<br />

on geological fault slip-rate data provide a way to<br />

visualize this by depicting the long-term deformation<br />

pattern of slip.<br />

Fig.1.Seismic hazard map of Italy derived from the historical<br />

earthquake record (GNDT-SSN, 2001). This map shows<br />

intensities that have a 90% probability of not being<br />

exceeded in a certain time period (e.g. 50 years) assuming<br />

a 475-year return period. White box shows the area of the<br />

seismic hazard map constructed during this study, whereas<br />

the dash white circle shows the Pollino seismic gap.<br />

In this paper we present such hazard maps for the<br />

region of the southern Apennines. Then we weigh<br />

our map against the historical record by comparing


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

the number of recorded earthquakes in the last 5-7<br />

centuries with that implied by slip-rates on active<br />

normal faults averaged over 18 kyrs, a time period<br />

which contains numerous seismic cycles. This<br />

comparison demonstrates that the long history of<br />

earthquakes in the Italian Apennines, which is<br />

certainly complete for 5-7 centuries (Boschi et al.<br />

1995) may indeed contain evidence for earthquake<br />

clustering, implying that the historical record is<br />

unrepresentative of the long-term deformation<br />

pattern.<br />

data from published trench-site palaeoseismic data<br />

were also utilised (e.g. Pantosti et al. 1993, Michetti<br />

et al. 1997). In these maps each fault ruptures in<br />

"floating" earthquakes (e.g. WGCEP 2002), which<br />

are distributed around a mean magnitude of Ms=6.5.<br />

THE AREA OF THE SOUTHERN APPENNINES<br />

AND THE POLLINO SEISMIC GAP<br />

A so-called “seismic gap” appears on a historical<br />

seismicity map in the Pollino region in the southern<br />

Italy (Fig.1). In the Pollino area where several active<br />

faults exist, no large magnitude historical<br />

earthquakes are known, but palaeoseismological<br />

studies clearly demonstrate the occurrence of<br />

surface faulting events (Michetti et al., 1997; Cinti et<br />

al., 1997; Michetti et al. 2000). However, seismic<br />

hazard and probabilistic maps based on the historical<br />

record (Slejko et al., 1998; Faenza and Pierdominici,<br />

2007) regard this area as an area of genuinely low<br />

hazard, disregarding the geological evidence. We<br />

suggest that knowledge of the long-term rates of<br />

surface faulting derived from offset geology is<br />

essential before such an area can be defined as low<br />

or high hazard.<br />

SEISMIC HAZARD MAP FROM GEOLOGICAL<br />

FAULT SLIP-RATE DATA<br />

Methodology<br />

Fault throw-rates are firstly converted into earthquake<br />

frequencies, assuming that each fault ruptures in<br />

"floating" earthquakes, which are distributed around a<br />

mean magnitude of fixed size. Then, this information<br />

is turned into a hazard map after using: i) empirical<br />

relationships between coseismic slip values, rupture<br />

lengths and earthquake magnitudes (Wells and<br />

Coppersmith 1994), ii) empirical relationships<br />

between earthquake magnitudes and intensity<br />

distributions that define the shapes and sizes of IX<br />

and VIII isoseismals (Grandori et al. 1991, D’ Amico<br />

et al. 1999) and iii) a simple attenuation/amplification<br />

functions for seismic shaking on bedrock compared<br />

to flysch and basin-filling sediments. The final<br />

product is a high spatial resolution seismic hazard<br />

map showing how many times each location has<br />

been shaken at a certain intensity value (e.g.<br />

intensity IX) over a fixed time period (e.g. since the<br />

last glaciation), which can be easily transformed into<br />

a map of recurrence intervals (see Papanikolaou<br />

2003 and Roberts et al., 2004).<br />

Data input<br />

Fault geometry and fault throw-rates have been<br />

extracted from Papanikolaou and Roberts (2007) and<br />

references therein (Fig.2). Throw rates are mainly<br />

extracted from post-glacial fault scarps, using as a<br />

reference the last major glacial retreat phase that<br />

initiated 18,000 years ago and left a clear imprint in<br />

the topography (Palmentola et al. 1990); other input<br />

175<br />

Fig. 2: (a) Location map of the study area, showing active<br />

faults that have displaced Holocene deposits, (b) Map of<br />

major active faults of the Southern Apennines in UTM<br />

coordinate frame (Papanikolaou and Roberts 2007). The<br />

bedrock map was modified and simplified from the CNR<br />

(1990) map. VF - Volturara Fault; IrF - Irpinia Fault; AntIrF -<br />

Antithetic Irpinia Fault; SGrF - San Gregorio Fault; AlF -<br />

Alburni Fault; VDF - Vallo di Diano Fault; VAF - Val' D Agri<br />

Fault; MaF - Maratea Fault; MAF - Monte Alpi Fault; MeF -<br />

Mercure Fault; PF - Pollino Fault, CaF – Castrovillari fault<br />

(from Cinti et al., 2002), CiF – Civita fault (from Michetti et<br />

al., 1997).<br />

Results<br />

Figure 3 shows the frequency of shaking for intensity<br />

VIII assuming homogeneous bedrock geology, a<br />

circular pattern of energy release and a 25 km radius<br />

for the VIII intensity. The highest hazard is observed<br />

in the centre of the map which receives enough<br />

energy to shake at intensity VIII or higher about 80<br />

times and decreases smoothly towards the tips of the<br />

fault array (about 20-30 times).


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

normalised long-term shaking values and the<br />

historical record both for intensities IX and VIII.<br />

Instead, there is a geographic variation, where<br />

earthquake shaking from historical seismicity for<br />

towns in the north area is ahead of the rate expected<br />

given modeling of the 18 kyrs record of fault slip, with<br />

the opposite relationship in the south close to the socalled<br />

"Pollino Gap". This suggests earthquakes<br />

have been clustered in time in the north during<br />

historical times and anticlustered in the south<br />

(Scenario c of the Introduction).<br />

Fig. 3: Seismic hazard map showing the frequency of<br />

shaking for intensity VIII assuming homogeneous bedrock<br />

geology, a circular pattern of energy release and a 25 km<br />

radius for the VIII isoseismal.<br />

Figure 4 shows how many times a locality receives<br />

enough energy to shake at intensities IX in 18 kyrs,<br />

assuming a 12.5 km radius of isoseismal IX and a<br />

simple attenuation/amplification function, where<br />

bedrock Mesozoic-Cenozoic limestones shake at a<br />

single intensity value less than the<br />

Quaternary/flysch/foredeep deposits at the same<br />

epicentral distance. In particular, the area of highest<br />

shaking frequency is located in the hangingwall<br />

centre of the Val’ D’ Agri fault, which will receive<br />

enough energy to shake at intensity IX or higher up<br />

to 36 times in 18 kyrs. On the other hand, the<br />

hangingwall centres of the Volturara and the Pollino<br />

distal faults will shake only about 10-15 times at<br />

intensities IX because these faults have lower sliprates.<br />

Therefore, the recurrence interval for intensity<br />

IX or higher is estimated to be as short as 500 yrs for<br />

localities in the hangingwall centre of the Val' D' Agri<br />

fault and as long as 1200-1800 yrs in the hangingwall<br />

centres of the Pollino and the Volturara faults.<br />

COMPARISON WITH THE HISTORICAL RECORD<br />

To investigate whether rates of earthquake<br />

occurrence measured using the historical record (e.g.<br />

442 years) are representative of longer time periods<br />

(18 kyrs) where the effects of clustering may be<br />

averaged out, the recurrence intervals extracted from<br />

these maps are compared with the historical record.<br />

Overall, 24 towns were chosen that are uniformly<br />

distributed in the entire area and most of them have<br />

numerous records of earthquake shaking dating back<br />

to the sixteenth and seventeenth century (Table 1).<br />

There is no correlation (R 2


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

not ruptured during this time period and other faults<br />

that could be in a temporal clustering period.<br />

Town<br />

(see also<br />

Fig. 3)<br />

Predictions<br />

from Fig.3 and<br />

Fig.4:<br />

Number of<br />

times in 18<br />

kyrs<br />

Table 1<br />

Normalised values<br />

Number of times<br />

In 442 yrs In<br />

654yrs<br />

Historical<br />

Record<br />

(Number of<br />

times each<br />

locality was<br />

shaken<br />

at stated<br />

intensity)<br />

IX VIII IX VIII IX VIII IX VIII<br />

Buccino 28 49 1 1 1 2 1 3<br />

Balvano 12 44 0 1 0 2 3 1<br />

Padula 31 68 1 2 1 2 1 0<br />

Atena<br />

20 70 1 2 1 3 2 2<br />

Lucana<br />

Tito 0 25 0 1 0 1 4 0<br />

Salerno 0 0 0 0 0 0 0 0<br />

Potenza 0 0 0 0 0 0 0 3<br />

Sicignano d’<br />

Alburni<br />

25 45 1 1 1 2 1 1<br />

Eboli 0 10 0 0 0 0 0 0<br />

Cassano allo<br />

Ionio<br />

3 15 0 0 0 1 0 0<br />

Maratea 18 48 0 1 1 2 0 1<br />

Laino<br />

Castello<br />

18 37 0 1 1 1 0 0<br />

Rotonda 13 37 0 1 0 1 0 0<br />

Castrovillari 12 22 0 1 0 1 0 0<br />

Sant'<br />

0 0 0 0 0 0 1 0<br />

Arcangelo<br />

Montemurro 8 42 0 1 0 2 1 0<br />

Sanza 8 65 0 2 0 2 0 1<br />

Lagonegro 18 70 0 2 1 3 1 0<br />

Lauria 20 55 1 1 1 2 0 2<br />

Lioni 23 32 1 1 1 1 3 1<br />

Teora 23 35 1 1 1 1 4 0<br />

Viggiano 21 46 1 1 1 2 1 0<br />

Montella 0 15 0 0 0 1 0 3<br />

Polla 15 65 0 2 1 2 2 2<br />

It is evident that the northwestern part of the study<br />

area (Irpinia and northern Basilicata) is ahead of its<br />

longer-term shaking record, indicating that this area<br />

may be in a temporal earthquake cluster. Towns like<br />

the Balvano, Atena Lucana, Tito, Lioni, Teora, Polla<br />

have experienced more intensity IX (in particular) and<br />

VIII events during the last five to seven centuries<br />

compared to their longer-term shaking record derived<br />

from the maps (Table 1, Fig. 4). On the other hand,<br />

the southeastern part of the study area (Pollino,<br />

southern Basilicata) may be in a temporary<br />

anticlustering process. Towns like the Lauria,<br />

Lagonegro, Sanza, Castrovillari, Rotonda, Laino<br />

Castello and Maratea have experienced less events<br />

(particularly for intensity VIII) compared to the longerterm<br />

shaking record (Table 1, Fig.3, Fig. 4). The<br />

latter can potentially explain why paleoseismic data<br />

extracted from the Pollino region (Michetti et al.,<br />

1997; Cinti et al., 1997) are in conflict with the<br />

historical record, with paleo surface ruptures found in<br />

a region with no major recorded historical events.<br />

References<br />

Boschi, E., et al. (1995). Catalogo dei forti terremoti in Italia<br />

dal 461 a.C. al 1980, Istituto Nazionale di Geofisica -<br />

177<br />

SGA Storia Geofisica Ambiente, printed by Grafica<br />

Ragno, Ozzano Emilia, BO, Italy, 973 pp.<br />

Cinti, F., L. Cucci, D. Pantosti, G. D’Addezio and M.<br />

Meghraoui (1997). A major seismogenic fault in a “silent<br />

area”: the Castrovillari fault (southern Italy), Geophysical<br />

Journal International, 130, 595-605.<br />

CNR (1990). Structural model of Italy, 1:500000, edited by<br />

Bigi,G., Cosentino, D., Parotto, M., Sartori, R., and<br />

Scandone, P., Consiglio Nazionale delle Richerche.<br />

S.E.L.C.A. Florence.<br />

D’Amico, V., Albarello, D. and Mantovani, E. (1999). A<br />

distribution-Free Analysis of Magnitude-Intensity<br />

Relationships: an Application to the Mediterranean<br />

Region. Phys. Chem. Earth 24, 517-521.<br />

GNDT-SSN Albarello, D.et al. (2001). New seismic hazard<br />

maps of the Italian territory. Gruppo Nazionale per la<br />

Difesa dai Terremoti- Servizio Seismico Nazionale<br />

Website:http://www.dstn.it/ssn/PROG/2000/carte_pericol<br />

osita/mcs_e.gif.<br />

Grandori, G., Drei, A., Perotti, F. and Tagliani, A. (1991).<br />

Macroseismic intensity versus epicentral distance: the<br />

case of Central Italy. Tectonophysics 193, 165-171.<br />

Faenza, L. and Pierdominici, S. (2007). Statistical<br />

occurrence analysis and spatio-temporal distribution of<br />

earthquakes in the Apennines (Italy). Tectonophysics<br />

439, 13-31.<br />

Marco, S., Stein, M., Agnon, A., and Ron, H. (1996). Longterm<br />

earthquake clustering: a 50000 year paleoseimic<br />

record in the Dead Sea Graben. Journal of Geophysical<br />

Research 101, 6179-6191.<br />

Michetti, A.M., L. Ferreli, L. Serva and E. Vittori (1997).<br />

Geological evidence for strong historical earthquakes in<br />

an “aseismic” region: the Pollino case (Southern Italy),<br />

Journal of Geodynamics 24, 1-4, 67-86.<br />

Michetti. A.M., Ferreli, L., Esposito, E., Porfido, S., Blumetti,<br />

A-M, Vittori, E., Serva, L., and Roberts, G.P. (2000).<br />

Ground effects during the 9 September 1998, Mw=5.6,<br />

Lauria earthquake and the seismic potential of the<br />

“Aseismic” Pollino region in Southern Italy. Seismological<br />

Research Letters 71, 31-46.<br />

Palmentola, G., Acquafredda, and Fiore, S. (1990). A new<br />

correlation of the Glacial Moraines in the Southern<br />

Apennines, Italy. Geomorphology 3, 1-8.<br />

Pantosti, D., Schwartz, D.P., and Valensise, G. (1993).<br />

Paleoseismology along the 1980 surface rupture of the<br />

Irpinia fault: implications for the Earthquake recurrence in<br />

the Southern Apennines, Italy. Journal of Geophysical<br />

Research 98, 6561-6577.<br />

Papanikolaou, I.D., (2003). Generation of high resolution<br />

seismic hazard maps through integration of earthquake<br />

geology, fault mechanics theory and GIS techniques in<br />

extensional tectonic settings. Unpublished Ph.D thesis,<br />

University of London, 437pp.<br />

Papanikolaou, I.D., and Roberts, G.P. (2007). Geometry,<br />

kinematics and deformation rates along the active normal<br />

fault system in the southern Apennines: Implications for<br />

fault growth. Journal of Structural Geology 29, 166-188.<br />

Roberts, G.P., Cowie, P., Papanikolaou, I., and Michetti,<br />

A.M., (2004). Fault scaling relationships, deformation<br />

rates and seismic hazards: An example from the Lazio-<br />

Abruzzo Apennines, central Italy. Journal of Structural<br />

Geology 26, 377-398.<br />

Slejko, D., Peruzza, L., and Rebez, A. (1998). Seismic<br />

hazard maps of Italy. Annali di Geofisica 41, 183-214.<br />

Yeats, R.S., and Prentice, C.S. (1996). Introduction to<br />

special section: Paleoseismology. Journal of Geophysical<br />

Research 101, 5847-5853.<br />

Working Group on California Earthquake Probabilities,<br />

(2002). Earthquake probabilities in the San Francisco bay<br />

region: 2002-2031. USGS Open-File Report 03-214.<br />

Wells, D.L., Coppersmith, K., (1994). New Empirical<br />

Relationships among Magnitude, Rupture Length,<br />

Rupture Width, Rupture Area, and Surface Displacement.<br />

Bull. Seismological Society of America 84, 974-1002.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE SPARTA FAULT, SOUTHERN GREECE: TECTONIC GEOMORPHOLOGY, SEISMIC<br />

HAZARD MAPPING AND CONDITIONAL PROBABILITIES<br />

Papanikolau Ioannis D. (1,2), Gerald Roberts (2), Georgios Deligiannakis (1,3), Athina Sakellariou (1), Emmanuel Vassilakis (3)<br />

(1) Mineralogy-Geology Laboratory, Department of Earth and Atmoshperic Sciences, Agricultural University of Athens, Iera Odos<br />

75, 118-55, Athens, Greece<br />

(2) AON Benfield UCL Hazard Research Centre, Department of Earth Sciences Birkbeck College and University College<br />

London, WC 1E 6BT,London UK, Email: (i.papanikolaou@ucl.ac.uk)<br />

(3) Laboratory of Natural Hazards, Faculty of Geology and Geoenvironment, National and Kapodistrian University of Athens,<br />

Panepistimioupolis, 15784, Athens, Greece<br />

Abstract (The Sparta fault, southern Greece: Tectonic geomorphology, seismic hazard mapping and conditional<br />

probabilities): The Sparta Fault system is a major structure approximately 64km long that bounds the eastern flanks of the<br />

Taygetos Mountain front (2.400m) and shapes the present-day Sparta basin. This fault is examined and described in terms of its<br />

geometry, segmentation, drainage pattern and postglacial finite throw, emphasizing also how these parameters vary along strike<br />

the fault. Based on fault throw-rates and the bedrock geology, geological data can offer both a qualitative and quantitative<br />

approach of the expected hazard distribution. This is achieved by the construction of a seismic hazard map based on fault throwrates<br />

that shows the number of times a locality receives enough energy to shake at a certain intensity value, extracting a locality<br />

specific long-term earthquake recurrence record. The Sparta fault was activated in 464 B.C., devastating the city of Sparta. Since<br />

no other major earthquake has been generated by this system since 464 B.C., a future event could be imminent. As a result, not<br />

only time-independent but also time-dependent probabilities, which follow the concept of the seismic cycle, have been calculated<br />

for the city of Sparta.<br />

Key words: Lakonia, slip-rates, time-dependent probabilities, Taygetos,<br />

INTRODUCTION<br />

In 464 B.C. a large earthquake devastated the city of<br />

Sparta (~20000 fatalities), causing great social unrest<br />

(Papazachos and Papazachou, 1997). This event is<br />

regarded as the oldest well-defined event in the<br />

Hellenic historical record (Papazachos and<br />

Papazachou, 1997). However, the area is<br />

characterized by low seismicity over the last 25<br />

centuries (Papanastassiou, 1999) and no other major<br />

event has occurred in the town of Sparta since 464<br />

B.C, suggesting that a future event could be<br />

imminent. This is also supported by cosmogenic<br />

dating techniques applied on the Sparta bedrock<br />

scarp showing that this fault ruptured repeatedly (six<br />

times over the last 13kyrs), with time intervals<br />

ranging from 500-4500yrs (Benedetti et al., 2002).<br />

This fault is studied in detail based on its postglacial<br />

scarp, the analysis of the drainage network and the<br />

major catchments that are influenced by footwall<br />

uplift.<br />

most impressive imprint in the topography showing<br />

sings of recent intense activity. The southern<br />

segment can be divided into two patches that in the<br />

past were probably two individual structures that are<br />

now hard-linked. This fault exhibits an impressive<br />

postglacial scarp that can be traced for many km<br />

(Fig.3). In particular, from the village of Anogia up to<br />

the area of Mystras, it is continuous and has an 8-<br />

12m high scarp (Fig.4).<br />

THE SPARTA FAULT SYSTEM<br />

The Sparta fault system (Fig. 1), bounds the eastern<br />

part of the Taygetos Mt (2.407m) and shapes the<br />

western boundary of the Sparta basin (Fig.2). It<br />

trends NNW-SSE and has a length of 64km. Its<br />

southern tip is located close to the Gerakari<br />

catchment approximately 2-3km SW from the<br />

Potamia village, whereas its northern tip towards the<br />

Alfios river, a couple of km westwards from the<br />

Kamaritsa village in the Megalopolis basin. Two<br />

major faults are traced within this structure (Fig.1).<br />

The northern segment is about 14 km long and<br />

characterized by lower slip-rates. On the other hand,<br />

the southern segment is 50km long and leaves the<br />

Fig. 1: Simplified geological map of the study area. It shows<br />

the segmentation pattern as well as the localities of the<br />

studied catchments profiles.<br />

178


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig.2: Distant view of the Sparta fault. It uplifts the Taygetos<br />

Mt on its footwall and shapes the basin of Sparta towards its<br />

hanging wall.<br />

through time. Whittaker et al. (2008) showed that<br />

rivers with drainage areas greater than 10km 2 and<br />

crossing faults that have undergone an increase in<br />

throw rate within the last 1Myrs have significant longprofile<br />

convexities. They also established that this<br />

relationship holds for throw rate variation along strike<br />

the same fault segment, as well as between faults.<br />

Moreover, Boulton & Whittaker (2009) suggested that<br />

rivers crossing active faults are undergoing a<br />

transient response to ongoing tectonic uplift and this<br />

interpretation is supported by typical signals of<br />

transience such as gorge formation and hill slope<br />

rejuvenation within the convex reach.<br />

Fig.3: View of the post-glacial scarp of the Sparta fault in<br />

the Kalyvia-Sochas locality.<br />

Fig.4: Topographic profile perpendicular to the postglacial<br />

scarp near the village of Anogia. It exhibits an 8.2m high<br />

scarp.<br />

TECTONIC GEOMORPHOLOGICAL ANALYSIS;<br />

CATCHMENTS AND TECTONIC UPLIFT<br />

The Evrotas river flows through the Sparta Basin,<br />

parallel to the Sparta Fault, trending NNW-SSE,<br />

while the secondary branches of this fluvial system<br />

consist of transient rivers which flow perpendicular to<br />

the main structure, indicating a strong tectonic<br />

influence to the drainage pattern. The combination of<br />

fault parallel and fault perpendicular flow is<br />

characteristic of active normal faulting settings.<br />

Kirby & Whipple (2003) demonstrated that<br />

tectonically unperturbed “equilibrium” fluvial long<br />

profiles are typically smooth and concave-up.<br />

However, upland rivers are also sensitive to alongstream<br />

variations in differential uplift (potentially<br />

leading to changes in the profile concavity or<br />

steepness index) and also to changes in uplift rate<br />

179<br />

Nine fluvial long profiles of the transient rivers<br />

crossing different segments of the Sparta fault were<br />

studied in order to examine the longitudinal convexity<br />

and its variation along strike. Such profiles were also<br />

compared to the longitudinal profiles of rivers that are<br />

not influenced by any fault. Furthermore, in order to<br />

examine the transience of the streams across the<br />

Sparta fault, cross sections perpendicular to the river<br />

flow in the headwaters and within the downstream<br />

convex reach were analyzed.<br />

Elevation(m)<br />

Elevation(m)<br />

Elevation(m)<br />

Elevation(m)<br />

Elevation (m)<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

F3AgiosKonstantinos<br />

0 2 4 6 8 10 12<br />

DownstreamDistance(km)<br />

F4Kastorio<br />

0 2 4 6 8 10 12<br />

DownstreamDistance(km)<br />

F7KalyviaSochas<br />

0 2 4 6 8 10 12<br />

DownstreamDistance(km)<br />

1200<br />

1000<br />

F9Potamia<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0 2 4 6 8 10 12 14 16 18 20<br />

0<br />

DownstreamDistance(km)<br />

F3Ag.Konstantinos<br />

F4Kastorio<br />

F7KalyviaSochas<br />

F9Potamia<br />

0 5 10 15 20 25<br />

Downstream Distance(km)<br />

Fig.5: a) Long profiles of catchments crossing perpendicular<br />

the Sparta Fault. Locality names are shown geographically<br />

in Fig.1. b) Comparison of long profiles in the same graph of<br />

the above catchments showing significant differences in<br />

longitudinal convexities along strike the faults.<br />

a<br />

b


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

While river long profile convexities can lead to the<br />

observation of the ongoing tectonic uplift, differential<br />

erosion of geological formations may also cause the<br />

same pattern. Thus, in order to exclude the latter<br />

phenomena and emphasize the tectonic uplift, the<br />

rivers to be examined must fulfill some restrictions.<br />

Whittaker, et al. (2008) suggest that the selected<br />

rivers should discharge a drainage basin larger than<br />

10km 2 above fault and the upstream length should be<br />

at least 5km. These restrictions are applicable in the<br />

south and central segments of Sparta fault. However,<br />

in the northern segment of Sparta fault the streams<br />

are not long enough to fulfill these criteria, due to the<br />

proximity of the watershed of Evrotas Basin to the<br />

SW segment of the fault.<br />

Catchments crossing the North, Central and South<br />

parts of the Sparta Fault were grouped and studied<br />

separately. Qualitative analysis of long profiles<br />

showed a significant difference in longitudinal<br />

convexity between the Central and both the South<br />

and North parts of the fault, leading to the conclusion<br />

of varying uplift rate along strike. A convex reach of<br />

205m height in Potamia catchment long profile<br />

(southeast part of the Sparta Fault) can be observed<br />

although it seems to have propagated upstream in<br />

relation to the fault. This could happen as the<br />

channel successively adjusts to the imposed uplift<br />

field (Whipple & Tucker, 2002). Moreover, Kalyvia<br />

Sochas fan deposits were extensively examined by<br />

Pope et al. (2003). Kalyvia Sochas catchment long<br />

profile revealed a convex reach of 246m height,<br />

which is in contact with the Sparta fault, in contrast to<br />

Potamia catchment’s convex reach. Agios<br />

Konstantinos catchment seems to have a concaveup<br />

channel profile, possibly indicating a constant slip<br />

rate (Whittaker, et al. 2008). On the other hand,<br />

towards the central part of the Sparta fault, were no<br />

fan deposits and talus cones appear and the finite<br />

throw is smaller, the Kastorio catchment’s convex<br />

reach height outreaches 590m, as measured from<br />

the fault (Fig. 5a,5b).<br />

In addition, the normalized steepness index, k sn ,<br />

using a reference concavity of 0.45, was calculated<br />

for six catchments crossing all Sparta fault parts, as<br />

well as for the two catchments crossing the antithetic<br />

structure and two catchments that cross no fault<br />

(localities shown in Figure 1). The k sn rates for the<br />

catchments closer to the tips of the Sparta fault (F3-<br />

Agios Konstantinos and F9-Potamia) were 81 and<br />

82.7 respectively, while in the central part the<br />

steepness rates are higher and vary from 98.5 to 114<br />

(98.5


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

town of Sparta. A considerably higher timedependent<br />

probability of 2,14% over the next 30<br />

years and 3,55% over the next 50 years has been<br />

calculated. The time dependent probability follows<br />

the seismic cycle concept (WGCEP 2002) and<br />

exhibits higher values because the elapsed time<br />

since the last event (2475yrs) has exceeded the<br />

mean recurrence interval (1792±458yrs). However,<br />

due to the irregularity of earthquake time intervals of<br />

the Sparta fault (Benedetti et al. 2002) and the<br />

introduction of a high sigma value (=0.64) this<br />

difference is noteworthy, but not substantial.<br />

Fig.6: Seismic hazard map for the Sparta Basin, showing<br />

how many times a locality receives enough energy to shake<br />

at intensities IX in 15±3kyrs, after considering the bedrock<br />

geology and assuming a circular pattern of energy release,<br />

with a 18 km radius of isoseismal IX.<br />

Fig. 7: Diagram showing the probability density for the town<br />

of Sparta immediately after the 464 B.C. event and the<br />

conditional probability density considering that no event<br />

occurred since 2011. A sigma value of 0.64 is used.<br />

Acknowledgements: The Institution of State Scholarships<br />

of Greece is thanked for support.<br />

References<br />

Benedetti, L, et al. (2002). Post-glacial slip history of the<br />

Sparta fault (Greece) determined by 36 Cl cosmogenic<br />

dating: Evidence for non-periodic earthquakes. Geophys.<br />

Res. Letters 29, 10.1029/2001GL014510.<br />

Boulton, S.J., Whittaker, A.C., (2009). Quantifying the slip<br />

rates, spatial distribution and evolution of active normal<br />

faults from geomorphic analysis: Field examples from an<br />

oblique-extensional graben, southern Turkey.<br />

Geomorphology 104, 299-316.<br />

Kirby, E., Whipple K.X., (2003). Distribution of active rock<br />

uplift along the eastern margin of the Tibetan Plateau:<br />

Inferences from bedrock channel longitudinal profiles.<br />

Journal of Geophysical Research, 108(B4/2217), pp.1-24<br />

Papaioannou, Ch., (1984). Attenuation of seismic intensities<br />

and seismic hazard in the area of Greece. Ph.D. Thesis,<br />

University of Thessaloniki, 200pp.<br />

Papanastassiou, D. (1999). Seismic hazard assessment in<br />

the area of Mystras-Sparta, south Peloponnesus,<br />

Greece, based on local seismotectonic, seismic, geologic<br />

information and on different models of rupture<br />

propagation. Natural Hazards 18, 237-251.<br />

Papanikolaou, I.D., (2003). Generation of high resolution<br />

seismic hazard maps through integration of earthquake<br />

geology, fault mechanics theory and GIS techniques in<br />

extensional tectonic settings. Unpublished Ph.D thesis,<br />

University of London, 437pp.<br />

Papazachos, B. C., Papaioannou, (1997). The<br />

macroseismic field of the Balkan area. Journal of<br />

Seismology 1, 181-201.<br />

Papazachos B. C. and Papazachou C., (1997). The<br />

earthquakes of Greece. Thessaloniki: Ziti Publications.<br />

Pope, R.J., Wilkinson, K.N., and Millington, A.C. (2003).<br />

Human and climatic impact on Late Quaternary<br />

deposition in the Sparta basin piedmont: Evidence from<br />

alluvial fan systems. Geoarcheology: An International<br />

Journal 18, 685-724.<br />

Psonis, K., Latsoudas, C., (1983). 1:50000 Geological Map<br />

“Xirokampion”, IGME.<br />

Psonis, K., (1990). 1:50000 Geological Map “Sparti”, IGME.<br />

Roberts, G.P., Cowie, P., Papanikolaou, I., and Michetti,<br />

A.M., (2004). Fault scaling relationships, deformation<br />

rates and seismic hazards: An example from the Lazio-<br />

Abruzzo Apennines, central Italy. Journal of Structural<br />

Geology 26, 377-398.<br />

Theodoulidis, N.P. (1991). Contribution to the study of<br />

strong motion in Greece. Ph.D. Thesis, University of<br />

Thessaloniki, 500pp.<br />

Whipple, K.X., Tucker, G.E., (2002). Implications of<br />

sediment-flux-dependent river incision models for<br />

landscape evolution. Journal of Geophysical Research,<br />

107(B2), pp.1-20.<br />

Whittaker, A.C., Attal, M., Cowie, P.A., Tucker, G.E. &<br />

Roberts, G., (2008). Decoding temporal and spatial<br />

patterns of fault uplift using transient river long profiles.<br />

Geomorphology 100, 506-526.<br />

Working Group on California Earthquake Probabilities,<br />

(2002). Earthquake probabilities in the San Francisco bay<br />

region: 2002-2031. USGS Open-File Report 03-214.<br />

Wells, D.L., Coppersmith, K. (1994). New Empirical<br />

Relationships among Magnitude, Rupture Length,<br />

Rupture Width, Rupture Area, and Surface Displacement.<br />

Bull. Seismological Society of America 84, 974-1002.<br />

Zindros, G., Exindavelonis, P. (2002). 1:50000 Geological<br />

Map “Kollinae”, IGME.<br />

181


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ACTIVE FAULTING TOWARDS THE EASTERN TIP OF THE CORINTH CANAL: STUDIED<br />

THROUGH SURFACE OBSERVATIONS, BOREHOLE DATA AND<br />

PALEOENVIRONMENTAL INTERPRETATIONS<br />

Papanikolau Ioannis D. (1, 4, Maria Triantaphyllou (2), Aggelos Pallikarakis (1,3), Georgios Migiros (1)<br />

(1) Mineralogy-Geology Laboratory, Department of Earth and Atmoshperic Sciences, Agricultural University of Athens, Iera Odos<br />

75, 118-55, Athens, Greece<br />

(2) Hist. Geology-Paleontology Department, Faculty of Geology and Geoenvironment, National and Kapodistrian University of<br />

Athens, Panepistimiopolis 15784, Athens, Greece<br />

(3) Laboratory of Natural Hazards, Faculty of Geology and Geoenvironment, National and Kapodistrian University of Athens,<br />

Panepistimioupolis, 15784, Athens, Greece<br />

(4) AON Benfield UCL Hazard Research Centre, Department of Earth Sciences, University College London, WC 1E 6BT, London<br />

UK, Email: i.papanikolaou@ucl.ac.uk<br />

Abstract (Active faulting towards the Eastern tip of the Corinth Canal: Studied through surface observations, borehole<br />

data and paleoenvironmental interpretations): The most important active fault that crosses the eastern tip of the Corinth Canal<br />

is studied in detail, involving surface observations, borehole data and peleoenvironmental interpretations. This fault intersects<br />

and/or is parallel and at short distances with major infrastructure facilities such as the Athens-Corinth highway, the railway and the<br />

Corinth Canal. A positive remark referring to the seismic hazard assessment is that this fault has a limited length (~5km) and thus<br />

it can not produce extensive and large displacement (>20cm) from primary surfaces ruptures. Moreover, borehole data and<br />

correlation between the footwall and hangingwall horizons, show no significant offset over the last 200ka, confirming that no<br />

significant displacements have been accumulated and/or has a very low slip-rate. Paleoenvironmental interpretations based on<br />

borehole data show a very complex sedimentation history during the Upper Pleistocene that involves subaerial exposure,<br />

backshore, paralic, lagoonal, shallow marine environments and possibly even some lake sediments.<br />

Key words: Isthmus, Saronic Gulf, seismic hazards, Isthmia - Kalamaki<br />

INTRODUCTION<br />

The Gulf of Corinth is one of the fastest extending<br />

regions worldwide with up to 20mm/yr that diminishes<br />

to 8 and 4 mm/yr towards its eastern end (Billiris et<br />

al. 1991; Briole et al. 2000). Our study area lies<br />

eastwards from the major active faults that shape the<br />

present-day eastern part of the Corinth Gulf, towards<br />

the Corinth Isthmus. In particular, it lies towards the<br />

eastern tip of the Corinth Canal (Fig.1). The area of<br />

the Isthmus divides the Corinth and Saronic Gulfs.<br />

The Corinth Canal is a major infrastructure project<br />

6.3km long, 8m deep and with up to 60m high slopes<br />

that often experience landslides (Marinos and<br />

Tsiambaos 2008, Papantoniou et al. 2008).<br />

The Canal was constructed 120 years ago and offers<br />

a unique opportunity to visualise the faults, forming<br />

an impressive mega-trench. More than 40 faults can<br />

be recognised, most of them normal and oblique<br />

normal (Freyberg 1973). Some of them are overlain<br />

by undeformed late Pleistocene strata and are clearly<br />

not a hazard threat. The majority of them do not<br />

displace the topography and are of limited length,<br />

therefore are considered as secondary structures<br />

with low slip-rates. However, towards the eastern tip<br />

of the Canal we traced a significant ENE-WSW<br />

trending fault that downthrows towards the SSE and<br />

appears to displace the modern ground surface. The<br />

hazard potential of this fault is studied in detail based<br />

on surface observations and the analysis of borehole<br />

data and discussed in this paper.<br />

182<br />

Fig.1: Digital elevation map showing the major and<br />

secondary faults in the study area (modified from Bornovas<br />

et al. 1972, Gaitanakis et al. 1985, Papanikolaou et al.<br />

1989, Papanikolaou et al. 1996, Roberts et al. 2009).<br />

FAULT MAPPING AND SURFACE<br />

OBSERVATIONS<br />

This fault lies in the hangingwall of the Kechries and<br />

Loutraki faults and in the footwall of the South<br />

Alkyonides fault that was activated during the 1981<br />

earthquake sequence. Our 1:5000 mapping shows<br />

that the study fault is extended for at least 1km<br />

southwards and 4km northwards of the Canal (Fig.1,<br />

Fig.2). This fault controls the topography of the area,<br />

producing 150m of topographic variation towards its


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

characteristics we suggest that the fault activity has<br />

progressively migrated eastwards towards our<br />

studied fault, which forms the predominant present<br />

day structure that influences the topography.<br />

Moreover, we identified some secondary active fault<br />

planes of the studied fault in the immediate<br />

hangingwall that offset the entire sedimentary column<br />

(Fig.2, Fig.4). These secondary fault planes<br />

outcropping in the immediate hangingwall, extend the<br />

width of the fault zone up to 50m.<br />

Fig.2 Simplified geological map showing the studied fault. A<br />

detailed map showing the main and secondary fault planes<br />

and the boreholes near the Canal crossing.<br />

Fig.4: View of the secondary fault plane of the principal fault<br />

on the southern part of the Canal. It also created a landslide<br />

in the past.<br />

Fig. 3: View of the main fault plane about 50m northwards<br />

the Canal (right). Close up view of the free face and the<br />

striations developed on the fault plane (left). This is a high<br />

angle normal fault dipping 65 o towards the SSE.<br />

centre, between the footwall and the hangingwall,<br />

and bounds recent alluvial sediments. Moreover, it<br />

offsets Upper Pleistocene sediments, exhibits a clear<br />

fault plane and a free face with striations (Fig.3),<br />

indicating recent activation. It strikes at 075 o -255 o<br />

(ENE-WSW trending) and is a high angle normal fault<br />

dipping at 65 o towards the SSE. Striations measured<br />

on the fault plane are plunging at 55 o towards the SE<br />

(145 o ), confirming that it is an almost pure normal<br />

fault. Strata exposed on the immediate footwall are<br />

Upper Pleistocene sediments, dipping about 10-20 o<br />

to the NNW, indicating also that this tilt is due to the<br />

recent fault activity.<br />

The fault plane has also been traced about 650m<br />

southwards the Canal where it trends at 070 o -250 o<br />

and dips 70 o towards the SSE (Fig.5). The crossing<br />

of the Isthmus and its intersection with the fault, is<br />

characterized by the presence of semicohesive<br />

Upper Pleistocene sediments of variable grain sizes<br />

(from cm thick pebbles to clay) with an abundance of<br />

ostracods that indicate a recent brackish marine<br />

shallow water environment (e.g. Krstic & Dermitzakis,<br />

1981). In addition, to the NW of the main fault plane<br />

and over a 350m distance we identified several fault<br />

planes parallel to the main fault trace. Based on their<br />

183<br />

Fig. 5: View of the fault about 650m southwards the Canal.<br />

Arrows point to the fault plane displayed in the photo<br />

towards the right. It trends 070 o -250 o and dips 70 o towards<br />

the SSE.<br />

Fig. 6: Location map showing the levelling route, the<br />

benchmarks and their post-1981 displacements (redrawn<br />

from Mariolakos and Stiros 1987). Benchmark 77 the only<br />

one that subsided by 4cm, lies at the immediate hangingwall<br />

of the fault and benchmark 79 that was uplifted by 2cm on its<br />

immediate footwall, implying that our studied fault may have<br />

been passively ruptured during the 1981 earthquake.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

It is interesting to note that the area of the Canal<br />

uplifted by approximately 2cm from the 1981<br />

earthquake based on a levelling campaign<br />

(Mariolakos and Stiros 1987) except for one<br />

benchmark that subsided by 4cm and lies at the<br />

immediate hangingwall of our studied fault (Fig. 6).<br />

The latter implies that our studied fault may have<br />

been passively ruptured during the 1981 earthquake.<br />

PALEOENVIRONMENTAL INTERPRETATIONS<br />

Borehole data<br />

Seven boreholes have been recovered on the<br />

easterm side of the Corinth Canal; three of them at<br />

the hanging wall up to 70 m deep and four at the<br />

footwall up to 57 m deep (Figure 2). Two of the<br />

longest boreholes were selected for detailed<br />

micropaleontological-paleoenvironmental analyses,<br />

while the rest were maintained for geotechnical<br />

analyses. Approximately 100 samples were collected<br />

from these boreholes (45 from borehole Bh-3 and 57<br />

from borehole Bh-7). Each sample (20 gr dry weight)<br />

was treated with H 2 O 2 to remove the organic matter,<br />

and subsequently was washed through a 125m<br />

sieve and dried at 60 o C. A subset of each sample<br />

was obtained using an Otto microsplitter to obtain<br />

aliquots of at least 200 benthic foraminifers. The<br />

microfauna were identified under a Leica APO S8<br />

stereoscope. A scanning electron microscope<br />

analysis (SEM Jeol JSM 6360, Dept. of Hist.<br />

Geology-Paleontology) was used for taxonomical<br />

purposes.<br />

Lithostratigraphic units<br />

In each borehole lithological alterations of sand, clay,<br />

clayey sand, conglomerate, marl, fractions of<br />

limestone, even topsoil have been recognized. It is<br />

important that among these layers no significant<br />

correlation was found except for a conglomerate<br />

horizon that lies on top of a coral colony. The latter<br />

strongly indicates that there are major lateral<br />

alterations and stratigraphic variations. Between the<br />

two boreholes we identified a layer that can be easily<br />

correlated in both of them. This is a conglomerate<br />

layer 5.5 m thick in Bh-3 and 3 m thick in Bh-7.<br />

Micropaleontological analysis<br />

Samples from Bh-3 and Bh-7 were analyzed in order<br />

to identify foraminiferal assemblages (Figure 7). The<br />

studied benthic foraminiferal assemblages are<br />

relatively abundant and moderately preserved. The<br />

identified foraminiferal species were grouped in<br />

euryhaline forms mainly represented by Ammonia<br />

spp. large and small sized, Elphidium spp.,<br />

Haynesina spp., Aubignyna perlucida and marine<br />

foraminiferal indicators, mostly including miliolids,<br />

and in a lesser degree full-marine species grouping<br />

Asterigerinata, Neoconorbina, Rosalina spp.,<br />

Planorbulina mediterranensis. A series of different<br />

depositional environments were recognized through<br />

micropaleontological analysis of the benthic<br />

foraminiferal fauna (e.g. Triantaphyllou et al., 2003;<br />

Goiran et al., 2011) in both boreholes. In Bh-3 and in<br />

Bh-7 we found approximately the same pattern of<br />

alternations, but with slight differences.<br />

184<br />

In Bh- 3 the microfauna indicates a shallow marine<br />

environment within the upper 10 m thick layer,<br />

partially influenced by fresh water, that turns to an<br />

open lagoon. Beneath this layer, the next 6 m were<br />

described as a paralic or even a backshore<br />

environment, with no or few broken foraminiferal<br />

specimens. The underlying layers were characterized<br />

as shallow marine to open lagoon environment,<br />

indicating that sea level had changed once more. In<br />

these sediments apart from foraminiferal fauna, a<br />

colony of Cladocora corals was found. This layer is<br />

approximately 6 m thick. Beneath these sediments<br />

the environment changes to paralic-backshore for the<br />

next 9 m and beneath them, again a 6 m thick layer<br />

is described as shallow marine. Down to the base of<br />

Bh-3 at 70 m depth, we encounter two more<br />

alternations of the depositional environment from<br />

shallow marine to coastal.<br />

At the top layers of Bh-7 the microfauna indicates a<br />

shallow marine to paralic environment. These<br />

sediments were 4 m thick while beneath them we<br />

encountered a 3 m thick layer with no microfauna at<br />

all; indicating the presence of a paralic or backshore<br />

environment. Below this layer we determine a 6 m<br />

thick layer described as shallow marine/paralic. The<br />

micropaleontological analysis provides clear<br />

evidence for the presence of a closed lagoon in the<br />

underlying 1 m thick layer. In the following 10 m we<br />

encountered at least 4 alternations between the<br />

depositional environment, from shallow marine to<br />

closed lagoon and back to shallow marine-paralic<br />

conditions. At approx. 23 m depth the environment<br />

changed again to closed lagoon, that followed once<br />

more by shallow marine with fresh water input, that<br />

turned to a closed lagoon for a short interval. Below<br />

this point no microfauna was found indicating a<br />

paralic/backshore environment. It is significant that at<br />

18m depth we found a colony of Cladocora corals,<br />

occurring approximately at the same height as in Bh-<br />

3. The beginning of fault’s deformation zone was<br />

observed at ca. 33m depth and ended at<br />

approximately 45m depth. In the following 12m,<br />

dense alternations of cataclastites and fragments of<br />

limestones occur. From that point until the end of Bh-<br />

7 at 56 m depth, no foraminifera specimens are<br />

traced, indicating the lack of full marine conditions.<br />

The presence of ostracods (Cyprideis spp.) at this<br />

level implies a brackish –oligohaline<br />

paleoenvironment.<br />

Presence of corals and other shallow marine<br />

foraminifera shows that marine sediments were<br />

deposited at glacioeustatic sea level highstands.<br />

Thus, we try to correlate these with the global<br />

glacioeustatic sea level curve and the known uplift<br />

rate from neighbouring corals (0.3mm/yr uplift rate<br />

from 200ka corals in the centre of the Canal about 2<br />

km westwards from our locality based on Collier et<br />

al. 1992). Within our boreholes we have a succession<br />

of lowstand and highstand deposits. Lowstand<br />

deposits such as subaerial deposits and top soil, are<br />

marking unconforminites with marine highstand<br />

deposits. Relative dates for these lowstand and<br />

highstand deposits can be derived from the sea level


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

curve. Therefore, it is clear from our records that the<br />

marine to subaerial transitions form when the location<br />

emerges above sea level either due to uplift and/or<br />

due to falling global sea level. As a result, marine<br />

sediments are expected at 125ka for several<br />

thousand years, then possibly during a short period<br />

at 175ka and then at a prolong period at 200ka<br />

(Fig.8).<br />

Fig.7: View of the stratigraphic columns and the<br />

depositional environment of the two boreholes (Bh3 located<br />

on the immediate footwall and Bh7 located on the<br />

immediate hangingwall of the studied fault).<br />

Fig.8: Sea level curve from Sidall et al. 2003 and the<br />

expected depositional environment based on the 0.3mm/yr<br />

uplift rate from the neighbouring dated corals of Collier et al.<br />

(1992).<br />

DISCUSSION- CONCLUSIONS<br />

We have studied a normal fault that trends ENE-<br />

WSW and downthrows towards the SSE. The current<br />

study confirms that the studied fault is active and<br />

forms a secondary structure that accommodates<br />

displacement between major E-W trending faults.<br />

This fault crosses an area, where major infrastructure<br />

facilities are based, such as the Corinth Canal, the<br />

highway and the railway. A positive remark referring<br />

to the seismic hazard assessment and design is that<br />

this fault has a limited length (~5km) and thus it can<br />

not produce extensive primary surfaces ruptures.<br />

Primary surface ruptures from this fault are not<br />

expected to exceed 20cm of displacement. In<br />

addition, correlation of the same horizons based on<br />

borehole data from the footwall and hanging wall of<br />

the fault, show an offset of 4±2m of the corals that<br />

based on the expected depositional environment are<br />

probably 200kyrs old. Therefore, no significant<br />

displacements have been accumulated over the last<br />

200kyrs, implying that it is a very low slip-rate fault.<br />

Finally, borehole data exhibit a complex<br />

paleoenvironment with major lateral and temporal<br />

alterations and stratigraphic variations as a result of<br />

the interplay between the sea level fluctuations and<br />

the tectonic activity that resembles to the nearby<br />

Perachora peninsula (Roberts et al. 2009).<br />

References<br />

Billiris, H., et al. (1991), Geodetic determination of tectonic<br />

deformation in central Greece from 1900 to 1988. Nature,<br />

350, 124– 129, doi:10.1038/350124a0.<br />

Bornovas, J., Lalechos, N. and Filipakis, N. (1972).<br />

1:50.000 scale geological map, Sheet “Korinthos”.<br />

Institute of Geology and Mineral Exploration<br />

Briole, P., A. Rigo, H. Lyon-Caen, J. C. Ruegg, K.<br />

Papazissi, C. Mitsakaki, A. Balodimou, G. Veis, D.<br />

Hatzfeld, and A. Deschamps, Active deformation of the<br />

Corinth rift, Greece: Results from repeated Global<br />

Positioning System surveys between 1990 and 1995, J.<br />

Geophys. Res., Solid Earth, 21, 25,605– 25,625, 2000.<br />

Collier, R.E.L., Leeder, R.M., Rowe, P. and Atkinson, T.<br />

(1992). Rates of tectonic uplift in the Corinth and Megara<br />

basins, Central Greece, Tectonics, 1159-1167.<br />

Gaitanakis, P., Mettos, A., and Fytikas, M. (1985). 1:50.000<br />

scale geological map, Sheet “Sofikon”. Institute of<br />

Geology and Mineral Exploration.<br />

Goiran, J.P., Pavlopoulos, K., Fouache, E., Triantaphyllou,<br />

M.V., Etienne, R., 2011. Piraeus, the ancient island of<br />

Athens: Evidence from Holocene sediments and<br />

historical archives. Geology, doi:10.1130/G31818.1.<br />

Freyberg, V. (1973). Geologie des Isthmus von Korinth,<br />

Erlangen Geologische Ablhandlungen, Heft 95, Junge<br />

und Sohn, Universitats Buchdruckerei Erlangen, 183pp.<br />

Krstic, N., Dermitzakis, M.D., (1981). Pleistocene Fauna<br />

from a section in the channel of Corinth (Greece). Ann.<br />

Geol. du Pays Hellen. 30 (2): 473-499.<br />

Mariolakos, I. and Stiros, S.C. (1987). Quaternary<br />

deformation of the Isthmus and Gulf of Corinth (Greece).<br />

Geology 15, 225-228.<br />

Marinos, P. and Tsiambaos, G. (2008). The Geotechnics of<br />

Corinth Canal: A review. Bulletin of the Geological<br />

Society of Greece XXXXI/I, 7-15.<br />

Papanikolaou, D., Chronis, G., Lykousis, V., Pavlakis, P.,<br />

Roussakis, G., and Syskakis, D. (1989). 1:100.000 scale,<br />

Offshore Neotectonic Map the Saronic Gulf. Earthquake<br />

Planning and Protection Organization, National Centre for<br />

Marine Research, University of Athens.<br />

Papanikolaou, D., Logos, E., Lozios, S. and Sideris, Ch.<br />

(1996). 1:100.000 Neotectonic Map of Korinthos Sheet,<br />

Earthquake Planning and Protection Organization,<br />

Athens.<br />

Papantoniou, L., Rozos, D., Migiros, G. (2008). Engineering<br />

geological conditions and slope failures along the Corinth<br />

Canal. Bull. Geol. Soc. Greece, Vol. XXXVI/I, 17-24.<br />

Roberts, G. P., S. L. Houghton, C. Underwood, I.<br />

Papanikolaou, P. A. Cowie, P. van Calsteren, T. Wigley,<br />

F. J. Cooper, and J. M. McArthur (2009). Localization of<br />

Quaternary slip rates in an active rift in 10 5<br />

years: An<br />

example from central Greece constrained by 234U-230Th<br />

coral dates from uplifted paleoshorelines, Journal of<br />

Geophys. Res. 114, B10406, doi:10.1029/2008JB0058<br />

Triantaphyllou, M.V., Pavlopoulos, K., Tsourou, Th. &<br />

Dermitzakis M.D., 2003. Brackish marsh benthic<br />

microfauna and paleoenvironmental changes during the<br />

last 6.000 years on the coastal plain of Marathon (SE<br />

Greece). Rivista Italiana Paleontologia et Stratigafia, 109<br />

(3), 539-547.<br />

Siddall, M., E. J. Rohling, A. Almogi-Labin, C. Hemleben, D.<br />

Meischner, Schmelzer, and D. A. Smeed (2003), Sealevel<br />

fluctuations during the last glacial cycle. Nature 423,<br />

853 – 858.<br />

185


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ROMAN AQUEDUCTS AS INDICATORS OF HISTORICALLY ACTIVE FAULTS<br />

IN THE MEDITERRANEAN BASIN<br />

Passchier, Cees W. (1), Gilbert Wiplinger (2), Gül Sürmelihindi (1), Paul Kessener (3), Talip Güngör (4)<br />

(1) Department of Earth Sciences, <strong>Gutenberg</strong> University , 55099 <strong>Mainz</strong>, Germany. Email: cpasschi@uni-mainz.de,<br />

surmelih@uni-mainz.de<br />

(2) Österreichisches Archäologisches Institut, Vienna, Austria. Email: gilbert.wiplinger@gmail.com<br />

(3) van Slichtenhorststraat 13, NL 6524 JH Nijmegen, Netherlands. Email: lenl@euronet.nl<br />

(4) Department of Geological Engineering, Dokuz Eylül University, Izmir, Turkey. Email: talip.gungor@deu.edu.tr<br />

Abstract (Roman aqueducts as indicators of historically active faults in the Mediterranean Basin): Roman aqueducts<br />

are a potential source of information on infrequently active faults in the Mediterranean Basin. We have localised more than<br />

1300 such aqueducts from the literature. Carbonate deposits in aqueduct channels can help to date and characterise fault<br />

activity in detail. As an example, the Deirmendere aqueduct near Ephesos, Turkey is described, which was apparently<br />

displaced and tilted by up to 2 metres a short time after it came into use. The aqueduct was subsequently repaired by<br />

constructing a second channel at the correct elevation running next to the original one over a length of several km.<br />

Key words: roman aqueduct, sinter, carbonate deposits, fault, fault scarp<br />

INTRODUCTION<br />

Major fault zones such as the San Andreas and<br />

North Anatolian fault systems have been extensively<br />

studied, and the local risk of earthquakes and their<br />

frequency is understood to some extend. Smaller<br />

faults, however, which may produce strong<br />

earthquakes at very long time intervals, may go<br />

undetected and may pose a similar or greater risk, as<br />

local building regulations are not adapted. Examples<br />

are the Basel earthquake of 1356 (Lambert et al.<br />

2005) and the Lisbon earthquake of 1755 (Mendez-<br />

Victor 2008). For a better understanding of<br />

earthquake risks and neotectonic processes, it is<br />

useful to determine which faults show this kind of<br />

long-term activity.<br />

Several attempts have been made in the past to<br />

localize infrequently active faults mainly by the study<br />

of historical records and paleoseismic events. Here,<br />

Roman aqueducts 1 are presented as a new tool for<br />

earthquake studies to enhance our knowledge of<br />

infrequently active faults in the Mediterranean Basin.<br />

ANCIENT AQUEDUCTS<br />

Ancient aqueducts are gravity driven water transport<br />

systems consisting of roofed masonry channels or<br />

piped conduits running from perannial springs to<br />

cities, villages or farms (Fig. 1,2). Meandering along<br />

the contour lines of the terrain, they gradually<br />

descend from source to destination (Hodge 1992,<br />

Wikander 2000, Grewe 1986). Aqueducts that served<br />

large cities frequently acquired considerable lengths,<br />

the 250 km Constantinople aqueduct at the top of the<br />

list (Grewe 1986, Hodge 1992, Çeçen 1996).<br />

1 We use the term “aqueduct” to represent the entire water<br />

transporting system of channel, piped conduit, intermediate<br />

basins, tunnels and substructures, not just the bridges.<br />

186<br />

Fig. 1. Presently known roman aqueducts in central<br />

Italy, showing the dense network of long channels,<br />

several of which cross major faults. ROMAQ Database.<br />

Aqueducts can be valuable for the study of<br />

historically active faults for several reasons:<br />

(1) they are long, anastomosing objects that<br />

cover a large area, and are much more likely to<br />

be cut by an active fault than any other structure;<br />

(2) they comprise bridges, inverted siphons and<br />

tunnels which are vulnerable to earthquake<br />

damage; (3) they may serve as an earthquake<br />

recording device for over 2000 years, i.e. from the<br />

construction date to present day; (4) they<br />

generally slope down with a minor and rather<br />

constant gradient allowing a detailed<br />

reconstruction of the uplift pattern produced by<br />

movement on faults; (5) since most aqueducts<br />

are linked with cities where archaeological work is<br />

in progress or has been done, the building and<br />

operation of the aqueduct can usually be dated


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

with an accuracy of decades or less; (6) many<br />

aqueducts contain carbonate deposits (Fig. 2) which<br />

record seasonal and abrupt changes in water<br />

chemistry, temperature and depth. Even if an<br />

aqueduct is not directly cut by an active fault, a<br />

seismic event can leave traces in the carbonate<br />

deposits that can be dated.<br />

Ancient aqueducts are found on all continents,<br />

including the Aztec and Inca Empires of South<br />

America (Wright et al. 2006), but the largest<br />

concentration is in the Mediterranean Basin. Most are<br />

of Roman Imperial age (50-300 AD) when hundreds<br />

of aqueducts were built within a short period of time,<br />

often close together (Fig. 1)(Bedon 1997; Hodge<br />

1992; Wikander 2000). More than 1300 Greek and<br />

Roman aqueducts are known from the literature.<br />

A number of aqueducts are known to have been<br />

damaged by earthquakes. An aqueduct near<br />

Venafro (Italy) was apparently damaged by three<br />

major earthquakes on what seemed to be a minor<br />

fault (Galli and Naso 2009; Galli et al. 2010); the<br />

major 1349 earthquake that damaged many cities<br />

in central Italy, known from medieval records, can<br />

now be traced to this fault (Galli et al. 2010). The<br />

aqueduct of Nimes (France) was damaged by an<br />

earthquake in the third or fourth century in a zone<br />

that was thought to be free of recent tectonic<br />

activity (Levret et al. 2008; Volant et al. 2009).<br />

Marra et al. (2004) reported on a small aqueduct<br />

near Rome that was deformed by an earthquake.<br />

Damage to the aqueducts of Petra (Jordan) could<br />

be attributed to an earthquake of AD 363<br />

(Bellwald et al. 2003). The aqueduct at Aphamea<br />

(Syria) was destroyed by earthquakes in 526 and<br />

528 AD and subsequently repaired (Balty 1987).<br />

At Al Hativ (Syria) the Dead Sea fault displaced<br />

an aqueduct by up to 13 metres in several<br />

seismic events (Meghraoui et al. 2003; Sbeinati et<br />

al. 2010).<br />

Fig. 2. Cross section of the roman aqueduct channel<br />

of Cologne, showing typical elements. The buried<br />

masonry channel is covered on the inside by a red<br />

layer of waterproof cement (“opus signinum”).<br />

Carbonate on the walls tapers upwards because of<br />

variable and gradually rising water level. In the<br />

Cologne case, the deposits indicate that the aqueduct<br />

ran for approximately 170 years (Fig. 3).<br />

187<br />

Fig. 3. Sinter from the Cologne aqueduct. Note typical<br />

laminated deposition. Height of sample 12 cm.<br />

THE ROLE OF CARBONATE DEPOSITS<br />

Many aqueducts were damaged by earthquakes<br />

during or after their active life, but it is usually<br />

hard to date these events, or to understand how<br />

many times the channels were damaged and<br />

repaired (cf. Meghraoui et al. 2003; Sbeinati et al.<br />

2010). In such cases, carbonate deposits in<br />

roman aqueducts may give clues.<br />

Many aqueducts carried water from karstic<br />

aquafers containing carbonates, which deposited<br />

on the walls of the channel. (Fig. 3)(Grewe 1986,<br />

Guendon & Vaudoir 2000, Garczynski et al. 2005,<br />

Guendon & Leveau 2005, Dubar 2006). This<br />

travertine like material, also known as sinter,


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

could, over time, acquire considerable thickness<br />

corresponding to the period that the aqueduct was<br />

running, from decades to as long as 800 years<br />

(Sbeinati et al. 2010). Earthquakes have interfered<br />

with these deposits in several ways.<br />

1 – Fault motion when an aqueduct is running.<br />

In the most dramatic case, the channel is cut by the<br />

fault and water flow is interrupted. Even if the<br />

channel was not destroyed, fracturing of the walls<br />

and sinter deposits could occur, with subsequent<br />

growth of new sinter over the fractures (Carbon et al.<br />

2005). In this way, the seismic event can be precisely<br />

localised in the sinter stratigraphy. Careful<br />

examination of the damaged site can reveal much<br />

about earthquake displacement and dating<br />

(Meghraoui et al. 2003; Volant et al. 2009; Galli et al.<br />

2010; Carbon et al. 2005). Volant et al. (2009) found<br />

deformation twins in calcite in sinter of the Nimes<br />

aqueduct, which they attribute to earthquake<br />

damage.<br />

Seismic events may also leave marks in sinter away<br />

from the damaged site, or where the active fault did<br />

not cross the aqueduct. Local damage to the channel<br />

or to bridges can lead to water loss and thinner<br />

deposits downstream after the earthquake (Levret et<br />

al. 2008; Carbon et al. 2005). In case a section of<br />

aqueduct was ruptured and subsequently repaired,<br />

sinter deposits in this section will differ from those<br />

upstream from the damaged site. Fragments of<br />

building material and speleothems fallen into the<br />

channel, and clastic sediments may be covered by<br />

sinter. Commonly, a new layer of waterproof cement<br />

is put on top of the deposits when repairing the<br />

channel. If the rupture site is lost due to later<br />

destruction or erosion, deposits downstream from the<br />

site may still differ from those upstream, and a<br />

relative dating of the earthquake can thus be<br />

established. A further item to be explored is the<br />

possibility that an earthquake may temporarily or<br />

permanently cause changes in the chemical<br />

composition of the water due to alterations in the<br />

cave aquifer system that fed the aqueduct.<br />

2 – Fault motion after the aqueduct has stopped<br />

running.<br />

If faults cut an aqueduct after it ceased to<br />

operate, dating of the event is possible by<br />

classical trenching and dating techniques used for<br />

active faults (Sbeinati et al. 2010). In this case,<br />

the aqueduct serves to give a minimum age for<br />

fault slip, and careful geodetic survey of the<br />

deformed channel can give information about<br />

vertical ground motion on both sides of the fault.<br />

THE DEIRMENDERE AQUEDUCT OF<br />

EPHESOS<br />

In recent years Gilbert Wiplinger of the<br />

Österreichisches Archaeologisches Institut (ÖAI)<br />

and an international team of researchers<br />

investigated the Deirmendere aqueduct of<br />

Ephesos, Turkey. This aqueduct, one of seven<br />

that served the ancient city, took its waters from a<br />

spring along a nappe contact to the SE (Fig. 4;<br />

Wiplinger 2008). The 37,5 km aqueduct was built<br />

as a masonry vaulted channel 80 cm wide and<br />

over 240 cm high and was equipped with tunnels<br />

and bridges. For a stretch of over 10 km, two<br />

channels appeared to run parallel to each other<br />

(Figs. 4,5). It was found that the channels were<br />

not built simultaneously but one some years after<br />

the other, the younger channel being partly built<br />

on top of the older one, so that both channels<br />

cannot have carried water at the same time.<br />

Halfway this 10 km stretch the aqueduct crosses<br />

a valley by means of the two story Bahçecikboaz<br />

bridge, where the level of both channels coincide.<br />

However, at the upstream end the younger<br />

channel’s floor is positioned 2 meters above the<br />

older one, while at the downstream end, close to<br />

where an extended tunnel section starts, it is 2<br />

meters below it. (Wiplinger 2008, Figs.<br />

4,5).<br />

Fig. 4. Double channel of the Deirmendere aqueduct<br />

looking north. The lower channel is the older one and<br />

was partly demolished to build the upper channel on the<br />

right. Location shown in Figure 5.<br />

188<br />

Fig. 5. a - Schematic geological map of the area south<br />

of Ephesos. The double channel section lies in the<br />

central marble unit; b - schematic height profile of the<br />

Deirmendere aqueduct showing the section with two<br />

channels. Geology after Çetnkaplan (2002).


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Initially it was thought that survey errors by the<br />

Roman engineers would explain this feature<br />

(Wiplinger 2008). However, recent work has shown<br />

that the earlier channel has a layer of sinter several<br />

mm wide, which indicates that it must have carried<br />

water for several years. This makes survey errors<br />

unlikely. The younger channel has carbonate<br />

deposits up to 30 cm wide accounting for many<br />

decades of running water up to the final stage of the<br />

aqueduct. Geological reconaissance early 2011 has<br />

localised a major fault scarp near the upstream<br />

starting point of the twin section. Whether the<br />

northern, downstream end of the twin section also<br />

coincides with a fault is currently under investigation.<br />

We presently think of the following scenario.<br />

The Deirmendere aqueduct was cut by at least one<br />

major fault after it has been in operation for only a<br />

few years, probably in the second half of the second<br />

century AD. The fault produced a scarp of<br />

approximately 2 metres high, interrupting the water<br />

flow and altering the slope of the aqueduct over a<br />

stretch of 10 km (Fig. 5). To restore the aqueduct, a<br />

new channel parallel to the disrupted one had to be<br />

built at the proper level. The aqueduct channel and<br />

the geology surrounding the Deirmendere aqueduct<br />

will be further investigated in 2011.<br />

Acknowledgements: The authors thank the Geocycles<br />

Cluster and the University of <strong>Mainz</strong> for financial support.<br />

The Österreichisches Archaeologisches Institut provided<br />

crucial logistic and general support.<br />

References<br />

Balty, J. C., (1987). Problèmes de l'eau a Apamée de Syrie.<br />

In: Les Travaux de la Maison de l'Orient. L'Homme et<br />

l'eau en Mediterranée et au Proche Orient. (Louis, P,<br />

Métral, F., Métral , J. eds). 4, 9-23.<br />

Bedon, R., (1997). Les aqueducs de la gaule romaine et<br />

des regions voisines. Pulim, Limoges.<br />

Bellwald,U., M. Al-Huneidi, A. Salihi, D. Keller, R. Naser, &<br />

D. Al-Eisawi, (2003). The Petra Siq, Nabataean hydrology<br />

uncovered. Petra National Trust.<br />

Carbon, D., G. Fabre, P. Volant, J.-L. Fiches, A. Levret & P.<br />

Combes, (2005). L’aqueduc de Nîmes dans la haute<br />

Vistrenque: analyse interdisciplinaire d’un tronçon<br />

souterrain. Gallia 62, 69-86.<br />

Çeçen, K., (1996). Roma su yollarnn en uzunu (The<br />

Longest Roman Water Supply Line).<br />

Çetnkaplan, M., (2002).Tertiary high pressure / low<br />

temperature metamorphism in the Mesozoic cover series<br />

of the Menderes massif and correlation with the cycladic<br />

crystalline complex. Unpublished PhD. Thesis, Izmir.<br />

Dubar, M., (2006). Recherche et interprétation climatique<br />

des microcycles du concrétionnement travertineux de<br />

l'aqueduc romain de fréjus (var, France) Quaternaire 17,<br />

79-85<br />

Galli, P.A.C., Naso J.A., (2009), Unmasking the 1349<br />

earthquake source (southern Italy); paleoseismological<br />

and archaeoseismological indications from the Aquae<br />

Iuliae Fault, Journal of Structural Geology 31,128–149.<br />

Galli, P.A.C., Giocoli, A., Naso, J.A., Piscitelli, S., Rizzo, E.,<br />

Capini, S. & Scaroina, L., (2010). Faulting of the<br />

Roman aqueduct of Venafrum (southern Italy):<br />

Methods of investigation, results, and seismotectonic<br />

implications. In: Ancient Earthquakes (Sintubin, M.,<br />

Stewart, I.S., Niemi, T.M., Altunel, E. eds). Geological<br />

Society of America Special Paper 471, 233-242.<br />

Garczynski, P, J. Foucras & M. Dubar, (2005).<br />

L´aqueduc d´Antipolis dit de la Boullide (Alpes<br />

Maritimes). Gallia 62, 13-34<br />

Grewe, K., (1986). Atlas der römischen<br />

Wasserleitungen nach Köln. Rheinland Verlag, Koeln.<br />

Guendon, J-L., Vaudoir, J., (2000). Concrétions et<br />

fonctionement de l´aqueduc: étude morphostratigraphique.<br />

In: L´aqueduc de Nîmes et le Pont du<br />

Gard. (Fabré, G., Fiches, J-L., Paillet, J-L. eds).<br />

CNRS Editions. 263-271<br />

Guendon, J-L., Leveau, Ph., (2005). Dépots carbonatés<br />

et fonctionement des aqueducs romains. Gallia, 62,<br />

87-96<br />

Hodge, A. T., (1992). Roman aqueducts and water<br />

supply, Duckworth, London.<br />

Lambert J., T. Winter, T.J. Dewez & P. Sabourault,<br />

(2005). New hypotheses on the maximum damage<br />

area of the 1356 Basel earthquake (Switzerland),<br />

Quaternary Science Reviews, 24, 383–401.<br />

Levret, A, Ph. Volant, D. Carbon, D Combescure, T.<br />

Verdel, A. Piant, O. Scotti & Ph. Laurent, (2008). L’<br />

aqueduc romain de Nimes (France): nouveaux<br />

résultats en archéosimicité. In: Archéosismicité et<br />

Vulnérabilité: patrimoine bâti et société. (Levret, A.<br />

ed),165-188<br />

Marra, F., P. Montone, M. Pirro & E. Boschi, (2004).<br />

Evidence of tectonics on a Roman aqueduct system<br />

(II-III century AD) near Rome, Italy. Journal of<br />

Structural Geology 26, 679-690.<br />

Meghraoui, M., F. Gomez, R. Sbeinati, J. Van der<br />

Woerd, M. Mouty, A. Nasser Darkal, Y. Radwan, I.<br />

Layyous, H. Al Najjar, R. Darawcheh, F. Hijazi, R. Al-<br />

Ghazzi & M. Barazangi, (2003). Evidence for 830<br />

years of seismic quiescence from palaeoseismology,<br />

archaeoseismology and historical seismicity along the<br />

Dead Sea fault in Syria. Earth and Planetary Science<br />

Letters 210, 35-52.<br />

Mendez-Victor, L. A., C. S. Oliveira & J. Azevedo,<br />

(2008). The 1755 Lisbon Earthquake: revisited.<br />

Springer Verlag.<br />

ROMAQ. Unpublished database. Passchier, C. W., Van<br />

Opstal, D., Schram, W. (eds.) Internet publication<br />

pending.<br />

Sbeinati, M., M. Meghraoui, G. Suleyman, F. Gomez,<br />

P. Grootes, M. Nadeau, H. Al Najjar & R. Al Ghazzi,<br />

(2010). Timing of earthquake ruptures at the Al Harif<br />

Roman aqueduct (Dead Sea fault, Syria) from<br />

archaeoseismology and paleoseismology. Geological<br />

Society of America Special Papers 471, 243-267.<br />

Volant, Ph., A. Levret, D. Carbon, O. Scotti, D.<br />

Combescure, T. Verdel, A. Piant, Ph. Laurent, (2009).<br />

An archaeo-seismological study of the Nîmes Roman<br />

aqueduct, France: indirect evidence for an M6 seismic<br />

event? Natural Hazards 49, 53-77.<br />

Wikander, O., (2000). Handbook of ancient water<br />

technology. Brill, Leiden.<br />

Wiplinger, G., (2008). The Deirmendere Aqueduct to<br />

Ephesus. Anodos 8, 393-400.<br />

Wright, K. R.,G. F. McEwan & R. M. Wright, (2006).<br />

Tipon: water engineering masterpiece of the Inca<br />

empire. ASCE Press.<br />

189


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

TESTING ARCHAEOSEISMOLOGICAL TECHNIQUES WITH INSTRUMENTAL SEISMIC<br />

DATA CAUSED BY THE Mw 5.1 LORCA EARTHQUAKE (5-11-2011, SE OF SPAIN)<br />

Pérez-López (1)*, R., J.L. Giner-Robles(2), M.A. Rodríguez-Pascua (1), F. Martín-González(3), J. García Mayordomo (1)<br />

J.A. Álvarez-Gómez(4)(5), M.J. Rodríguez-Peces(4), J.M. Insua-Árévalo (4), J. J. Martínez-Díaz (4) and P.G. Silva(6)<br />

(1) IGME – Instituto Geológico y Minero de España, Área de Investigación en Riesgos Geológicos. Madrid, SPAIN. Email:<br />

r.perez@igme.es , m.rodríguez@igme.es, julian.garcia@igme.es<br />

(2) Departamento de Geología. UAM, Universidad Autónoma de Madrid. SPAIN. Jorge.giner@uam.es<br />

(3) Universidad Rey Juan Carlos I, Madrid. SPAIN<br />

(4) Departamento de Geodinámica. Facultad de Ciencias Geológicas. Universidad Complutense de Madrid. SPAIN. Email:<br />

jmartinez@geo.ucm.es<br />

(5) Instituto de Hidráulica Ambiental "IH-Cantabria", E.T.S.I. Caminos, Canales y Puertos, Universidad de Cantabria. Spain.<br />

(6) Departamento de Geología. Universidad de Salamanca. Email: pgsilva@usal.es<br />

* Corresponding author<br />

Abstract (Testing archaeoseismological techniques with instrumental seismic data caused by the Mw 5.1 Lorca<br />

Earthquake, 5-11-2011, SE of Spain): the earthquake of the 11 th of May of 2011 was located near of the village of Lorca (SE of<br />

Spain) and with a magnitude (Mw) of 5.1. Despite of the small sized- earthquake, it generated a relative high seismic intensity (VII<br />

EMS), nine people were killed and almost 14,000 were left from their homes. The total estimated cost of this earthquake is 75<br />

millions of Euros. A systematic analysis of damaged ancient buildings was performed with the aim to check archaeoseismological<br />

new techniques. Namely, we have described different Earthquake Archaeological Effects (EAE) affecting both monumental<br />

buildings as modern edifices. The interest is the possibility to compare EAEs with focal parameters of the earthquake and the<br />

geometry of the seismic source, the Alhama de Murcia Fault (FAM).The preliminary results displays an almost N-S seismic wave<br />

anysotropy obtained from different EAE recorded at eight different monumental buildings.<br />

Key words: archaeoseismology, instrumental earthquake, EAE, Lorca, SE of Spain<br />

THE LORCA EARTHQUAKE OF THE 5-11-2011<br />

The 11 th of May, 2011, a small and shallow<br />

earthquake struck the village of Lorca (92,000<br />

inhabitants approximately). This earthquake hit the<br />

city at 18:50 h (local time), such as you can see in<br />

the broken clock of the tower of the San Francisco<br />

church (also so-called the church of “El Paso Azúl”).<br />

The recorded magnitude by the Spanish Instituto<br />

Geográfico Nacional, IGN, was 5.1 Mw (www.ign.es).<br />

Nevertheless, the maximum value of the peak ground<br />

acceleration (PGA) was 0.41g, recorded at the<br />

basement of the ancient jailhouse of the city. Also, 5<br />

mm of GPS movement northward was recorded in<br />

the basement of the Fireman Station. Outliers values<br />

for this small magnitude.<br />

Unfortunately, this small earthquake killed nine<br />

people, two from the collapse of one modern building<br />

located at La Viña neighbourhood, southward of the<br />

city, and the others due to the falling of the hanging<br />

decoration from the outer façade. Besides, almost<br />

14,000 people were moved from their homes and<br />

300 of them still live in camp sites. The city was<br />

dramatically collapsed and during the following<br />

twelve hours the fatality ruled the city.<br />

A few hours after the main shock, a group of<br />

earthquake geologists grew up from diverse Spanish<br />

institutions, and they went to the disaster zone with<br />

the aim of finding answers to these questions: (1)<br />

what was the seismic source of this earthquake? Or<br />

better, which is the causative fault of the earthquake?<br />

(2) Is there evidence of surface rupture for such a<br />

minor event?<br />

190<br />

Fig. 1: Up. Seismic serie of the Lorca earthquake recorded<br />

by the IGN. The size of the point is related to the<br />

earthquake magnitude. The colour indicates the day of the<br />

occurrence. The focal mechanism solutions show two<br />

strike-slips with reverse component. The red lines indicate<br />

the main active faults. The FAM (Alhama de Murcia major<br />

Fault) is a NE-SW reverse fault with oblique component,<br />

with a total length of ca 90 km (Martínez-Díaz, 2002).<br />

Down: frequency grey bars indicate the number of<br />

earthquakes recorded in 3 h.<br />

Empirical models predict small rupture area (i.e. ca<br />

12 km 2 , Wells & Coppersmith, 1994), (3) Can we<br />

compile an Earthquake Environmental Effect


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

catalogue from this shaking? Rock fall, ground cracks<br />

affecting soils and rocks, changes of temperature in<br />

hot springs and thermal aquifers, liquefactions, and<br />

so, and finally, (4) Can we do an archaeoseismictype<br />

analysis from the monumental building damage<br />

within the village? In this case we tried to make a<br />

compilation of the maximum number of the<br />

Earthquake Archaeological Effects (see Rodríguez-<br />

Pascua et al., 2011 for a comprehensive description<br />

of EAEs). The interest of this last analysis is the<br />

correlation between EAEs and focal and geometrical<br />

parameters of the earthquake.<br />

Despite the low magnitude of the earthquake, we<br />

have enough geological and seismological data to<br />

find a preliminary answer to these questions: high<br />

seismic intensity (VII EMS, www.ign.es), high<br />

building damage (almost 200 damaged buildings are<br />

catalogued), we have an active fault with different<br />

outstanding palaeoseismic studies (the Alhama de<br />

Murcia Fault, FAM)(Silva et al., 1997, Martínez-Díaz<br />

1998)(see location in Fig.1).<br />

Consequently, the aim of this work is to check EAEs<br />

usefulness by comparing the damage anisotropy<br />

caused by the main shock of the earthquake (Mw<br />

5.1), and performed along the old town of the historic<br />

village of Lorca. The technique used here is widely<br />

described in Giner-Robles et al. (2009).<br />

SEISMOTECTONICS OF THE EARTHQUAKE<br />

A precursor event of Mw 4.5 (www.ign.es) occurred 3<br />

hours before of the main shock of 5.1. The epicentre<br />

location of both earthquakes coincides with the<br />

mapped active fault of Alhama de Murcia (FAM)(<br />

Bousquet & Montenat, 1974))(Fig.1). The location of<br />

the epicentres on the FAM suggests that a small<br />

segment of 3 x 3 km could have broken. Moreover,<br />

the focal mechanism solutions indicate either a NE-<br />

SW fault plane (in coincidence with the FAM<br />

trending) or NW-SE fault plane. The geometry of both<br />

mechanisms shows an oblique fault with reverse<br />

component. This geometry is in agreement with the<br />

geology and palaeoseismological studies on the FAM<br />

(Silva et al., 1997; Martínez Díaz et al., 2001). The<br />

estimated depth of the earthquake lies between 3<br />

and 5 km (www.ign.es).<br />

In the following six days, more than 100 aftershocks<br />

with a maximum magnitude of 3.9 were recorded<br />

(www.ign.es). The temporal occurrence of<br />

aftershocks shows cluster behaviour in small groups<br />

instead of a slight decay of the number of aftershocks<br />

from the main shock (Fig. 1).<br />

GEOLOGICAL EFFECTS (EEE) OF THE<br />

EARTHQUAKE<br />

A shallow earthquake (2 km deep) and high seismic<br />

intensity (VII EMS) suggested the possibility to have<br />

a surface rupture associated to the earthquake. With<br />

this aim, we performed a field trip across the fault<br />

trace of FAM to catalogue Earthquake Environmental<br />

Effects (EEE) according to the ESI 07 scale (Michetti<br />

et al., 2007).<br />

191<br />

Fig. 2. Rock fall triggered by the Lorca earthquake and<br />

located at Las Estancias Sierra, southwestwardly of the<br />

epicentre location.<br />

The main effect was rock falling located in steep<br />

carbonatic cliffs in the nearby (


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

casualties, particularly on the habitations located<br />

right under the cliff. A quick visual inspection<br />

suggested the perfect state of the measures,<br />

although this observation should be confirmed after a<br />

systematic and detailed inspection. The approximate<br />

volume of rock falling blocks is lesser than 10 3 m 3 .<br />

Fig. 3. Partial collapse of the easternmost point of Lorca’s<br />

castle cliff. A detached rock-block produced damage in a<br />

near house.<br />

ANALISIS OF EAE: DAMAGE DESCRIPTION<br />

The earthquake of Lorca (SE Spain) was responsible<br />

for a large amount of damage and seismic intensity<br />

on a wide range of buildings in the city of Lorca,<br />

including the historical buildings. Aerial view shows a<br />

concentration of damage in the highest towers,<br />

mainly affecting the arches, buttresses canopies,<br />

bollards etc. Rotations also appear in decorative<br />

elements such as bollards and obelisks, like the<br />

obelisk in the San Francisco church.<br />

The most relevant key of this archaeoseismic study is<br />

the correlation between the instrumental information<br />

of this earthquake with damage in historical and<br />

modern buildings. This information also can be<br />

correlated with seismic and geological parameters,<br />

such as the magnitude and focal mechanism of<br />

earthquake, the seismogenic fault and site effects<br />

related to the geology of Lorca.<br />

Lorca's earthquake has not produced widespread<br />

buildings collapses (only two buildings collapsed),<br />

during the inspection we have recognized, classified<br />

and described more than one hundred effects of the<br />

earthquake on buildings and structures, similar to<br />

those described in the work of Giner-Robles et al.,<br />

2009, Rodríguez-Pascua et al., 2011. These authors<br />

defined the Earthquake Archaeological Effects,<br />

(commonly known as EAE).<br />

The EAE describes and quantifies the coseismic<br />

deformation in archaeological sites and historic<br />

buildings. After Giner-Robles et al., 2009 and<br />

Rodríguez-Pascua et al., 2011, they classified<br />

according to EAE: (1) permanent deformation of the<br />

surface (2) temporary deformation by the seismic<br />

shaking during the earthquake.<br />

192<br />

Fig. 4. Damage in the tower and main buildings of the<br />

Clarisas´ Convent. (a) Destroyed tower, (b), collapsed<br />

roof of the main edifice, (c) and (d) X-fractures in the East<br />

and West sides of the annexed building.<br />

Figure 4 shows an example of the type of analysis<br />

that we have performed in the village. We have<br />

located the damaged building and we have described<br />

all of EAEs recognized. In this case, the Clarisas’s<br />

Convent displayed two types of EAEs: collapsed roof<br />

and walls, and penetrative conjugated faults affecting<br />

oriented sides of annexed buildings. Other type of<br />

study involves the orientation of fallen key stones in<br />

arches. Fig. 5 is a clear example of arches showing<br />

evidence of damage anisotropy obeying the<br />

mechanism proposed by Giner et al., 2009.<br />

All this preliminary information has been represented<br />

in a rose diagram (Fig. 6), showing the main<br />

orientation in a NW-SE trend, N150º-160ºE. We<br />

interpret these direction NW-SE with azimuth from<br />

the SE.<br />

PRELIMINARY RESULTS<br />

- Alhama-Murcia Fault (FAM) is the structure with<br />

greater evidence of Quaternary activity in the area:<br />

paleoseismic activity (Mw > 6.0) over the last 1000<br />

years, associated with thermal springs and a wellrecognized<br />

surface trace. Destructive historical<br />

seismicity located along the trace during the XVII,<br />

XVIII and XIX centuries were reported in chronicles.<br />

FAM has a trace that it is parallel to one of the nodal<br />

planes of focal mechanisms obtained for the<br />

earthquake. The oblique (reverse - sinistral<br />

movement) movement of the fault is consistent with<br />

the focal mechanism solution.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 6. Rose diagram of the seismic shaking, using the<br />

EAEs. Preliminary results (After Martínez-Díaz et al.,<br />

2011).<br />

References<br />

Fig. 5. Example of analysis of deformation structures and<br />

the results of the orientation assessment of the strain (San<br />

Juan’s Church) (After Martínez-Díaz et al., 2011).<br />

- The high seismic intensity experienced by the town<br />

of Lorca (intensity VII EMS-98 scale, data IGN)<br />

associated with a magnitude 5.1 Mw, may be due to<br />

the earthquake spread from the Sierra de la Tercia<br />

(epicentral area) to the SW. The lack of geological<br />

effects to the east of the epicenter (La tercia Range<br />

and Lower Guadalentin Valley) support the possible<br />

existence of rupture directivity southwest.<br />

-The seismic wave propagation supports the<br />

directionality of the FAM rupture spread from the<br />

epicentral area, crossing the city of Lorca. This<br />

reason associated with the shallowness of the<br />

earthquake, would explain the high seismic intensity<br />

and peak accelerations of 0.36 g (IGN data) recorded<br />

in the accelerometer of the old prison of Lorca<br />

(located in the downtown). However, the high value<br />

of PGA could be also related to site effects.<br />

- The archaeoseismic data (more than a hundred<br />

values) suggest an origin of the deformation<br />

associated with a nearby seismic field, implying that<br />

the main earthquake rupture occurred beneath the<br />

historic city of Lorca because the faulting subsurface<br />

rupturing runs below the Lorca village.<br />

- The ESI 07 macroseismic classification (Michetti et<br />

al. 2007) for this earthquake is between VI and VII,<br />

according to the features observed. Hence, the<br />

interest of this small-sized earthquake is that it could<br />

be used as a lower limit for the instrumental<br />

calibration of the scale ESI 07.<br />

Acknowledgements: Thanks are given to the Spanish<br />

Instituto Geografico Nacional (IGN) for providing<br />

instrumental data of the earthquake. Also a special mention<br />

to the people of Lorca which helped us even with their<br />

homes collapsed and all people which showed friendly and<br />

kindness against the earthquake disaster.<br />

Bousquet, J.C. y Montenat, C. (1974). Presence décrochements<br />

NE-SW plio-quaternaires dans les Cordillères<br />

Bétiques Orientales (Espagne). Extension et signification<br />

général. C. R. Acad. Sci. Paris 278: 2617-2620.<br />

Giner-Robles, J.L. M.A. Rodríguez-Pascua, R. Pérez-<br />

López, P.G. Silva, T. Bardají, C. Grützner and K.<br />

Reicherter. (2009). Structural Analysis of Earthquake<br />

Archaeological Effects (EAE): Baelo Claudia Examples<br />

(Cádiz, South Spain), vol. 2,130p. Instituto Geológico y<br />

Minero de España, Madrid. Dep. Leg. M-27206-2009.<br />

Martínez Díaz, J.J., E. Masana , J.L. Hernández-Enrile & P.<br />

Santanach. (2001). Evidence for coseismic events of<br />

recurrent prehistoric deformation along the Alhama de<br />

Murcia Fault, Southeastern Spain. Geol. Acta. 36 (3-4):<br />

315-327.<br />

Martínez Díaz, J.J. (1998). Neotectónica y Tectónica Activa<br />

del Oeste de Murcia y sur de Almería (Cordillera Bética).<br />

Tesis Doctoral. Universidad Complutense Madrid. 470p.<br />

Martínez-Díaz, J. J. (2002). Stress field variety related to<br />

fault interaction in a reverse oblique-slip fault: the Alhama<br />

de Murcia Fault, Betic Cordillera, Spain. Tectonophysics,<br />

356: 291-305.<br />

Martínez Díaz J. J.; et al. (2011). Geological Preliminary<br />

Field Report of The Lorca Earthquake (5.1 Mw, 11th May<br />

2011). Instituto Geológico y Minero de España (IGME),<br />

Spanish Group of Active Tectonics, Paleosismicity and<br />

Risks, Universidad Complutense de Madrid (UCM),<br />

Universidad Autónoma de Madrid (UAM) and Universidad<br />

Rey Juan Carlos de Madrid (URJC). 45p.<br />

Michetti A.M., Esposito E., Guerrieri L., Porfido S., Serva L.,<br />

Tatevossian R., Vittori E., Audemard F., Azuma T.,<br />

Clague J., Comerci V., Gurpinar A., Mc Calpin J.,<br />

Mohammadioun B., Morner N.A., Ota Y. & Roghozin E.<br />

(2007). Intensity Scale ESI 2007. In. Guerrieri L. & Vittori<br />

E. (Eds.): Memorie Descrittive Carta Geologica. d’Italia.,<br />

vol. 74, Servizio Geologico d’Italia – Dipartimento Difesa<br />

del Suolo, APAT, Roma, 53 p.<br />

Rodríguez-Pascua M.A., R. Pérez-López, J.L. Giner-<br />

Robles, P.G. Silva, V.H. Garduño-Monroy and K.<br />

Reicherter. (2011). A Comprehensive Classification of<br />

Earthquake Archaeological Effects (EAE) in<br />

Archaeoseismology: application to ancient remains of<br />

Roman and Mesoamerican cultures. Quaternary<br />

International, doi:10.1016/j.quaint.2011.04.044.<br />

Silva, P.G., Goy, J.L., Zazo, C., Lario, J., Bardají, T. (1997).<br />

Paleoseismic indications along "aseismic" fault segments<br />

in the Guadalentín depresion (SE Spain). J.<br />

Geodynamics 24(1-4), 105-115.<br />

Wells, D.L. & Coppersmith, K.J. (1994). New empirical<br />

relationships among magnitude, rupture length, rupture<br />

width, rupture area, and surface displacement. Bull.<br />

Seismol. Soc. Amer. 84, 974-1002.<br />

193


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

FRONTIERS OF EARTHQUAKE ARCHAEOLOGY: THE OLYMPIA AND SAMICUM CASES<br />

(PELOPONNESE, GREECE)<br />

Reicherter, Klaus (1)<br />

(1) RWTH Aachen University, Inst. of Neotectonics and Natural Hazards, Aachen, 52056, Germany, email:<br />

k.reicherter@nug.rwth-aachen.de<br />

Abstract (Frontiers of earthquake archaeology: the Olympia and Samicum cases (Peloponnese, Greece)): The ancient<br />

Olympia on the western Peloponnese/Greece is worldwide known as the place where for 1170 years the classical Olympic Games<br />

took place. Furthermore, to domino-like toppled column drums of the Zeus temple are famous and widely accepted as earthquakerelated<br />

structural damage the 522/551 AD events. The Peloponnese peninsula is characterized by frequent seismicity within the<br />

Euroasian-African convergence zone (Hellenic arc and backarc system) by mainly E-W trending normal faults.<br />

Herein, some examples of the heterogeneous structural deformation are given observed in the ancient Olympia including an<br />

interpretation. Also, we show deformation patterns at Samicum, a Hellenistic fortified village (acropolis) with a cyclopean masonry<br />

situated on top of the coastal Lapithas mountain ridge, close to Olympia. Both examples may be interpreted as archaeoseismic<br />

evidence, however, an on-fault palaeoseismological approach and assessment is missing in this part of Greece.<br />

Key words: earthquake geology and archaeology, Olympia, Samicum, structural deformation<br />

INTRODUCTION<br />

This paper deals with the combination of well-known<br />

historic sites and their earthquake-related damage.<br />

By a number of ancient historians (see references)<br />

we know about the destruction of ancient Olympia by<br />

several earthquakes (early 4 th cent. AD, 522 AD, 551<br />

AD) and by warfare and vandalism. On the other<br />

hand, the close-by fortified village Samicum was<br />

described as abandoned by Pausanias in the 2 nd<br />

cent. AD (Fig. 1). This seems rather strange taking<br />

into account the strategic position of the village on<br />

top of the mountain Lapithas. We apply the<br />

classification of Earthquake Archaeological Effects of<br />

Rodríguez-Pascua et al. (2011; EAE) to describe and<br />

classify the observed damage. The investigation of<br />

past earthquakes is commonly split into several<br />

disciplines, which cover Instrumental Seismology,<br />

Historical Seismology, Archaeoseismology and<br />

Earthquake Geology or Palaeoseismology depending<br />

on information available (e.g. Caputo and Helly,<br />

2008). However, the ambiguity of data and results of<br />

distinct disciplines (Niemi, 2008), e.g. of written<br />

sources and recent observations, results in<br />

limitations of the seismic hazard assessment of an<br />

area (Nur, 2008).<br />

The Elis region is characterized by a huge number of<br />

neotectonic faults, which have been intensely studied<br />

(Fig. 2; Lekkas et al., 1993; Lekkas et al., 2000,<br />

Papanikolaou et al., 2007; Fountoulis and<br />

Mariolakos, 2008). The seismicity in the study area<br />

on the western Peloponnese peninsula is high. The<br />

last earthquake with M 5.5 struck the Pyrgos area on<br />

the 23 rd of March in 1993 with maximum intensities of<br />

VIII (e.g. Lekkas et al., 2000) associated with EEE<br />

(Earthquake Environmental Effects on the INQUA<br />

ESI-scale, Papanikolaou et al., 2009) like<br />

liquefaction, landslides, ground fractures and<br />

changes in the aquifer level.<br />

STUDY AREAS<br />

Fig. 1: Study area of ancient sites in Greece, inset<br />

shows location of Fig. 2.<br />

194


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: Study area of ancient sites on the Peloponnese peninsula, Greece, with major faults (compiled from Lekkas et al., 1993;<br />

Lekkas et al., 2000, Papanikolaou et al., 2007; Fountoulis and Mariolakos, 2008).<br />

Ancient Olympia<br />

The ancient Olympia is a famous World Heritage site<br />

located in the Peloponnesian province of Elis in the<br />

west of the peninsula (Figs.1 and 2). Known as the<br />

place of the classical Olympic Games, the sanctuary<br />

(Altis) consists of several buildings, which spread<br />

between the Kladeos creek and Mount Kronos. The<br />

Altis resembles a variety of buildings of different use<br />

was forgotten by people until 1829 AD. One of the<br />

most famous examples of EAE are the domino-style<br />

fallen columns of the Temple of Zeus (5 th cent. BC),<br />

which made it on the cover of a book entitled<br />

“Archaeoseismology” (see Stiros, 1996). The former<br />

Temple of Zeus, today known as Temple of Hera (it<br />

was rededicated in the 5 th cent. BC), is reported to<br />

have been destroyed by an earthquake in the early<br />

Date Event<br />

590-580 BC Construction of the Temple of Hera<br />

472-456 BC Construction of the Temple of Zeus<br />

c. 175 BC Earthquake? and repair of the Zeus<br />

statue, roof and columns (Pausanias,<br />

IV,31,6 and Dinsmoor, 1941)<br />

sculptures of the Temple of Zeus<br />

(Dinsmoor, 1941), which are replaced<br />

36 BC Earthquake? Second-hand roof tile –<br />

repair work (Dinsmoor, 1941)<br />

3 th cent. AD Series of earthquakes?<br />

early 4 th<br />

Earthquake (destroys Temple of Hera)<br />

cent. AD<br />

394 AD Pillage of Olympia by the Goths<br />

426 AD Theodosius orders destruction of the<br />

Temple of Zeus<br />

522 AD Earthquake (destroys Temple of Zeus)<br />

in combination with floods of the<br />

Alfeios and Kladeos, and landslide<br />

from the Kronos hill? (date unclear<br />

could be 551 AD)<br />

551 AD Earthquake (destroys Temple of Zeus)<br />

in combination with floods of the<br />

Alfeios and Kladeos, and landslide<br />

from the Kronos hill?<br />

Tab. 1: Timetable of con-/ and destruction observed in<br />

Olympia (compiled with different sources)<br />

and age during more than 3000 years (since c. 2500<br />

BP until 6 th cent. AD). Finally, the area was flooded<br />

and covered by sediments of the Kladeos creek and<br />

Alfeios river (Fouache and Pavlopoulos, 2010) and<br />

195<br />

Fig. 3: Fallen columns of the Temple of Zeus, Olympia<br />

(Classical period)<br />

4 th cent. AD, and was never rebuilt. Table 1 sums the<br />

construction and destruction events of the Zeus and<br />

Hera temples, several dates of possible destruction<br />

are ambiguous and not stated by exact dating,<br />

furthermore the causative fault(s) are not known.<br />

Samicum<br />

The fortified village of Samicum (or Macistus, ancient<br />

Kato Samia, however there is discussion, Pausanias<br />

and Polybius mention only Samicum, and Xenophon<br />

only Macistus) is situated in the southern part of the<br />

Elis region (Ilia, Triphylia) on top of the Lapithas<br />

mountain ridge, south of mouth of the Alfeios river


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

(Perseus Digital Library, 2011). Most probably<br />

Samicum was founded in the last 5 th cent. BC<br />

(classical Hellenistic period) and was occupied until<br />

the 2 nd cent. AD (Roman period). Samicum was<br />

occupied by the Aetolian Polysperchon against the<br />

Arcadians in 244 BC, and was later taken by Philip,<br />

in 219 BC. According to archaeologists the Samicum<br />

site has to be considered as one of the most<br />

important cities in Elis because of the strategic<br />

position on the mountain top, controlling the Kaiafas<br />

pass, lagoons and springs.<br />

The village is surrounded by a so-called pseudopolygonal<br />

wall of c. 1500 m length in a trapezoid<br />

shape with several towers, thus forming an acropolis.<br />

The cyclopean masonry of the wall surrounding<br />

Samicum is characterized by polygonal limestones,<br />

which fit each other with precision and minimal<br />

clearance between the stones and no use of mortar,<br />

little gaps are filled with shaped small stones.<br />

via the vaulted archway of c. 200 BC, dropped key<br />

stones can be observed (Fig. 4).<br />

Finally, the Phillipeion, a 338 BC construction, shows<br />

nice corner breakouts (Fig. 5), described as “dipping<br />

corners” by Rodríguez-Pascua et al. (2010) in the<br />

EAE.<br />

OBSERVATIONS<br />

Ancient Olympia<br />

Walking around in the Altis of Ancient Olympia<br />

reveals a good opportunity to also study earthquakerelated<br />

structural damage. On the other hand, a lot of<br />

reconstruction work and replacement limits full<br />

pleasure, which is further diminished by “slow”<br />

deformation due to landslides in the Kronos hill area.<br />

The most prominent features are the domino-like<br />

fallen columns of the Temple of Zeus (Fig. 3).<br />

Already Stiros (1996) stated that one of the columns<br />

is known to have been fallen during a storm (possibly<br />

after its reconstruction in the 19 th cent. AD). Another<br />

peculiar observation is that the colonnade columns<br />

fell towards the N- and the S-side of the temple´s<br />

long axis. The spaces between individual column<br />

drums are filled with sediment yielding ceramics.<br />

Fig. 5: Phillipeion, Olympia, c. 338 BC (Hellenistic<br />

period)<br />

However it is remarkable that the Roman ruins do not<br />

show such peculiar features of deformation. After an<br />

earthquake in the early Byzantine period (6 th cent.<br />

AD), the Altis was covered by a “mass of yellow earth<br />

spread over the entire area after a cloudburst<br />

perhaps in the 7 th cent. AD” (Dinsmoor, 1973).<br />

Samicum<br />

After reaching the ruins of Samicum on the Lapithas<br />

mountain enjoy the nice view around including the<br />

Volax and Kaifas lagoons. Little of the central part is<br />

excavated, the most impressive feature of Samicum<br />

is the cyclopean masonry of the city wall with several<br />

fallen towers and gates. Towards the south a steep<br />

cliff frames the fortified village. The wall and its<br />

blocks show various indicators of seismic damage.<br />

Among them are moved and rotated blocks, corner<br />

break-outs, collapsed watch towers (Figs. 6 and 7).<br />

Up to now these damages are not described and<br />

related to earthquakes, also there is no historic<br />

account for earthquake damage in Samicum. But<br />

Pausanias refers to his journey to Samicum, where<br />

Fig. 4: Vaulted entrance to the stadium Olympia (Crypt),<br />

approx. 200 BC (Hellenistic period)<br />

Dinsmoor (1941) observed to different styles of<br />

column drum beds, on which the connection to each<br />

other is observable: one with Lewis Holes, one<br />

without (but with the Empolion, the central hole). Of<br />

175 drums analyzed only 30 drums yielded Lewis<br />

Holes (Dinsmoor, 1941). This and repair work at<br />

columns as well as repair clamps at the western<br />

corners, lead Dinsmoor (1941) to postulate a 175 BC<br />

earthquake damage. Further on, entering the stadium<br />

196<br />

Fig. 6: Samicum, wall damage (Hellenistic period)


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

he found the village abandoned and destroyed in the<br />

2 nd cent. AD.<br />

Fig. 7: Samicum, wall damage and tower collapse<br />

(Hellenistic period)<br />

CONCLUSIONS<br />

We presented EAE of the ancient Greek sites of<br />

Olympia and Samicum, both have in common that<br />

the structural damage up to now has not been<br />

mapped, characterized and catalogued in detail.<br />

Also, the structural damage has not been classified<br />

and differentiated in the cause of damage. The<br />

Samicum case is most probably a “one-event”<br />

earthquake deformation. Whereas Ancient Olympia<br />

has obviously suffered from several earthquakes and<br />

other destructions (pillage, landslides and flooding).<br />

Both cases have also in common that the causative<br />

fault(s) for the earthquakes have not been detected<br />

yet, palaeoseismological studies are missing. Recent<br />

seismicity, such as the 1993 Pyrgos event (EEE<br />

description in Papanikolaou et al., 2009) shows that<br />

even moderate earthquakes (M 6.5) can cause<br />

severe structural damage and secondary effects as<br />

landslides and liquefaction. So, as a conclusion both<br />

cases reveal clearly the frontiers of<br />

archaeoseismology, and are at the moment that what<br />

Sintubin et al. (2008) pointed out: just a good story.<br />

References<br />

Caputo, R., Helly, B. (2008). The use of distinct disciplines<br />

to investigate past earthquakes. Tectonophysics,<br />

453(14), 719.<br />

Dinsmoor, W.B. (1941). An archaeological earthquake at<br />

Olympia. Am. J. Archaeol., 45/3, 399-427.<br />

Dinsmoor, W.B. (1973). The architecture of Ancient Greece.<br />

An account of its historic development. Biblio and Tannen,<br />

424 pp.<br />

Fouache, E., Pavlopoulos, K. (2010). The interplay between<br />

environment and people from Neolithic to Classical times<br />

in Greece and Albania. In: Martini, I.P. (ed.). Landscapes<br />

and Societies: Selected Cases, Chapter 10, 155-166.<br />

Fountoulis, I., Mariolakos, I. (2008). Neotectonic folds in the<br />

central-western Peloponnese, Greece. Z. dt. Ges.<br />

Geowiss., 159/3, 485-494.<br />

Lekkas, E., Papanikolaou, D., Fountoulis, L. (1993).<br />

Neotectonic map of Greece, Pyrgos-Tropaia sheets, scale<br />

1:100.000, Research Project of eth Univ. of Athens, Dept.<br />

of Geology, Division of Dynamic, Tectonic and Applied<br />

Geology, Athens. Available online at:<br />

http://elekkas.gr/en/research/mapping.html<br />

Lekkas, E., Fountoulis, I., Papanikolaou, D. (2000). Intensity<br />

distribution and neotectonic macrostructure Pyrgos<br />

Earthquake data (26 March 1993, Greece). Nat. Haz., 21,<br />

19-33.<br />

Niemi, T. (2008). Historical Earthquake Catalogues and<br />

Archaeological Data: Avoiding Circular Reasoning.<br />

Seismological Research Letters, 79(2), 289.<br />

Nur, A. (2008). Apocalypse. Earthquakes, Archaeology, and<br />

the Wrath of God. Princeton University Press, Princeton.<br />

Papanikolaou, D., Fountoulis, I., Metaxas, C. (2007): Active<br />

faults, deformation rates and Quaternary paleogeography<br />

at Kyparissiakos Gulf (SW Greece) deduced from<br />

onshore and offshore data. Quat. Int., 171-172, 14–30.<br />

Papanikolaou, I.D., Papanikolaou, D.I., Lekkas, E. (2009).<br />

Advances and limitations of the Environmental Seismic<br />

Intensity scale (ESI 2007) regarding near-field and farfiled<br />

effects from recent earthquakes in Greece:<br />

implication for the seismic hazard assessment. In:<br />

Reicherter, K., Michetti, A.M., Silva, P.G. (eds.).<br />

Paleoseismology: Historical and prehistorical records of<br />

earthquake ground effects for seismic hazard<br />

assessment. J. Geol. Soc. London Spec. Publ., 316: 11-<br />

30.<br />

Pausanias (c. 115-180 AD). Pausanias Description of<br />

Greece with an English Translation by W.H.S. Jones,<br />

Litt.D., and H.A. Ormerod, M.A., in 4 Volumes.<br />

Cambridge, MA, Harvard University Press; London,<br />

William Heinemann Ltd. 1918. Online at<br />

http://www.perseus.tufts.edu.<br />

Perseus Digital Library, lasted visit 5/2011:<br />

http://www.perseus.tufts.edu<br />

Polybius (c. 200 – 120 BC). Histories. Evelyn S.<br />

Shuckburgh (translator). London, New York. Macmillan.<br />

1889. Reprint Bloomington 1962. Online at<br />

http://www.perseus.tufts.edu.<br />

Rodríguez-Pascua, M.A., Pérez-López, R., Giner-Robles,<br />

J.L., Silva, P.G., Garduño-Monroy, V.H., Reicherter, K.<br />

(2011, in press). A comprehensive classification of<br />

Earthquake Archaeological Effects (EAE) in<br />

Archaeoseismology: application to ancient remains of<br />

Roman and Mesoamerican cultures. Quat. Int.<br />

Sintubin, M., Stewart, I. S., Niemi, T. and Altunel, E. (2008).<br />

Earthquake Archaeology Just a Good Story?<br />

Seismological Research Letters, 79(6), 767768.<br />

Stiros, S.C. (1996). Identification of Earthquakes from<br />

Archaeological Data: Methodology, Criteria and<br />

Limitations. In: Archaeoseismology (edited by Stiros, S. C.<br />

and Jones, R. E.). Fitch Laboratory Occasional Paper 7.<br />

Institute of Geology and Mineral Exploration and the<br />

British School at Athens, Athens, 129152.<br />

Strabo (c. 63 BC - 29 AD), Geography. Online at<br />

http://www.perseus.tufts.edu.<br />

Xenophon (c. 430-354 BC), Hellenica, 3.2. Online at<br />

http://www.perseus.tufts.edu.<br />

197


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

REGIONAL STRAIN-RATES ON ACTIVE NORMAL FAULTS AND VARIABILITY IN THE<br />

SEISMIC CYCLE: AN EXAMPLE FROM THE ITALIAN APENNINES<br />

Roberts, Gerald (1, Joanna Faure Walker (1), Patience Cowie (2), Richard Phillips (3), Ken McCaffrey (4), Ioannis Papanikolaou<br />

(5), Max Wilkinson (4), Alessandro Michetti (6), Peter Sammonds (1).<br />

(1) Department of Earth and Planetary Sciences, Birkbeck/UCL, University of London, Gower Street, WC1E 7HX. UK.<br />

gerald.roberts@ucl.ac.uk<br />

(2) Department of Earth Sciences, University of Bergen, Postboks 7800, NO-5020 BERGEN<br />

(3) Institute of Geophysics and Tectonics, School of Earth and Environment, University of Leeds, LS2 9LT, UK<br />

(4) Department of Earth Sciences, Durham University, Science Labs, Durham DH1 3LE. UK.<br />

(5) Mineralogy-Geology Lab of the Agricultural University of Athens, Greece.<br />

(6) Università degli Studi dell'Insubria - Sede di Como, Facoltà di Scienze MM. FF. NN., Dipartimento di Scienze Chimiche e<br />

Ambientali, Via Valleggio, 11 - 22100 Como – Italy.<br />

Abstract(Regional strain-rates on active normal faults and variability in the seismic cycle: an example from the Italian<br />

Apennines): The rate at which a fault slips fundamentally determines the seismic hazard because average earthquake recurrence<br />

intervals tend to decrease as slip rates increase. Slip-rates on different faults within a growing fault system are not distributed<br />

randomly in space and time, because slip-rates must accommodate the regional strain-rate. Here we show how slip-rates vary<br />

temporally and spatially along the Italian Apennines, constrained by offsets across fault scarps dated with 36 Cl in situ cosmogenic<br />

dating. We map regional variations in strain-rate, investigating natural variability in the seismic cycle.<br />

Key words: Fault scarps, regional strain-rates, cosmogenic dating, seismic-cycle.<br />

INTRODUCTION<br />

Palaeoseismologists and earthquake geologists<br />

should try to move forward from simply characterising<br />

slip on single faults. Instead, they should try to<br />

characterise slip across systems of faults as it is<br />

clear that faults interact through stress transfer, and<br />

this controls regional patterns of slip-rate (the<br />

geography of seismic hazard), and the recurrence<br />

intervals for earthquakes on specific faults. The<br />

regional approach is needed to complement the<br />

regional strain-rate databases provided through GPS<br />

geodesy and instrumental seismicity. We should<br />

attempt to constrain how the geography of seismic<br />

hazard implied over 10 2 -10 3 years (palaeoseismology<br />

and earthquake geology) compares with that over 10-<br />

100 years (geodesy and instrumental seismicity). We<br />

expect differences between data from these different<br />

timescales to reveal temporal and spatial variability in<br />

the seismic cycle (e.g. Faure Walker et al. 2010).<br />

Fault scarps and 36 Cl cosmogenic dating<br />

The Italian Apennines are characterised by fault<br />

scarps that offset and deform deposits and landforms<br />

formed during the last glacial maximum. The offsets<br />

have accumulated since 15 ±3 ka shown by studies<br />

of tephrachronology and 14 C dating (Giraudi and<br />

Frezzotti 1995, 1997) and this is confirmed by 36 Cl in<br />

situ cosmogenic dating (Palumbo et al. Schlagenhauf<br />

2009, Schlagenhauf et al. 2010), and our own<br />

ongoing cosmogenic dating. The scarps are<br />

widespread and allow a regional study of strain rates<br />

and natural variability in the seismic cycle.<br />

198<br />

Figure 1. A regional strain-rate map for the Italian<br />

Apennines constrained by offsets of 15 ±3 ka<br />

features across fault scarps.<br />

Regional Strain rates<br />

A key unknown is how far a fault can stray from its<br />

long-term slip rate, both at a timescale equivalent to<br />

the interseismic period (10 2 -10 3 years), and over<br />

timescales equivalent to several seismic cycles (10 3 -<br />

10 4 years). The lack of such knowledge impedes our<br />

ability to perform probabilistic seismic hazard


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

assessments and understand the underlying physics<br />

that controls repeated earthquake slip. In order to<br />

study the existence of possible deficits or surpluses<br />

of geodetic and earthquake strain in the Italian<br />

Apennines compared to 15 ±3 kyr multi-seismic-cycle<br />

strain-rates, horizontal strain-rates are calculated<br />

using slip-vectors from striated faults and offsets of<br />

Late Pleistocene-Holocene landforms and sediments,<br />

using an adaptation of the Kostrov equations (Faure<br />

Walker 2010).<br />

Strain-rates calculated over 15 ±3 kyr within 5km x<br />

5km grid squares vary from zero up to 2.34 ±0.54 x<br />

10 -7 yr -1 , 3.69 ±1.33 x 10 -8 yr -1 , and 1.20 ±0.41 x 10 -7<br />

yr -1 in the central Apennines Lazio-Abruzzo region,<br />

the Molise-North Campania region, and the southern<br />

Apennines South Campania-Basilicata region,<br />

respectively. The data resolve variations in strain<br />

orientations and magnitudes along the strike of<br />

individual faults. Strain-rates over a time period of 15<br />

±3 kyrs from 5km x 5km grid squares integrated over<br />

an area of 1.28 x 10 4 km 2 (80 km x 160 km), show<br />

the horizontal strain-rate of the central Apennines is<br />

1.18 (+0.12/-0.04) x 10 -8 yr -1 parallel to the regional<br />

principal strain direction (043-223 o ±1 o ). In Molise<br />

and North Campania, the horizontal principal strainrate<br />

calculated over an area of 5 x 10 3 km 2 (50 km x<br />

100 km) is 2.11 (+1:14/-0:16) x 10 -9 yr -1 along the<br />

principal horizontal strain direction (039-219 o ±3 o ).<br />

Within the southern Apennines region within an area<br />

of 8 x 10 3 km 2 (50 km x 160 km), the average<br />

horizontal principal strain-rate is 3.70 ±0:26 x 10 -9 yr -1<br />

along the horizontal principal strain direction (044-<br />

224 o ±2 o ).<br />

Strain-rates calculated within 5 x 5 km and 20 x 20<br />

km grid squares, and at a regional scale, are highest<br />

in the central Apennines, medial in the southern<br />

Apennines and lowest Molise and North Campania.<br />

At the regional length-scale, the strain-rates are<br />

comparable in direction and magnitude to strain-rates<br />

calculated using GPS. Smaller areas (~2000-7000<br />

km 2 ), corresponding to polygons defined by geodesy<br />

campaigns (126 years) and seismic moment<br />

summations (700 years) show higher 10 2 yr strainrates<br />

than 10 4 yr strain-rates in some areas, with the<br />

opposite situation in other areas where seismic<br />

moment release rates in large (> Ms 6.0) magnitude<br />

historical earthquakes have been reported to be as<br />

low as zero. High strain-rates over 15 ±3 kyr in<br />

places occur where instrumental seismicity rates<br />

have been low. This demonstrates that strain-rates<br />

vary spatially on the length-scale of 10-100 km and<br />

on a timescale between 10-100 yrs and 10 4 yrs in the<br />

Italian Apennines.<br />

The multi seismic cycle strain-rates are used to<br />

calculate earthquake recurrence intervals for a given<br />

earthquake slip magnitude, at the scale of individual<br />

seismic sources; these value are compared to<br />

palaeoseismic data. The results are used to discuss<br />

spatial and temporal earthquake clustering and the<br />

natural variability of the seismic cycle.<br />

Acknowledgements: We thank the Natural Environment<br />

Research Council for funding. NERC Urgency Grant<br />

NE/H003266/1. A LiDAR and field study of surface rupture<br />

and post-seismic slip for the 6 th<br />

April 2009 L’Aquila<br />

Earthquake (M6.3). Dr. K. McCaffrey, Dr. G. P. Roberts,<br />

Prof. P. Cowie. (April 2009-May 2010). NERC Standard<br />

Grant NE/E01545X/1. Testing Theoretical models for<br />

Earthquake Clustering using 36 Cl Cosmogenic Exposure<br />

Dating of Active Normal Faults in Central Italy. Prof. P.<br />

Cowie, Dr. G. P. Roberts, Dr. K. McCaffrey (October 2007-<br />

2010).<br />

References<br />

Faure Walker, J., Mechanics of continental extension from<br />

Quaternary strain field in the Italian Apennines.<br />

Unpublished PhD Thesis, University College London.<br />

2010.<br />

Faure Walker, J., Roberts, G. P., Sammonds, P., Cowie, P.<br />

A., Comparison of earthquakes strains over 10 2 and 10 4<br />

year timescales: insights into variability in the seismic<br />

cycle in the central Apennines, Italy. Journal of<br />

Geophysical Research VOL.115, B10418,<br />

doi:10.1029/2009JB006462, 2010<br />

Giraudi, C., Frezzotti, M., 1995. Paleoseismicity in the Gran<br />

Sasso Massif (Abruzzo, Central Italy). Quaternary<br />

International, 25, 81-93.<br />

Giraudi, C., Frezzotti, M., 1997. Late Pleistocene glacial<br />

events in the central Apennines, Italy. Quaternary<br />

Research, 48, 280-290.<br />

Palumbo, L., L. Benedetti, D. Bourles, A. Cinque, R. Finkel,<br />

2004. Slip history of the Magnola fault (Apennines,<br />

Central Italy) from<br />

36 Cl surface exposure dating:<br />

evidence for strong earthquakes over the Holocene,<br />

Earth and Planetary Science Letters, 225, 163-176.<br />

Schlagenhauf, A., Identification des forts seismes passes<br />

sur les failles normales actives de la region Lazio-<br />

Abruzzo (Italie centrale) par “datations cosmogeniques”<br />

(36Cl) de leurs escarpments. These, Docteur de<br />

l’Universite Joseph Fourier, Grenoble, France. 2009.<br />

Schlagenhauf, A., et al. Using Chlorine-36 cosmonuclides to<br />

recover earthquake histories on limestone normal fault<br />

scarps : a reappraisal of methodology and<br />

interpretations. Geophys. J. Int., doi:10.1111/j.1365-<br />

246X.2010.04622.x 2010.<br />

199


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE VARIABILITY OF ALONG-STRIKE CO-SEISMIC SLIP: A NEW EXAMPLE FROM THE<br />

IMPERIAL FAULT OF SOUTHERN CALIFORNIA<br />

Rockwell, Thomas K., (1,2, Yann Klinger (2)<br />

(1) Geological Sciences, San Diego State University, San Diego, CA 92182 USA, trockwell@geology.sdsu.edu<br />

(2) Institute du Physique du Globe, Paris<br />

Abstract (The variability of along–strike co-seismic slip: A new example from the Imperial Fault of Southern California):<br />

The 1940 M7 earthquake on the Imperial fault in southern California and northern Baja California produced 70 km of surface<br />

rupture along a N37W striking right lateral fault. This rupture crossed cultivated fields with long alignments of crop and tree rows,<br />

along with an orthogonal pattern of canals and roads, and 15 km of the rupture was photographed from the air in high-resolution<br />

stereo immediately after the earthquake. We determined the precise scale of these vintage aerial photographs and used them to<br />

assess the short spatial variability of slip along strike. We also analyzed 3 km of very high-resolution aerial photography taken the<br />

day after the 1979 rupture of the northern half of the Imperial fault. We find that lateral slip varies substantially along strike by<br />

more than 30% over distances of tens to hundreds of meters. These results are similar to the variability determined after the 1999<br />

Izmit and Duzce ruptures in Turkey, and the 2010 El Mayor-Cucapah rupture in northern Baja California.<br />

Key words: surface rupture, Imperial Fault, rupture variability, rupture distribution<br />

INTRODUCTION<br />

The short spatial variability of co-seismic slip along<br />

strike for large strike-slip fault ruptures has generally<br />

been attributed to the inability of geologists to<br />

measure off-fault deformation. However, after the<br />

1999 Izmit and Duzce earthquakes in Turkey,<br />

Rockwell et al. (2002) demonstrated from offset<br />

groves of trees that both off-fault near-field<br />

deformation and lateral variations of co-seismic slip<br />

along fault strike were significant variables in making<br />

field measurements of displacement after large<br />

earthquakes. New optical correlation techniques<br />

applied to the 2010 El Mayor-Cucapah surface<br />

rupture in northern Baja California (Hudnut et al.,<br />

2010) confirm that there are both short and long<br />

wavelength along-strike variations in lateral slip as<br />

well as substantial off-fault deformation such that<br />

field measurements of lateral slip consistently<br />

underestimated the amount of strike-slip, even where<br />

basement rock is close to the surface and alluvial<br />

cover is thin.<br />

In this work, we present new measurements for the<br />

1940 and 1979 surface ruptures along the Imperial<br />

fault in southern California (Figure 1), made from<br />

high-resolution aerial photography. We measured<br />

over 600 new displacements along 15 km of the 1940<br />

rupture, and over 300 new displacements along 3 km<br />

of the 1979 rupture, which are the extents of highresolution<br />

aerial photography for these ruptures.<br />

Measurements are mostly on long, linear cultural<br />

features and crop rows, and are spaced between one<br />

to tens of meters apart, allowing for resolution of<br />

variations in lateral slip. We present these new<br />

measurements and conclude that slip variations are<br />

real aspects of co-seismic slip along large strike-slip<br />

faults. We also show that the slip distribution reported<br />

for the 1940 earthquake underestimates the<br />

maximum displacement, as field measurements after<br />

the earthquake were made on widely spaced cultural<br />

200<br />

features, such as roads and canals, and the zone of<br />

maximum displacement was not visited.<br />

These observations point to the need for field<br />

geologists to over sample displacement data after an<br />

earthquake, rather then make random, widely spaced<br />

measurements that may not represent the actual coseismic<br />

slip distribution.<br />

DETERMINATION OF SCALE OF THE AERIAL<br />

PHOTOGRAPHY<br />

We determined the scale for the aerial photographs<br />

by measuring the distance between field boundaries<br />

(~400 m) and other man-made structures and then<br />

comparing them to the identical features in Google<br />

Earth. We made measurements only in the central<br />

(50%) portion of the stereo pairs, as the affects of<br />

parallax are minimal in this part of the imagery. We<br />

then constructed a scale that is accurate to within<br />

1%. We then used the scaling tool in Adobe<br />

Illustrator to construct a smaller scale to measure<br />

offset cultural features.<br />

The largest uncertainty in both the 1940 and 1979<br />

data sets is the determination of the alignment of a<br />

crop row, tree row, road, canal, or other feature. We<br />

used a straight-line segment to place along a feature<br />

on one side of the fault, and copied the segment to<br />

maintain a parallel line for the same feature across<br />

the fault. The line width is about 10 cm, which is<br />

about the smallest increment of displacement that<br />

could be measured for the 1940 imagery. For the<br />

1979 photography, we could commonly measure to<br />

about 5 cm. A significant source of uncertainty is the<br />

actual placement of the line along a feature, which is<br />

dependent on how straight the feature was, and how<br />

far it could be extended beyond the near-field fault<br />

zone. In most cases with crop rows, the plow lines<br />

were found to be remarkably straight outward from<br />

the fault for tens to hundreds of meters, and straight<br />

line segments were easy to place along the middle or


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 1: Location map of the Imperial fault in southern California and northern Baja California. Note that the 1940 earthquake<br />

ruptured the entire length of the Imperial fault, whereas the 1979 earthquake only ruptured the northern half. Also note the<br />

paleoseismic sites along the fault, and the other historical surface ruptures.<br />

edge of a crop row. For the 1979 imagery, we<br />

estimate that small crop rows had placement<br />

uncertainties of only 10-15 cm, whereas for the 1940<br />

imagery, the uncertainties are larger and perhaps as<br />

much as 0.5 m. In affect, this provided hundreds of<br />

alignment arrays to be measured along the strike of<br />

the fault to assess lateral slip variability, with<br />

reasonably small uncertainty estimates.<br />

OBSERVATIONS<br />

edges of a road adjacent to a field was offset about 7<br />

m (Figure 3). Slip decreased<br />

in both directions to between 5 and 6.5 m, similar to<br />

the variability in lateral displacement that we<br />

observed along the entire photographed portion of<br />

the rupture.<br />

Comparison to J.P. Buwalda’s field measurements<br />

from his field notes<br />

As a first test of our methods, we located the sites of<br />

Buwalda’s field measurements that were made<br />

immediately after the 1940 rupture, of which there<br />

were only six along the 15 km of rupture captured in<br />

the aerial photography. Figure 2 shows our slip<br />

estimates versus Buwalda’s field measurements;<br />

they are in close agreement except for the All<br />

American Canal (site B1a) where Buwalda reports<br />

only a single strand, whereas two strands are clearly<br />

evident in the imagery and cause displacement of the<br />

canal, resulting in our larger estimate.<br />

Maximum Displacement<br />

Maximum displacement for the 1940 earthquake<br />

exceeded the reported 6 m of lateral displacement by<br />

about a meter. In the same general vicinity as<br />

Sharp’s 6 m measurement, we determined that both<br />

201<br />

Fig. 2: Comparison of J.P. Buwalda’s field<br />

measurements with our estimates from analysis of<br />

aerial photography.<br />

Slip Distribution<br />

Figure 4 shows the slip distribution for the 1940<br />

earthquake from our new measurements, along with


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

J.P. Buwalda’s 1940 field measurements in red<br />

(corrected to a bearing of 323 o ). We also include<br />

measurements by R.V. Sharp on preserved features<br />

north and south of the international border (black<br />

Fig. 3: The 1940 rupture along the Imperial fault north of<br />

the international border. The All American Canal is just<br />

meters north of the border. About 2 km NW from the<br />

border, a road/field boundary is offset about 7 m, the<br />

maximum displacement that we found along the 1940<br />

rupture. R.V. Sharp measure about 6 m near the border,<br />

whereas J.P. Buwalda measured only about 5 m across<br />

one of two strands. The detail shows the offset, with<br />

scale included, at our measurement points 38 and 39.<br />

dots). Finally, we surveyed a line of telephone poles<br />

that were established prior to the 1940 earthquake at<br />

Tamaulipas (formerly Cucapa) and resolved 2.7 m of<br />

right-lateral strike-slip (blue star). Our new<br />

observations increase the number of slip<br />

observations for the 1940 surface rupture by over an<br />

order of magnitude, although the majority are on a<br />

section of the fault that represents only 20-25% of the<br />

full length of the rupture.<br />

An important observation on slip distribution is that<br />

by Rockwell et al. (2002) for the 1999 Izmit and<br />

Duzce earthquakes. This degree of variability has<br />

been observed for many previous earthquakes,<br />

including the 1999 Hector Mine earthquake (Treiman<br />

et al., 2002), the 1992 Landers earthquake (Sieh et<br />

al., 1993), the 1987 Superstition Hills earthquake<br />

(Sharp et al., 1989), and many others. However, in<br />

most previous cases, slip measurements were made<br />

on small geomorphic features such as channel<br />

margins, channel thalwegs, alluvial bars, canyon<br />

walls, etc., none of which were linear for any distance<br />

from the fault. Consequently, near-field off-fault<br />

deformation could not be assessed and it was<br />

generally assumed that some of the variability was<br />

the result of non-quantified offset. In contrast,<br />

measurement of long crop rows, tree lines, roads,<br />

fences and other long cultural features along the<br />

1940 surface rupture demonstrates that significant<br />

lateral variations in displacement are real, and that<br />

they occur over short spatial dimensions. For<br />

instance, we determined displacement for adjacent<br />

rows of trees and crops for entire fields. In some<br />

cases, as in figure 5, the variability occurs at about<br />

our estimated resolution of displacement uncertainty,<br />

a half meter. In this case, we measured offset of<br />

individual tree lines to vary by about 1 m over a<br />

lateral distance of a few hundred meters, similar to<br />

the variability on offset tree lines in the Izmit<br />

earthquake, but one could argue that within the<br />

stated uncertainty, these measurements agree.<br />

However, other examples, such as the offset crop<br />

rows in figure 6, we measured the offsets to +10 cm,<br />

and estimate the uncertainty to about +20 cm. In this<br />

case, lateral slip varied from zero to over a meter<br />

along a several hundred meter section of rupture.<br />

The overall degree of variability along strike is<br />

Fig. 4: Slip distribution for the 1940 Imperial fault rupture. Our new observations are between the international border (zero<br />

point) and about km 14 north of the border. J.P. Buwalda’s field measurements are plotted in red, R.V. Sharp’s estimates of<br />

displacement are plotted as black dots, and our measurement of an offset telephone line is the blue star.<br />

there are significant lateral variations in displacement<br />

over short spatial dimensions, similar to that reported<br />

202<br />

evident in the slip distribution curve in figure 4. Areas


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

of larger offset tend to have greater variations in<br />

displacement, even though resolution of displaced<br />

features and estimates of uncertainty remain about<br />

the same. From these observations, we conclude<br />

Fig. 5: Offset grove of trees between 2.1 and 2.4 km NW<br />

of the international border. A dot was placed on the<br />

center of each tree, and the dots were regressed to<br />

resolve lateral displacement. Uncertainty is estimated at<br />

about 0.5 m.<br />

much of this is rarely accounted for in the<br />

measurement of non-linear geomorphic features but<br />

it is clearly evident for the Imperial ruptures.<br />

Based on the collection of nearly 1000 new<br />

displacement measurements from the 1940 and 1979<br />

surface ruptures along the Imperial fault, several<br />

generalizations can be made. First, the sparse field<br />

data from the 1940 earthquake, nearly all of which<br />

was collected at convenient road crossings, missed<br />

the maximum displacement of about 7m, as well as<br />

the maximum displacement along strike, although<br />

most of the field observations fall near the average<br />

displacement for a section of fault. It is evident from<br />

this analysis that dense collection of displacement<br />

data is required to quantify the maximum<br />

displacement along a rupture, and that sparse data<br />

likely miss such displacement peaks and other<br />

details of the rupture.<br />

Second, lateral variations along strike are substantial<br />

and show similar variability to that documented with<br />

survey data after several recent earthquakes. Offfault<br />

warping is evident in the bending of crop rows,<br />

and some of this bending may be missed in near-field<br />

field mapping of a rupture after an earthquake. Using<br />

long agricultural features as closely spaced<br />

alignment arrays allows for both the assessment of<br />

total slip at a point along the fault, as well as the<br />

lateral variability of displacement along the fault. The<br />

variability that we document here is similar in<br />

magnitude and spatial scales to that documented<br />

after the Izmit and Duzce earthquakes, as well as<br />

more recently with optical imaging techniques after<br />

the 2010 El Mayor-Cucapah earthquake in northern<br />

Baja California (Hudnut et al., 2010).<br />

References<br />

Fig. 6: Offset crop rows between 11.7 and 12 km NW of<br />

the international border. Crops rows could be measured<br />

to 10-15 cm resolution, but offsets vary by over a meter.<br />

that lateral variability of displacement typically varies<br />

by as much as 30% along a section of rupture.<br />

DISCUSSION AND CONCLUSIONS<br />

Until the advent of pre- and post-earthquake<br />

comparisons of LiDAR and optical imagery data,<br />

most measurements of lateral displacement along<br />

strike-slip faults after large earthquakes were<br />

conducted on nearfield, non-linear features such as<br />

rills, stream channels, channel margins, bars, and<br />

other common geomorphic features (Clark et al.,<br />

1972; Sharp, 1982; Sieh et al., 1993; Treiman et al.,<br />

2002; Barka et al., 2002). Surveying of long cultural<br />

features such as tree rows, telephone pole arrays,<br />

and fence lines demonstrate that significant off-fault<br />

warping can account for a substantial amount of the<br />

near-field strike slip (Rockwell et al., 2002), and<br />

Barka, A., Akyuz, H. S. et al. 2002. The surface rupture and<br />

slip distribution of the 17 August 1999 Izmit earthquake<br />

(M7.4), North Anatolian fault. Bulletin of the<br />

Seismological Society of America (Special Issue on the<br />

1999 Izmit and Duzce, Turkey, Earthquakes, N. Toksoz<br />

(ed.)), 92(1), 43–60.<br />

Clark, M.M., 1972, Surface rupture along the Coyote Creek<br />

fault, in The Borrego Mountain Earthquake of April 9,<br />

1968, U.S. Geol. Sur. Prof. Pap. 787, p. 55-87.<br />

Hudnut, K.W., J.M. Fletcher, T.K. Rockwell, J.J. Gonzalez-<br />

Garcia, O. Teran, and S.O. Akciz, 2010, Earthquake<br />

rupture complexity evidence from field observations. EOS<br />

Transactions, Fall AGU, T51E-02.<br />

Rockwell, T. K., S. Lindvall, T. Dawson, R. Langridge, W.<br />

Lettis, Y. Klinger, 2002, Lateral offsets on surveyed<br />

cultural features resulting from the 1999 Izmit and Duzce<br />

earthquakes, Turkey. Bulletin of the Seismological<br />

Society of America, v. 92, no. 1, pp. 79-94.<br />

Sharp, R.V., 1982, Comparison of 1979 surface faulting with<br />

earlier displacements in the Imperial Valley, in The<br />

Imperial Valley, California earthquake of October 15,<br />

1979: U.S. Geological Survey Professional Paper 1254,<br />

p. 213-221.<br />

Sieh, K., L. Jones, et al., 1993, Near-field investigations of<br />

the Landers earthquake sequence, April to July 1992,<br />

Science v. 260, pp. 171–176.<br />

Treiman, J.A., Kendrick, K.J., Bryant, W.A., Rockwell, T.K.,<br />

and McGill, S.F., 2002, Primary surface rupture<br />

associated with the Mw 7.1 16 October 1999 Hector Mine<br />

earthquake, San Bernardino County, California. Bulletin<br />

of the Seismological Society of America, v. 92, no. 4, pp.<br />

1171-1191.<br />

203


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKE ARCHAEOLOGICAL EFFECTS GENERATED BY THE LISBON<br />

EARTHQUAKE (FIRST OF NOVEMBER 1755) IN THE CORIA´S CATHEDRAL (CÁCERES,<br />

WESTERN SPAIN)<br />

Rodríguez-Pascua M.A. (1), P.G. Silva (2), Perucha Atienza, M.A. (1), J.L. Giner-Robles (3), R. Pérez-López (1)<br />

(1) Instituto Geológico y Minero de España. Street Ríos Rosas, 23. 28003-Madrid. SPAIN. E-mail: ma.rodriguez@igme.es,<br />

ma.perucha@igme.es ; r.perez@igme.es<br />

(2) Dpto. Geología, Escuela Politécnica Superior de Ávila, Universidad de Salamanca. Avda. Hornos Caleros, 50. 05003-Ávila.<br />

pgsilva@usal.es<br />

(3) Dpto. Geología. Facultad de Ciencias. Universidad Autónoma de Madrid. Cantoblanco. Tres Cantos. Madrid. SPAIN. E-mail:<br />

jlginer@gmail.es<br />

Abstract (Earthquake Archaeological Effects generated by the Lisbon Earthquake (first of November 1755) in the Coria´s<br />

Cathedral (Cáceres, western Spain): The Lisbon Earthquake was the most destructive earthquake in the Western European<br />

history. This earthquake affected the entire Iberian Peninsula and the city of Lisbon completely collapsed. The intensity of this<br />

earthquake was X (EMS-1998) and damaged the historical buildings of Spain. These effects are preserved in historical buildings,<br />

like the Coria’s Cathedral. The damage observed in this cathedral is described by using the new classification of Earthquake<br />

Archaeological Effects (EAE), with the aim to study both strain structures and the seismic wave pattern.<br />

Key words: Lisbon Earthquake (1755), Coria´s Cathedral, Earthquake Archaeological Effects (EAE).<br />

INTRODUCTION<br />

The Lisbon earthquake (November 1 st , 1755) is the<br />

largest earthquake that struck Western Europe in<br />

historic times. This earthquake affected the<br />

population in a physical sense and also changed the<br />

knowledge of the scientific origin of earthquakes.<br />

This date was the starting point of foundation of<br />

modern seismology. The Lisbon earthquake affected<br />

the entire Iberian Peninsula and North Africa. Its<br />

shaking was felt in Central European countries like<br />

Germany (Martínez Solares, 2001). Nowadays the<br />

epicentre of the earthquake is still the subject of<br />

scientific debate (Gutscher, 2005), although the<br />

approximate location is widely assumed at the<br />

southwest of San Vicente Cape, regardless of the<br />

exact fault that produced it. The maximum intensity<br />

of this earthquake was X (EMS-1998) (Martínez<br />

Solares and Mezcua, 2002), being located in<br />

southern Portugal (Algarve Coast). Some of those<br />

seismic intensity effects from the earthquake are still<br />

visible in the historic heritage all around the Iberian<br />

Peninsula, such as churches and cathedrals. This is<br />

the case study of the Coria´s Cathedral (Cáceres,<br />

central part of Spain). The building suffered structural<br />

damage, even collapse the cupola of the tower.<br />

GEOGRAPHICAL SETTING<br />

The town of Coria is located to the NW of the city of<br />

Caceres in the central Western part of Spain and<br />

near of the border with Portugal. The isoseismal map<br />

of the Lisbon Earthquake (Martínez-Solares, 2001)<br />

suggests that Coria and surroundings were affected<br />

by a seism intensity VI (Fig. 1). The main damage in<br />

the city was produced in the cathedral, in which the<br />

collapse of the cupola of the tower killed thirteen<br />

people.<br />

Fig. 1: Location of the town of Coria on the isoseismal<br />

map of the 1755 Lisbon earthquake (Intensity scale<br />

EMS-1998) (after Martínez Solares, 2001).<br />

METHODOLOGY<br />

Some of the effects of the earthquake in the<br />

cathedral of Coria were documented by the Dean of<br />

the Cathedral (Martínez-Vázquez, 1999) and are<br />

available for public consult in the archive of the<br />

Cathedral. Hence, we have elaborated a list of<br />

structural damage using the classification of<br />

Earthquake Archaeological Effects (EAE)<br />

(Rodríguez-Pascua et al., 2011). However, we have<br />

to bear in mind that the Cathedral is also affected by<br />

geotechnical problems that could mask the effects of<br />

the earthquake of Lisbon. Those have to be<br />

discriminated by using existing documentation, both<br />

the earthquake and previous geotechnical studies<br />

(Martínez-Vázquez, 1999).<br />

204


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Rotated and displaced drums in columns and<br />

pinnacles:<br />

12- Clockwise rotations of the pinnacles<br />

(hexagonal pyramidal base) of the north<br />

front of the Cathedral (Fig. 6).<br />

13- Collapse of the tower’s cupola.<br />

Displaced masonry blocks:<br />

14- Sinistral displacement of masonry blocks in<br />

the columns of the organ (Fig. 7).<br />

15- Displacement of masonry blocks in the<br />

tower’s balcony.<br />

Fig. 2: Plan of the Coria´s Cathedral. The red<br />

numbers are the location of the different EAEs (see<br />

the text for further explanation).<br />

DATA<br />

According to the classification of damage proposed<br />

by the EAE scale (Rodríguez-Pascua et al., 2011),<br />

the effects of the Lisbon earthquake catalogued in<br />

the Cathedral of Coria are the following (see Fig 2 for<br />

spatial location):<br />

Penetrative fractures in masonry blocks:<br />

1- Cracks in the East front. Existing prior to the<br />

earthquake and enhanced by this.<br />

2- Cracks in the lintel of the north entrance and<br />

in the north front of the Cathedral.<br />

3- Cracks in the west main front separated by<br />

the central column which differentiates the<br />

two entrance arches of the door (Fig. 3).<br />

4- Cracks in the ceilings and vaults of the<br />

central nave of the building.<br />

Dropped key stones in arches or lintels in windows<br />

and doors:<br />

5- Arches of the bell tower with horizontal<br />

displacement of key stones in the segments<br />

to the arcs of the N front of the tower.<br />

6- Arch of the north entrance with dropped<br />

lintel.<br />

Collapsed walls or balustrades:<br />

7- Collapsed balustrade and pinnacles of the<br />

“Relics Balcony” (Fig. 4).<br />

8- Fall down of the balustrade’s pinnacles in<br />

the south terrace of the Cathedral (Fig. 5A).<br />

9- Fall down of the balustrade’s pinnacles in<br />

the north and south front of the Cathedral<br />

(Fig. 5B).<br />

10- Collapsed balustrades of the balconies in<br />

the south front subsequently repaired with<br />

bricks.<br />

11- Fall down of the tower’s pinnacles.<br />

Fig. 3: Cracks in the W main front of the Coria´s<br />

Cathedral. A) View before restoration; B) view after<br />

restoration in 2009.<br />

DISCUSSION AND CONCLUSIONS<br />

The intense seismic damage of the Coria Cathedral<br />

could be attributed to its topographic location: on the<br />

cliff edge overlooking the Alagón River’s flood plain.<br />

The average orientation of cracks (NE-SW) to about<br />

45° to the central axis of the building, suggest a<br />

sinistral shear in the building. All of this data<br />

(checked by epoch documents) could be applied in<br />

other buildings affected by the Lisbon Earthquake.<br />

205


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

As a principal geological effect of the Lisbon<br />

Earthquake within the area, the river modified his<br />

course and changed the channel by moving to the<br />

south. Consequently, the medieval “Stone Bridge”<br />

was out of use and today you can ask yourself why<br />

medieval people had built a heavy stone bridge in the<br />

middle of the cultivation land?<br />

Fig. 5: A) Fall down of the balustrade´s pinnacles in<br />

the south terrace of the Cathedral B) decapitated<br />

pinnacles in the north front of the Cathedral.<br />

Fig. 4: Collapsed balustrade and pinnacles of the<br />

“Relics Balcony”.<br />

Acknowledgements: Thanks are given to the project<br />

ACI2009-1037 (MICINN).<br />

References<br />

Gutscher, M.A. 2005. What caused the Great Lisbon<br />

Earthquake? Science, 305: 1247-1248.<br />

Martínez-Solares, J.M. 2001. Los efectos en España del<br />

terremoto de Lisboa. Instituto Geográfico Nacional.<br />

Ministerio de Fomento. 756 pp.<br />

Martínez-Solares, J.M. & Mezcua, J. 2002. Catálogo<br />

sísmico de la Península Ibérica (880 a.C.-1900). Instituto<br />

Geográfico Nacional. Ministerio de Fomento. 756 pp.<br />

Martínez-Vázquez, F. 1999. El terremoto de Lisboa y la<br />

Catedral de Coria (Vicisitudes del Cabildo) 1755-1759.<br />

Colección Temas Cauriaciences. Vol. V. Ed.<br />

Ayuntamiento de Coria. 187 pp.<br />

Rodríguez-Pascua, M.A.; Pérez-López, R.; Giner-Robles,<br />

J.L.; Silva, P.G.; Garduño-Monroy, V.H. & Reicherter, K.<br />

2011. A comprehensive classification of Earthquake<br />

Archaeological Effects (EAE) in archaeoseismology:<br />

Application to ancient remains of Roman and<br />

Mesoamerican cultures. Quaternary International, DOI:<br />

10.1016/j.quaint.2011.04.044.<br />

Fig. 6: Clockwise rotations of the pinnacles<br />

(hexagonal pyramidal base) of the north front of the<br />

Cathedral.<br />

Fig. 7: Sinistral displacement of masonry blocks in the<br />

columns of the organ.<br />

206


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

NEOTECTONIC OF THE LONGITUDINAL FAULT SYSTEM IN SOUTHERN<br />

COSTA RICA<br />

Rojas, Wilfredo (1), Nury Simfors-Morales (2), Luis Sanez (3), Åke Sivertun (4)<br />

(1, 3) Escuela Centroamericana de Geología, University of Costa Rica. Email: wrojas@geologia.ucr.ac.cr<br />

(2, 4) Department of Computer Science, University of Linköping, Sweden. Email: nury.simfors2@comhem.se<br />

Abstract (Neotectonic of the longitudinal fault system in Southern Costa Rica): Southern Costa Rica is one of the most<br />

seismically active areas in Costa Rica because the subduction between Cocos, Caribbean and Nazca Plates. Several active fault<br />

zones go also across the study area. One of this fault zones is the WNW-ESE Longitudinal fault system (LFS) which is the longest<br />

fault in southern Costa Rica. The goal of this study is to find a suitable place to use it as a sanitary landfill site. The results indicate<br />

that there are several geological structures in the study area that indicate deformation due active faulting but more detail studies<br />

are necessary in order to understand better the seismogenic potential of this fault. Future work in southern Costa Rica will take<br />

into account these issues.<br />

Keywords: Trenching, Neotectonics, Longitudinal, Subduction<br />

INTRODUCTION<br />

Most of the seismic activity in southern Costa Rica is<br />

due to the subduction between Cocos, the Caribbean<br />

and the Nazca Plates. However, the area is also<br />

characterized by several active fault systems which<br />

cut across southern Costa Rica, causing also seismic<br />

activity and serious damage during strong<br />

earthquakes. The Longitudinal Fault System (LFS) is<br />

the longest fault system in southern Costa Rica with<br />

a length of about 167.2 km (Mann & Corrigan, 1990),<br />

(Kolarsky et. al., 1995), (Cowan et al., 2001) and<br />

(Montero, 1998) (Fig. 1). It is an ESE-striking obliquereverse<br />

fault, parallel to the volcanic arc in Costa<br />

Rica and western Panama (Montero, 1994). The<br />

recurrence of earthquake events of this fault is not<br />

well known, while the fault slip rate may be roughly<br />

estimated at 15 mm/year in Costa Rica and 10<br />

mm/year in Panamá (Cowan et al., 2001).<br />

The objective of this preliminary work is to find a<br />

suitable place to use it as a sanitary landfill site. Two<br />

trenches were open on the outskirts of Rio Claro,<br />

southern Costa Rica (Fig. 1). The results indicate that<br />

there are several geological structures in one of the<br />

excavated trenches that indicate deformation due to<br />

active faulting. Two other fault scarps on the way to<br />

Ciudad Nelly cited in this work were also visited.<br />

They could be potential trenching targets in case<br />

more studies will be done in the future. Taking into<br />

account the importance of this fault in southern Costa<br />

Rica and western Panama is recommended to do<br />

more studies and apply dating techniques in order to<br />

better understand the seismogenic potential of the<br />

fault and the damage it could pose to the<br />

communities. The University of Costa Rica (UCR)<br />

has a seismic monitoring project in southern Costa<br />

Rica which is studying the spatial and temporal<br />

seismicity and earthquake rupture processes in the<br />

region. The LFZ and other important faults in the<br />

region will be studied in more detail in a close future.<br />

Fig. 1: Digital Elevation Model of Costa Rica showing the<br />

location of southern Costa Rica. In the map two triangles<br />

show the location of Rio Claro and Ciudad Nelly. The<br />

trenches were excavated in Rio Claro (see: Fig.3). The fault<br />

scarp follows toward the NE and ca. 1 kilometer from the<br />

trench side the fault scarp can be seen clearly along the<br />

smooth plain of southern Costa Rica ( Fig. 4 and 5). On the<br />

outskirts of Ciudad Nelly was also found two interesting<br />

places where the LFZ could be studied in the near future<br />

(Fig. 6). (Map courtesy of University of Costa Rica).<br />

207


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

HISTORIC SEISMICITY IN SOUTHERN COSTA<br />

RICA<br />

The spatial distribution of the main historical<br />

subduction related earthquakes (1803, 1854, 1904,<br />

1941 and 1983) which have affected southern Costa<br />

Rica over the past 200 years have been generated<br />

by the subduction of the Cocos plate along its<br />

interface with Panama Microplate. Although most of<br />

the historic earthquakes are subduction related, few<br />

of them are due to faulting. Subduction earthquakes<br />

rupture at shallow focal depths between 28 and 34<br />

km and magnitudes around 7.0 to 7.5 (Rojas et al.,<br />

1993) and (Morales-Simfors et al., submitted).<br />

According to the historic earthquakes there is a<br />

medium-term probability of a strong earthquake Mw<br />

7.3 in the cited area in the near future. The<br />

recurrence interval of strong events (Mw ≥ 7.0) is in<br />

the order of 45 years. In case a big earthquake<br />

occurs in southern Costa Rica, damage is expected<br />

with a Mercalli Intensity of VIII (Rojas 2008) and this<br />

may cause significant casualties in the communities.<br />

Fig. 2: Simplified geological map of southern Costa Rica. It<br />

shows sedimentary rocks in brown (Mesozoic-Paleozoic),<br />

Volcanic Rocks in green (Miocene-Pliocene), Volcanic<br />

Quaternary rocks in dark green. The small map in the left<br />

corner shows the major fault zones in the study area (Maps<br />

courtesy of UCR).<br />

THE LONGITUDINAL FAULT SYSTEM (LFZ)<br />

The WNW-ESE trending Longitudinal fault system<br />

extends from western Panama to central Costa Rica<br />

in the west (Mann & Corrigan, 1990) (Fig.2). The slip<br />

rate of the Longitudinal fault zone may be


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

One of the trenches presents a clear evidence of<br />

geological structures due to active faulting.<br />

Fig. 5: A. View of the fault scarp at NE of the trench site at<br />

Rio Claro (Coordinates: 569.405 291.319, Piedras Blancas<br />

Sheet, 1.50000). B. The scarp is smaller than the scarp<br />

close to the trench in figure 2. This fault scarp is ca. 2<br />

meters high and it extends several kilometers along the<br />

plain in Rio Claro. C. The fault trace goes almost<br />

perpendicular across the road. It is possible to see a narrow<br />

change in the dip of the road.<br />

The Fig. 3B and C shows a close view of the grayblue<br />

deformed clay layer in the trench which is<br />

overturned and faulted by two thrust faults and other<br />

secondary normal faults but in order to do a better<br />

interpretation of the tectono-stratigraphic sequence of<br />

this site, a detail study of the fault is necessary. The<br />

LFZ and other faults in the region appears to be in<br />

obvious proximity with important infrastructure such<br />

as hospitals, schools, roads, highways etc. in<br />

southern Costa Rica and western Panama, In order<br />

to understand the active tectonics in the area, more<br />

detail studies are necessary to evaluate the<br />

seismogenic potential of the fault and answer other<br />

questions about the age of the fault scarps,<br />

earthquake recurrence and seismic hazard in<br />

southern Costa Rica.<br />

Acknowledgements: The authors thank Peter Bryer for<br />

helping us with the English manuscript. We are also grateful<br />

to the Municipality of Golfito, southern Costa Rica and the<br />

Osa-Golfo Dulce Research Program (PIOSA), University of<br />

Costa Rica<br />

References<br />

Fig. 6: View of fault scarp at Abrojo, Costa Rica<br />

(Coordinates: 582.785 and 286146, Canoas topographic<br />

sheet, 1 50000). The scarp goes almost parallel to the<br />

highway from Costa Rica to Panama.<br />

The second place was 1 kilometer away from the<br />

trench site, where the fault scarp can be seen clearly<br />

and follows towards the NE several kilometers along<br />

the smooth plains in southern Costa Rica (Fig. 4).<br />

The displacement of this scarp is ca. 2 meters (Fig.<br />

5). The third point was outside Ciudad Nelly 30<br />

kilometers to the SE of the trench side where it is<br />

possible to see several geological structures from the<br />

main highway that communicate Costa Rica with<br />

Panama. In this site the scarp is ca 2 meters high<br />

(Fig. 6).<br />

RESULTS AND RECOMMENDATIONS<br />

Two trenches were excavated in the study area in<br />

order to find a suitable site for a sanitary landfill site.<br />

Cowan H, M., N. Machette., K.M. Haller & R.L. Dart (2001).<br />

Map and Database of Quaternary Faults and Folds in<br />

Panama and Its Offshore Regions. Open-File Report 98-<br />

779. 41 pp.<br />

Fisher, D. M., T.W, P. Gardner., P. Sak; J. Marshall & M.<br />

Protti (2001). Uplift patterns in the Forearc of the Middle<br />

America Trench, Costa Rica: implications for mass<br />

balance and fore-arc kinematics. Eos, Transactions of the<br />

American Geophysical Union, 82, F1151.<br />

Kolarsky, R.A. & P. Mann (1995). Structure and<br />

neotectonics of an oblique-subduction margin,<br />

southwestern Panama. In: Geologic and tectonic<br />

development of the Caribbean Plate boundary in<br />

Southern Central America: (Mann, P. ed.). Geological<br />

Society of America Special Paper 295, p. 131-157.<br />

Mann, P. & J.D, Corrigan (1990). Model for late Neogene<br />

deformation in Panama: Geology, 8, 558-562.<br />

Montero, W. (1994). Neotectonics and related stress<br />

distribution in a subduction-collisional zone, Costa Rica.<br />

In: Geology of an evolving island arc (Seyfried, H. and<br />

Hellmann, W, eds.). Profil (University of Stuttgart,<br />

Germany, 7, 125-141.<br />

Montero, W., P. Denyer., R. Barquero., G.E. Alvarado., H,<br />

Cowan., M.N, Machette., K.M, Haller & R.L.Dart (1998).<br />

Map and database of Quaternary faults and folds in<br />

Costa Rica and its offshore regions: U.S. Geological<br />

Survey Open-File Report 98-481, 63 p., 1 plate<br />

(1:750,000 scale).<br />

Mora, S (1979). Estudio geológico de una parte de la región<br />

sureste del Valle del General, Provincia Puntarenas,<br />

Costa Rica: San José, University of Costa Rica, unpubl.<br />

Senior Undergraduate thesis, 185 p.<br />

Morales-Simfors, N; et. al. A contribution to the study of the<br />

Osa-Golfo Dulce Seismicity, Costa Rica: March 11-14<br />

earthquakes. Int. J. Earth Sci. (Submitted).<br />

Rojas, W., Bungum, H & C, Lindholm (1993). Historical and<br />

recent earthquakes in Central America. Revista<br />

Geológica de América Central, 16, 5-21.<br />

Rojas, W. (2008). Informe anual del proyecto de<br />

investigación de monitoreo sísmico de Zona Sur de<br />

Costa Rica VI-113-A7-170, RSN, San José, 13 pp.<br />

209


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

GEODETIC STUDIES IN THE ZAFARRAYA FAULT (BETIC CORDILLERAS)<br />

Ruano, Patricia (1,2), Antonio J. Gil (3), Jesús Galindo-Zaldívar (1,2), Gracia Rodríguez-Caderot (4), María Clara de Lacy (3),<br />

Antonio M. Ruiz (3), María Jesús Borque (3), Juan A. Armenteros (3), Antonio Herrera (3), Antonio Jabaloy (1), Angel C. López-<br />

Garrido (2), Antonio Pedrera (5), Carlos Sanz de Galdeano (2)<br />

(1) Dept. Geodinámica, Universidad de Granada. Campus Fuentenueva 18071-Granada, Spain. Email: pruano@ugr.es<br />

(2). Instituto Andaluz de Ciencias de la Tierra, Facultad de Ciencias, 18071 Granada, Spain<br />

(3). Dept. Ing. Cartográfica, Geodésica y Fotogrametría. E.P.S., Universidad de Jaén, Campus Las Lagunillas, 23071 Jaén,<br />

Spain.<br />

(4). Dept. Astronomía y Geodesia, Universidad Complutense de Madrid, Ciudad Universitaria s/n, 28040 Madrid, Spain.<br />

(5). Instituto Geológico y Minero de España. C/ Alcázar del Genil, 4. 18006 Granada, Spain.<br />

Abstract (Geodetic studies in the Zafarraya fault (Betic cordilleras): The Zafarraya Fault Zone constituted the seismic source<br />

of the 1884 Andalucian earthquake with an estimated magnitude of 6.5 – 7. It is a normal fault zone located in the northern limb of<br />

the Sierra Tejeda antiform and intersects the Internal and External Zone boundary of the Betic Cordillera. Two non- permanent<br />

GPS networks made up of 16 sites have been installed and surveyed in September 2004 and February and July 2010 in order to<br />

constrain the present-day motion of these structures. Result in 6 years elapsed time may help to discuss about coseismic and<br />

aseismic motions of the fault and elastic deformation before faulting.<br />

Key words: GPS network, Zafarraya Fault, Sierra Tejeda antiform, present-day deformation<br />

THE ZAFARRAYA FAULT AND SIERRA TEJEDA<br />

ANTIFORM<br />

The Betic Cordillera is located in the Western<br />

Mediterranean built up by the distributed<br />

deformations related to the Eurasian-African plate<br />

boundary and characterized by a present day<br />

moderate seismicity.<br />

The Zafarraya Fault Zone is located in the Southwest<br />

of Granada Basin, which is one of the most important<br />

Neogene intramountain basins in the Cordillera that<br />

cover the Internal and External zones boundary. It is<br />

placed at the northern limb of the Sierra Tejeda<br />

antiform, and could be interpreted as a collapse<br />

structure developed along the external arch of the<br />

uplifted fold.<br />

The main normal fault trends roughly E-W and<br />

includes other minor faults like Llanos de la Dona<br />

Fault. The whole fault system constitutes a set of<br />

fractures extending 23 km in length and 4 km in width<br />

that runs obliquely to the boundary between the<br />

External and Internal zones. The main fault segment<br />

has associated a rectilinear mountain front formed in<br />

Jurassic limestones of the External Zone and Triassic<br />

marbles of the Internal Zone. This mountain front is<br />

divided into several segments of E-W to NW-SE<br />

orientations. The kinematics of the fault is mainly<br />

normal with a minor dextral component as can be<br />

deduced by the observed slicken-lines. This fault<br />

controls the main features of the Zafarraya polje, an<br />

endorheic area developed in its hanging wall, and<br />

filled by Tortonian to Quaternary sediments (López-<br />

Chicano et al., 2002). The throw of this fault zone is<br />

1500 m, considering the displacement of the<br />

External-Internal zone boundary. Taking into account<br />

210<br />

the regional geological setting, most of the slip<br />

probably occurred since the Tortonian. The Zafarraya<br />

Fault was the causative fault of the largest historical<br />

earthquake registered in Spain (the Andalusian<br />

Earthquake, December 25, 1884) with a maximum<br />

intensity of X (MSK scale) from which a magnitude of<br />

6.5 - 7 has been calculated (Muñoz and Udías, 1981)<br />

with an estimated total rupture of 16 km.<br />

The slip-rate of the fault calculated from geological<br />

markers (10 Ma, Tortonian) is 0.125 mm/yr (Sanz de<br />

Galdeano et al., 2003) and 0.17 mm/yr (Reicherter et<br />

al., 2003). Whereas, from paleoseismological studies<br />

the slip-rate estimated is 0.35 mm/yr (Reicherter et<br />

al., 2003), 0.3 and 0.45 mm/yr with a recurrence<br />

period of 2-3 kyr for major, surface rupturing<br />

earthquakes (Reicherter et al., 2010).<br />

THE GPS NETWORKS<br />

Two non-permanent GPS networks were installed in<br />

Zafarraya Fault Zone and Sierra Tejeda antiform in<br />

2004. The 16 GPS sites (Fig. 1) are located in a local<br />

and a regional network that extends up to the coast<br />

line in order to study the local motion along the<br />

Zafarraya fault and the regional development of the<br />

Sierra Tejeda antiform. These networks were made<br />

up of sixteen reinforced concrete pillars anchored to<br />

rock with an embedded forced centring system to<br />

assure that the antennas are placed exactly at the<br />

same position in different reoccupations. The local<br />

network comprises the sites 811, 812 and 816,<br />

located on the hanging wall of the fault. Most of them<br />

were built up on Jurassic limestones of the External<br />

Zones. The sites 810, 813, 814 and 815 are located<br />

in the footwall, and were build-up mainly on<br />

limestones and marbles. Although these sites are


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

located across the contact between External and<br />

Internal Zones, this contact is inactive at Present,<br />

and the southern part of the network may constitute a<br />

reference for the activity of the most recent Zafarraya<br />

Fault. The regional network that extends from the site<br />

800, located on Jurassic limestones of the External<br />

Zones, crosses southwards the Zafarraya Fault and<br />

the contact between External and Internal Zones and<br />

reaches the uppermost part of Sierra Tejeda in the<br />

site 850, located on Triassic Alpujarride marbles. The<br />

southern part of the network is located along the<br />

southern limb of Sierra Tejeda, which is deformed by<br />

NW-SE oriented normal faults, and reaches the coast<br />

line (sites 890 and 880, builds up respectively on<br />

Alpujarride metapelites and marbles). The regional<br />

network also covers the WNW periclinal end of the<br />

Sierra Tejeda antiform (Fig. 1).<br />

All sites meet the following requirements: no<br />

obstruction above 15 degrees; no high power lines<br />

nearby; easily accessible.<br />

The first survey was done in September 2004<br />

(Borque et al. 2005) and a. second survey in<br />

February 2010, for local network and in July 2010 for<br />

regional one.<br />

The GPS constellation was tracked throughout a<br />

three-day campaign with 24-hour sessions per day in<br />

the local network and a six-day campaign in two<br />

settles with 2 shared sites in the regional one.<br />

For data acquisition we used 6 dual frequency carrier<br />

phase GPS receivers Leica System 1200, consisting<br />

of GX1230 receivers and AX1202 antennas.<br />

The GPS data processing was performed by using<br />

Bernese 5.0 software in the following way: single<br />

sessions were computed in multibaseline mode. The<br />

first step (preprocessing) related to receivers clocks<br />

calibration, performed by code pseudoranges, and<br />

detection and repair of cycle slips and removal of<br />

outliers, was carried out simultaneously for L1 and L2<br />

data. The final solution for each session was<br />

obtained using the iono-free observable with precise<br />

ephemeris and absolute antenna phase centre<br />

variation files. The fixed solution of the coordinates<br />

was estimated using the QIF method to fix integer<br />

ambiguities. Troposphere parameters every two<br />

hours were estimated.<br />

Fig. 1: Geological map and GPS network locations of Sierra Tejeda and Zafarraya fault.<br />

211


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

RESULTS AND DISCUSSION<br />

These geodetic surveys provide new insights on the<br />

Zafarraya fault (ZF) and Sierra Tejeda antiform<br />

(STA). Slow-motion faults need long periods of<br />

measurements to have suitable information.<br />

The obtained results for both networks suggest a<br />

slow to moderated motion of the structures during the<br />

last 6 years. In Figure 2 are shown the present-day<br />

crustal deformation rates in terms of the annual<br />

WNW trend. There are small differences due to the<br />

local activity of tectonic structures. To facilitate the<br />

interpretation of the presented results apart from a<br />

Eurasia fixed reference frame the results are<br />

presented (blue arrows, Fig. 2) in a frame were we fix<br />

the station 810, shared by both networks and located<br />

southward ZafarRaya fault. This way, it is easy to<br />

appreciate the motion of ZF and STA. South of the<br />

ZF most of the statiofs show very small northward<br />

motion that sugest a very slow activity of STA that<br />

constitutes the footwall of the fault. However, in the<br />

Fig. 2: GPS Velocity vectors in mm/yr with standard error ellipses. Black arrows: Eurasia fixed reference frame; Blue<br />

arrows: Site 810 fixed reference frame. The seismicity from September 2004 to July 2010 from IGN data base is shown.<br />

velocity vectors with Eurasia fixed reference frame.<br />

These deformation rates are in order of several mm<br />

per year. Although velocity vectors must be quoted<br />

with its standard ellipses even largest of some annual<br />

velocity vector, all the vectors are consistent with a<br />

hanging wall the behavior od the sites is variable.<br />

While site 812 and the western end underline its<br />

dextral character, sites 811 and 816 may show a<br />

combination of motion with a cover NW-SE normal<br />

fault producing active NW-SE extension. The sites<br />

212


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

820 and 830 located in its eastern area show its<br />

normal motion. Toward the North, the vector of site<br />

800 could be related to the dextral motion of the<br />

Llanos de la Dona Faults.<br />

From these results a NW-SE to N-S extension up to<br />

2 mm/yr could be estimated for the Zafarraya fault<br />

zone. These rates are higher than those estimated by<br />

paleoseismological studies. Moreover, the<br />

paleoseismic data suggests that recent activity is<br />

higher that neotectonic activity. However, it is not<br />

well-constrained the importance of creep motions in<br />

the fault and the accommodation of elastic<br />

deformation in the area before seismic activity.<br />

The historical and instrumental (Fig. 2) data support<br />

that Zafarraya fault has a seismic activity with<br />

discontinuous slip-rate. The occurrence of at least<br />

two seismic series in the area (Sep 2005 and Aug-<br />

Sep 2007), during elapsed time between surveys,<br />

could be the responsible of these highest rates that<br />

confirm the coseismic behavior of the ZF.<br />

The present and future results will be very significant<br />

for a better understanding of the active tectonic<br />

structure interaction providing short-term mode of<br />

deformation and slip rates of one the most active<br />

sectors of the Betic Cordilleras.<br />

Acknowledgements: Projects CSD2006-00041, CGL-<br />

2008-03474-E/BTE, CGL2010-21048, P09-RNM-5388<br />

AYA2010-15501, and research groups of Junta de<br />

Andalucía RNM282, RNM370 and RNM148 are<br />

acknowledged.<br />

References<br />

Borque, M.J., Galindo-Zaldívar, J., Gil, A.J., Jabaloy, A.,<br />

Lacy, M.C., López, A.C., Rodríguez-Caderot, G., Ruiz,<br />

A.M., Sanz de Galdeano, C. (2005): Establishment of a<br />

non-permanent GPS network to monitor the deformation<br />

in Zafarraya Fault and Sierra Tejeda Antiform (Spain).<br />

Física de la Tierra, 17, 23-31.<br />

López Chicano, M.; Calvache, M. L.; Martín-Rosales, W. y<br />

Gisbert, J. (2002). Conditioning factors in flooding of<br />

karstic poljes—the case of the Zafarraya polje (South<br />

Spain). Catena, 49, 331-352.<br />

Muñoz, D., Udias, A. (1981) - El Terremoto de Andalucía<br />

del 25 de Diciembre de 1884. Instituto Geográfico<br />

Nacional, Madrid.<br />

Reicherter, K., Jabaloy, A., Galindo-Zaldívar, J., Ruano, P.,<br />

Becker-Heidmann, P., Morales, P., Reiss, S. & González-<br />

Lodeiro, F. (2003) Repeated palaeoseismic activity of the<br />

Ventas de Zafarraya fault (S Spain) and its relation with<br />

the 1884 Andalusian earthquake. Int. J. Earth Sci., 92,<br />

912-922.<br />

Reicherter, K., Jabaloy, A., Galindo-Zaldívar, J., Becker-<br />

Heidmann, P. & Sanz de Galdeano, C. (2010). Trenching<br />

results of the Ventas de Zafarraya Fault (Southern<br />

Spain). In: Contribución de la Geología al Análisis de la<br />

Peligrosidad Sísmica (J.M. Insua y F. Martín-González,<br />

eds.). Sigüenza (Guadalajara, España). 125.<br />

Sanz de Galdeano, C., Pelaéz Montilla, J.A. & López<br />

Casado, C. (2003) Seismic Potential of the Main Active<br />

Faults in the Granada Basin (Southern Spain). Pure and<br />

Applied Geophysics, 160, 1537-1556.<br />

213


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

NEOTECTONIC ACTIVITY OF THE GRANADA BASIN –<br />

NEW EVIDENCE FROM THE PADUL-NIGÜELAS FAULT ZONE<br />

Rudersdorf, Andreas, Jochen Hürtgen, Christoph Grützner, Klaus Reicherter<br />

Lehr- und Forschungsgebiet Neotektonik und Georisiken, RWTH Aachen, Lochnerstraße 4-20, 52056 Aachen (Germany).<br />

E-Mail: andreas.rudersdorf@rwth-aachen.de , jochen.huertgen@rwth-aachen.de<br />

Abstract (Neotectonic activity of the Granada Basin – New evidence from the Padul – Nigelas fault zones): The Padul-<br />

Nigüelas Fault Zone (PNFZ) is situated at the south-western mountain front of the Sierra Nevada (Spain) in an extensive regime<br />

and belongs to the internal zone of the Betic Cordilleras. The aim of this study is a collection of new evidence for neotectonic<br />

activity of the fault zone with classical geological field work and modern geophysical methods, such as ground penetrating radar<br />

(GPR). Among an apparently existing bed rock fault scarp with triangular facets, other evidences, such as deeply incised valleys<br />

and faults in the colluvial wedges, are present in the PNFZ. The preliminary results of our recent field work have shown that the<br />

synsedimentary faults within the colluvial sediments seem to propagate basinwards and the bed rock fault is only exhumed due to<br />

erosion for the studied segment (west of Marchena). We will use further GPR data and geomorphologic indices to gather further<br />

evidences of neotectonic activity of the PNFZ.<br />

Key words: Active faulting, ground penetrating radar (GPR), neotectonics, Granada Basin<br />

INTRODUCTION<br />

The Granada Basin is a seismically active<br />

intramontane basin in the Betic Cordillera of<br />

Andalucía, Southern Spain. Earthquakes have been<br />

documented both in the historic and the instrumental<br />

record (Reicherter (2001) among others). Situated in<br />

an extensive regime, normal faults delimit basins of<br />

Neogene and Quaternary ages. In this study, modern<br />

geophysical methods like ground penetrating radar<br />

(GPR) have been combined with classical geologic<br />

field work to gain new insight into neotectonic activity<br />

of the Padul-Nigüelas Fault Zone.<br />

sedimentary succession from Tortonian ages (11.6 -<br />

7.2 Ma BP) on and consist of calcarenites,<br />

evaporites, terrigenous clastics as well as carbonate<br />

intercalations. The youngest sequences are<br />

composed of Quaternary colluvial and alluvial<br />

deposits. The footwall, the Alpujárride complex,<br />

comprises several units, such as metapelites,<br />

metapsammites and quartzites at the base and<br />

carbonate rocks as uppermost formation.<br />

(Azañón et al., 2002)<br />

GEOLOGY<br />

The convergence of the Eurasian and African plates<br />

resulted in coupled extension and compression<br />

during the Oligocene and Miocene whereas recent<br />

stress fields designate a NW-SE oblique<br />

convergence with a velocity of 4 mm/a (McClusky et<br />

al., 2003). As a result the Betic Cordillera orogeny<br />

formed a highly folded mountain belt simultaneously<br />

with associated normal faults. Uplift and exhumation<br />

resulted in erosion of stacked lithological units<br />

(Galindo-Zaldívar et al., 2003). The Betic Cordilleras<br />

are subdivided into an External Zone and an Internal<br />

Zone, which is subdivided into the Malaguide-<br />

Complex, the Alpujárride-Complex and the Nevado-<br />

Filábride-Complex because of different metamorphic<br />

facies.<br />

The Padul-Nigüelas Fault Zone (PNFZ), situated in<br />

the internal zone of the Betic Cordilleras, is part of a<br />

NW-SE trending system of normal faults adjacent to<br />

sedimentary basins. They usually show a<br />

Fig. 1: Geological Map of the southern Iberian region (after<br />

Reicherter & Peters, 2005).<br />

The PNFZ (Fig. 2) is well documented in the<br />

westernmost and easternmost parts whereas faults<br />

are circumspectly indicated in the central part north<br />

of Dúrcal. Due to the hard-rock carbonate lithology of<br />

the Alpujarrian basement, active faults in this<br />

segment are commonly preserved as bedrock faultscarps.<br />

214


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: Geological Map of the Padul-Nigüelas Fault Zone, simplified and composed from Sanz de Galdeano et al. (1975) and<br />

González Donoso et al. (1978).<br />

METHODS<br />

Apart from classical field methods, such as<br />

geological mapping, we used geophysical survey<br />

methods like ground penetrating radar and laser<br />

distance measurements to get a detailed image of<br />

both subsurface and outcropping structures in order<br />

to give evidence for historic and recent seismic<br />

Fig. 3: Illustration of a laser distance measurement.<br />

Exhumed scarp near the natural monument “Nigüelas<br />

Fault”.<br />

215<br />

activities in the Padul-Nigüelas Fault Zone.<br />

Laser distance measurements<br />

Laser distance measurements (LDM) have been<br />

done with a compact rangefinder (LTI TruPulse 360)<br />

with an integrated tilt sensor and compass to<br />

measure slope, horizontal distance, vertical distance,<br />

inclination and azimuth of points in outcrops. LDM<br />

was used (1) to describe undulating scarp<br />

geometries and their dimensions and (2) to get exact<br />

relationships between several faults in outcrops of<br />

colluvial wedges and alluvial fans. The method of<br />

LDM is illustrated in Fig. 3. The rangefinder is<br />

mounted on a tripod and the position of the tripod is<br />

stationary during a measurement. With a narrow grid,<br />

it is possible to get detailed information about<br />

exhumed scarp planes or faults in colluvial wedges.<br />

We have recorded these measurements at several<br />

sites in the PNFZ.<br />

Ground penetrating radar<br />

Ground penetrating radar is a non-invasive<br />

geophysical method which uses electromagnetic<br />

waves for shallow subsurface surveys. Due to<br />

changing magnetic and electric properties in the<br />

underground, reflections of transmitted waves can be<br />

registered and the two way travel time (TWT, in ns)<br />

of the waves gives information about the depth of a<br />

reflector. We used a 270 MHz antenna to record the<br />

raw data, which have to be interpreted by various<br />

filtering and correction tools. According to subsurface<br />

conditions (permittivity, conductivity, presence of<br />

water-saturated or clay-rich sediments), the<br />

penetration depth varies between 5 and 10 m with<br />

high-resolution data. In our study area, we especially


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

investigated the colluvial wedges and alluvial fans of<br />

the Padul-Nigüelas Fault Zone with the aim to find<br />

basinwards trending faults in the subsurface.<br />

PRELIMINARY RESULTS<br />

The 3-week field work for this study has just been<br />

completed in the end of April 2011 and the evaluation<br />

and processing of recorded data is still in progress.<br />

Therefore, we present some preliminary results,<br />

especially from the central part of the Padul-Nigüelas<br />

Fault Zone on the Marchena fan.<br />

During our field work, we have found several obvious<br />

evidences which suggest geologically recent fault<br />

activity in the PNFZ, such as triangular facets, deeply<br />

incised channels, faultward dipping layers and faults<br />

in both colluvial wedges and alluvial fans at<br />

numerous sites within the PNFZ. These features<br />

were also observed and documented by other<br />

authors (e.g. Alfaro et al. (2001), Galindo-Zaldívar<br />

(2003)).<br />

The morphology of the PNFZ is characterized by the<br />

exhumed fault scarp, developed on the Alpujarrian<br />

bedrock. We could confirm the observations from<br />

Alfaro et al. (2001), who has already mentioned that<br />

the scarp is exhumed due to erosion and not due to<br />

fault activity. The exhumed fault scarps of the PNFZ<br />

are specified by lateral as well as downslope<br />

undulations of the main fault plane surface in a<br />

widespread range. The dip angle varies in a range<br />

from 20° to 65° and the dip direction from 190° to<br />

260°. These undulations were documented and<br />

combined with laser distance measurements on<br />

several sites along the exhumed fault scarp of the<br />

entire fault zone.<br />

Due to the apparently inactive main fault, our<br />

investigations in the field brought the colluvial<br />

wedges and alluvial fans in front of the PNFZ into<br />

focus. We gathered multiple profiles of faults in the<br />

colluvium and alluvium by classical field methods and<br />

by laser distance measurements. We have found<br />

active as well as buried faults in the sediments (see<br />

also Hamdouni et al. (2008)). Most of the outcrops<br />

are created naturally by strong river incision into the<br />

sediments and are situated adjacent to the bedrock<br />

fault scarp.<br />

For this state of our study, we have concentrated the<br />

field-work on the widespread alluvial fan deposits<br />

located between the two segments of the Padul-<br />

Nigüelas Fault Zone developed on the Alpujarrian<br />

bedrock north of Marchena, which is axially dissected<br />

Fig. 4: Outcrop of an incised channel in the alluvial and<br />

colluvial fan in the north of Marchena. A: Photograph of the<br />

colluvial sediments and existing faults. B: Sketch of the<br />

same outcrop with a few traceable layers.<br />

by the river Dúrcal. We collected several GPR<br />

profiles and documented several outcrop profiles in<br />

the proximal colluvial wedges developed at fault-fan<br />

contacts. One example is located in the northern part<br />

of the Marchena fan (Fig. 2). Fig. 4 shows a<br />

photograph of the outcrop (A) with interpreted faults<br />

in the sediments and a sketch (B) with the simplified<br />

geometry and certain marker horizons. On the basis<br />

of these horizons, we can preliminary determine the<br />

offset of the larger faults between a few cm and<br />

1.8 m. Noticeable features in the faults are aligned<br />

clasts with their longitudinal axis along slip direction<br />

and carbonate-coated clasts, which indicate water<br />

circulation on the faults. Some of the faults are<br />

traceable up to the surface and build small scarps in<br />

the topology.<br />

Fig. 5: GPR profile on the upper surface of the outcrop (Fig. 4). The profile is not topographically corrected and a velocity of<br />

0.1 m/ns is assumed for the time-depth conversion. Grey box: excerpt of the outcrop. Black lines: interpreted faults.<br />

216


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

In addition to the documentation of the faults in the<br />

colluvial wedge, we achieved a GPR profile on the<br />

upper surface of the outcrop (Fig. 5). The processed<br />

data is not topographically corrected, but first<br />

interpretations allow inferring a few faults in the<br />

profile. An exact correlation between the outcrop and<br />

the radar profile was not done so far. This profile as<br />

well as profiles in other studies (Reicherter et al.<br />

(2001)) demonstrates that GPR surveys are a useful<br />

method to locate faults and faultward dipping layers<br />

within colluvial wedges along presumably active<br />

range-front faults, such as the case of the Padul-<br />

Nigüelas Fault Zone.<br />

The above mentioned features also occur at sites<br />

NW of El Puntal, Padul and Nigüelas which points to<br />

similar fault activity in this stage of our research.<br />

DISCUSSION & OUTLOOK<br />

The Padul-Nigüelas Fault Zone has been examined<br />

extensively for decades and is considered as active<br />

since instrumental and historical seismicity indicate<br />

several earthquakes in the Granada Basin. In our<br />

field study we found the fault scarps exposed<br />

because of erosion instead of fault activity. The<br />

colluvial wedges are not undisturbed, they show both<br />

buried and active faults, therefore we consider<br />

basinwards trending fault activities for this sector, as<br />

well as for the entire PNFZ.<br />

Since our study is not finished, GPR data and<br />

geomorphologic indices will be used to give further<br />

evidence for either active or inactive faults in the<br />

Padul-Nigüelas Fault Zone.<br />

References<br />

Alfaro, P., J. Galindo-Zaldívar, A. Jabaloy, A.C. López-<br />

Garrido & C. Sanz de Galdeano, (2001). Evidence for the<br />

activity and paleoseismicity of the Padul fault (Betic<br />

Cordillera, southern Spain). Acta Geologica Hispanica 36<br />

(3-4), 283-295.<br />

Azañón, J.M., J. Galindo-Zaldívar, V. García-Dueñas and A.<br />

Jabaloy (2002). Alpine tectonics II. Betic Cordillera and<br />

Balearic Islands. In: W. Gibbons and T. Moreno (eds.),<br />

The Geology of Spain 16, 401-416.<br />

Galindo-Zaldívar, J., A.J. Gil, M.J. Borque, F. González-<br />

Lodeiro, A. Jabaloy, C. Marín-Lechado, P. Ruano, C.<br />

Sanz de Galdeano (2003). Active faulting in the internal<br />

zones of the central Betic Cordilleras (SE, Spain). Jorunal<br />

of Geodynamics 36, 239-250.<br />

González Donoso, J.M., V. Garcá-Dueñas, J.A. Gallegos, J.<br />

Avidad Castañeda (1978). Mapa Geologico de España -<br />

1041 Dúrcal. Instituto Geológico y Minero de España.<br />

Hamdouni, R., Irigaray, C., Fernández, T., Chacón, J.,<br />

Keller, E.A. (2008). Assessment of relative active<br />

tectonics, southwest border of the Sierra Nevada<br />

(southern Spain). Geomorphology 96, 150-173.<br />

McClusky, S., R. Reilinger, S. Mahmoud, D. Ben Sari, A.<br />

Tealeb (2003). GPS constraints on Africa (Nubia) and<br />

Arabia plate motions. Geophysical Journal International<br />

155 (1), 126-138.<br />

Reicherter, K.R. (2001). Paleoseismologic advances in the<br />

Granada basin (Betic Cordilleras, southern Spain). Acta<br />

Geologica Hispanica 36 (3-4), 267-281.<br />

Reicherter, K., G. Peters (2005). Neotectonic evolution of<br />

the Central Betic Cordilleras (Southern Spain).<br />

Tectonophysics 405,191-212.<br />

Sanz de Galdeano, C., J.M. González Donoso, J.A.<br />

Gallegos (1975). Mapa Geologico de España - 1026<br />

Padul. Instituto Geológico y Minero de España.<br />

217


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

HOLOCENE SEAFLOOR FAULTING IN THE GULF OF CORINTH: THE POTENTIAL FOR<br />

UNDERWATER PALEOSEISMOLOGY<br />

Sakellariou, Dimitris (1), Lykousis Vasilis (1), Rousakis Grigoris (1)<br />

(1). Institute of Oceanography, Hellenic Centre for Marine Research, 47km Athens-Sounio Ave., 19013 Anavyssos, Greece.<br />

Email: sakell@ath.hcmr.gr, vlikou@ath.hcmr.gr, rousakis@ath.hcmr.gr<br />

Abstract (Holocene seafloor faulting in the Gulf of Corinth: the potential for underwater paleoseismology): The techniques<br />

used for the marine geological – geophysical investigation of the seafloor of the Gulf of Corinth were suitable for the mapping of<br />

the offshore faults and for the detection of recent, Holocene faulting activity. The available seismic data provide clear evidence that<br />

several faults have moved repeatedly in Holocene times and have produced cumulative offsets of up to several meters during the<br />

last 14-13 kyrs. The next step in the investigation of the offshore faulting in the Gulf of Corinth will be to use higher resolution<br />

methods to perform on-fault seismic profiling and to recognize individual earthquake ruptures along the faults on the seafloor of the<br />

Gulf of Corinth.<br />

Key words: seafloor faults, Holocene earthquakes, recent movements<br />

INTRODUCTION<br />

The Gulf of Corinth is an active continental rift<br />

developed within the stretching Aegean region,<br />

perpendicular to the alpine Pindos mountain chain. It<br />

is a 100 km-long rift characterized by high extension<br />

rates in N-S direction (currently up to 20 mm/yr,<br />

Clarke et al. 1998), localized mostly within the<br />

narrow, 15-20 km wide, marine basin. With a very<br />

high seismicity and more than ten earthquakes of<br />

magnitude M>6 in the last 50 years, the Gulf of<br />

Corinth is an ideal site to study active tectonics and<br />

recent fault movements.<br />

The offshore fault pattern shown in Fig. 1 has been<br />

recognized on numerous seismic reflection profiles<br />

Fig. 1: Landsat image with main active offshore faults in the Gulf of Corinth rift. Swath bathymetry after Alexandri et al. (2003)<br />

and Nomikou et al. (this volume)<br />

218


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

(airgun, boomer, 3.5 kHz) acquired during several<br />

cruises of R/V Aegaeo between 1995-2005 and R/V<br />

Alkyon recently (Sakellariou et al., 2001; 2004; 2007;<br />

Lykousis et al., 2007). More marine geophysical<br />

studies have been conducted by other Greek or<br />

international teams and have contributed to the<br />

understanding of the structure of the Gulf of Corinth<br />

(Stefatos et al., 2002; Moreti et al., 2004; Zelt et al.,<br />

2004; McNeill et al., 2005; Bell et al., 2008)<br />

continuous sequence of Holocene mud and sand<br />

turbidites. Correlation of the seismic data with the<br />

sedimentary sequence recovered in the cores<br />

enabled absolute dating of basin-wide reflectors,<br />

which are frequently offset by faults. Careful and<br />

detailed analysis of the shallow seismic stratigraphy<br />

led Lykousis et al. (2007) to estimate Holocene slip<br />

rates of the basin bounding faults.<br />

Although the scope of the above campaigns was<br />

mostly to study the neotectonic structure of the rift<br />

with relatively low to medium resolution techniques,<br />

many seafloor faults were identified to have moved in<br />

Holocene or even in historic times. The aim of this<br />

paper is to shed light these faults and explore the<br />

potential for high resolution underwater paleoseismological<br />

studies in the Gulf of Corinth.<br />

LECHAION GULF<br />

Sakellariou et al. (2004b) showed that the Lechaion<br />

Gulf is currently a half graben developed on the<br />

hangingwall of the south-facing on/offshore Loutraki<br />

fault. Antithetic, north facing faults run E-W on the<br />

shallow southern shelf of the Gulf. Recent Boomer<br />

profiles and older Airgun 10 in 3 and 5 in 3 seismic<br />

profiles show that these faults cut through the 10-<br />

15m thick Holocene deposits.<br />

The Boomer profile of Fig. 2 shows one of the northfacing<br />

athithetic faults of the Lechaion Gulf<br />

crosscutting the submerged Last Glacial Maximum<br />

(LGM) landscape and the Holocene drape. The offset<br />

of 6-7 m must have been accumulated after the<br />

inundation of the LGM Corinth lake about 14 ka ago.<br />

Fig. 2: Boomer profile, 175 Joules, 1-2,5 kHz, Lechaion<br />

Gulf.<br />

CENTRAL GULF OF CORINTH<br />

A dense grid of Airgun 10 in 3 single channel<br />

reflection profiles revealed the shallow structure and<br />

the seismic stratigraphy of the Central Gulf of Corinth<br />

deep basin (Lykousis et al., 2007). Gravity coring<br />

from R/V Aegaeo and long piston coring from R/V<br />

Marion Dufresne (Moreti et al. 2004) validated the<br />

interpretation of the seismic data and recovered a<br />

219<br />

Fig. 3. Airgun 10in 3<br />

single channel reflection profile<br />

through the deep basin of the Central Gulf of Corinth,<br />

south of Itea bay. Left=South, right=North. Top: Raw<br />

profile. Down: Interpreted profile and line drawing. The<br />

yellow line marks the 14 ka old interface between LGM<br />

lacustrine and Holocene marine sedimentation. Slip<br />

rates of the individual faults have been estimated from<br />

the vertical offset (Lykousis et al., 2007).<br />

Vertical slip rates of 0.6-0.7 m/ka have been<br />

calculated for the intra-basin faults, while rates of 2.4-<br />

3.7 m/ka were estimated for the basin bounding<br />

faults. Still, the resolution of the technique is not<br />

sufficient high to allow recognition of individual<br />

earthquakes and rupturing events.<br />

The deep basin of the Gulf of Corinth is an ideal site<br />

for on-fault and off-fault underwater paleoseismological<br />

studies. The geophysical record of the<br />

shallow subseafloor sedimentary sequence with the<br />

continuous succession of mud and sand turbidites,<br />

potential well defined seismic reflectors, may resolve<br />

characteristic structures associated with individual<br />

fault ruptures. High resolution seismics with deeptowed<br />

vehicles in combination with carefully selected<br />

coring sites on the hangingwall and footwall of the<br />

faults will provide recognition and dating of dislocated<br />

layers and thus define individual earthquakes.<br />

WESTERN GULF OF CORINTH<br />

After the 1995 Aegion earthquake the western part of<br />

the Gulf of Corinth has been the site of intensive onand<br />

offshore surveys by Greek and international<br />

teams. Numerous field surveys and marine<br />

geological-geophysical campaigns have been<br />

conducted and yielded very wealth data sets on


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

active tectonics and faulting. Onshore<br />

paleioseismological studies on the Heliki Fault and<br />

the broad region of Aegion provide new knowledge<br />

and understanding of the fault behaviour onshore.<br />

Systematic offshore seismic reflection profiling<br />

enabled detailed mapping of the seafloor faults of the<br />

western Gulf of Corinth. Nevertheless, as in the rest<br />

part of the Gulf, more precise data on the activity of<br />

specific offshore faults has been only occasionally<br />

gained. Two examples are given here below.<br />

Medium resolution Airgun 5 in 3 single channel<br />

seismic reflection profiling has been conducted on<br />

the shelf and upper slope off Aegion – Diakofto<br />

region. A couple of the shot profiles crossed the<br />

eastward offshore prolongation of the well studied<br />

onshore Aegion fault (Fig. 4).<br />

the western Gulf of Corinth and show significant fault<br />

movements in Holocene in the basin.<br />

More precisely, significant fault movements have<br />

been detected on the south facing South Eratini fault,<br />

between Psaromyta Cape and Aegion (Fig. 5). The<br />

Airgun profile of Fig. 5 shows the multiplefaulting of<br />

the basin between Psaromyta Cape (North) and<br />

Aegion (South). The 3.5 kHz profiles through the<br />

South Eratini Fault indicate that the stratified<br />

Holocene drape of the basin has been vertically<br />

dislocated by about 5 m.<br />

Fig. 4: Medium resolution Airgun 5 in 3<br />

single channel<br />

seismic reflection profiling crosscutting the Aigion fault.<br />

Upper image: Map of Aigion area with the trace of the<br />

Aigion fault, the location of the profile and the<br />

sedimentological description and radiometric ages of an<br />

offshore drilling (Schwartz & Tziavos, 1979). Middle: raw<br />

profile. Down: Interpretation of the seismic profile. The<br />

north-facing Aigion fault and an antithetic, south-facing<br />

fault are marked in red. The yellow reflector marks the<br />

interface between clayey silt above and cobbles-pebbles<br />

below, dated to about 1o ka. Note that the Aegion fault<br />

offset vertically the yellow reflector by about 8 m and the<br />

seafloor by about 4 m.<br />

Careful interpretation of the seismic profiles from the<br />

Aegion shelf along with the sedimentological data<br />

from the offshore drillings (Schwartz & Tziavos,<br />

1979) indicate that the Aegion fault has produced a<br />

cumulative vertical offset of at least 4m during the<br />

last 10 kyrs.<br />

Fig. 5: Airgun 10 in 3 seismic profile (top) and 3.5 kHz<br />

profile (bottom) through the South Eratini Fault. Note<br />

that the 3.5 kHz profile shows vertical offset of the<br />

Holocene drape by about 5.5-4.5 m.<br />

CONCLUSIONS<br />

The techniques used for the marine geological –<br />

geophysical investigation of the seafloor of the Gulf<br />

of Corinth were suitable for the mapping of the<br />

offshore faults and for the detection of recent,<br />

Holocene faulting activity. The available seismic data<br />

provide clear evidence that several faults have<br />

moved repeatedly in Holocene times and have<br />

produced cumulative offsets of up to several meters<br />

during the last 14-13 ka. The next step in the<br />

investigation of the offshore faulting in the Gulf of<br />

Corinth will be to use higher resolution methods to<br />

perform on-fault seismic profiling and to recognize<br />

individual earthquake ruptures along the faults on the<br />

seafloor of the Gulf of Corinth.<br />

Further on, Airgun 10 in 3 seismic profiles and 3.5 kHz<br />

profiles have been acquired from the deep basin of<br />

220


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

References<br />

Alexandri M., Nomikou P., Ballas D., Lykousis V. &<br />

Sakellariou D. (2003): Swath bathymetry map of Gulf of<br />

Corinth. Geoph. Res. Abstracts, Vol. 5, 14268, EGS 2003<br />

Bell, R., McNeill, L.C., Bull, J.M., and Henstock, T.J., 2008.<br />

Evolution of the western Gulf of Corinth continental rift,<br />

Greece. Geological Society of America Bulletin, 120, 156-<br />

178.<br />

Clarke, P.J., Davies, R.R., England, P.C., Parson, B.,<br />

Billiris, H., Paradissis, D., Veis, G., Cross, P.A., Denys,<br />

P.H., Ashkenazi, V., Bingley, R., Kahle, H.-G., Muller, M.-<br />

V., and Briole, P., 1998, Crustal strain in central Greece<br />

from repeated GPS measurements in the interval 1989-<br />

1997: Geophysical Journal International, 135, 195-214.<br />

Lykousis, V., Sakellariou, D., Moretti, I., and Kaberi, H.,<br />

2007, Late Quaternary basin evolution of the Gulf of<br />

Corinth: Sequence stratigraphy, sedimentation, fault slip<br />

and subsidence rates: Tectonophysics, 440, 29-51<br />

Leeder, M.R., Portman, C., Andrews, J.E., Collier, R.E.L.,<br />

Finch, E., Gawthorpe, R.L., McNeill, L.C., Perez-Arlucea,<br />

M., and Rowe, P., 2005, Normal faulting and crustal<br />

deformation: Alkyonides Gulf and Perachora peninsula,<br />

eastern Gulf of Corinth rift basin, Greece: Journal<br />

Geological Society of London, 162, 549-561.<br />

McNeill, L., Cotterill, C., Stefatos, A., Henstock, T., Bull, J.,<br />

Collier, R., Papatheoderou, G., Ferentinos G., and Hicks,<br />

S. 2005, Active faulting within the offshore western Gulf<br />

of Corinth, Greece: implications for models of continental<br />

rift deformation: Geology, 33, 241-244.<br />

Moretti, I., Lykousis, V., Sakellariou, D., Reynaud, J.-Y.,<br />

Benziane, B., and Prinzhoffer, A., 2004, Sedimentation<br />

and subsidence rate in the Gulf of Corinth: what we learn<br />

from Marion Dufresne’s long-piston coring: Comptes<br />

Rendus Geoscience, 336, 291-299.<br />

Nomikou, P., Alexandri, M., Lykousis, V., Sakellariou, D. &<br />

Ballas, D. (2011). Swath bathymetry and morphological<br />

slope analysis of the Corinth Gulf. 2 nd INQUA-IGCP-567<br />

International Workshop on Active Tectonics, Earthquake<br />

Geology, Archaeology and Engineering, Corinth, Greece<br />

(2011), this volume<br />

Sakellariou, D., Lykousis, V. & Papanikolaou, D. (2001)<br />

Active faulting in the Gulf of Corinth, Greece. In: 36th<br />

CIESM Congress <strong>Proceedings</strong>, 36 pp. 43.<br />

Sakellariou, D.,Kaberi, H. & Lykousis,V. (2004) Infuence of<br />

active tectonics on the recent sedimentation of the Gulf of<br />

Corinth basin. In: 10th International Congress of Greek<br />

Geological Society, Abstracts 15-17 April, p. 232-233,<br />

Thessaloniki.<br />

Sakellariou D., Fountoulis I. & Lykousis V. (2004b).<br />

Lechaion Gulf: the last descendant of the Proto-Gulf-of-<br />

Corinth basin. 5 th<br />

Int. Symp. On East. Mediterranean<br />

Geology, <strong>Proceedings</strong> 2/881-884, Thessaloniki<br />

Sakellariou, D., Lykousis, V., Alexandri, S., Kaberi, H.,<br />

Rousakis, G., Nomikou, P., Georgiou, P., and Ballas, D.,<br />

2007, Faulting, seismic-stratigraphic architecture and late<br />

Quaternary evolution of the Gulf of Alkyonides basin –<br />

East Gulf of Corinth, Central Greece: Basin Research,<br />

19, 273-295.<br />

Schwarz M.L. & Tziavos C. (1979): Geology in the search<br />

for Ancient Helice. J. Field Archaeology, 6, p. 243-252<br />

Stefatos, A., Papatheoderou, G., Ferentinos, G., Leeder,<br />

M., and Collier, R., 2002, Active offshore faults in the Gulf<br />

of Corinth, Greece: Their seismotectonic significance:<br />

Basin Research, 14, 487-502.<br />

Zelt, B.C., Taylor, B., Weiss, J.R., Goodliffe, A.M.,<br />

Sachpazi, M., and Hirn, A., 2004. Streamer tomography<br />

velocity models for the Gulf of Corinth and Gulf of Itea,<br />

Greece. Geophys. J. Int., 159, 333-346.<br />

221


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

DAMAGE ASSESSMENT IN ARCHAEOSEISMOLOGY:<br />

METHODS AND APPLICATION TO THE ARCHAEOLOGICAL ZONE COLOGNE,<br />

GERMANY<br />

Schreiber,Stephan (1, Klaus-G. Hinzen (1)<br />

(1) Earthquake Geology and Archaeoseismology Group, Cologne University, Vinzenz-Pallotti-Str. 26, 51429 Bergisch<br />

Gladbach. GERMANY. Email: stephan.schreiber@uni-koeln.de<br />

Abstract (Damage Assessment in Archaeoseismology: Methods and Application to the Archaeological Zone Cologne,<br />

Germany): A work flow to investigate archaeoseismological problems including the use of modern surveying equipment and<br />

quantitative methods was developed. We exemplify a comprehensive damage assessment in the historic city center of Cologne,<br />

Germany. During the construction of an underground museum (Archeological Zone Cologne), parts of the Roman and medieval<br />

city are being excavated. The remains exhibit structural damages spread over an area of 150 x 200 m. The structures were<br />

mapped using a phase-based 3D laser scanner. The acquired data were analyzed and the results lead to a quantitative damage<br />

database for the investigation area.<br />

Key words: Damage Analysis, Laser Scanning, Quantitative Methods, Roman<br />

INTRODUCTION<br />

One of the first steps in modern archaeoseismology<br />

is the precise mapping of damages with potential<br />

seismogenic origin. In addition to the archaeoseismologic<br />

working scheme proposed by Galadini et<br />

al. (2006) and specified by Hinzen et al. (2009) and<br />

Hinzen (2011), Schreiber & Hinzen (2011) presented<br />

a site-specific modification for the situation in<br />

Cologne (Fig.1). The investigation of the archaeological<br />

site is structured into four major categories:<br />

(1) investigation of the constructions, (2) surface<br />

topography, (3) subsurface conditions, and (4) geological<br />

setting of the region.<br />

Fig.1: Working scheme for the investigation of an<br />

archaeological site with four major thematic groups (after<br />

Schreiber & Hinzen, 2011.<br />

METHODS<br />

This contribution focuses on the first category and<br />

shows how a comprehensive quantitative dataset is<br />

collected, analyzed, and prepared for further modelling<br />

and estimation of the damage cause.<br />

Data Acquisition & Processing<br />

222<br />

In this study the constructions were mapped using a<br />

phase-based 3D laser scanner (FARO Photon80).<br />

The scanner emits a permanent bundled infrared<br />

beam towards the measurement target via a rotating<br />

mirror. The target reflects the beam and the reflected<br />

signal is detected by the measuring device. The<br />

scanner records the phase shift between transmitted<br />

and received signal and calculates the distance to<br />

the target. Combined with instrumental parameters<br />

including the position of the rotating mirror during the<br />

measurement and the recording position of the<br />

scanner, the Cartesian coordinates are calculated for<br />

each discrete reflection point. The resolution of<br />

0.00076° in horizontal and 0.009° in the vertical<br />

direction, which correlates to 0.13 mm and 1.57 mm<br />

at a distance of 10 m, allows a distance resolution in<br />

the range of 1-2 millimeters (Schreiber et al., 2011b).<br />

Due to the fast data acquisition rate of 120.000 pt/s<br />

measurement times are short and ongoing<br />

excavations are not significantly disturbed. The data<br />

were processed with the software JRC 3D<br />

Reconstructor 2 (Sequieira et al., 1999, Sgrenzaroli &<br />

Wolfart, 2002) which is a capable tool to handle large<br />

3D point clouds.<br />

After the application of different automatic filters to<br />

remove erroneous points (e.g. points at edges or<br />

points acquired in the open sky) and the manual<br />

cleaning of the point clouds to remove vegetation or<br />

modern structures, the scans were merged into<br />

models of substructures of the investigation area.<br />

Due to smaller file sizes (5-20% of the overall data<br />

volume) these “submodels” are easier to handle in<br />

the analysis phase.<br />

Damage Analysis<br />

The substructure models were used to perform a<br />

detailed analysis of the scanned structures and their


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

damages. The structural damages were (1) identified,<br />

(2) located within the global reference system, (3)<br />

classified (rotation, displacement, tilting, cracking),<br />

and (4) quantified with damage-dependent methods.<br />

All results were merged into a damage database.<br />

The damage database is linked to a construction<br />

inventory, where information about the buildings from<br />

(1) previous and recent archaeological excavations,<br />

(2) city annals and historical reports and (3) historical<br />

maps were collected. In addition to these archaeological<br />

and historical sources, information about the<br />

construction type, the building material and the<br />

function of the building was collected.<br />

WORKING AREA<br />

The working area is located in the historic city center<br />

of Cologne, on the top of the slope to a former side<br />

arm of the Rhine River (Fig. 2). The excavations are<br />

part of the Archaeological Zone Cologne (AZC), a<br />

large museum area, which will exhibit remains from<br />

2000 years of history after its completion. During<br />

several building phases the constructions expanded<br />

over the edge of the slope, towards the Rhine River,<br />

so the excavated buildings are founded partially in<br />

artificial fillings. The topography, the complexity of<br />

the construction ground, the position next to a large<br />

river and the seismic potential of the region suggest<br />

several possible causes for the observed damages.<br />

The following paragraphs present three examples of<br />

the data acquisition and damage analysis methods<br />

applied to different structures in a rather complex<br />

excavation environment. Since 2008 over 200 laser<br />

scans were collected in the AZC resulting in a point<br />

cloud of 2.4 billion individual points. This data were<br />

combined to ten models of substructures within the<br />

AZC. The analysis of the data leads to a damage<br />

database with currently 2000 detected damages.<br />

Example: Praetorium<br />

The praetorium, the palace of the Roman governor<br />

was the main administrative building of the province<br />

germania inferior. The first building phase of the<br />

praetorium started with the beginning of the first<br />

century AD followed by three more phases (end of<br />

the 1st century, end of the 2nd century and 4th<br />

century). The structure was completely demolished in<br />

the 9th century (Gechter & Schütte, 2000). The<br />

foundations of the northern half of this former 90 m<br />

long and 25 m high building were excavated in the<br />

1950s and 1970s. We mapped the remains of the<br />

construction in 97 scans from different height levels<br />

and combined them to four models of the main parts<br />

of the building.<br />

The models were used to identify, locate and classify<br />

different damages. The main damage type in the<br />

praetorium is the tilting of walls; 822 tilt<br />

measurements were made on virtual cross-sections<br />

of the point cloud, which were exported to CAD<br />

software. Figure 3 shows the distribution of<br />

directional tilting found in the praetorium.<br />

Fig.2:(Left) Aerial photo of the historic city center of Cologne<br />

with the main structures of the excavations. The white line<br />

indicates the planned museum area. (Right) 3D surface model<br />

of the natural ground. The position of the Roman city wall is<br />

indicated and the black rectangle marks the excavation area<br />

(Schreiber et al., 2011a). Circles give the location of borings<br />

used to construct the model. The former sidearm of the Rhine<br />

River is marked with the dotted white line.<br />

223<br />

Fig.3: Floor plan of the praetorium with directional tilting<br />

vectors. Color indicates the orientation sector (blue: west,<br />

green: east, red: north, yellow: south). The vector length<br />

gives the amount of tilting.<br />

The eastern part of the praetorium is tilted eastward.<br />

The distribution of tilting correlates well to a change<br />

in the subsurface where the constructions expand<br />

over the edge of the former slope towards the Rhine<br />

River and are partly founded in artificial fillings. The<br />

maximum observed tilt is 15.1° towards east. Average<br />

tilt of all eastward tilted walls is 3.1°± 2.4°.<br />

Maximum tilt towards west is 12.9° with an average<br />

of 1.5°± 1.8°, and the maximum tilts towards north<br />

and south are 5.6° and 11.4°, respectively. In


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

addition to tilting, 104 cracks with opening widths of<br />

up to 25 cm were quantified from the point cloud; 39<br />

horizontal displacements in the walls with opening<br />

width up to 9 cm were detected. It is the first time the<br />

documentation focuses on the damage of the<br />

buildings. Before only some of the damages were<br />

documented as a side product in archaeological<br />

drawings; however, not quantified.<br />

Example: Roman Well<br />

A 12.5 m deep Roman well is located west of the<br />

Cologne city hall (Fig.2). The well was mapped in two<br />

campaigns with 20 scans. The scans were acquired<br />

using a special retaining device. The scanner was<br />

installed upside-down and lowered down the well<br />

with an electrical winch; the position was fixed using<br />

air-pressured stamps (Fleischer et al. 2010).<br />

block using the position of the plane normal. Figure 4<br />

shows the tilting vectors for each block on top of a<br />

unrolled cylindrical scan of the inside of the well. The<br />

damage analysis showed that the block layers are<br />

tilted up to 11° within the damage zone (layers<br />

between –8 m and –9 m in Fig. 4). The average tilting<br />

of the discrete blocks is 4.2°± 3.3°. The maximum tilt<br />

is 15.8°.<br />

Example: Roman Sewer<br />

The 180 m long sewer located north of the excavation<br />

area is part of the complex Roman dewatering<br />

system of the city. The eastern and western sections<br />

of the sewer were built with different construction<br />

techniques. The latter is made of large tuff blocks.<br />

This 120 m long section was scanned with 25 scans.<br />

Damages include displaced blocks in the roof (Fig. 5)<br />

and shell-like spallings on the sidewalls.<br />

Fig.4: Flat projection of the wall of the Roman well. Arrows<br />

give the direction of tilting and rotation of the normal of<br />

virtual planes, fitted to the front side of each block. The<br />

length of the arrows gives the amount of tilting. The white<br />

points are the centre points of the fitted planes connected<br />

by the black lines indicating the trend of the layer centers<br />

(after Fleischer et al., 2010). The bars at the right give the<br />

locations of the repaired section, the deformed section and<br />

the main damage area.<br />

The structure shows two sections (Fig. 4): The upper<br />

5.5 m were reconstructed probably during medieval<br />

times, while the lower 7 m remained deformed. For<br />

the analysis of the orientation, virtual planes were<br />

fitted in a least square sense to the front sides of 338<br />

blocks in 33 layers. This technique allowed the<br />

quantification of the tilting and the rotation of each<br />

224<br />

Fig.5: (Upper) Displacement of blocks in the vault of the<br />

western part of the Roman sewer. (Lower) Cracking of the<br />

vault in the eastern part of the sewer. In the bottom, the top<br />

of the filling of the sewer is still in place. The inset shows a<br />

cross-section of the point cloud when the sediment filling of<br />

the sewer was still in place. The green curve shows the<br />

original shape of the vault, the red line follows the rotated<br />

and displaced vault in the damaged area.<br />

The eastern section, which ends in an outlet through<br />

the eastern Roman city wall of Cologne, was<br />

scanned in an emergency campaign before the<br />

excavations were stopped due to critical static conditions<br />

of the sewer. Therefore only four scans were


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

made so far. In this part the vault of the sewer is<br />

made of opus caementicium. Here massive shell-like<br />

spallings were found on the bottom parts of the<br />

sidewalls. The roof shows a large crack with<br />

displacements of up to 15 cm on a length of minimum<br />

10 m (Fig. 5). Currently the damages in the western<br />

part are investigated in detail using virtual crosssections<br />

in 1 m intervals. For the eastern part<br />

additional scans are necessary to perform a<br />

complete damage analysis.<br />

CONCLUSIONS<br />

The application of 3D laserscanning in the AZC<br />

shows that this technique is an appropriate tool<br />

especially when complex structures are analyzed.<br />

The main challenge during the mapping of the<br />

Cologne praetorium was the widespread and complex<br />

damage pattern, which required a large amount<br />

of scans. In the Roman well the vertical shaft,<br />

required a complete different approach. The scanner<br />

was mounted upside down in order to avoid shadowing<br />

effects and the time consuming construction of<br />

different plateaus in the shaft. The scanning of the<br />

sewer required a careful selection of scanning<br />

positions to avoid shadowing effects due to steep<br />

angles in the only 1.6 m wide structure. It could be<br />

shown that the analysis of the damages needs a sitespecific<br />

approach in order to quantify the different<br />

types of damages. In comparison to traditional<br />

mapping techniques laser scanning provides a large<br />

amount of high-resolution data in a short period of<br />

time. In case of fragile conditions of the excavated<br />

objects, repeated scanning can virtually conservate<br />

the original conditions before these are alterated due<br />

to ongoing restoration and/or excavation. The<br />

accuracy of the measurements allows a detailed<br />

analysis even when the original finding is no longer<br />

accessible.<br />

The results of the damage analysis are the base for<br />

further studies on the damage cause. The type and<br />

the amount of the damages can be used to narrow<br />

down possible damage scenarios. However, as<br />

shown in Figure 1, the mapping and the analysis of<br />

the damages are one of the first steps in the<br />

workflow. Further steps, e.g. the surface and<br />

subsurface analysis (Schreiber et. al 2011a) and the<br />

modelling of scenarios (e.g. Hinzen et al., 2010,<br />

2011b) are necessary for an accurate investigation of<br />

an archaeoseismological problem.<br />

Acknowledgements:<br />

We are grateful to the staff of the Archaeological Zone<br />

Cologne. We thank the staff of the Seismological Station<br />

Bensberg, especially Claus Fleischer for his dedication to<br />

the fieldwork<br />

This work is beeing financed by Deutsche Forschungsgemeinschaft<br />

DFG (HI 660 / 2-1).<br />

References<br />

Fleischer, C., K.-G. Hinzen and S. Schreiber, (2010).<br />

Laserscanning eines römischen Brunnens in der<br />

Archäologischen Zone Köln. AVN Allgemeine<br />

Vermessungs Nachrichten 5/2010, 176-181.<br />

Galadini, F., K.-G. Hinzen and S. Stiros, (2006).<br />

Archaeoseismology: Methodological Issues and<br />

Procedure. Journal of Seismology 10, no. 4, 395-414.<br />

Gechter, M. and S. Schütte, (1999). Ursprung und<br />

Voraussetzung des Mittelalterlichen Rathauses und<br />

seiner Umgebung. Stadtspuren 22, 69-195.<br />

Hinzen, K.-G., C. Fleischer, S.K. Reamer, S. Schreiber and<br />

S. Schütte, (2009). Quantitative Methods in<br />

Archaeoseismology. In: Archaeoseismology and<br />

Palaeoseismology in the Alpine-Himalayan Collisional<br />

Zone (Pérez-López, R., Grützner, C., Lario, J.,<br />

Reicherter, K., Silva, P.G. eds). Baelo Claudia, Spain, 50-<br />

51.<br />

Hinzen, K.-G., S. Schreiber and B.Yerli, (2010). The Lycian<br />

Sarcophagus of Arttumpara, Pinara (Turkey) -Testing<br />

Seismogenic and Anthropogenic Damage Scenarios.<br />

Bulletin of the Seismological Society of America 100 (6),<br />

3148-3164.<br />

Hinzen, K.-G., (2011). Earthquake, Archaeoseismology. In:<br />

Encyclopedia of Solid Earth Geophysics (Gupta, H. K.,<br />

ed.) Springer, Heidelberg, DOI 10.1007/978-90-481-<br />

8702-7.<br />

Hinzen, K.-G., H. Kehmeier, S. Schreiber and S.K. Reamer,<br />

(2011). Test Of Earthquakes And Rockfall As Possible<br />

Cause Of Damage To A Roman Mausoleum In Pinara,<br />

Sw Turkey. In: <strong>Proceedings</strong> of 2nd INQUA-IGCP-567<br />

International Workshop on Active Tectonics, Earthquake<br />

Geology, Archaeology and Engineering (Papanikolaou,<br />

I.D., Reicherter, K., Vött, A., Silva, P.G. eds.) Corinth,<br />

Greece, submitted.<br />

Schreiber, S., I. Wiosna, M. Wegner and K.-G. Hinzen,<br />

(2011a). Archeoseismological Study in the Historic City<br />

Center of Cologne, Germany. 71. Jahrestagung der<br />

Deutschen Geophysikalischen Gesellschaft, 21.-<br />

24.2.2011, Cologne, Germany.<br />

Schreiber, S., K.-G. Hinzen and C. Fleischer (2011b).<br />

Excavation Parallel Laser Scanning of a Medieval<br />

Cesspit in the Archaeological Zone Cologne, Germany.<br />

Journal of Computing and Cultural Heritage, submitted.<br />

Schreiber, S. and K.-G. Hinzen, (2011). Archeoseismological<br />

Investigations in the Historic Center of<br />

Cologne, Germany. Seismological Research Letters 82,<br />

No. 2, 334.<br />

Sequieira, V., K. Ng, E. Wolfart, J.G.M. Goncalves and D.<br />

Hogg, (1999). Automated Reconstruction of 3D Models<br />

from Real Environments. ISPRS J. Photogramm. 54, 1-<br />

22.<br />

Sgrenzaroli, M. and E. Wolfart, (2002). Accurate Texture-<br />

Mapped 3D models for Documentation, Surveying and<br />

Presentation Purposes. In: CIPA WG 6 International<br />

Workshop on Scanning For Cultural Heritage Recording,<br />

Greece, Corfu, 1-2 September 2002.<br />

225


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

EARTHQUAKE TRIGGERING, CLUSTERING, AND THE SYNCHRONIZATION OF FAULTS<br />

Christopher H. Scholz<br />

Lamont-Doherty Earth Observatory, Columbia University. Palisades, NY. USA<br />

Abstract (Earthquake Triggering, Clustering, and the Synchronization of Faults): Large earthquakes are sometimes<br />

observed to trigger other large earthquakes on nearby faults. The magnitudes of the calculated Coulomb stress transfers<br />

presumed to cause the triggering are 10 -2 – 10 -3 the earthquake stress drop and the triggering delay times are similarly small with<br />

respect to the natural recurrence time of the earthquakes. This requires that both faults be simultaneously very close to the ends of<br />

their seismic cycles. Paleoseismological data show that for the same regions prior earthquakes have occurred in clusters in space<br />

and time separated by long quiescent periods. Both observations suggest that synchronization is occurring between faults. Theory<br />

and observations indicate that synchronization can occur between nearby faults with positive stress coupling and intrinsic slip<br />

velocities within an entrainment threshold. In the south Iceland seismic zone, the central Nevada seismic belt and the eastern<br />

California shear zone several synchronous clusters that apparently act independently, can be recognized. This behaviour is the 3D<br />

equivalent of the phase locking those results in the seismic cycle being dominated by large characteristic earthquakes, and for<br />

synchronization of fault segments along a single fault. Rupture patterns of repeated individual earthquakes or earthquake clusters<br />

are not identical in either the 2D or 3D cases. The state of this system, which exhibits strong indications of synchrony without exact<br />

repetition, may be called fuzzy synchrony.<br />

226


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

RELIEF PRODUCTION, UPLIFT AND ACTIVE TECTONICS IN THE GIBRALTAR ARC<br />

(SOUTH SPAIN) FROM THE LATE TORTONIAN TO THE PRESENT<br />

Silva, Pablo G. (1, Alex Ribó (1), Moises Martín Betancor (2), Pedro Huerta (1), M. Ángeles Perucha (3), Cari Zazo (4), Jose L.<br />

Goy (1), Cristino J. Dabrio (5), Teresa Bardají (6)<br />

(1) Dpto. Geología, Escuela Politécnica Superior de Ávila, Universidad de Salamanca. Avda. Hornos Caleros, 50. 05003-Ávila.<br />

SPAIN. pgsilva@usal.es, aribo@usal.es, phuerta@usal.es, joselgoy@usal.es<br />

(2) Dpto. Cartografía y Expresión Gráfica en Ingeniería, Universidad de Las Palmas de Gran Canaria, Campus Tafira Baja<br />

35017-Las Palmas de Gran Canaria, SPAIN. mmartin@dcegi.ulpgc.es<br />

(3) Instituto Geológico y Minero de España. Ríos Rosas, 23. 28003-Madrid. SPAIN. ma.perucha @igme.es<br />

(4) Dpto. Geología, Museo Nacional de Ciencias Naturales (CSIC), 28006-Madrid SPAIN<br />

(5) Dpto. Estratigrafía. Facultad CC. Geológicas. Universidad Complutense de Madrid. 28040-Madrid. SPAIN.<br />

(6) Dpto. Geología, Universidad de Alcalá de Henares, 28871- Alcalá de Henares, Madrid SPAIN<br />

Abstract (Relief production, uplift and active tectonics in the Gibraltar Arc (South Spain) from the Late Tortonian to the<br />

present): This work analyzes the topographic evolution of the Gibraltar Arc (Betic Cordillera) from the Tortonian to the present.<br />

This first preliminary approach is based on the morphometric analysis of drainage basins and the identification of stratigraphic<br />

markers for ancient positions of the sea-level. Computed data indicate that mean thickness of eroded materials by fluvial erosion is<br />

593 m, generating an overall isostatic rebound of about 495 m. These data fit with computed mean elevations for individual basins,<br />

suggesting a quasi-isostatic equilibrium in the zone. However uplift is differentially distributed indicating that the main tectonic<br />

structures outcropping in the zone (ancient Betic thrusts and main tectonic contacts) are accommodating differential uplift, and<br />

therefore have been actives throughout Late Neogene to Quaternary times.<br />

Key words: Differential Uplift, Geophysical relief, Erosional unloading, Gibraltar Arc, Betic Cordillera, Spain.<br />

INTRODUCTION<br />

Isostatic uplift in response to erosional unloading of<br />

single faulted range fronts, mountain ranges and<br />

entire cordilleras has been proved to be a key<br />

process for the long-term landscape evolution in<br />

tectonically active areas (e.g. Molnar and England,<br />

1990; Bishop, 2007; Fernández-Ibáñez et al., 2010).<br />

Denudational unloading promotes relevant<br />

lithospheric upward flow, and subsequent surface<br />

uplift, controlling the topographic evolution of active<br />

orogens and backfeeding the rhythms of mass rocktransfer<br />

between ranges, local sedimentary basins<br />

and oceans/seas (e.g. Watts, 2001). Some authors<br />

envisage this process to explain regional surface<br />

uplift in relation to active fluvial erosion in mountain<br />

chains (Gilchrist et al., 1994; Small and Anderson,<br />

1998; Brocklerhurst & Whipple, 2002), oceanic<br />

islands (Menéndez et al., 2008) or even single<br />

faulted-fronts (Fernández-Ibáñez et al., 2010). On the<br />

other hand, aside of the backfeeded contribution of<br />

tectonic and isostatic uplift, cyclic or unique events of<br />

base-level drop can affect to the related rates of<br />

erosion and uplift. This study is a preliminary<br />

analysis focused on the topographic development of<br />

the Gibraltar Arc (Betic Cordillera, South Spain) since<br />

Late Neogene times, and it is part of a large project<br />

within the framework of the EUROCORES-<br />

TOPOMED Program on the topographic evolution of<br />

the Betic-Rif Cordillera in the Western Mediterranean<br />

Basin.<br />

GEOLOGICAL BACKGROUND<br />

The development of the Betic–Rif orogen on top of<br />

the Africa–Iberia plate boundary across the transition<br />

227<br />

zone between the Atlantic Ocean and the<br />

Mediterranean Sea resulted in a complex structural<br />

framework mostly developed during the Alpine<br />

orogeny. The role of fluvial unloading within the<br />

northern segment of the Gibraltar Arc is analyzed by<br />

means of the morphometric analysis of the main<br />

“direct” river basins draining towards the Gulf of<br />

Cádiz (west) and the Alboran Basin (East) from the<br />

fluvial outlets of the Guadalete river (Cádiz)) up to<br />

Motril (Granada). The study comprises the analysis<br />

of the computed bulk volume of rock-mass removed<br />

by erosion since the Tortonian until the present, as<br />

well as the analysis of this same parameter for<br />

discrete time-windows: Tortonian, Messinian,<br />

Pliocene and Quaternary. This last sequenced<br />

analysis will allow establishing time-dependent<br />

differential uplift, but also discrete temporal pulses of<br />

uplift for the whole studied zone, especially before<br />

and after the “Messinian Salinity Crisis”. This event<br />

occurred between 5.96 and 5.33 Myr ago triggering<br />

the disconnection of the Mediterranean Sea from the<br />

Atlantic Ocean in response to the final stages of<br />

build-up of the Gibraltar Arc, as well as a rapid<br />

episode of desiccation in the Mediterranean basin<br />

with a related sea-level drop of about 1000 metres in<br />

the western Mediterranean (e.g. Blanc, 2006). In fact,<br />

recent studies (i.e. Iribarren et al., 2009) indicate that<br />

the continentalization (i.e. emersion or uplift) of this<br />

sector of the Betic Cordillera took place around 5.3<br />

Myr ago, probably linked to this event. In fact<br />

computed data of sedimentary budget (5,490<br />

km 3 /Myr) and sedimentation rates (0.18 mm/yr)<br />

within the Alboran Basin during the Late Neogene<br />

were higher than those recorded for previous and


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

subsequent time-periods. The Messinian sea-level<br />

drop generated an unusual topographic scenario<br />

along the growing emerged tectonic wedge of the<br />

Gibraltar Arc, with an oceanic (Atlantic) foreland<br />

basin to the West and a large desiccated basin to the<br />

East. This situation gave place to an abrupt<br />

asymmetry in the base-level at both sides of the<br />

growing tectonic wedge and therefore, triggered<br />

asymmetric fluvial erosion, probably backfeeding<br />

differential uplift in mountain building. The<br />

subsequent Zanclean Flooding (García Castellanos<br />

et al., 2010) inundated again the entire<br />

Mediterranean basin in an instantaneous time-span<br />

of about 2 years dissipating a gravitational potential<br />

energy of about 1.63x10 22 J and balancing again<br />

base-levels at both sides of the Gibraltar Arc. Sea<br />

level reached a maximum of ca. + 70 m during the<br />

Pliocene previous to the eventual growing of the<br />

Antarctica ice sheet.<br />

ANALYSIS OF RELIEF AND UPLIFT<br />

A digital elevation model of 40x40m pixel resolution<br />

has been used in order to extract main ridgelines and<br />

drainage basins of the Gibraltar Arc from Cádiz<br />

(West) to Motril (East), differentiating 50 individual<br />

drainage basins larger than 18 km 2 over a bulk area<br />

of 14552 km 2 , which excludes the western slope<br />

basins draining towards the Guadalquivir Basin. 11<br />

drainage basins dissect the western Atlantic slope of<br />

the Gibraltar Arc with an area of 5849 km 2 , and 39<br />

ones are located in the Mediterranean slope over an<br />

area of 9063 km 2 . For a preliminary estimation of the<br />

volume of removed materials by fluvial erosion a preincision<br />

surface was computed for each individual<br />

basin from the present elevation of their ridgelines<br />

perimeters and representative erosional surfaces.<br />

The results indicate eroded budgets of 8632.9 km 3<br />

for the bulk of the studied section of the Gibraltar Arc,<br />

of which 836.7 km 3 corresponds to the Atlantic slope<br />

and 7765.2 km 3 for Mediterranean one. Dividing the<br />

obtained removed volumes by the present drained<br />

areas results in the theoretical equivalent thickness<br />

of removed materials or “geophysical relief” (Small<br />

and Andersson, 1998), which for the bulk studied<br />

area is of 593 m.<br />

In a second step hypsometric curves were computed<br />

for each individual drainage basin in order to obtain<br />

the distribution and values of mean and maximum<br />

present elevations for the studied basins, resulting in<br />

a mean value for the whole studied area of 411.4 m<br />

above the sea-level, but with a clear asymmetry<br />

between the Atlantic (108 m) and Mediterranean<br />

(492.3 m) slope of the Gibraltar Arc, with computed<br />

percentual weights over the total analyzed area<br />

always above of 70% and therefore statistically<br />

representative. Comparing the values of mean<br />

elevations and computed geophysical reliefs offer<br />

well fit linear regressions, with correlation coefficients<br />

close to 0.9 for the whole studied zone and their<br />

zonations. This fact implies a good correlation<br />

between the present elevation and the theoretical<br />

thickness of removed materials by fluvial erosion,<br />

suggesting near-isostatic equilibrium conditions for<br />

the zone. From this approach, in a third step the zone<br />

was divided in individual crustal/lithospheric sectors<br />

corresponding to the analyzed basins in order to<br />

calculate the consequent uplift in response to<br />

erosional unloading following the Airy-Heiskannen<br />

isostatic model. Different approaches have been<br />

applied in this study in order to obtain the isostatic<br />

response (e.g. Gilchrist et al., 1994; Fernández-<br />

Ibáñez et al., 2010), but the best fitting ratios<br />

between computed uplift and present mean elevation<br />

Fig. 1: Digital Elevation Model of the Gibraltar Arc (Western Betic Cordillera) displaying the main elements of the present relief,<br />

drainage basins and stratigraphic markers (littoral sediments) analyzed in this study. Mapped Betic thrusts and normal faults<br />

(conventional symbols) are those controlling differential uplift from the Messinian. A: Atlantic Slope; M: Mediterranean Slope,<br />

subdivided in Frontal Flysch (Mf), Eastern (Me) and Western (Mw) analyzed sectors. Gub: Guadalete Neogene basin; Rb:<br />

Ronda Neogene basin; S-Az: Setenil-Antequera zone; Grb: Granada Neogene basin; Gfb: Guadalfeo Neogene basin. Numbers<br />

corresponds to the individual analyzed drainage basins.<br />

228


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

correspond to the application of the Gilchrist et al.<br />

(1994) isostatic constant, which offer linear<br />

regression lines with correlation indexes up to 88%.<br />

This fact suggests that isostatic compensation is<br />

more related to crust/mantle models rather to<br />

lithosphere/ asthenosphere models, but also that the<br />

studied zone should be in near-isostatic equilibrium<br />

at present. The application of crust/mantle isostatic<br />

constant yields a value of +495 m for the mean uplift<br />

of the whole investigated area of, but differentially<br />

distributed from the Atlantic flysch sector (+74.8 m),<br />

the Mediterranean flysch sector (+170.4 m), the<br />

Mediterranean mantle-core complex of the Ronda<br />

Range sector (+417.2 m) and the Mediterranean<br />

sector west to the Málaga basin along the Mijares<br />

Range (+592.8 m). However, taking into account the<br />

present mean maximum elevation of the analyzed<br />

area (1195.8 m), maximum denudational isostatic<br />

rebound represent the 41-49% of the recorded uplift<br />

since the Tortonian, the remaining 59-51 % should<br />

be achieved by the African-Eurasian plates<br />

convergence.<br />

Implementation of digital geological data obtained<br />

from the IGME data base (GEODE) in the digital<br />

elevation model allows the extraction of<br />

representative linear and punctual stratigraphic<br />

markers for different Late Neogene sea-level<br />

positions. The geological data implemented in the<br />

DEM only correspond to littoral, shallow marine and<br />

proximal delta units and the intersection of their<br />

upper contacts with the present relief sketch or depict<br />

approximate paleo-shorelines with a confidence of ±<br />

40m forced by the DEM resolution. These relative<br />

paleoshorelines has been obtained for different timewindows<br />

corresponding to the Late Tortonian,<br />

Messinian, Lower Pliocene, Upper Pliocene and<br />

Lower Pleistocene. These appear identified by<br />

means of different colours in the map of figure 1.<br />

The analysis of these stratigraphic markers is still in<br />

progress, but the preliminary analysis of their<br />

spatial/elevation distribution clearly indicates that the<br />

main stage of relief production in the studied zone<br />

was attained after the relevant Messinian sea-level<br />

drop (5.96 Myr). Most of the littoral Messinian<br />

depostis identified in the GEODE data base in the<br />

Mediterranean slope of the studied zone may mainly<br />

correspond to younger Pliocene sediments.<br />

Therefore “true” Messinian deposits in the studied<br />

area are only present northwards and westwards of<br />

the main ridgeline of the Gibraltar Arc (Fig. 1) along<br />

the ancient Atlantic front of the growing tectonic<br />

wedge. In this zone Messinian, but also Tortonian<br />

deposits, occur at ridgeline locations on the<br />

headwaters of the present drainage basins, which<br />

obviously indicate that main fluvial dissection stages<br />

are post-Messinian.<br />

The occurrence of sediments of these ages in the<br />

present headwaters of the Mediterranean slope is<br />

consequence of Messinnian and post-Messinian<br />

aggressive headward erosion of the largest fluvial<br />

basins trespassing the ancient main ridgeline of the<br />

Gibraltar Arc (Fig. 2) such as those corresponding to<br />

the Guadalfeo, Guadairo, Guadarranque and<br />

229<br />

Guadalhorce rivers (Fig.1). This ridgeline was<br />

already emerged during the Tortonian limiting the<br />

Noegene basins of Ronda and Granada by the south.<br />

The main emerged ridgeline was only interrupted<br />

(inundated) during the pre-Messinian times in the<br />

Málaga Basin along the present Guadalhorce<br />

corridior, the unique location in which there are<br />

outcrops of littoral/shallow marine Tortonian<br />

sediments eastwards the ancient ridgeline of the<br />

Gibraltar Arc within the studied zone (Fig. 1). The<br />

Pliocene and Pleistocene littoral materials in the<br />

Mediterranean simply partially filled the main fluvial<br />

basins generated during the Messinian. These<br />

deposits never appear in ridgeline positions along the<br />

entire Mediterranean slope.<br />

The elevation distribution of post-Messinian sealevels<br />

markers strongly support that the<br />

aforementioned differential uplift is being conducted<br />

along the more relevant NNE-SSW or N-S thrust<br />

contacts developed during the final stages of the<br />

construction of the Gibraltar Arc broadly delineated in<br />

Fig. 2: Conceptual Model for the interaction of fluvial<br />

unloading and gravitational unloading triggering the<br />

subsequent differential uplift along the Gibraltar Arc during<br />

the Messinian sea-level drop (rates from Iribarren et al.,<br />

2009).<br />

Figure 1. These tectonic structures can be<br />

considered as active ones in which a vertical partition<br />

of tectonic and isostatic movements seems to occur<br />

in order to distribute regional differential uplift. In the<br />

central sector of the Gibraltar Strait these tectonic<br />

structures controlled the generation of N-S subsiding<br />

troughs (controlled by normal faulting) during the<br />

Pliocene, clearly control the distribution of differential<br />

uplift of the Last Interglacial marine terraces and are<br />

presently related with instrumental seismicity (Zazo<br />

et al., 1999; Silva et al., 2006).<br />

CONCLUSIONS<br />

The analysis of relief of the Gibraltar Arc presented in<br />

this work is a preliminary approach to decode the<br />

evolution of the onshore topography of this sector of<br />

the Betic Cordillera from the Tortonian to the present.<br />

Initial computed data indicate that the equivalent<br />

thickness of erosional unloading (geophysical relief)<br />

is of 593 m for the whole studied area. Isostatic uplift<br />

response to fluvial unloading is of 495 m, but<br />

differentially distributed, with maximum values in the<br />

eastern studied sector south of Sierra Nevada<br />

(Motril) and minimum values in the western Atlantic<br />

sector (Cádiz). Subsequent analyses will be also


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

focused in the relationships of seismotectonic activity<br />

of the more outstanding North-South tectonic<br />

structures accommodating differential uplift in the<br />

central-axial sector of the Gibraltar Arc.<br />

The sequential analysis of eroded volumes during<br />

discrete periods will allow to establish theoretical<br />

values of uplift for the different studied drainage<br />

basins and sectors. The preliminary results in this<br />

study indicate an accumulated mean uplift of +495 m<br />

since the end of the Tortonian (c.a. 7.2 Myr) which<br />

suggest broad mean uplift rates of 0.068 mm/yr for<br />

this time-period, clearly decreasing towards the<br />

Gibraltar Strait area where uplift rates are of ca. 0.01-<br />

0.02 mm/yr. The obtained values are within the range<br />

of the estimated uplift rates by other authors, and<br />

other methods, for the whole Betic Cordillera (i.e.<br />

Iribarren et al., 2009) which is of 0.05 mm/yr, but<br />

decreasing to 0.02 mm/yr to the Gibraltar Strait (Fig.<br />

2).<br />

A final preliminary remark is that isostasy triggered<br />

by mass-rock transfer between the growing tectonic<br />

wedge of the Gibraltar Arc and two the adjacent<br />

marine basins can explain about the 40-50% of the<br />

recorded uplift since the end of the Tortonian. On the<br />

other hand, the volume of post-Tortonian sediments<br />

accumulated in the Alboran basin indicates mean<br />

rates of sedimentation of about 0.18-0.24 mm/yr (Fig.<br />

2). The conversion of these sedimentation rates to<br />

erosion rates will allow checking the uplift response<br />

by means of different isostatic crust/mantle and<br />

lithosphere/asthenosphere conceptual models.<br />

These models will also take into account the isostatic<br />

rebound related to sea-water unloading in the<br />

Mediterranean slope triggered by the relevant sealevel<br />

drop occurred during the Messinian (Fig.2).<br />

Acknowledgements: This work has been supported by the<br />

EUROCORES-TOPOMED PROGRAM through the Spanish<br />

research project CGL2008-03474-E/BTE (CSIC) produced<br />

by the TopoMed-Spain Onshore Research Group of the<br />

University of Salamanca. Also financial support has been<br />

obtained from the research projects CGL08-03998BTE,<br />

CGL08-04000BTE<br />

References<br />

Blanc, P. L. (2006). Improved modelling of the Messinian<br />

Salinity Crisis and conceptual implications. Palaeogeogr.<br />

Palaeoclimatol. Palaeoecol. 238, 349–372.<br />

Bishop, P. (2007). Long-term landscape evolution: Linking<br />

tectonics and surface processes. Earth Surface<br />

Processes and Landforms, 32, 329–365.<br />

Brocklehurst, S.H., Whipple, K.W. (2002). Glacial erosion<br />

and relief production in the Eastern Sierra Nevada,<br />

California. Geomorphology 42 , 1-24.<br />

Fernández-Ibáñez, F., Pérez-Peña, J.V., Azor, A., Soto,J.<br />

Azañón, J.M. (2010). Normal faulting driven by<br />

denudational isostatic rebound. Geology, 38, 643-646.<br />

Garcia-Castellanos, D., Estrada, F., Jiménez-Munt, I.,<br />

Gorini,C., Fernández, M., Vergés, J. & De Vicente, R.<br />

(2009). Catastrophic flood of the Mediterranean after the<br />

Messinian salinity crisis. Nature, 462, 778-782.<br />

Gilchrist, A.R., Kooi, H. & Beaumont, C. (1994). Post-<br />

Gondwana geomorphic evolution of southwestern Africa:<br />

Implications for the controls on landscape development<br />

from observations and numerical experiments: Journal of<br />

Geophysical Research, 99, 12,211–12,228.<br />

Iribarren, L., Vergés, J. & Fernandez, M. (2001). Sediment<br />

supply from the Betic–Rif orogen to basins through<br />

Neogene. Tectonophysics, 475, 68–84<br />

Menéndez, I., Silva, P.G., Martín Betancor, M., Pérez<br />

Torrado, F.J., Guillou, H. & Scaillet; S. (2008). Fluvial<br />

dissection, isostatic uplift, and geomorphological<br />

evolution of volcanic islands (Gran Canaria, Canary<br />

Islands, Spain). Geomorphology, 102, 189 – 203.<br />

Molnar, P. & England, P. (1990). Late Cenozoic uplift of<br />

Mountain ranges and global climate change: Chicken or<br />

egg?: Nature, 346,,29– 34.<br />

Silva, P.G., Goy, J.L., Zazo, C., Bardají, T., Lario, J.,<br />

Somoza, L., Luque, L. & González Fernández, F.M.<br />

(2006). Neotectonic fault mapping at the Gibraltar Strait<br />

Tunnel area, Bolonia Bay (South Spain). Eng. Geology,<br />

84, 31–47<br />

Small, E. & Anderson, R. (1998). Pleistocene relief<br />

production in Laramide mountain ranges, western United<br />

States: Geology, 26, 123–126.<br />

Watts, A.B. (2001). Isostasy and flexure of the lithosphere.<br />

Cambridge University Press, Cambridge, UK, pp. 458.<br />

Zazo, C., Silva, P.G., Goy, J.L., Hillaire-Marcel, C. Lario, J.<br />

Bardají, T. & González, A. (1999). Coastal uplift in<br />

continental collision plate boundaries: Data from the Last<br />

interglacial marine terraces of the Gibraltar Strait area<br />

(South Spain). Tectonophysics, 301, 95-119.<br />

230


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

REASSESSING ANCIENT EARTHQUAKES ON MINOAN CRETE<br />

GETTING RID OF CATASTROPHISM<br />

Sintubin, Manuel (1, Simon Jusseret (2,1), Jan Driessen (2)<br />

(1) Geodynamics & Geofluids Research Group, Department of Earth & Environmental Sciences, Katholieke Universiteit Leuven,<br />

Celestijnenlaan 200E, B-3001 Leuven. BELGIUM. Email: manuel.sintubin@ees.kuleuven.be<br />

(2) Aegean Interdisciplinary Studies Research Group (AegIS-CEMA-INCAL), Université Catholique de Louvain, Place B. Pascal<br />

1, B-1348 Louvain-la-Neuve. BELGIUM. Email: jan.driessen@uclouvain.be; simon.jusseret@uclouvain.be<br />

Abstract(Assessing ancient earthquake on Minoan Crete. Getting rid of catastrophism): Early in the 20 th century Arthur<br />

Evans invoked catastrophic earthquakes to explain the destruction encountered in the Palace of Knossos. Ever since, these<br />

earthquake catastrophes have been used as reference events to which structural damage or destruction layers on Minoan sites<br />

are indiscriminately attributed in archaeological and archaeoseismological publications. However, structural damage to<br />

archaeological remains cannot be unequivocally attributed to earthquakes. A detailed analysis of Late Minoan contexts moreover<br />

reveals that multiple moderate earthquakes occurred during this 200-year period (c. 1600 – 1400 BC). All evidence suggests that<br />

earthquakes did not play a crucial role in Minoan history and did certainly not cause the decline of Minoan civilization. This<br />

reassessment of the Minoan case clearly demonstrates that earthquakes in itself are incapable of causing the collapse of a<br />

civilization.<br />

Key words: Minoan Crete, Late Bronze Age, catastrophism, earthquake archaeology<br />

Early in the 20 th century Arthur Evans invoked<br />

catastrophic earthquakes to explain the destructions<br />

encountered during the excavations of the Minoan<br />

Palace at Knossos, Crete (Fig. 1). Such seismic<br />

catastrophes were considered to have ended the<br />

Protopalatial period (c. 1700 BC), to have been<br />

responsible for the “great destruction” during the<br />

Neopalatial period (c. 1600 BC), and to have<br />

ultimately caused the collapse of Minoan civilization<br />

around 1450 BC (Evans, 1928).<br />

Ever since, these earthquake catastrophes have<br />

been taken for granted and used as reference events<br />

to which structural damage to buildings and other<br />

cultural remains or earthquake-related destruction<br />

layers are indiscriminately attributed in<br />

archaeological (e.g., Sakellarakis & Sapouna-<br />

Sakellaraki, 1991) and archaeoseismological<br />

publications (e.g., Monaco & Tortorici, 2004;<br />

Vallianou, 1996). But is there any reliable evidence to<br />

support the existence of such catastrophic<br />

earthquakes and can we parameterize them?<br />

Fig. 1: The “house of the fallen blocks” at the palatial<br />

site of Knossos, a particular damage interpreted by<br />

Evans (1928) to have been caused by a catastrophic<br />

earthquake.<br />

231<br />

First, it is extremely difficult to attribute unequivocally<br />

structural damage to Minoan archaeological remains<br />

to earthquakes. In most cases it cannot be excluded<br />

that other physical and/or anthropogenic agents have<br />

generated the damage observed (cf., Driessen,<br />

1995). A macroseismological parameterization of<br />

these ancient earthquakes based on the detailed<br />

archaeological record remains a very challenging<br />

prospect.<br />

Secondly, a detailed analysis of Late Minoan (c. 1600<br />

– 1400 BC) archaeological contexts (cf., Driessen &<br />

Macdonald, 1997) shows that earthquake-related<br />

damage, repairs, adjustments (e.g., Driessen, 1987)<br />

and/or abandonment are all rather isolated and local<br />

phenomena within and not necessarily<br />

contemporaneous between the different sites. This<br />

evidence reveals that most probably multiple<br />

moderate earthquakes occurred during this 200-year<br />

time period, rather comparable to today’s seismicity<br />

of the island.<br />

There is seemingly only consistent archaeological<br />

evidence for widespread, earthquake-related damage<br />

on Crete, as well as on Thera (Santorini), Kos, and


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Rhodes at c. 1600 BC (transition of ceramic stage<br />

MM IIIB to LM IA), known as the “great destruction”<br />

(Evans, 1928). This potentially “catastrophic” event<br />

is, however, followed by a sudden increase of the<br />

number of secondary sites, in particular in eastern<br />

Crete (cf., Driessen & Macdonald, 1997), new palatial<br />

architecture with new architectural (anti-seismic?)<br />

elements, such as ‘pier-and-door partitioning’ (cf.,<br />

Driessen, 1987), and the greatest construction<br />

program of any prehistoric era in the Aegean. The<br />

heyday of Minoan civilisation followed this major<br />

seismic event.<br />

Between c. 1520 and c. 1480 BC (ceramic stage LM<br />

IA) this prolific building program came to an abrupt<br />

end. Monumental buildings were left unfinished. This<br />

heralded the demise of Minoan society. In a period of<br />

one to two generations – from c. 1480 to c. 1425 BC<br />

(ceramic stage LM IB) – a wave of fire destruction<br />

raged over the island. Settlements were abandoned,<br />

population density declined. Both a “crisis<br />

architecture” (cf., Driessen, 1995) and a “crisis cult”<br />

took hold of Minoan society. Evidence for<br />

earthquake-related damage in that period indicates<br />

multiple moderate earthquakes affecting local<br />

communities rather than island-wide destructive<br />

events. Most of the destruction was indeed caused<br />

by man. All evidence indicates that the socio-political<br />

and economic landscape of Minoan society<br />

completely disintegrated and collapsed in that period,<br />

leaving behind a “failed state”. The power vacuum<br />

was later – during ceramic stage LM II (c. 1425 to<br />

c.1400 BC) – filled by the Mycenaens.<br />

Although Late Minoan society can clearly be<br />

characterized as a society in crisis (cf. Driessen &<br />

Macdonald, 1997), no hard evidence exists to link<br />

this societal decline with (catastrophic) earthquakes.<br />

Even the existence of such seismic catastrophes<br />

during Minoan history – except for the c. 1600 BC<br />

event – should be questioned. Minoans lived with<br />

earthquakes, very much as modern Cretans do.<br />

Earthquakes did not play a crucial role in Minoan<br />

history and did definitively not cause the decline of<br />

Minoan civilization. At most, they added some extra<br />

stress to a society already in crisis.<br />

This reassessment of the Minoan case illustrates that<br />

earthquakes, irrespective of their magnitude and<br />

recurrence, provoke different societal responses,<br />

largely depending on the political, social, economic<br />

and military context. Earthquakes in itself are<br />

incapable of causing the collapse of a community, let<br />

alone a civilization. It’s therefore time to get rid of the<br />

catastrophism that has burdened earthquake<br />

archaeology for too long (e.g., Nur, 2008; Nur &<br />

Cline, 2000).<br />

Acknowledgements: This contribution frames in the<br />

UNESCO-IUGS funded International Geoscience<br />

Programme IGCP 567 Earthquake Archaeology. S.<br />

Jusseret is currently a Intercommunity Scientific<br />

Postdoctoral Collaborator of the Francqui Foundation<br />

(Belgium) at the K.U.Leuven.<br />

References<br />

Driessen, J. (1987). Earthquake-Resistant Construction and<br />

the Wrath of the "Earth-Shaker". Journal of the Society of<br />

Architectural Historians 46, 171-178.<br />

Driessen, J. (1995). "Crisis Architecture". Some<br />

observations on Architectural Adaptations as Immediate<br />

Responses to Changing Socio-Cultural Conditions. Topoi<br />

5, 63-88.<br />

Driessen, J., Macdonald, C. F. (1997). The Troubled Island.<br />

Minoan Crete before and after the Santorini Eruption,<br />

Liège & Austin.<br />

Evans, A. (1928). The Palace of Minos, part II., London.<br />

Monaco, C., Tortorici, L. (2004). Faulting and effects of<br />

earthquakes on Minoan archaeological sites in Crete<br />

(Greece). Tectonophysics 382, 103-116.<br />

Nur, A. (2008). Apocalypse. Earthquakes, Archaeology, and<br />

the Wrath of God. Princeton University Press, Princeton.<br />

Nur, A, Cline, E. H. (2000). Poseidon's Horses: Plate<br />

Tectonics and Earthquake Storms in the Late Bronze Age<br />

Aegean and Eastern Mediterranean. Journal of<br />

Archaeological Science 27, 43-63.<br />

Sakellarakis, J. A., Sapouna-Sakellaraki, E. (1991).<br />

Archanes. Ekdotike Athenon S.A., Athens.<br />

Vallianou, D. (1996). New Evidence of Earthquake<br />

Destructions in Late Minoan Crete. In:<br />

Archaeoseismology (edited by Stiros, S. C. & Jones, R.<br />

E.). Fitch Laboratory Occasional Paper 7. Institute of<br />

Geology & Mineral Exploration & The British Scholl at<br />

Athens, Athens, 153-167.<br />

232


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PALEOTSUNAMIS EVIDENCE FROM A COMBINED INLAND AND OFFSHORE STUDY IN<br />

THE AUGUSTA BAY AREA (EASTERN SICILY, ITALY)<br />

Smedile, Alessandra (1, Paolo Marco De Martini (1), Daniela Pantosti (1)<br />

(1) Istituto Nazionale di Geofisica e Vulcanologia. Via di Vigna Murata, 605. 00143- Roma. ITALY.<br />

Email: alessandra.smedile@ingv.it, paolomarco.demartini@ingv.it, daniela.pantosti@ingv.it<br />

Abstract (Paleotsunamis evidence from a combined inland and offshore study in the Augusta bay area (eastern Sicily,<br />

Italy): We present the geological evidence for a 4000 year-long record of repeated tsunami events along the coast of the Augusta<br />

Bay derived from the combination of inland and offshore cores data. The research was carried out through a multi-theme approach<br />

which benefited from an long historical record that we used to guide detailed geomorphologic, geologic, and geophysical surveys,<br />

that were performed both inland and offshore. These surveys served to locate and characterize the best sites were to perform the<br />

coring collection that where studied through and several laboratory analyses. The Augusta Bay represents a unique case study<br />

because besides showing a long record of paleotsunami events, it contains the first inland and offshore evidence for the Crete AD<br />

365, and Santorini Late Minoan tsunamis in the central-western Mediterranean area.<br />

Key words: Eastern Sicily, tsunami deposits, AD 365 Crete and Santorini events.<br />

Eastern Sicily (Italy) was repeatedly affected by<br />

tsunami waves related to large local historical<br />

earthquakes (e.g. 1908, 1693, 1169) (CPTI Working<br />

group, 2004) as well as to far-field sources (e.g. AD<br />

365 Crete earthquake) (Jerome, 380). Along the<br />

eastern Sicily coasts, we selected the Augusta Bay, a<br />

natural gulf about 15 km wide and with a 25 km-long<br />

shoreline (Fig.1), as the key area of this study. In<br />

fact, it is one of the locations where both information<br />

available from historical written reports on tsunami<br />

effects (hit localities, inundated areas and run-up<br />

distribution) (Gerardi et al., 2008) and local<br />

geomorphology suggest it is very favorable for the<br />

research of the geological signature of past tsunamis.<br />

Well-targeted sediment samples have been collected<br />

both inland and offshore through coring at different<br />

depths. Small ponds, marshes and lagoons<br />

characterize the coastal area, while a relatively wide<br />

continental shelf with a thick late-Holocene record<br />

has been investigated offshore through the<br />

acquisition of a tight grid of CHIRP-sonar profiles.<br />

The integrated interpretation of the geophysical and<br />

geological data has been carried out in order to<br />

recognize, date and correlate key-layers in the<br />

sediment column that may be directly or indirectly<br />

related to tsunami events.<br />

A total number of 22 cores were collected inland at<br />

two different sites with a maximum distance of 530 m<br />

from the present coastline (De Martini et al., 2010).<br />

The dominant fine to very fine stratigraphy is<br />

intercalated by at least 6 high-energy depositional<br />

layers, repeatedly found in several cores. These<br />

relatively thin (about 10 cm) single massive and<br />

structureless beds with abrupt erosional lower<br />

contact are made of coarse to fine sand and present<br />

a bioclastic component (sometimes predominant)<br />

made of microfauna (benthic and planktonic<br />

233<br />

foraminifera, from both shallow and open marine<br />

environment) and shell fragments both suggestive of<br />

a marine origin. Chronological constraints on the age<br />

of these deposits are based on 8 AMS radiocarbon<br />

datings and on the attribution of a tephra layer to the<br />

122 BC Etna eruption. For the marine shell samples<br />

(details in De Martini et al., 2010 and Smedile et al.,<br />

2011), measured C14 ages were dendrochronologically<br />

corrected using a marine calibration<br />

curve that incorporate a time-dependent global ocean<br />

reservoir correction of about 400 yrs (Reimer et al.,<br />

2009). Moreover, the marine palaeo-reservoir effect<br />

was subjected to the local effect (R offset) that, in<br />

the Mediterranean Sea, appears to be constant for<br />

the past 6 or 7 ka (Reimer and McCormac, 2002).<br />

The appropriate R offset can be selected from the<br />

Chrono Marine reservoir Database (Reimer and<br />

Reimer, 2001).<br />

Figure 1: The Augusta Bay area in eastern Sicily, Italy.<br />

The Augusta Hospital and Priolo Reserve sites are<br />

marked with white empty rectangles, while a white box<br />

locates the offshore coring site (MS-06). Two<br />

panoramic pictures of the in-land investigated sites are<br />

also shown.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

On the basis of the combination of all the data<br />

collected, the inland sequence spans the last 4100<br />

yrs. In terms of tsunami timing, we could list them as<br />

follow (PR= Priolo Reserve site; AU= Augusta<br />

Hospital site): younger than 1420–1690 AD (PR-01),<br />

650–770 AD (AU-00), 160–320 AD (PR-02), 600–400<br />

BC (AU-01), 800–600 BC (PR-03), 975–800 BC (AU-<br />

02) and 2100–1635 BC (PR-04).<br />

Three of the tsunami deposits found at the Priolo<br />

Reserve site may be associated with historical<br />

tsunamis: PR-01 with the 1693 local event, PR-02<br />

with the 365 AD Crete event and PR-04 with the ca.<br />

3600 BP Santorini event.<br />

The offshore record was derived from a 6.7 m-long<br />

piston-core sampled 2 km offshore the Augusta<br />

harbor at 72 m depth (Fig.1). The core study includes<br />

X-ray imaging, isotopic dating, tephrachronology,<br />

grain-size and foraminiferal analyses (Smedile et al.,<br />

2011). The homogeneous sequence of dark grey<br />

mud is interrupted at -2.9 m b.s.f. (below sea floor) by<br />

the same Etna tephra deposit found inland. Through<br />

the analysis of tephrostratigraphy, radiocarbon dating<br />

and radioactive tracers, the entire core sequence has<br />

been dated back to the last 4500 yrs. Furthermore,<br />

we estimated an age range for the 12 high energy<br />

intervals as follows: E1 (AD 1820-1920), E2 (AD<br />

1430-1810), E3 (AD 930-1170), E4 (AD 590-800), E5<br />

(AD 430-660), Ex (AD 90-370), E6 (BC 350-130), E7<br />

(BC 580-320), E8 (BC 660-400), E9 (BC 800-560),<br />

E10 (BC 1130-810), E11 (BC 1720-1200). Moreover,<br />

the quantitative micropaleontological (on the benthic<br />

foraminifera assemblage) and sedimentological<br />

analyses highlighted 12 anomalous layers marked by<br />

high concentration of displaced epiphytic foraminifera<br />

(species growing in vegetated substrates like the<br />

Posidonia oceanica) and subtle grain size changes.<br />

These anomalous layers are likely to have been<br />

caused by high-energy events, with tsunamis (backwash<br />

wave) as best candidates. This hypothesis is<br />

also supported by the fact that the ages of 5 of these<br />

peculiar layers coincide with that of historical<br />

tsunamis [1908 (E1), 1693 (E2), 1169 (E3), AD 365<br />

Crete (Ex) and ca. 3600 BP Santorini (E11)].<br />

Moreover, to better detail and replicate the MS06<br />

results at least for the recentmost sequence, new<br />

cores were collected in the northern part the Augusta<br />

Bay. These cores were sampled in order to define a<br />

W-E oriented transect along the shelf from 58 to 110<br />

m water depth.<br />

affected by intermittent erosional and sedimentation<br />

events as well as by antrophic activities .<br />

Acknowledgements: This work was funded by the Italian<br />

Dipartimento della Protezione Civile in the frame of the<br />

2004–2006 and 2007–2009 agreements with Istituto<br />

Nazionale di Geofisica e Geofisica e Vulcanologia — INGV<br />

with contribution of the EU Transfer project. We have to<br />

warmly thank M.S. Barbano, F. Gerardi, C. Pirrotta from<br />

Catania University, L. Gasperini, L. Bellucci, A. Polonia from<br />

ISMAR-CNR Bologna and P. Del Carlo and E. Boschi from<br />

INGV for their enthusiastic and productive collaboration.<br />

References<br />

CPTI Working group, 2004. Catalogo Parametrico dei<br />

Terremoti Italiani, version 2004 (CPTI04). INGV,<br />

Bologna. http://emidius.mi.ingv.it/CPTI/<br />

De Martini, P.M., Barbano, M.S., Smedile, A., Gerardi, F.,<br />

Pantosti, D., Del Carlo, P., Pirrotta, C, (2010). A 4000 yrs<br />

long record of tsunami deposits along the coast of the<br />

Augusta Bay (eastern Sicily, Italy): paleoseismological<br />

implications. Marine Geology, 276, 42-57, doi:<br />

10.1016/j.margeo.2010.07.005<br />

Gerardi, F., Barbano, M.S., De Martini, P.M., Pantosti, D.,<br />

2008. Discrimination of tsunami sources (earthquake vs.<br />

landslide) on the basis of historical data in eastern Sicily<br />

and southern Calabria. Bulletin of the Seismological<br />

Society of America, 98 (6), 2795–2805.<br />

Smedile A., P.M. De Martini, D. Pantosti, L. Bellucci, P. Del<br />

Carlo, L. Gasperini, C. Pirrotta, A. Polonia, E. Boschi<br />

(2011). Possible tsunamis signatures from an integrated<br />

study in the Augusta Bay offshore (Eastern Sicily–Italy),<br />

Marine Geology, 281, 1-13, doi:<br />

10.1016/j.margeo.2011.01.002.<br />

Reimer, P.J., McCormac, F.G., 2002. Marine radiocarbon<br />

reservoir corrections for the Mediterranean and Aegean<br />

Seas. Radiocarbon 44, 159–166.<br />

Reimer, P.J., Reimer, R.W., 2001. A marine reservoir<br />

correction database and on-line interface. Radiocarbon<br />

43, 461–463 suppl. mat.URL: http://www.calib.org.<br />

Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck,<br />

J.W., Blackwell, P.G., Bronk Ramsey, C., Buck, C.E.,<br />

Burr, G.S., Edwards, R.L., Friedrich,M., Grootes,<br />

P.M.,Guilderson, T.P.,Hajdas, I., Heaton, T.J., Hogg,<br />

A.G., Hughen, K.A., Kaiser, K.F., Kromer, B., McCormac,<br />

F.G., Manning, S.W., Reimer, R.W., Richards, D.A.,<br />

Southon, J.R., Talamo, S., Turney, C.S.M., van der<br />

Plicht, J., Weyhenmeyer, C.E., 2009. IntCal09 and<br />

Marine09 radiocarbon age calibration curves, 0–50, 000<br />

years cal BP. Radiocarbon 51 (4), 1111–1150.<br />

The Augusta Bay represents a unique case study<br />

because it allows a comparison between geological<br />

(both inland and offshore) and historical records. For<br />

the 365 AD Crete tsunami and the Late Minoan<br />

Santorini event, our findings represent the first<br />

inland-offshore evidence in the central-western<br />

Mediterranean area. On the basis of these results we<br />

can propose, for the past 4 ka in the Augusta Bay, an<br />

inland and offshore geologic average tsunami<br />

recurrence interval of about 550-650 and 320 years,<br />

respectively. This difference is conceivably due to the<br />

better preservation of the stratigraphic record in the<br />

offshore with respect to coastal areas, commonly<br />

234


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

BULGARIAN NATIONAL DIGITAL SEISMOLOGICAL NETWORK<br />

Solakov D. (1), L.Dimitrova (1), S.Nokolova (1), St. Stoyanov (1), S. Simeonova (1), L. Zimakov (2), L.Khaikin (3)<br />

(1) Dpt. Seismology, National Institute of Geophysics, Geodesy and Geography. Acad. G. Bonchev str.,bl. 3, 1113 – Sofia,<br />

BULGARIA, Emails: dimos@geophys.bas.bg, lidim@geophys.bas.bg, sbnik@geophys.bas.bg, stoyan@geophys.bas.bg,<br />

stelas@geophys.bas<br />

(2) Refraction Technology, Inc., USA. l.zimakov@reftek.com<br />

(3) Russia, leonidh@ramblert.ru<br />

Abstract (Bulgarian national digital seismological network): The Bulgarian National Digital Seismological Network (BNDSN)<br />

consists of: National Data Centre (NDC); 13 stations equipped with RefTek Seismic Recorders DAS130 and 1 station equipped<br />

with Quanterra 680; broadband seismometers and accelerometers. The real-time data acquisition and processing are performed<br />

by a system for signal detection; evaluation of the signal parameters; phase identification and association; source estimation.<br />

Seismic interactive analysis system and Early warning system are running in the NDC also. Modern digital equipment installed at<br />

Bulgarian seismic stations, careful selection of the software packages running in the data centre proved to be suitable choice for<br />

the purposes of BNDSN – to ensure reliable automatic localization of the seismic events and rapid notification of the governmental<br />

authorities in case of felt earthquakes on the territory of Bulgaria, to provide a modern basis for seismological studies in Bulgaria.<br />

Key words: Bulgarian National Digital Seismological Network, digital equipment, data processing<br />

INTRODUCTION<br />

In 2005, the Geophysical Institute (with governmental<br />

support) performed overall modernization of the<br />

Bulgarian National Seismological Network (BNDSN)<br />

(Fig.1). Modern digital equipment and broadband<br />

seismometers were installed in all stations.<br />

Data acquisition<br />

Real-time data transfer from seismic stations to NDC<br />

is realized via Virtual Private Network (VPN) of the<br />

Bulgarian Telecommunication Company (BTC)<br />

(Fig.2) with the following characteristics: 64kbps<br />

baud rate from each individual site to the digital<br />

network of BTC; high security – closed Internet<br />

access; broadband 2Mbps optical line established<br />

between the NDC and the Centre of Communication<br />

Company in Sofia. This solution proved to be very<br />

stable and has ensured a reliable communication<br />

system through the six years of exploitation.<br />

Fig. 1: Bulgarian National Seismological Network<br />

At present the upgraded network consists of a<br />

National Data Centre (NDC), 13 stations equipped<br />

with RefTek High Resolution Broadband Seismic<br />

Recorders – model DAS 130-01/3, 1 station<br />

equipped with Quanterra 680 (installed in 1996 in<br />

station VTS by project PLATO1/MEDNET as a<br />

station from VBB Mediterranean Network) and<br />

sensors: very broadband - STS2, STS1, KS2000,<br />

RefTek151/120; broadband - CMG 3ESPC,<br />

CMG40T; short-period - S13 and accelerometers<br />

RefTek 131/03.<br />

Fig.2. Real-time data transfer from seismic stations to NDC<br />

via Virtual Private Network (VPN) and three layer local<br />

network<br />

235


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The network availability exceeded 99.99%. The<br />

communication interruptions did not cause any data<br />

loss at the NDC. The data is backed up in the field<br />

station recorder 4Mb RAM memory and is<br />

retransmitted to the NDC immediately after the<br />

communication link is re-established. The recorders<br />

are equipped with 2 compact flash storages able to<br />

save more than 1 month long data. The data from the<br />

flash disks can be downloaded remotely using FTP.<br />

The data acquisition and processing hardware<br />

redundancy at the National Data Center is achieved<br />

by two clustered SUN Fire V240 servers with<br />

swappable software module and two Blade 1500<br />

Workstations. In the case of server failure the<br />

swappable module re-directs the real-time data flow.<br />

To secure the acquisition, processing and data<br />

storage processes a three layer local network is<br />

designed at the NDC. First layer incorporates<br />

acquisition and real-time processing equipment, the<br />

second layer consists of interactive processing and<br />

archiving modules and external users belong to the<br />

third layer (Fig.2). Real-time data acquisition is<br />

performed using REFTEK’s full duplex errorcorrection<br />

protocol RTPD. Data from the Quanterra<br />

recorder and foreign stations are fed into RTPD in<br />

real-time via SeisComP/SeedLink protocol. Using<br />

SeisComP/SeedLink software the NDC transfers<br />

real-time data to INGV, Roma and NEIC, USA. The<br />

BNDSN is a part of European Virtual Broadband<br />

Seismograph Network and the NDC transmits realtime<br />

data to the ORFEUS Data Centers. Regional<br />

real-time data exchange with Romania, Macedonia,<br />

Serbia and Greece is established at the National<br />

Data Center also. The data flow from neighbor<br />

countries is incorporated in the real-time data stream<br />

in the NDC and is used together with BNDSN data<br />

for localization of the local, regional and distance<br />

seismic events.<br />

Data processing<br />

Data processing is performed by the Seismic<br />

Network Data Processor (SNDP) (Haikin and<br />

Kushnir, 2005; Haikin et al., 2009) software package<br />

running on the both Servers. SNDP includes<br />

subsystems:<br />

Real-time subsystem (RTS_SNDP) – for signal<br />

detection; evaluation of the signal parameters;<br />

phase identification and association; source<br />

estimation;<br />

Seismic analysis subsystem (SAS_SNDP) – for<br />

interactive data processing;<br />

Early warning subsystem (EWS_SNDP) - based<br />

on the first arrived P-phases.<br />

The signal detection process is performed by<br />

traditional STA/LTA detection algorithm. Input data<br />

streams are band-pass filtered with band limits<br />

described in a parametric file. The filter parameters of<br />

the detectors are defined on the base of previously<br />

evaluated ambient noise at the seismic stations.<br />

The annual noise distribution at the BNDSN stations<br />

is presented on Fig. 3 (the 14 figures around the<br />

map).<br />

Fig. 3. Evaluation of ambient seismic noise at BNDSN stations<br />

236


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

For the analysis purposes, the range of recorded<br />

periods from 0.1s to 100s is divided into two subperiods:<br />

1. The periods longer than 1s, where the<br />

microseisms are spread;<br />

2. The periods shorter than 1s, where the most<br />

common noise source is the man-made one.<br />

The results show that:<br />

– For the periods greater then 1s the mode<br />

curve of Power Spectral Density (PSD) closely tracks<br />

the New Low Noise Model (NLNM) (Fig.3) and even<br />

low magnitude events should be reliably detected.<br />

There is no need to filter the signal in this frequency<br />

range and no need to add parameters in the<br />

parametric detector file.<br />

– For the periods lower then 1s the<br />

distribution of PSD has unique model for every<br />

station depending on specific noise sources<br />

(Dimitrova, 2009). The most significant noise sources<br />

are often related to human activities at or near the<br />

Earth´s surface. In this frequency range the filter<br />

parameters of detectors are selected on the base of<br />

the corresponding station noise model (Fig.3).<br />

The output of the Real-time subsystem data<br />

processing is a daily bulletin with the hypocentral<br />

determinations. The event localizations are<br />

visualized on a map and an e-mail is sent to the list<br />

of subscribers. In case of felt (M>=2.5) or damaging<br />

earthquake on the territory of Bulgaria the NDC<br />

issues information to the government authority, mass<br />

media and broad public. The NDC in close<br />

cooperation with the Civil defense carries out<br />

macroseismic investigation in the epicentral region.<br />

The interactive processing of the seismic event<br />

parameters and magnitude determinations are<br />

performed by manual graphic analysis of the<br />

seismogram (Fig. 4). An advantage of SAS_SNDP is<br />

easy access to the automatic arrivals and waveforms<br />

collected in a disk loop with capacity more than 365<br />

days. Additional advantage of the subsystem is the<br />

ability to operate with the waveforms for rapid<br />

manual relocation of the events in time interval close<br />

to real time.<br />

Fig. 4. Interactive data processing by means of Seismic analysis subsystem – SAS_SNDP - tuning of the seismic phase<br />

parameters<br />

The Early warning subsystem provides messages on<br />

the based of first arrived P-phases within 5 to 8<br />

seconds. An alarm module is switched on and the<br />

epicenter of the strong event is visualized on a map.<br />

Some extra modules for network command/control,<br />

state-of-health network monitoring and data archiving<br />

are running as well in the National Data Centre.<br />

Three types of archives are produced in the NDC -<br />

two continuous - miniSEED format and RefTek<br />

237<br />

PASSCALl format; and one event oriented in CSS3.0<br />

scheme format.<br />

CONCLUSIONS<br />

Modern seismological equipment installed at<br />

Bulgarian seismic stations, carefully selected and<br />

developed software packages for data processing<br />

proved to be suitable choice for the purposes of<br />

BNDSN and NDC. Currently, the NDC and BNDSN<br />

allow reliable automatic localization of magnitude


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

events ML>=1.5 within the network, and ML>=2.5 at<br />

regional distances.<br />

Presently BNDSN is a world-class broadband digital<br />

network providing reliable, high quality real-time<br />

seismic monitoring and rapid earthquake information<br />

to both scientific communities and authorities in<br />

Bulgaria and international community for seismic<br />

hazard mitigation and research.<br />

Acknowledgements: The modernization of the BNDSN<br />

was carried out within the projects: Decision SB -<br />

3/04.05.2005 of Permanent Commission for Prevention of<br />

the Population from Natural Disasters, Technological<br />

Accidents and Catastrophes (PCPNDTAC), Bulgaria,<br />

Contract N IKI -11/01.09.2005 with Bulgarian National<br />

Science Fund, Ministry of Education and Science:<br />

“Environmental Monitoring Implement for Risk Assessment<br />

of natural and man-made hazard (EMIRA)”.<br />

References:<br />

Dimitrova L., (2009). Noise level on selected digital stations<br />

of the National Operative Teleseismic System for Seismic<br />

Information (NOTSSI). Comptes rendus de l’Academie<br />

bulgare des Sciences. Vol 62, No4, pp 515-520.<br />

Haikin, L. and A. Kushnir, (2005). Seismic Network Data<br />

Processor (SNDP) - Comprehensive Software for UNIX<br />

Networks. SYNAPSE Science Center, . 122.<br />

Haikin L., L. Begun and M. Morskov, (2009). Haikin-SNDM-<br />

Applications. Seismic Network Data<br />

Monitoring.Comprehensive Software for UNIX Networks.<br />

Introductory Software manual. Version 2.0.<br />

238


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

PALEOSEISMIC STUDY OF THE SUDETIC MARGINAL FAULT AT THE LOCALITY BÍLÁ<br />

VODA (BOHEMIAN MASSIF)<br />

Štpaníková Petra (1, Nývlt Daniel (2), Hók Jozef (3), Dohnal Jií (4)<br />

(1). Dpt. Engineering, Institute of Rock Structure and Mechanics, Academy of Sciences of the Czech Republic, V Holešovikách<br />

41, 182 09 Prague 8.CZECH REPUBLIC. E-mail: stepancikova@irsm.cas.cz<br />

(2). Czech Geological Survey, Brno Branch, Leitnerova 22, 658 69 Brno, CZECH REPUBLIC, E-mail: daniel.nyvlt@geology.cz<br />

(3). Dpt. Geology and Paleontology, Faculty of Natural Sciences, Comenius University in Bratislava, Mlynská dolina, 842 15<br />

Bratislava 4, SLOVAKIA, E-mail: hok@fns.uniba.sk<br />

(4). Faculty of Sciences, Charles University in Prague, Inst. Hydrogeology, Engineering Geology and Applied Geophysics,<br />

Prague Albertov 6, Prague 2, 128 43, CZECH REPUBLIC, E-mail: dohnalj@natur.cuni.cz<br />

Abstract (Paleoseismic study of the Sudetic Marginal Fault at the locality Bílá Voda (Bohemian massif): The study area is<br />

situated in the north-eastern part of the Bohemian Massif and comprises SE part of the Sudetic Marginal Fault zone (SMF) with<br />

pronounced mountain front, which borders the Sudeten Mts. Paleoseismological trenches at the locality of Bílá Voda municipality<br />

performed across the SMF were carried out in order to study potential prehistoric morphogenic earthquakes responsible for the<br />

present-day morphology. Geomorphologic analysis and electrical resistivity tomography enabled to position the trenches at<br />

suitable site. The trenches revealed prevailing strike-slip character of the late Quaternary faulting. Faulted deposits were dated by<br />

OSL and radiocarbon dating methods and the results are discussed here. At least four or five events creating a surface rupture<br />

during late Quaternary (Holocene) were inferred. Due to strike-slip character, the horizontal displacement and the related slip-rate<br />

on the SMF is still under investigation. From the previous study, minimum magnitude M 6.3 is expected for the SMF.<br />

Key words: active tectonics, trenching, paleoseismology, Sudetic Marginal Fault<br />

The study area is situated in the north-eastern limit of<br />

the Bohemian Massif in the Czech Republic (central<br />

Europe). The studied Sudetic Marginal Fault (SMF) is<br />

a part of the Elbe Fault System, which comprises<br />

WNW-striking zone from SE North Sea to the front of<br />

Outer Carpathian nappes and disrupts the Variscan<br />

central Europe at the length of several hundreds of<br />

kilometers. The SMF is about 200km long and<br />

divides the Variscan Sudetic Mountains block from<br />

upland-like Fore-Sudetic block hosting Cenozoic<br />

cover. For a length of 130, km it controls the<br />

pronounced mountain front of the Sudetic Mountains<br />

(Fig. 1).<br />

Quaternary activity of the fault has been<br />

demonstrated e.g. by Middle and Upper Pleistocene<br />

fluvial terraces that are truncated by the SMF and<br />

expressed by 5 - 20 m high scarp in their longitudinal<br />

profiles in the Polish portion of the fault (e.g.<br />

Krzyszkowski et al., 1995). Local historic<br />

earthquakes recorded within the SMF had epicentral<br />

intensity estimated to reach only I 0 =4 -7 (MSK),<br />

which would correspond to macroseismic magnitude<br />

M M =3-4.9 (Guterch and Lewandowska-Marciniak,<br />

2002). They include the events from years 1594 and<br />

1778 (Zlotoryja/Legnica), and 1615 and 1786 near<br />

Bardo/Dzieroniów. Other historic earthquakes with<br />

Fig. 1. Digital elevation model of the morphologically distinctive Sudetic Marginal Fault (SMF) that controls the Sudetic Mountains front.<br />

Red stars - historic earthquakes, I 0 = epicentral intensity; white square – study area.<br />

239


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

more distant epicentres or those being felt in the<br />

zone parallel to the SMF (Fig. 1) include events in<br />

1496, 1562, 1786, 1823, 1877, and 1895 (Kárník et<br />

al., 1958; Olczak, 1962; Pagaczewski, 1972; Guterch<br />

and Lewandowska-Marciniak, 2002).<br />

As the historic earthquakes were not large enough to<br />

create the morphology (I 0 =4-7 MSK; cf. Guerrieri and<br />

Vittori, 2007), the study of potential presence of<br />

larger prehistoric earthquakes responsible for the<br />

growth of the mountain front was carried out. This<br />

paper presents results from investigation of the<br />

mountain front at the locality Bílá Voda municipality<br />

(Czech Republic). The results from three parallel<br />

trenches (C, D, F) performed across the SMF are<br />

discussed here.<br />

In order to determine the position of the fault for the<br />

following trenching, electric resistivity tomography<br />

with electrode spacing 2 m was performed (Fig. 2).<br />

The fault appeared to be expressed by the presence<br />

of a remarkable resistivity gradient, which divided<br />

high-resistivity of anticipated metamorphic rocks on<br />

the footwall and low-resistivity of Miocene deposits<br />

on the hanging-wall (Štpaníková et al., 2011). This<br />

was confirmed in the trenches C, D, F (Fig 2b). All<br />

the trenches revealed a subvertical fault zone striking<br />

135°-150° dividing Paleozoic crystalline rocks<br />

(phyllites, schists, granitic aplite) and late Quaternary<br />

colluvial deposits overlaying warped Miocene<br />

sediments. The SMF has been generally considered<br />

as a normal fault having experienced inversion during<br />

Late Cretaceous/Paleogene and extensional<br />

reactivation since Miocene during Alpine cycle (c.f.<br />

Badura et al., 2007). Nevertheless, our trenches<br />

revealed that the movements on the fault, at least the<br />

youngest ones, had very probably prevailing strikeslip<br />

character. The strike-slip on the main marginal<br />

fault with strike and dip of 135°/75°NE is suggested<br />

by a flower structure within the 4-m wide zone of<br />

tectonic breccia and fault gouge displayed both on<br />

the trench walls and on the floor, and by completely<br />

different lithologies on the both sides without any<br />

matching points.<br />

The down-thrown block is composed by Miocene<br />

(unit A; Fig 3), strongly kaolinised clayey silty sand<br />

with a small pebble admixture of local lithology. The<br />

clay originated by chemical weathering of feldspars<br />

and granite groundmass. These sediments may be<br />

correlated with Member C described few km to the<br />

SE by Štpaníková et al. (2010) and interpreted as<br />

to be deposited in a fluvio-limnic environment during<br />

the Carpatian and lower Badenian (i.e. early to mid<br />

Miocene, ~18–15 Ma). These deposits are<br />

upwarped, which was indicated also by preceding<br />

ERT profiles (see Fig 2), and covered by colluvial<br />

deposits (unit B). The colluvial deposits represent<br />

matrix-supported intermediate to sandy diamicts with<br />

gravel clast content between 5 and 25% and sandy<br />

to silty matrix. Largest clasts are up to 20–25 cm and<br />

are made of angular to subangular gneiss clasts,<br />

granodiorite and quartz clasts. The colluvial<br />

sediments have a sharp erosional base and their<br />

lowermost part is enriched by Mn oxides as a result<br />

of manganese precipitation from percolating water<br />

just above the impermeable Miocene clayey sand.<br />

Fig. 2. Model resistivity of ERT at the locality Bílá Voda, Wenner–Schlumberger electrode array with logged geology in<br />

Trench C. a) Resistivity in ohm m; vertical exaggeration=1; iteration 3; RMS error=2.3%. The SMF lies within the<br />

expressive resistivity gradient between Stations 143 and 145. b) Geology in the trench 1 – Late Quaternary colluvium; 2<br />

– Miocene lacustrine deposits, 3 – fault zone with tectonic breccia and fault gauge, 4 – micashists, 5 – granitic aplite, 6<br />

– gneisses, 7 – Late Holocene colluvium.<br />

240


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 3. Log of the trench F. See text for explanation of units A to E. Unit F – fault zone with tectonic breccia, G – metamorphic<br />

rocks.<br />

The colluvium close to the fault is obviously faultrelated<br />

and is interpreted to be derived from the fault<br />

scarp created after an earthquake (c.f. McCalpin,<br />

2009); it includes material from the fault zone and<br />

particularly clasts of the fault breccia. Although the<br />

colluvium does not display completely clear discrete<br />

sedimentological units that could be ascribed to<br />

individual events, there are some indications; in<br />

trench F (Fig 3), the part of colluvium that includes<br />

blackish matrix with breccia clasts have wedge shape<br />

(B1) and seems to be covered by another wedgeshaped<br />

colluvial unit (B2), which includes the fault<br />

rocks at lesser amount. This would indicate at least<br />

two faulting events, which resulted in the origin of the<br />

colluvial wedges. This fault-related colluvium is then<br />

covered by 35–40 cm thick beige to grey lens<br />

composed of beds of slightly clayey silt, sand and<br />

granule (unit C). They bear a very fine lamination,<br />

especially in silty layers. The upper part of the lens is<br />

dominated by bands of Mn and Fe oxides, which<br />

percolate fine to medium sand with some granule<br />

clasts. The sedimentary body is fault-parallel oriented<br />

and it was very probably a small pool emerging in<br />

front of the fault scarp. After its deposition the lens<br />

were tilted and displaced by minor faults (Fig 4),<br />

which displayed various sense of movement on the<br />

both walls of trench F. In trench C the deposit is only<br />

tilted. This must have been a result of another<br />

younger faulting (3 rd event). Unfortunately, no pollen<br />

Fig. 4. Tilted and displaced unit C - lens composed of beds<br />

of clayey silt, sand and granule; trench F.<br />

241<br />

or macro-remains were found within the silty<br />

sediment and only residual sediment carbon was<br />

dated by AMS 14 C dating.<br />

We have a set of 3 OSL dates from this colluvial unit<br />

ranging from 25.8 ± 1.6 ka to 9.5 ± 0.9 ka. We accept<br />

only the youngest one as reliable age, because the<br />

age corresponds to ~1 m long lens of sorted sand,<br />

which imply better zeroing before burial of this<br />

material. The older dates are probably influenced by<br />

residual signal. This is supported by 2 radiocarbon<br />

dates (all radiocarbon ages given here are calibrated<br />

ages using IntCal09 calibration dataset; Reimer et<br />

al., 2009) falling into the early Holocene – 7,997 ± 23<br />

a and 6,433 ± 29 a. Both ages originate from residual<br />

sediment carbon, as no plant macro-remains were<br />

found in this unit. The exact age may therefore be<br />

slightly older than the ages given above.<br />

The above-mentioned fault-derived colluvium is then<br />

cut by the youngest faults of the flower structure,<br />

which would indicate the following (4 th ) faulting event<br />

unless it was simultaneous with the deformation of<br />

the silty lens (3 rd event), which is ambiguous. Both<br />

the colluvium (incl. the silty lens) and the fault zone is<br />

sealed by banded layers due to superficial flow<br />

processes (unit D). They create individual bands 3 to<br />

10 cm thick, which continue >5 m downhill. The<br />

thickness of bands is much smaller, than it was<br />

described in other trenches studied along the Sudetic<br />

Marginal Fault at Vlice (Štpaníková et al., 2010),<br />

where up to 20 cm thick layers were interpreted to be<br />

deposited by a sheet gelifluction and ascribed to the<br />

Late Glacial, which was also supported by<br />

radiocarbon dating. So, the banded layers here in<br />

Bílá Voda were considered to be of similar origin and<br />

age, i.e. from Late Glacial, which would have<br />

important implication for inferring the age of the<br />

faulting (c.f. Štpaníková et al., 2009, 2011).<br />

However, new data on both OSL and radiocarbon<br />

age of the banded layers here give similar ages of<br />

2.6 ± 1.2 ka and 2,610 ± 108 a respectively. The<br />

thickness and dating of the layers here may show on<br />

the frost creeping of this material, which does not<br />

imply flow processes connected with seasonal<br />

thawing of the active permafrost layer, which was not<br />

present here during the Holocene. Frequent and


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

usually rapid diurnal frost creep processes may affect<br />

the superficial layer typically 5–10 cm thick even at<br />

the warm margin of the solifluction-affected<br />

environment (Matsuoka, 2001). Therefore, the<br />

presence of these processes at low altitude as here<br />

(~360 m) during the Holocene is feasible.<br />

In trench C on the footwall, these banded layers<br />

display around 35cm high step just coinciding with<br />

minor faults with opposite dip. These faults also do<br />

not show any displaced matching points, so<br />

horizontal movements with minor reverse component<br />

might be inferred here. On the hanging wall, these<br />

banded layers are slightly up-warped, which might<br />

have been related to a transpression responsible for<br />

the above–described step in the banded layers. This<br />

would imply another, potentially 5 th or 4 th event.<br />

The banded layers are covered by late Holocene<br />

colluvial deposits usually 0.5 to 1.5 m thick (unit E).<br />

These are matrix-supported, with the matrix<br />

composed predominantly of sandy fraction originating<br />

from the weathering of local Javorník granodiorite,<br />

which is mixed with silt and mostly pebble to fine<br />

cobble clasts. Gravel clast typically first cm large are<br />

mostly subangular to less subrounded and are made<br />

of Javorník granodiorite, quartz, (ortho)gneiss and<br />

schist, rather rare are clast of Nordics (originating<br />

from tills and glaciofluvial deposits left here by<br />

Pleistocene ice sheets; Nývlt et al., 2011) and pieces<br />

of bricks pointing to young age. Radiocarbon dates of<br />

charcoals from these young Holocene colluvial<br />

deposits range from 1,145 ± 89 a to 757 ± 30 a. On<br />

the top of these colluvia a cambisol type of recent soil<br />

is developed with 25–30 cm thick light grey-brown<br />

A-horizon with ploughed erosional base.<br />

To summarize, the results from the trenches at the<br />

Bílá Voda locality show potentially four to five<br />

movements on the SMF during late Quaternary<br />

(Holocene). Due to lack of kinematic indicators in the<br />

trenches perpendicular to prevailing movements, the<br />

sense of the strike-slip is still under investigation.<br />

Since the perpendicular trenches across strike-slips<br />

do not show the amount of horizontal slip, also sliprate<br />

could not be assessed with reasonable<br />

confidence. Nevertheless, based on the previous<br />

trenching (Štpaníková et al., 2010), according to<br />

empirical relationship ‘magnitude versus maximum<br />

vertical displacement’ (Wells and Coppersmith,<br />

1994), the minimum moment magnitude M 6.3 is<br />

expected on the SMF.<br />

Acknowledgements: The research was supported by the<br />

Grant Agency of the Czech Republic No. Czech Science<br />

Foundation No. 205/08/P521. The work has been<br />

elaborated within the Institute Research Plan of the Institute<br />

of Rock Structure and Mechanics, Academy of Sciences of<br />

the Czech Republic, No. AVOZ30460519.<br />

of the Bohemian Massif, Central Europe. Acta<br />

Geodynamica et Geomaterialia 4 (148), 7–29.<br />

Guerrieri, L., Vittori, E. (Eds). (2007). Intensity scale ESI<br />

2007. Mem. Descr. Carta Geologica d´Italia, 74, Servizio<br />

Geologico d´Italia – Dipartimento Difesa del Suolo,<br />

APAT, Rome, Italy, 41pp.<br />

Guterch, B., Lewandowska-Marciniak, H., (2002). Seismicity<br />

and seismic hazard in Poland. Folia Quaternaria 73, 85–<br />

99.<br />

Kárník, V., Michal, E., Molnár, A., (1958). Erdbebenkatalog<br />

der Tschechoslowakei bis zum Jahre 1956. Geofysikální<br />

Sborník 69, 411–598.<br />

Krzyszkowski, D., Migo, P., Sroka, W., (1995). Neotectonic<br />

Quaternary history of the Sudetic Marginal fault, SW<br />

Poland. Folia Quaternaria 66, 73–98.<br />

Matsuoka, N., (2001). Solifluction rates, processes and<br />

landforms: a global review. Earth-Science Reviews 55,<br />

107–134.<br />

McCalpin, J. (Ed.) (2009). Paleoseismology. Second<br />

Edition. Academic Press. 613 pp.<br />

Nývlt, D., Engel, Z., Tyráek, J., (2011). Pleistocene<br />

Glaciations of Czechia. In: Ehlers, J., Gibbard, P. L.,<br />

Hughes, P. D., (Eds): Quaternary Glaciations – Extent<br />

and chronology, A closer look. Developments in<br />

Quaternary Science 15, 37–46, Elsevier.<br />

Olczak, T., (1962). Seismiczno Polski w okresie 1901–<br />

1950. Acta Geophysica Polonica 10, 3–11.<br />

Pagaczewski, J., (1972). Catalogue of Earthquakes in<br />

Poland in 1000–1970 years. Mat. I Prace Inst. Geofiz.,<br />

vol. 51. 36 pp.<br />

Reimer, P.J., Baillie, M.G.L., Bard, E., Bayliss, A., Beck,<br />

J.W., Bertrand, C., Blackwell, P.G., Buck, C.E., Burr, G.,<br />

Cutler, K.B. Damon, P.E., Edwards, R.L., Fairbanks,<br />

R.G., Friedrich, M., Guilderson, T.P., Hughen, K.A.<br />

Kromer, B. McCormac, F.G., Manning, S. Bronk Ramsey,<br />

C., Reimer, R.W., Remmele, S., Southon, J.R., Stuiver,<br />

M., Talamo, S., Taylor, F.W., van der Plicht, J., &<br />

Weyhenmeyer, C.E., (2009). Intcal09 and Marine09<br />

Radiocarbon Age Calibration Curves 0-50,000 Years Cal<br />

BP. Radiocarbon 51, 1111–1150.<br />

Štpaníková P., Dohnal J., Pánek T., ój M., Smolková V.,<br />

Šilhán K., (2011). The application of electrical resistivity<br />

tomography and gravimetric survey as useful tools in an<br />

active tectonics study of the Sudetic Marginal Fault<br />

(Bohemian Massif, central Europe). Journal of Applied<br />

Geophysics 74, 69-80.<br />

Štpaníková, P., Hók, J., Nývlt, D., (2009). Trenching<br />

survey on the south-eastern section of the Sudetic<br />

Marginal Fault (NE Bohemian Massif, intraplate region of<br />

central Europe). In: Archaeoseismology and<br />

Palaeoseismology in the Alpine-Himalayan Collisional<br />

Zone (Pérez-López, R., Grützner, C., Lario, J.,<br />

Reicherter, K., Silva, P.G. eds). Baelo Claudia, Spain,<br />

149–151.<br />

Štpaníková, P., Hók, J., Nývlt, D., Dohnal, J., Sýkorová,<br />

I., Stemberk, J., (2010). Active tectonics research using<br />

trenching technique on the south-eastern section of the<br />

Sudetic Marginal Fault (NE Bohemian Massif, central<br />

Europe). Tectonophysics 485, 269–282.<br />

Wells, D.L., Coppersmith, K.J., (1994). Empirical<br />

relationships among magnitude, rupture length, rupture<br />

area, and surface displacement. Bulletin of the<br />

Seismological Society of America 82, 974–1002.<br />

References<br />

Badura, J., Zuchiewicz,W., Štpaníková, P., Przybylski, B.,<br />

Kontny, B., Caco, S., (2007). The Sudetic Marginal<br />

Fault: a young morphotectonic feature at the NE margin<br />

242


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

TOWARD DEVELOPMENT OF A LONG RUPTURE HISTORY OF THE IMPERIAL FAULT IN<br />

MESQUITE BASIN, IMPERIAL VALLEY, SOUTHERN CALIFORNIA<br />

Tsang, Rebecca Y. (1, Thomas K. Rockwell (2), Aron J. Meltzner (3), Paula M. Figueiredo (4)<br />

(1) Dept. of Geol. Sci., San Diego State University, San Diego, California 92182, U.S.A. Email: tsang.y.rebecca@gmail.com<br />

(2) Dept. of Geol. Sci., San Diego State University, San Diego, California 92182, U.S.A. Email: trockwell@geology.sdsu.edu<br />

(3) Earth Observatory of Singapore, Nanyang Technological University, Singapore. Email: meltzner@ntu.edu.sg<br />

(4) IDL / FCUL - Geology Department, Lisbon University, Lisbon, Portugal. Email: pmfigueiredo@fc.ul.pt<br />

Abstract (Toward Development of a Long Rupture History of the Imperial Fault in Mesquite Basin, Imperial Valley,<br />

Southern California): We conducted a paleoseismic study on the northern Imperial fault at the Dogwood site in Mesquite Basin,<br />

southern California, to extend the record of late Holocene surface ruptures. New trench exposures have revealed evidence for up<br />

to 17 events in the past 1300 years, yielding an average recurrence interval of 80 years, and the large CV suggests the fault<br />

ruptures in a non-episodic manner. Our data indicate that the connection between lake and earthquake cycles is either very weak<br />

or non-existent as many of the ruptures occurred during dry periods between lakes. However, there is a strong correlation between<br />

the earthquake chronologies in the southern San Andreas fault and those on the northern Imperial fault, suggesting surface<br />

rupture at the southern portion of San Andreas fault may have triggered surface slip on the northern Imperial fault, or vice versa.<br />

Key words: clustering, Imperial, paleoseismology, recurrence<br />

INTRODUCTION<br />

The Imperial fault is a northwest-striking, dextral fault<br />

located in the Mesquite Basin, an area in which the<br />

seismicity patterns have been interpreted in terms of<br />

spreading and transform faulting (Sharp, 1982). The<br />

North American-Pacific plate boundary rate is<br />

approximately 45±1 mm/yr for southern California<br />

(DeMets et al., 1994). The Imperial, along with the<br />

San Jacinto and Elsinore faults, are considered as<br />

part of the San Andreas fault system and<br />

accommodate a significant proportion of the plateboundary<br />

motion (Hill et al, 1990). The 70-km fault<br />

crosses the U.S.-Mexico Border and terminates in<br />

the south at a right stepover to the Cerro Prieto fault,<br />

and in the north at the Brawley Seismic Zone which<br />

Fig. 1: Map<br />

showing the<br />

major structures<br />

and the<br />

Dogwood site in<br />

the Imperial<br />

Valley (modified<br />

from Thomas<br />

and Rockwell,<br />

1996).<br />

243


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

extends farther north to the Salton Sea (Figure 1).<br />

The fault was not recognized until the 1940 M L 7.1<br />

earthquake that was caused by an end-to-end<br />

rupture propagating mostly to the southeast<br />

(Buwalda and Richter, 1941; Richter, 1958). Thirtynine<br />

years later, the fault ruptured again to produce a<br />

M L 6.6 earthquake (Chavez et al, 1982). Besides<br />

surface ruptures, the area is also characterized by<br />

regular aseismic creep and high levels of<br />

microseismicity (Cohn et al., 1982; Lyons et al.,2002;<br />

Shearer, 2002).<br />

In earlier work, we resolved the timing and<br />

displacement for the six most recent surface<br />

ruptures, all of which have occurred in the past 500<br />

years. In this study, we conducted new paleoseismic<br />

studies on the northern Imperial fault to extend the<br />

record of late Holocene surface ruptures to better<br />

understand the behaviour of this plate boundary<br />

strike-slip fault. Twelve new trench exposures at the<br />

Dogwood site in Mesquite Basin, near El Centro,<br />

California have revealed evidence for up to 17 events<br />

in the past 1300 years.<br />

The Dogwood Site<br />

Our study area is located about 10 km north of<br />

Interstate Highway 8, next to Dogwood Road which<br />

makes it easily accessible (Figure 2). El Centro has<br />

an arid climate and is covered by farmland, but the<br />

lack of vegetation particularly at our site makes<br />

excavation easier. The Imperial fault at this site is<br />

expressed surficially as a single strand, which<br />

Fig. 2: Satellite image of the study area (yellow box) in<br />

the Imperial Valley, southern California.<br />

reduces the structural complexity that might mask<br />

evidence for past surface ruptures (Figure 2). Most<br />

importantly, the Dogwood site has preserved a<br />

remarkable stratigraphy of Lake Cahuilla which is an<br />

ephemeral freshwater body formed when the<br />

floodwaters of the Colorado River intermittently filled<br />

the Salton depression during the Holocene (Sharp<br />

1982a). This allows us to identify event horizons and<br />

to obtain ages for past surface ruptures by<br />

radiocarbon dating of detrital charcoal that was<br />

embedded within the stratigraphy.<br />

Fig. 3: Composite trench logs of T9B-SE (top) and T4Deep-SE (bottom) showing the remarkable Lake Cahuilla<br />

stratigraphy and the event horizons identified in these two exposures.<br />

244


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Stratigraphy and Lake Chronology<br />

The stratigraphy consists of lacustrine, deltaic and<br />

fluvial sediments that were deposited during the filling<br />

and drying periods of Lake Cahuilla. Up to a total of 5<br />

m of stratigraphy below ground surface was exposed<br />

in our trenches which contains a record of 6 to 7 lake<br />

episodes. The lacustrine deposits consist of massive<br />

and bedded clay; the deltaic deposits are<br />

characterized by interbedded fine silt, silty clay and<br />

clay; and the fluvial deposits contain planar and<br />

cross-bedded silt and fine sand. In some exposures,<br />

the strata above the most recent lake deposition<br />

(post 1720 AD) comprises of localized flood deposits<br />

and spoil. A well-sorted sandy unit (Unit 390) of 1 to<br />

2 m in thickness was preserved in 10 trench<br />

exposures and was interpreted as a sand blow. Other<br />

sand blows also appear in five exposures of the<br />

lower section. The well-bedded stratigraphy at this<br />

site allows us to resolve the age of past earthquakes<br />

relative to the stratigraphy rather precisely (Figure 3).<br />

Calendar years of past earthquakes were determined<br />

by using the lake chronology that Meltzner and<br />

Rockwell had developed (unpublished data). They<br />

submitted over 180 detrital charcoal and peat<br />

samples from the Lake Cahuilla strata at different<br />

localities for 14 C analysis. About 50 14 C dates were<br />

used in the OxCal program to calculate calendar<br />

years of the different lake episodes. (For details of<br />

OxCal, please refer to Lienkaemper and Ramsey,<br />

2009). Identifying and locating an event horizon<br />

within the lake stratigraphy would then allow us to<br />

obtain calendar years of past surface ruptures.<br />

was deposited at the paleosurface rather horizontally,<br />

forming an angular unconformity with the strata<br />

below.<br />

A rating scheme of the paleoseismic indicators for<br />

each event horizon was devised to better qualify the<br />

likelihood of a past earthquake. Each indicator<br />

receives a rating of 1 to 4, with 1 being the weakest<br />

and 4 being the strongest. For example, an angular<br />

unconformity is a strong line of evidence hence it has<br />

a rating of 4. On the contrary, downward growth of<br />

displacement might be a result of fault dying out<br />

upward rather than cumulative displacement of<br />

multiple events on the same fault. Therefore, this<br />

indicator has the lowest rating of 1 due to the<br />

ambiguity in interpretation. The frequency (F) of<br />

these indicators is another factor to determine the<br />

likelihood of an event. In this method, the value of F<br />

is calculated by dividing the number of exposures<br />

where the indicator was observed by the number of<br />

exposures where the indicator should be observed.<br />

Finally, R is multiplied by F for the R*F value which<br />

indicates the likelihood of a paleoearthquake. The<br />

higher the R*F value, the more likely an event had<br />

occurred (Figure 5). Event 4 although receives a low<br />

R*F value, its occurrence was assured by a buried<br />

offset channel study by Meltzner and Rockwell<br />

(unpublished data). Among all 17 events, almost all<br />

events receive an R*F value higher than 1 except for<br />

Events 4, 8.5, 8.7 and 9. These events could be<br />

interpreted as unlikely or as events with small<br />

displacement.<br />

Evidence for Surface Ruptures<br />

Common paleoseismic indicators were used to<br />

identify event horizons for past surface ruptures,<br />

including upward fault termination, fissure fill, sand<br />

blow, angular unconformity, colluvial wedges and<br />

scarp-derived debris, liquefaction pipes and<br />

downward growth of displacement. For example, in<br />

Figure 4, several indicators and their geometrical<br />

relationships define the event horizon for Event 8. A<br />

fissure fill is bounded by two fault splays terminating<br />

upward at the same horizon, where a sand blow unit<br />

Fig. 5: Likelihood of an event based on quality and<br />

quantity of paleoseismic indicators as seen in the plot of<br />

average RF values of each event. Higher values<br />

indicate more likelihood.<br />

Fig. 4: Paleoseismic indicators used to identify the<br />

horizon of Event 8 include upward fault termination<br />

(UT), angular unconformity (AU) and fissure fill (FF).<br />

245<br />

CONCLUSION<br />

We conducted this paleoseismic study on the<br />

northern Imperial fault at the Dogwood site in<br />

Mesquite Basin, southern California, to extend the<br />

record of late Holocene surface ruptures to better<br />

understand the behavior of this plate boundary strikeslip<br />

fault. Event horizons in 12 trench exposures were<br />

identified within the Lake Cahuilla stratigraphy using<br />

the standard paleoseismic indicators such as angular<br />

unconformity, fissure fill and upward truncation of<br />

faults. By using the lake chronology model that<br />

Meltzner and Rockwell (unpublished data) developed<br />

for Lake Cahuilla in the Imperial Valley, we


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

developed an earthquake chronology of the northern<br />

Imperial fault. Our study has revealed evidence for<br />

up to 17 events in the past 1300 years, yielding an<br />

average recurrence interval of about 80 years, albeit<br />

with a large standard deviation and coefficient of<br />

variation (Figure 6). We tested the hypothesis of a<br />

connection between lake and earthquake cycles,<br />

however, our data suggest that the connection is<br />

either very weak or non-existent as the majority of<br />

the ruptures occurred during dry periods between<br />

lakes. On the other hand, we do find a strong<br />

correlation between the earthquake chronologies<br />

determined for the southern San Andreas fault and<br />

the rupture history on the northern Imperial fault,<br />

suggesting surface rupture at the southernmost<br />

portion of San Andreas fault may have triggered<br />

earthquakes on the northern Imperial fault (Figure 7).<br />

Fig. 6: Earthquake calendar ages obtained from OxCal.<br />

The higher the probability density, the more wellconstrained<br />

the ages are. The height of a peak does not<br />

give any information about the likelihood of a<br />

paleoearthquake.<br />

Acknowledgements: We are grateful for the help we<br />

received to make this project possible. We thank the<br />

landowner for permission to excavate on his property and<br />

Jim Little for the backhoe operation. Field assistants<br />

including Gayatri Marliyani, Nissa Morton, Barrett Salisbury,<br />

Mike Buga, Petra Štpaníková, Eulàlia Masana and Katie<br />

Anderson offered much help in the exposure preparation.<br />

This project was funded by the National Earthquake<br />

Hazards Reduction Program (NEHRP).<br />

References<br />

Buwalda, J. P., & C. F. Richter (1941). Imperial Valley<br />

earthquake of May 18, 1940. Geological Society of<br />

America Bulletin, 52, 1942-1943.<br />

Chavez, D., J. Gonzales, A. Reyes, M. Medina, C. Duarte,<br />

J. N. Brune, F. L. Verson, III, R. Simons, L. K. Hutton, P.<br />

T. German, and C. E. Johnson. (1982). Mainshock<br />

location and magnitude determination using combined<br />

U.S. and Mexican location and magnitude determination<br />

using combined U.S. and Mexican data. The Imperial<br />

Valley, California, Earthquake, October 15, 1979. United<br />

States Geological Survey Professional Paper, 1254, 51-<br />

54.<br />

Cohn, S. N., C. R. Allen, R. Gilman, & N. R. Goulty. (1982).<br />

Preearthquake and postearthquake creep on the Imperial<br />

fault and the Brawley fault zone. United States Geological<br />

Survey Professional Paper, 161-167.<br />

Fig. 7: Earthquake chronologies from paleoseismic<br />

sites on the southern San Andreas fault and the<br />

northern Imperial fault. The possible earthquakes<br />

Coa-3 and Coa-6, and the less likely events E8.5,<br />

E8.7 and E9 are shown in gray text. Correlations of<br />

events at the three paleoseismic sites are highlighted<br />

in pink.<br />

DeMets, C., R. G. Gordon, D. F. Argus, and S. Stein.<br />

(1994). Effects of recent revisions to the geomagnetic<br />

reversal time scale on estimates of current plate motions.<br />

Geophysical Research Letters, 21, 2191-2194.<br />

Hill, D. P., J. P. Eaton, and L. M. Jones (1990). Seismicity:<br />

1980-1986, in Wallace, R.E., ed. The San Andreas Fault<br />

System, U.S. Geol. Surv. Profess. Pap. 1515, 114-151.<br />

Lienkaemper, J. J., and C. B. Ramsey (2009). OxCal:<br />

Versatile Tool for Developing Paleoearthquake<br />

Chronologies. Seismol. Res. Lett., 80, 431-434.<br />

Lyons, S. N., Y. Bock, and & D. T. Sandwell (2002). Creep<br />

along the Imperial Fault, southern California, from GPS<br />

measurements. J. of Geophysical Research, 107 (B10).<br />

Richter, C. F. (1958). Elementary seismology: San<br />

Francisco. W. H. Freeman, 768.<br />

Sharp, R. V. (1982). Tectonic Setting of the Imperial Valley<br />

Region. United States Geological Survey Professional<br />

Paper, 5-14.<br />

Shearer, P. M. (2002). Parallel fault strands at 9-km depth<br />

resolved on the Imperial Fault, Southern California.<br />

Geophysical Research Letters, 29 (14).<br />

Thomas, A. P., and T. K. Rockwell (1996). A 300- to 550-<br />

year history of sip on the Imperial fault near the U.S.-<br />

Mexico border: Missing slip at the Imperial fault<br />

bottleneck. Journal of Geophysical Research, 101,<br />

5987-5997.<br />

246


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

MAPPING PALEO SHORELINES IN LESVOS ISLAND: NEW CONTRIBUTION TO THE<br />

LATE QUATERNARY RELATIVE SEA LEVEL CHANGES AND TO THE NEOTECTONICS<br />

OF THE AREA<br />

Vacchi, Matteo (1,2), Alessio Rovere (1,2), Nickolas Zouros (3), and Marco Firpo (1)<br />

(1) Department for the Study of the Territory and its Resources, University of Genova, Genova. (Italy)<br />

(2) SEAMap srl, Seascape Evaluation Assessment and Mapping, Genova (Italy)<br />

(3) Department of Geography, University of Aegean, Mytylene, Lesvos Island (Greece)<br />

Abstract (Mapping paleo-shorelines in Lesvos Island: new contribution to the late quaternary relative sea level changes<br />

and to the neotectonics of the area): This study attempts to delineate the late Quaternary morphotectonic evolution of the<br />

coastal zone of southern Lesvos Island (NE Aegean Sea) on the basis of detailed field mapping of different sea level indicators.<br />

The geomorphological analysis of the markers, coupled with radiometric dating, allowed to identify a Late Quaternary regional<br />

uplift trend controlled by the footwall of a large normal fault (Lesvos Fault) located between Lesvos and Chios Islands. At local<br />

scale, superimposed to this trend, the area is characterized by many indicators of rapid vertical displacements (uplift and<br />

subsidence) related to the high seismicity. The most important co-seismic event has been found to have happened between 3365<br />

and 3924 BP which uplifted of about 0.8 m a large sector of the southern Lesvos coastline.<br />

Key words: raised shorelines; coastal uplift, Lesvos Island; Relative sea level changes.<br />

INTRODUCTION<br />

Due to the geodynamic setting, several sectors of the<br />

Mediterranean basin exhibit evidence of differential<br />

vertical movements during the late quaternary<br />

(Stewart & Morhange, 2009). In particular, several<br />

papers investigated the coastal morphotectonics of<br />

the Aegean region, one of the most seismically active<br />

areas not only at Mediterranean scale (Stiros et al.,<br />

2000; 2009 Palyvos et al., 2008; Cundy et al., 2010).<br />

In many cases geomorphological approaches<br />

became very practical solutions to get quantitative<br />

information on late quaternary uplift and<br />

paleoseismicity of coastal areas (Palyvos et al.<br />

2008). For this reason, detailed mapping of paleosea<br />

level markers has been often used as a tool to<br />

quantify coastal uplift and relative sea level changes<br />

in several areas of the Aegean Sea (Pirazzoli et al.,<br />

2004; Stiros et al., 2009). In this study we analyzed<br />

the coastal geomorphology of the southern sector of<br />

Lesvos Island, located in the NE Aegean Sea. Here,<br />

despite several papers already dealt with the<br />

neotectonics of this area, detailed information on the<br />

coastal morphotectonics of the Island is lacking.<br />

Morphological, biological and sedimentary records of<br />

past sea level were mapped, analyzed and, where<br />

possible, sampled and dated.<br />

(Roumelioti et al., 2011).The study area is located in<br />

the south eastern part of the Island, in the coastal<br />

area comprised between Tarti and the entrance of<br />

the Kalloni gulf (Fig.1). From the geomorphological<br />

point of view, it is mainly characterized by steep cliffs<br />

alternated with gravel pocket beaches, often with<br />

beachrocks outcrops. The outcropping lithologies are<br />

composed by a basement of Alpidic and pre-Alpidic<br />

rocks, mainly Triassic limestones, marbles and<br />

schists. In the whole sector the tidal range does not<br />

exceed 0.2 m (Vousdoukas et al., 2009).<br />

Marine and coastal processes produce many kinds of<br />

geomorphologic markers, especially on rocky shores,<br />

that are related to the contemporary sea-level<br />

position (Pirazzoli, 2007). Among them tidal notches,<br />

and the inner edge of benches and shore platform<br />

are considered precise indicator of paleo sea level<br />

stands in microtidal environment such as the<br />

Mediterranean Sea. Endolithic organisms often leave<br />

morphological signs that can be used as sea-level<br />

indicator and, at Mediterranean scale, Lithophaga<br />

boreholes horizontal upper limit represent a good<br />

biological indicator of former m.s.l. (Laborel and<br />

Laborel-Deguen, 1994).<br />

Lesvos islandis located in a geotectonically complex<br />

area, because directly affected by the North<br />

Anatolian Fault Zone (NAFZ), its westward<br />

continuation in the Aegean Sea, known as the North<br />

Aegean Trough (NAT) and the West Anatolia Graben<br />

System (WAGS) in Asia Minor with significant<br />

historical seismicity (Papazachos and Papazachou,<br />

1997; Papazachos and Kiratzi, 1996). As a result of<br />

the interaction between those tectonic systems,<br />

Lesvos Island presents a strong diversity in fault<br />

setting and is presently characterized by the parallel<br />

activity of both normal and strike slip faults<br />

247


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Along the platform fragments of marine shells<br />

(vermetids, serpulids, gastropods) were found still<br />

preserved in growth position at about .0.5 m a.sl.<br />

AMS radiocarbon dating were performed on 3<br />

samples of marine shells (gastropods and serpulids)<br />

sampled at about 4 km of distance one from the<br />

other. Calibrated ages dated the samples<br />

respectively 3365 – 3648 BP (K1 beta); 3460 – 3812<br />

BP (AF1 beta); 3609 – 3924 BP (AF3 poznan).<br />

Fig.1. Geographical location and active faults of the study<br />

area. Focal mechanism represent the > 4M earthquakes<br />

after the 1928.<br />

A second level of uplifted wave cut platform was<br />

mapped along this sector at an elevation ranging<br />

from 5 to 6 m. This level is recurrent in 4 sites along<br />

the coastline and is always covered by beach<br />

deposits rich in marine fauna. Two samples of marine<br />

shells (Patella ferruginea) were sampled. AMS<br />

radiometric dating gave calibrated ages of 29160 ±<br />

150 (V1 beta) and 27410 ± 120 (Ag2 beta).<br />

A detailed mapping of paleo-shorelines including<br />

erosional, biological and depositional sea level<br />

markers was carried out in the island during three<br />

field campaigns between September 2009 and<br />

October 2010. Although debates in literature<br />

(Kellettat, 2006; Desruelles et al., 2009), in this study<br />

we considered beachrocks as markers of past<br />

relative sea level mainly because of their strict<br />

altimetric correlation with other sea level markers<br />

such as benches, tidal notches or wave cut<br />

platforms. In fact, the combined analysis of erosive<br />

and depositional paleo-shoreline indicators is of<br />

timely importance because, although marine<br />

erosional forms are more precise indicators of sea<br />

level in coastal settings, they rarely preserve<br />

dateable materials, which are more frequent in<br />

depositional landforms (Pirazzoli, 2005; Pirazzoli,<br />

2007).<br />

Markers elevations were measured using a 3 m<br />

metal bar with centimeter sub-division and in-built<br />

spirit level in order to achieve better vertical and<br />

horizontal accuracy. Elevations were measured with<br />

reference to SL at the time of measurement with a<br />

maximum error of 10 cm and the markers location<br />

was measured using a handheld GPS (± 5m).<br />

Underwater transects were carried out up to – 10 m<br />

to map and sample the submerged markers (mainly<br />

notches and beachrocks).<br />

RESULTS<br />

The most evident morphological marker was<br />

represented by a uplifted wave cut platform<br />

developing for about 6 km eastwards and westward<br />

to Agios Fokas cape (Fig. 2).<br />

The platform is continuous (despite the several<br />

changes of lithologies occurring in the area) and<br />

presented an inner margin often characterized by an<br />

abrasion notch with the roof positioned at about + 0.9<br />

m a.s.l. This level was confirmed by the presence of<br />

isolated limestones blocks lying on the platform<br />

showing the presence of lithophaga boreholes up to<br />

0.8 m above the present sea level.<br />

Fig. 2. Uplifted wave cut platforms in the sector of Agios<br />

Fokas and Vrisa. They often showed an inner margin often<br />

characterized by an abrasion notch with the roof at about +<br />

0,9 m a.s.l. as indicated by the dashed line in the photo<br />

above.<br />

On the vertical limestone cliffs, a modern tidal notch<br />

was always present (width 50- 70 cm). Recurrent<br />

uplifted relict of tidal notches were also observed at<br />

different elevation varying from + 0.7 up to +12 m<br />

a.s.l. (Fig.3c) locally interrupted where lithological<br />

conditions are unfavorable for their formation and<br />

preservation (i.e., schists).<br />

248


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Their lateral continuity, until the village of Melinta<br />

was however confirmed by examination by boat.<br />

Morphology and elevation of the notches showed<br />

discrepancies and altimetric correlations has been<br />

very difficult. However, a recurrent morphology was<br />

observed on tidal notches positioned at + 9 to +11 m.<br />

They always presented a double concavity that<br />

appeared smoothed in exposed areas and well<br />

preserved in the sheltered ones (Fig. 3b). In front of<br />

the village of Melinta a large limestone rock<br />

presented an uplifted wave cut platform<br />

characterized by an inner margin at +11.2 m clearly<br />

indicated by the upper limit of a well preserved<br />

lithophaga boreholes band (Fig.3a).<br />

The area of Tarti is characterized by almost vertical<br />

cliffs carved in Triassic limestones. Evidence of<br />

Lithophaga boreholes bands were observed on both<br />

sides of the Tarti cape. The upper limit is located at +<br />

1.8 m. Any evidence of uplifted erosional markers or<br />

modern tidal notches was observed despite the<br />

favorable lithology of the area. On the contrary, a<br />

submerged tidal notch was mapped all along the<br />

cape. The maximum concavity was measured at -0,7<br />

m below the modern sea level (Fig.4).<br />

DISCUSSION<br />

Raised sea stands have been identified along the<br />

whole south eastern coast. Altimetric distribution of<br />

the different markers pointed out some recurrent<br />

levels especially in the areas comprised between<br />

Agios Fokas cape and the village of Melinta.<br />

The first is positioned at + 0.8 ± 0.2 m. According to<br />

Pirazzoli, (2007) and Stiros et al., (2009), the good<br />

preservation of both morphological and biological<br />

markers suggests a very rapid uplift probably related<br />

to a co-seismic event.<br />

The second paleo-shoreline is positioned at + 4.80 ±<br />

0.2 m and, according with the dating, was referred to<br />

MIS 3. The highest recurrent sea level stand is<br />

positioned at 10.9 ± 0.2 m. No radiometric dating was<br />

possible because the lack of datable material.<br />

The geographical limit of these recurrent levels has<br />

been positioned in the village of Plomari. In fact,<br />

eastward to Plomari, although indicators of raised<br />

shorelines were still present, however their altimetric<br />

distribution was significantly different with respect to<br />

the Agios Fokas and Melinta sectors. The site of Tarti<br />

is the only one that showed both evidence of uplift<br />

and subsidence.<br />

The detailed analysis of morphologies and<br />

distribution of the markers, coupled with radiometric<br />

dating, allowed to identify a Late Quaternary regional<br />

uplift trend controlled by the footwall of a large<br />

normal fault (Lesvos Fault) located between Lesvos<br />

and Chios Islands (Mascle & Martin, 1990) (Fig.1).<br />

Fig.3. Raised shorelines mapped in the area of Kryfty and<br />

Melinta. a) wave cut platform with inner limit positioned at +<br />

11.2 and characterized by lithofaga boreholes. b) double<br />

concavity notch occurring in Kryfty cove. c) raised tidal<br />

notches carved in the Triassic limestones of Kryfty area.<br />

The maximum concavity is at + 0.8 m.<br />

Uplifted beachrocks outcrops were observed along<br />

the whole SE sector of the island. The beachrocks<br />

are often organized in multiple slabs reaching up to<br />

about + 3.1 m a.s.l. The only exception is<br />

represented by Tarti site where beachrocks<br />

developed from – 1.1 m up to intertidal zone.<br />

249<br />

Fig.4 . Underwater tidal notch occurring all along the Tarti<br />

cape.<br />

At local scale, superimposed to this trend, the area is<br />

characterized by many indicators of rapid vertical<br />

displacements (uplift and subsidence) related to the<br />

high seismicity. According to the radiomentric dating<br />

and geomorphological evidence,an important coseismic<br />

event has been found to have happened<br />

between 3365 and 3924 BP. This rapid event uplifted<br />

of about 0.8 m a large sector of the southern Lesvos<br />

coastline.<br />

According to Lambeck (1996) and Siddal et al.,<br />

(2008), the position of the MIS 3 shorelines at about<br />

+ 5 m on the present position allowed to quantify the


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

rate of uplift during the late quaternary. However, as<br />

pointed out by several indicators, the average rate is<br />

a sum of gradual and co-seismic uplifts and more<br />

accurate consideration strongly needs further<br />

radiometric dating. If further studies will confirm our<br />

data, Lesvos Fault is among one of the less<br />

documented active fault in this sector of the Aegean<br />

Sea to have ruptured in recurring earthquake within a<br />

relatively recent timespan.<br />

In conclusion, geomorphological markers of past sea<br />

levels are well known for being one of the main<br />

means to estimate neotectonic trends and historical<br />

recurrence and magnitude of co-seismic events. The<br />

scientific value of this kind of studies is flanked by the<br />

value assumed for society. The latter is twofold: on<br />

one side, very active sites like Lesvos Island are a<br />

potential source of data to tune global sea level<br />

models and therefore global estimates of future sea<br />

level rise; on the other side, they have a local value<br />

in terms of earthquake risk assessment.<br />

References<br />

Cundy, A.B., Gaki-Papanastassiou, K., Papanastassiou, D.<br />

Maroukian, H., Frogley, M.R., Cane, T.(2010) Geological<br />

and geomorphological evidence of recent coastal uplift<br />

along a major Hellenic normal fault system (the Kamena<br />

Vourla fault zone, NW Evoikos Gulf, Greece) Marine<br />

Geology 271, 156–164<br />

Desruelles, S., Fouache, É., Ciner, A. Dalongeville, R.,<br />

Pavlopoulos, K., Kosun, E., Coquinot, Y., Potdevin, J.L.<br />

(2009). Beachrocks and sea level changes since Middle<br />

Holocene: Comparison between the insular group of<br />

Mykonos–Delos–Rhenia (Cyclades, Greece) and the<br />

southern coast of Turkey. Global and Planetary Change<br />

Volume 66, 1-2, 19-33.<br />

Kellettat, D. 2006. Beachrock as Sea-Level Indicator?<br />

Remarks from a Geomorphological Point of View. Journal<br />

of Coastal Research 22 (6), 1558-1564.<br />

Mascle, J. ,Martin, L. (1990). Shallow structure and recent<br />

evolution of the Aegean Sea: a synthesis based on<br />

continuous reflection profiles. Marine Geology 94, 271–<br />

299.<br />

Lambeck, K. (1996) – Sea-level changes and shore-line<br />

evolution in Aegean Greece since Upper Paleolithic time.<br />

Antiquity, 70, 588-610.<br />

Palyvos, N., Lemeille, F., Sorel, D., Pantosti, D.<br />

Pavlopoulos, K. (2008). Geomorphic and biological<br />

indicators of paleoseismicity and Holocene uplift rate at a<br />

coastal normal fault footwall (western Corinth Gulf,<br />

Greece). Geomorphology 96, 16–38<br />

Papazachos, C.B. and Kiratzi, A.A., (1996). A detailed study<br />

of the active crustal deformation in the Aegean and<br />

surrounding area, Tectonophysics 253(1-2), 129-153.<br />

Papazachos, B., Papazachou, C., 1997. The earthquakes of<br />

Greece, Ziti editions. Thessaloniki<br />

Pirazzoli, P.A. (2005). A review of possible eustatic,isostatic<br />

and tectonic contributions in eight late-Holocene relative<br />

sea-level histories from the Mediterranean area.<br />

Quaternary Science Review 24, 1989–2001<br />

Pirazzoli, P.A., Stiros S.C., Fontugne M., Arnold M. (2004).<br />

Holocene and Quaternary uplift in the central part of the<br />

southern coast of the Corinth Gulf (Greece). Marine<br />

Geology 212, 35-44<br />

Pirazzoli, (2007). Sea level studies. Geomorphological<br />

Indicators. Encyclopedia of Quaternary Science, 2974-<br />

2983.<br />

Roumelioti, Z., Kiratzi, A., Benetatos, C., (2011). Time<br />

Domain Moment Tensors of earthquakes in the broader<br />

Aegean Sea for the years 2006-2007: the database of the<br />

Aristotle University of Thessaloniki, Journal of<br />

Geodynamics 51, 179-189.<br />

Siddal, M., Rohling, E.J., Thompson, W.G., Waelbroeck, C.<br />

(2008). Marine isotope stage 3 sea level fluctuation: data<br />

synthesis and new outlook. Reviews of Geophysics 46, 1-<br />

29.<br />

Stewart, I.S. Morhange, C. (2009), Coastal geomorphology<br />

and sea-level change, in J. C.Woodward (ed.), The<br />

Physical Geography of the Mediterranean. Oxford<br />

University Press, Oxford, 385–413<br />

Stiros, S.C., Laborel, J., Laborel-Deguen, F., Papageorgiou<br />

S., Evind J., Pirazzoli P.A.. 2000. Seismic coastal uplift in<br />

a region of subsidence: Holocene raised shorelines of<br />

Samos Island, Aegean Sea, Greece. Marine Geology<br />

170, 41-58<br />

Stiros, S.C. Pirazzoli, P.A. Fontugne, M., (2009). New<br />

evidence of Holocene coastal uplift in the Strophades<br />

Islets (W Hellenic Arc, Greece). Marine Geology 267,<br />

207–211.<br />

Vousdoukas, M.I., Velegrakis, A.F., Karambas, T.V. (2009).<br />

Morphology and sedimentology of a microtidal beach with<br />

beachrocks: Vatera, Lesbos, NE Mediterranean.<br />

Continental Shelf Research 29, 1937–1947.<br />

250


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

BOULDER DEPOSITS IN SOUTHERN LESVOS: AN EVIDENCE OF THE 1949’S CHIOS-<br />

KARABURUM TSUNAMI?<br />

Vacchi, Matteo (1), Alessio Rovere(1), Marco Firpo (1) and Nickolas Zouros (2)<br />

(1) Department for the Study of the Territory and its Resources, University of Genova, Genova. (Italy)<br />

(2) SEAMap srl, Seascape Evaluation Assessment and Mapping, Genova (Italy)<br />

(3) Department of Geography, University of Aegean, Mytylene, Lesvos Island (Greece)<br />

Abstract (Boulder deposits in southern Lesvos: an evidence of the 1949’s Chios-Karaburum tsunami?): Aim of this study<br />

was to understand the mechanism of deposition of clusters of large boulders, consisting of beachrock slabs, which were found on<br />

the southern coasts of Lesvos Island (NE Aegean Sea). An integrated study was carried out in order to detect whether transport<br />

and accumulation can be related to an exceptional storm or a tsunami event. The results provided evidence of a tsunami wave<br />

impact during the last century. Specifically, the boulder accumulation is likely to have been deposited by the tsunami that was<br />

triggered by the 6.7M Chios-Karaburum earthquake of 1949. Lesvos island, on the basis of the seismic sources, was already listed<br />

as tsunami affected area but field evidence was previously lacking. The results represent a new contribution that could be of<br />

primary importance in assessing the degree of vulnerability of the local coastal communities of southern Lesvos to future tsunami<br />

events.<br />

Key words: boulder deposits; catastrophic waves; Lesvos Island; coastal hazard.<br />

INTRODUCTION<br />

Identification of boulders displaced and eventually<br />

transported by tsunami or exceptional storm waves<br />

plays a crucial role in the assessment of the<br />

occurrence of past catastrophic events (Goto et al.,<br />

2009). Understanding the nature and impact of these<br />

events is fundamental in terms of evaluation,<br />

mitigation and management of current coastal<br />

hazards. In the Mediterranean, historical tsunamis<br />

were often reported as consequences of destructive<br />

earthquakes (Soloviev et al., 2000). In the Ionian and<br />

Aegean Seas, geomorphological records of tsunamis<br />

were detected in several areas (Mastronuzzi and<br />

Sansò., 2000; Scheffers and Kellettat, 2003;<br />

Scheffers and Scheffers, 2007, Scicchitano et al.,<br />

2007; Scheffers et al., 2008). In particular, the area<br />

surrounding western Turkey and Greece is among<br />

the most seismically active and rapidly deforming<br />

regions in the world (Nyst and Tatcher, 2004), and<br />

thus has historically been strongly affected by<br />

tsunami events (Papadopoulos & Chalkis, 1894;<br />

Soloviev, 1990; Soloviev et al., 2000,Fig. 1).<br />

The aim of this study is to understand the mechanism<br />

of deposition of clusters of large boulders (weighing<br />

up to 17 tons) which were found on the southern<br />

coasts of Lesvos Island (NE Aegean Sea). In<br />

particular, studies of boulder morphology, tectonic<br />

setting, wave climate and historical context were<br />

carried out in order to detect whether transport and<br />

accumulation can be related to an exceptional storm<br />

or a tsunami event.<br />

The study area (Fig. 1) is located in the southern part<br />

of the island, near the villages of Plomari and Agios<br />

Isidoros. From a geomorphological point of view, this<br />

area is characterized by cliffs alternating with sandy<br />

251<br />

to gravel beaches, often with beachrock outcrops.<br />

The main lithologies are represented by Triassic<br />

schists and limestones. Due to the southern facing,<br />

this costal sector is characterized by a relatively low<br />

wave regime because of the protection from the main<br />

swells of the Aegean sea generated by northern<br />

winds (Soukissian et al., 2007). The maximum fetch<br />

does not exceed 100 nautical miles, and main waves<br />

(coming from SE) reach maximum off-shore wave<br />

height of about 1,8 m (Vousdoukas et al., 2009).<br />

Fig. 1. Geographical location and seismo-tectonic setting of<br />

the study area. Focal mechanism of the > 4M earthquakes<br />

after the 1928. The white circle indicates the 1949’s M. 6.7<br />

event triggering a tsunami wave recorded in Chios and in<br />

the Karaburun peninsula. Black dots in the small square<br />

represent the historical tsunamigenic earthquakes<br />

according to Soloviev, 1990.<br />

A total of 47 boulders was found in the study area.<br />

Some of them are organized in clusters, others are<br />

scattered along the shoreline (Fig. 2). After a first<br />

general mapping, more detailed measures


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

concentrated on the 26 boulders showing a major<br />

axis 1 m. The boulders consisted of beachrock<br />

slabs (Fig. 2) of which having unit weight of 2.6 tm -3 .<br />

The boulders had a maximum size of 4.5 × 2.5 × 0.6<br />

m with a volume of about 7 m 3 and a maximum<br />

weight of about 17 t. Some peculiar features are<br />

constant on the majority of the beachrock boulders:<br />

i) the boulders are scattered up to about 20 m from<br />

the shoreline and their elevation does not exceed 0.5<br />

m asl;<br />

ii) many boulders are upside down, as proved by<br />

erosive features normally cut upon the upper surface<br />

of a beachrock and presently facing the ground (Fig.<br />

2);<br />

iii) Several boulders, especially in the area of Agios<br />

Isidoros, are completely buried in the sand (Fig. 2).<br />

The presence of biogenic encrustations (mainly<br />

vermetids and serpulids) suggests a mid-sublittoral<br />

pre-transport position of the boulders (as these<br />

organisms usually live within this environment,<br />

Laborel, 1987).<br />

These settings were validated by the transects<br />

carried out along the whole seaward extension of the<br />

beachrock outcrop: the high degree of fracturing<br />

coupled with the presence of scattered broken pieces<br />

mainly at the seaward edge of the beachrock outcrop<br />

(about -3 m) confirmed the hypothesis of an original<br />

submerged position or a submerged joint bounded<br />

scenario. In fact, geometrical analysis of the boulders<br />

and of large holes in the submerged part of the<br />

beachrock itself revealed a close correspondence<br />

between the shapes of the boulders and the shapes<br />

of the holes. On the basis of the pre-transport<br />

settings individuated, values of storm wave height<br />

(Hs) and tsunami wave height (Ht) theoretically<br />

required for the boulder displacement were<br />

calculated using the equations proposed by Nott,<br />

(2003), Pignatelli et al., (2009), Benner et al., (2010).<br />

Fig. 2 a) and b) examples of large boulders mapped in the<br />

site of Plomari; c) and d) examples of large boulders<br />

mapped in the site of Agios Isidoros<br />

DISCUSSION<br />

Significant discrepancies were observed between the<br />

different hydrodynamic approaches. For the<br />

submerged boulders (SMBS), the Nott equation<br />

calculated storm wave values (Hs) up to 15 m to<br />

252<br />

displace the largest boulder. The maximum storm<br />

wave values (Hs) computed using the Benner et al.<br />

approach are considerably smaller, slightly<br />

exceeding 8 m.<br />

Nott’s joint bounded equation (JBBS) computed<br />

storm waves values (Hs) exceeding 20 m whereas<br />

storm wave values derived from Pignatelli et al.<br />

equations reached maximum values of about 10 m.<br />

The Aegean Sea is characterized by relatively short<br />

fetch and relatively small swells, mainly generated<br />

from northern winds (Soukissian et al., 2007). It is<br />

very unlikely that waves exceeding 10 m could be<br />

generated in this sector of the Aegean, especially in<br />

a south-facing area, characterized by a maximum<br />

fetch not exceeding 100 nautical miles and by<br />

maximum off-shore significant wave height (Hs) not<br />

exceeding 2 m.<br />

These hydrodynamic results suggest that a tsunami<br />

could have been responsible for on-shore deposition<br />

of the boulders, but the discrepancies between the<br />

different approaches did not provide unambiguous<br />

evidence. Further analyses were then performed in<br />

order to test the reliability the tsunami hypothesis.<br />

Some considerations based both on<br />

geomorphological indicators and on seismotectonic<br />

sources, support the hypothesis that the boulder<br />

accumulation in southern Lesvos was caused by<br />

tsunami.<br />

The morphology and lithostructural settings of the<br />

coastal area play a crucial role in boulder detachment<br />

and transport by catastrophic waves. Mastronuzzi et<br />

al., (2006) indicated the presence of layered units<br />

and bedrock fracturing as important pre-conditions<br />

for boulder displacement. Coastal sectors<br />

characterized by beachrock outcrops and affected by<br />

tsunami wave impact often showed on-shore broken<br />

slab accumulation (Vött et al., 2007, 2009; Scheffers<br />

and Scheffers, 2007). In the study area, the slabs<br />

were probably torn out of the original beachrock unit<br />

which was already fractured by several seismic<br />

events affecting the whole coastal sector. Moreover,<br />

on the majority of boulders, a fragile layer of biogenic<br />

encrustation was observed. Its preservation is a clear<br />

indicator of short transport generated by a single<br />

wave (Mastronuzzi et al., 2006).<br />

Other evidence can be gathered by subdividing the<br />

a-axis orientation of the boulders on the basis of the<br />

relative Hs values. Most of the boulders presenting<br />

Hs values exceeding 7,5 m (i.e. not compatible with<br />

calculated extreme storm events) had their elongated<br />

axis almost perpendicular to the shoreline and<br />

oriented between 170°N and 200°N. However, other<br />

boulders are oriented mainly between 130° and<br />

150°N. This is the direction of the major swells in the<br />

area (Vousdoukas et al., 2009). This orientation<br />

pattern reflects a scheme already present in other<br />

Mediterranean boulder accumulations (Mastronuzzi<br />

and Sansò, 2004; Scicchitano et al., 2007) and was<br />

explained with re-orientation by storm waves after a<br />

tsunami event. According to the previous


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

considerations, we hypothesize that a single tsunami<br />

wave displaced all the boulders from the submerged<br />

position to the shore. Subsequent storm wave<br />

events, coming from S-SE, reworking the deposited<br />

boulders were able to re-orient only the smaller<br />

ones.Further confirmation was provided by the<br />

literature dealing with Mediterranean regions most<br />

frequently affected by tsunamis (Papadopoulos and<br />

Chalkis, 1984; Altinok and Ersoy, 2000; Soloviev et<br />

al., 2000; Papadopoulos and Foakefs, 2005; Yolsal<br />

et al., 2007; Papadopoulos, 2009). These papers<br />

provided spatial distribution of tsunami hazard in the<br />

Aegean area based on the historical occurrence of<br />

tsunami events as well as on the triggering seismic<br />

sources (Papadopoulos and Chalkis, 1984;<br />

Papazachos and Dimitriou, 1991; Papadopoulos and<br />

Foakefs, 2005). On the basis of the distribution of<br />

tsunamigenic earthquakes and of the historical data<br />

of past events, several authors listed Southern<br />

Lesvos as an area vulnerable to tsunami events<br />

(Fig.3) (Papadopoulos and Chalkis, 1984; Soloviev et<br />

al., 1990 Papadopoulos and Foakefs, 2005; Yolsal et<br />

al., 2007). This integrated study, ranging from<br />

geomorphological to seismotectonic data,<br />

corroborated the tsunami hypothesis and most<br />

probably ruled out extreme storms as the cause of<br />

transport and displacement of the boulders.<br />

Fig. 3 Maps of the tsunami affected area compiled by<br />

Papadopoulos & Foakefs, 2005; Papadopoulos & Chalkis,<br />

1984; Yolsal et al., 2007.<br />

The high seismic activity affecting this sector of the<br />

Aegean Sea often triggered tsunami waves that<br />

caused damage to the surrounding coastal areas<br />

(Papazachos and Papazachou, 1997; Altinok et al.,<br />

2005). Historical data are available on six main<br />

events which took place 20 th March 1389, 24 th<br />

November 1772, 13 th November 1856, 19 th January<br />

1866, 3 rd April 1881 and 23 rd July 1949 (Soloviev et<br />

al., 2000; Altinok et al., 2005).<br />

Radiocarbon age determinations were performed on<br />

two samples of biological encrustation, the first,<br />

(serpulids) from the Plomari cluster and the second<br />

(vermetids) from Agios Isidoros cluster. AMS<br />

radiocarbon age determination indicated that both<br />

were of recent origin (fraction of modern carbon,<br />

~1950). Because of this, it was possible to carry out<br />

an historical investigation with the help of the cultural<br />

association “to polion” of Plomari. Aged inhabitants of<br />

the village confirmed the sudden appearance of the<br />

boulders on the shoreline but the people who were<br />

interviewed could not recall precise dates. Important<br />

data were achieved by using an historical photograph<br />

of Plomari taken in 1896 (Fig. 4). In the same area<br />

where boulders are presently found the photograph<br />

shows no traces dislocated blocks.<br />

This photograph allows us to place a further temporal<br />

limit on the date of the boulders deposition. In 1896<br />

there is no evidence of broken beachrock slabs on<br />

the shoreline. This is consistent with the theory of a<br />

single pulse wave as the depositional mechanism as<br />

opposed to a continuous action of the waves on the<br />

beachrock outcrops. These historical data supported<br />

the results obtained by radiometric dating, indicating<br />

a tsunami event that cannot be earlier than the year<br />

1896.<br />

253<br />

Fig. 4 Plomari today and in the 1896. Boulder deposits are<br />

missing in the 1896’s photograph.<br />

Analysis of the historical tsunami catalogs isolated<br />

the Chios-Karaburum earthquake of 1949 as the only<br />

known event capable of creating the boulder deposit.<br />

The epicenter of the earthquake (M=6.7 Papazachos<br />

and Papazachou, 1997) was located offshore of<br />

Chios Island (Fig.1) at about 40 km from Plomari.<br />

The tsunami triggered by the earthquake affected the<br />

coasts of Chios Island, and Karaburum historical<br />

reports are available in both sectors (Altinok et al.,<br />

2005).<br />

This paper presents results of geomorphologic traces<br />

of catastrophic wave impact in a coastal sector


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

included in the lists of areas affected by tsunami, but<br />

where field evidence of paleo-tsunamis was<br />

previously lacking. The enigmatic large boulders<br />

observable on the southern coast of Lesvos Island<br />

were investigated without excluding, a priori, any<br />

possible cause of deposition. The integrated study of<br />

geomorphological, hydrodynamic, seismotectonic<br />

and historical data provided evidence of a tsunami<br />

wave impact during the last century. Specifically, the<br />

boulder accumulation is likely to have been deposited<br />

by the tsunami that was triggered by the 6.7M Chios-<br />

Karaburum earthquake of 1949.<br />

Plomari (founded in 1849) home to about 6000<br />

people, is the second largest town of Lesvos Island<br />

and is the only sizable settlement on the southern<br />

coast of Lesvos. Its geographical position justifies the<br />

creation of serious policies related to seismic hazard.<br />

In fact, Papazachos et al. (2004) indicated the<br />

possibility of a strong seismic event in the near future<br />

(up to M 6.5–6.7) in the area of Lesvos and Chios.<br />

The above results represent a new contribution that<br />

could be of primary importance in assessing the<br />

degree of vulnerability of the local coastal<br />

communities to future tsunami events.<br />

Acknowledgements: The authors would like to thank Prof.<br />

G. Mastronuzzi (Bari, IT) for his suggestions and to C.N.<br />

Bianchi (Genoa, IT) for the identification of biological marine<br />

encrustation.<br />

References<br />

Altinok, Y., Ersoy., S. (2000). Tsunamis Observed on and<br />

Near the Turkish Coast. Natural Hazards 21, 185–205.<br />

Altinok, Y., Alpar, B., Ozer, N., Gazioglu, C. 2005. 1881 and<br />

1949 earthquakes at the Chios-Cesme Strait (Aegean<br />

Sea) and their relation to tsunamis. Natural Hazard and<br />

Earth System Science, 5, 717–725.<br />

Benner, R.; Browne, T.;Brückner, H.; Kelletat, D.; Scheffers,<br />

A. (2010). Boulder Transport by Waves: Progress in<br />

Physical Modelling. Zeitschrift für Geomorphologie NF.<br />

Supp., 54, (3), 127-146.<br />

Goto, K., Okada, K., Imamura, F., (2009). Characteristics<br />

and hydrodynamics of boulders transported by storm<br />

waves at Kudaka Island, Japan. Marine Geology 262,<br />

14–24<br />

Laborel J, 1987. Marine biogenic constructions in the<br />

Mediterranean—A review. Scientific report of Port-Cros<br />

Natural. Park (France) 13: 97–126.<br />

Mastronuzzi, G., Pignatelli, C., Sanso, P., (2006). Boulder<br />

fields: a valuablemorphological indicator of Paleotsunami<br />

in the Mediterranean Sea. Zeitschrift für Geomorphologie,<br />

NF 146, 173–194<br />

Mastronuzzi, G., Sanso, P., (2000). Boulder transport by<br />

catastrophic waves along the Ionian coast of Apulia<br />

(Southern Italy). Marine Geology 170, 93–103.<br />

Mastronuzzi, G., Sansò, P., (2004). Large boulder<br />

accumulations by extreme waves along the Adriatic<br />

Coast of southern Apulia (Italy). Quaternary International<br />

120, 173–184.<br />

Nyst, M., Thatcher, W. (2004). New constraints on the<br />

active tectonic deformation of the Aegean. Journal of<br />

Geophysical Resources 109, B11406,<br />

doi:10.1029/2003JB002830.<br />

Nott, J., (2003). Tsunami or storm waves?—Determining<br />

the origin of a spectacular field of wave emplaced<br />

boulders using numerical storm surge and wave models<br />

and hydrodynamic transport equations. Journal of<br />

Coastal Research 19, 348–356.<br />

Papadopoulos, G. A. (2009). Tsunamis, in J. C. Woodward<br />

(ed.), The Physical Geography of the Mediterranean.<br />

Oxford University Press, Oxford, 493–512.<br />

Papadopoulos, G. A., Fokaefs, A. (2005). Strong tsunamis<br />

in the Mediterranean Sea: a re-evaluation. ISET Journal<br />

of Earthquake Technology, 42, 159–170.<br />

Papadopoulos, G. A., Imamura, F.: (2001). A proposal for a<br />

new tsunami intensity scale. ITS 2001 Proc., Session 5,<br />

N. 5-1, 569–577<br />

Papadopoulos, G.A., Chalkis, B.J., (1984). Tsunami<br />

observed in Greece and the surrounding area from<br />

antiquity up to present times. Marine Geology 56, 309–<br />

317<br />

Papazachos B. C., Dimitriou P. P., (1991). Tsunamis In and<br />

Near Greece and Their Relation to the Earthquake Focal<br />

Mechanisms. Natural Hazards 4, 161-170.<br />

Papazachos, B., Papazachou, C., (1997). The earthquakes<br />

of Greece, Ziti editions. Thessaloniki<br />

Papazachos, C.B., Karakaisis, G.F., Scordilis, E.M., and<br />

Papazachos, B.C., (2004). Probabilities of activation of<br />

seismic faults in critical regions of the Aegean area,<br />

Geophysical Journal International 159, 679–687.<br />

Pignatelli, C., Sansò, P., Mastronuzzi, G. (2009). Evaluation<br />

of tsunami flooding using geomorphologic evidence.<br />

Marine Geology 260, 6–18.<br />

Scheffers, A., Kelletat, D., (2003). Sedimentologic and<br />

geomorphic tsunami imprints worldwide—a review. Earth-<br />

Science Reviews 63, 83–92.<br />

Scheffers, A., Scheffers, S.R., (2007). Tsunami deposits on<br />

the coastline of West Crete (Greece). Earth and<br />

Planetary Science Letters 259, 613–624.<br />

Scheffers A., Kelletat, D., Vött A., May, S.M., Scheffers,<br />

S.R., (2008). Late Holocene tsunami traces on the<br />

western and southern coastlines of the Peloponnesus<br />

(Greece). Earth and Planetary Science Letters 269, 271-<br />

279.<br />

Scicchitano, G., Monaco, C., Tortorici, L., (2007). Large<br />

boulder deposits by tsunami waves along the ionian<br />

coast of south-eastern Sicily (Italy). Marine Geology 238,<br />

75–91.<br />

Soloviev, S.L., (1990). Tsunamigenic zones in the<br />

Mediterranean Sea. Natural Hazards 3, 183–202.<br />

Soloviev, S.L., Solovieva, O.N., Go, C.N., Kim, C.S.,<br />

Shchetnikov, N.A., (2000). Tsunamis in the<br />

Mediterranean Sea 2000 B.C.-2000 A.D. Kluwer,<br />

Dordrecht.<br />

Soukissian, T.H., Hatzinaki, M., Korres, G.,, Papadopoulos,<br />

A., Kallos, G., Anadranistakis, E. (2007). Wind wave atlas<br />

of the Hellenic seas. Hcmr publication, 300 pp.<br />

Vött, A., Brückner, H., May, M., Lang, F., Herd, R.,<br />

Brockmüller, S., (2007). Strong tsunami impact on the<br />

Bay of Aghios Nikolaos and its environs (NW Greece)<br />

during classical-Hellenistic times. Quaternary<br />

International 181, (1), 105-122<br />

Vött, A., Brückner, H., Brockmüller, S., Handl, M., May,<br />

S.M., Gaki-Papanastassiou, K., Herd, R., Lang, F.,<br />

Maroukian, H., Nelle, O,. Papanastassiou, D. (2009).<br />

Traces of Holocene tsunamis across the Sound of<br />

Lefkada, NW Greece. Global and Planetary Change 66,<br />

112-128.<br />

Vousdoukas, M.I., Velegrakis, A.F., Dimou, K., Zervakis V.,<br />

Conley, D.C. (2009). Wave run-up observations in<br />

microtidal, sediment-starved pocket beaches of the<br />

Eastern Mediterranean. Journal of Marine Systems 78,<br />

537–547.<br />

Yolsal, S., Taymaz, T., Yalciner, A., C. (2007).<br />

Understanding tsunamis, potential source regions and<br />

tsunami-prone mechanisms in the Eastern<br />

Mediterranean.In: Taymaz, T., Yilmaz, Y. Dilek, Y. (eds)<br />

The geodynamics of the Aegean and Anatolia. Geological<br />

Society, London, Special Publications, 291, 201–230.<br />

254


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

ACTIVE FAULTING AND EARTHQUAKE-INDUCED SLOPE FAILURES IN<br />

ARCHEOLOGICAL SITES: CASE STUDY OF DELPHI, GREECE<br />

Valkaniotis, Sotiris (1, George Papathanassiou, Spyros Pavlides<br />

(1) Dept of Geology, Aristotle University of Thessaloniki, Greece, email gpapatha@auth.gr<br />

Abstract (Active faulting and earthquake-induced slope failures in arcacheological sites: case study of Delphi, Greece):<br />

The archeological area of Delphi is one of the most visited places in Greece. However, as it is described in historical reports,<br />

severe ground deformations triggered by earthquakes took place. In order to assess the rockfall potential and the relevant hazard<br />

within this area, the parameters of the 1870 earthquake were taken into account. Afterwards, the GIS-based Newmark’s<br />

displacement method was applied in order to compute the permanent displacement values in the study area. The outcome<br />

provided by this study, indicates that the zone located at the archaeological area is characterized as very unstable. Moreover, the<br />

value of the critical acceleration within this zone is very low and even a small magnitude seismic event could generate rockfalls.<br />

Thus, the hazard and consequently the risk is extremely high and permanent mitigation measures should be developed.<br />

Key words: Delphi, earthquake, rockfall, hazard<br />

INTRODUCTION-GEOLOGICAL SETTING<br />

The WNW-ESE-trending Gulf of Corinth separates<br />

the northern coast of the Peloponnese from mainland<br />

Greece (Figure 1). To the east it is bounded by the<br />

Isthmus of Corinth through which the Corinth Canal<br />

passes into the Saronic Gulf. The Gulf of Corinth cuts<br />

across the NNW-SSE regional trend of the Hellenic<br />

mountain chain and overlies the Pindus, Parnassos<br />

and Othris zones of the internal Hellenides. This has<br />

resulted in a NNW-SSE and NW-SE structural grain<br />

to the area, which may have had some control on the<br />

later extensional deformation (McKenzie, 1972; 1978;<br />

Papazachos & Papadopoulos, 1977; Makris, 1976;<br />

1978; Le-Pichon & Angelier, 1979; Angelier et al.,<br />

1982; Papadopoulos et al., 1986; Doutsos et al.,<br />

1988; Jackson & McKenzie, 1988). In the northern<br />

part of the Gulf of Corinth Rift, the mountain chains of<br />

Elikonas, Parnassos – Giona and part of Pindos are<br />

situated.<br />

DELPHI-ARAHOVA-AMFISSA FAULT ZONE<br />

One of the most important and impressive fault zones<br />

of the Central Greece area is the Amfissa – Delphi –<br />

Arahova Fault Zone (Pèchoux, 1977; Sebrier, 1977;<br />

de Boer & Hale, 2000). Detailed field mapping along<br />

with a wealth of historical documents and references<br />

for strong earthquakes back to 373 BC unveil an<br />

important and complex active fault structure<br />

(Piccardi, 2000; Valkaniotis, 2009). The fault zone<br />

extends along the southern rims of Mt Parnassos for<br />

a total length of more than 25 km and terminates to<br />

the west in Mt Giona over the plain of Amfissa<br />

(Figure 2). Normal- to oblique-slip displacement in<br />

the fault zone (Delphi – Arahova faults) is undergoing<br />

a intriguing transition to strike-slip displacement in<br />

the western part (Agia Efthimia fault) documented<br />

during field mapping (Valkaniotis, 2009). The<br />

mountainous area between Kifissos Basin and<br />

Amfissa – Delphi – Arahova Fault zone contains<br />

smaller fault structures following pre-existing alpine<br />

fault zones reactivated in the present extensional<br />

stress regime.<br />

255<br />

Fig. 1: Active and possible active faults in Central Greece.<br />

Fault data from Pavlides et al. (2007, 2008)<br />

Delphi fault is a normal-oblique normal fault, with<br />

dipping to the south and strike direction NW-SE to<br />

WNW-ESE for the western part and W-E for the<br />

central part (Figure 2). Bedrock lithology is compiled<br />

by alpine (Jurassic-Creataceous) limestones of<br />

Parnassos-Giona geotectonic unit. In the hangingwall,<br />

these limestones are overlain by flysch<br />

sediments, and a thick unit of Quaternary consistent<br />

and loose breccias and scree.<br />

The Delphi fault cuts through the archaeological site<br />

of the Delphi Oracle. Clearence of recent (historical)<br />

depostis in the site due to archaeological excavations<br />

provide a detailed examination of the fault zone<br />

structure and earthquake fault displacement. The<br />

fault zone, reaching a thickness of 500-2000 m in<br />

Delphi-Arahova valley, consists of sub-parallel fault<br />

planes, with a mean dip of 70-80 0 , converging in a<br />

depth of ~2 km with the main fault surface of 60 0 dip<br />

(Valkaniotis, 2009).<br />

The entire set of monuments that constitute the<br />

archaeological area of Delphi (Stadium, Sanctuary of<br />

Apollo, Castalia spring, High School, Temple of<br />

Athena) are placed in front of the main surface of the<br />

fault at Delphi, which forms a striking morphological<br />

vertical scarp. The archaeological site of Delphi is<br />

situated in an active scree deposition area, with


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

continuous rock and debris due to the high relief and<br />

tectonic weakening of the bedrock. Severe damages<br />

and surface ruptures are reported in Delphi (Piccardi,<br />

2000; Papadopoulos, 2000; Valkaniotis, 2009) for<br />

numerous historical earthquakes (373BC, 278BC,<br />

326AD, 552AD, 1870AD). Rock falls in the vicinity of<br />

Delphi are reported during numerous other smalllarge<br />

earthquakes in the broader area. The last<br />

strong earthquake in the Delphi-Arahova-Amfissa<br />

fault zone was the 1870 event, with a magnitude of<br />

6.8 and a possible rupture of the whole fault zone<br />

(Schmidt, 1879; Ambraseys & Pantelopoulos, 1989;<br />

Papadopoulos, 2000).<br />

model of an infinite slope in material having both<br />

frictional and cohesive strength and is given by:<br />

c ' tan ' m<br />

tan '<br />

w<br />

FS (2)<br />

tsin a tan a tan a<br />

where ’ is the effective friction angle, c’ is the<br />

effective cohesion, is the slope angle, is the<br />

material unit weight, w is the unit weight of water, t is<br />

the thickness of the mass at right angles to the slope<br />

and m is the proportion of the slab thickness that is<br />

saturated.<br />

A Newmark analysis can be extended to regional<br />

analysis using GIS software (Miles & Keefer, 2000),<br />

Arcinfo, by applying equations 1 and 2 to raster data<br />

layers created for each input variable.<br />

APPLYING THE NEWMARK APPROACH TO<br />

DELPHI AREA<br />

Fig. 2: The Delphi-Arahova-Amfissa Fault Zone. Fault plane<br />

projections for western and central-east part. After<br />

Valkaniotis (2009).<br />

EVALUATING THE EARTHQUAKE-INDUCED<br />

LANDSLIDE HAZARD<br />

The potential of a slope failure triggered by an<br />

earthquake, can be evaluated by three main<br />

methods; pseudostatic, permanent or statistical<br />

analysis (Miles & Keefer, 2000). The latter accesses<br />

the landslide hazard by assuming the past predicts<br />

the future and the pseudostatic employs a static<br />

slope stability analysis with the addition of a<br />

horizontal force modeling the earthquake ground<br />

motion (Miles & Keefer, 2009). The second analysis,<br />

proposed by Newmark (1965), provides information<br />

regarding actual slope stability based on accepted<br />

characterization of the severity of the earthquake<br />

shaking (Miles & Keefer, 2009).<br />

The development of a Newmark analysis requires the<br />

evaluation of the parameters of the expected<br />

earthquake shaking and the capability of the<br />

geological unit to resist this dynamic effect. The latter<br />

parameter is quantified as the critical acceleration<br />

(a c ), a threshold ground acceleration necessary to<br />

overcome basal sliding resistance and initiate<br />

permanent downslope movement (Jibson, 2007).<br />

The computation of critical acceleration is based on<br />

the following equation proposed by Newmark, (1965)<br />

a ( FS 1) gsina<br />

(1)<br />

c<br />

where FS is the factor of safety, is the angle of the<br />

sliding surface (slope angle), g is the acceleration of<br />

gravity.<br />

According to Jibson et al. (1998), the factor of safety<br />

is evaluated using a relatively simple limit-equilibrium<br />

256<br />

The first step in the Newmark approach is the<br />

calculation of the factor of safety of the slope using<br />

equation 2. Therefore, the strength parameters of the<br />

geological units in the specific area should be<br />

evaluated. In our study, for the heavily jointed and<br />

weathered limestone in the study area, mean values<br />

of 30 o (angle of friction) and 15 kPa (cohesion) were<br />

employed and 30 o and 10 kPa for the formation of<br />

flysch, respectively, based on literature review.<br />

Information from the literature was used to estimate<br />

the thickness of the failed material parallel to the<br />

slope. Khazai & Sitar (2000) proposed a correlation<br />

between slope angle and the thickness of the failed<br />

mass in which a slope angle between 40-60 o<br />

corresponds to a thickness of 1m; this was adopted<br />

in the present study.<br />

Finally, the value of the topographic slope angle <br />

(figure 3) was obtained from the DEM of a 30 m grid<br />

prepared from contour lines on the 1:5000-scale<br />

topographic maps using ArcInfo Software. In the<br />

study area the slope angle varies from 0 o to 60 o ,<br />

however in this research we took into consideration<br />

only the areas where slope angle >20 o .<br />

Fig. 3: Slope map of the study area.<br />

Having computed the factor of safety, the next step<br />

was the estimation of the critical acceleration using<br />

equation 1. However, it was first necessary to modify<br />

estimated values of Fs


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

et al., 1998). In order to achieve this, the cohesion of<br />

the geological units was increased, mainly in steep<br />

areas, until a factor of safety equal to one was<br />

accomplished. In order to avoid unrealistic cohesion<br />

values, the procedure suggested by Jibson et al.<br />

(1998) was used and an Fs of 1.001 was assigned to<br />

the unstable cells (figure 4).<br />

necessarily precisely predict measured landslide<br />

displacements in the field. Rather, Newmark's<br />

displacement is a useful index of how a slope is likely<br />

to perform during seismic shaking.<br />

Fig. 6: Distribution of the generated PGA values based on<br />

the scenario of 1870 earthquake.<br />

Fig. 4: Map showing the distribution of the fact of safety<br />

against slope instability.<br />

The critical acceleration value for each 30 m-grid<br />

pixel was then computed and a relative map was<br />

compiled for the whole area of Delphi using the<br />

spatial analysis of ArcInfo Software. The criticalacceleration<br />

map (figure 5) can be characterized as a<br />

seismic landslide susceptibility map as it delineates<br />

areas prone to slope failure independent of any<br />

ground-shaking scenario (Jibson et al., 1998).<br />

The value of the peak ground acceleration (PGA) that<br />

was employed in our study was estimated using the<br />

attenuation relationship proposed by Skarlatoudis et<br />

al. (2003) and by taking into consideration as<br />

earthquake magnitude the relevant seismic event of<br />

1870.<br />

In the present study, the Newmark displacement was<br />

computed using equation [3] proposed by Jibson<br />

(2007), valid for a magnitude from 5.3 to 7.6. This<br />

equation estimates rigid-block Newmark<br />

displacement as a function of the critical acceleration<br />

and peak ground acceleration.<br />

2.335 1.478<br />

a<br />

log 2.710 log 1 c ac<br />

<br />

D <br />

n a <br />

max a <br />

max<br />

(3)<br />

0.424 M<br />

0.454<br />

where D n is Newmark displacement in centimeters,<br />

a c is critical (yield) acceleration in g's, and a max is the<br />

peak horizontal ground acceleration in g's.<br />

The distributions of the estimated values of Newmark<br />

displacement are shown in figure 7. The boundary<br />

values taken for the applied approach of Newmark's<br />

displacement were 2, 5, 10, 25, 50 and 100 cm. Sites<br />

where the estimated Dn is >5 cm can be considered<br />

as prone to slope failure while failure is unlikely<br />

where Dn < 5 cm. As it is shown in figure 7, most of<br />

the study area is considered as a low potential to<br />

earthquake-induced rockfalls zone since the<br />

computed newmark displacement ranges between 2<br />

and 5 cm. However, a “hotspot”, an area of high<br />

potential to slope failures, is delineated at the eastern<br />

part of the archeological area of Delphi where the<br />

estimated displacement is >100 cm.<br />

Fig. 5: Map showing the distribution of the value of critical<br />

acceleration<br />

As an outcome, a PGA contour map of the study<br />

area was developed based on the computed values<br />

of ground motion using the Euclidean distance model<br />

provided by ArcInfo software. As can be seen in<br />

figure 6, the value of PGA in the study area varies<br />

from 0.32 to 0.71g.<br />

The next step in the analysis developed by Newmark<br />

(1965) is the calculation of the cumulative permanent<br />

displacement of the slopes for the given level of<br />

ground shaking. According to Jibson et al. (1998),<br />

Newmark’s method is based on a fairly simple model<br />

of rigid-body displacement and thus does not<br />

257<br />

Fig. 7: map showing the estimated Newmark displacement<br />

values in the area of Delphi


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

CONCLUSIONS<br />

The goal of this study was the evaluation of the<br />

potential of rockfall occurrence generated by seismic<br />

events in the vicinity of the archaeological site of<br />

Delphi, Greece. In order to achieve this, the scenario<br />

of 1870 earthquake was taken into account. The<br />

outcome provided by this research indicates that the<br />

value of Newmark displacement close to the<br />

archaeological site is low and the area is considered<br />

as low potential to earthquake-induced rockfalls.<br />

However, high values of displacement were<br />

estimated at the archaeological area where the factor<br />

of safety against slope failure and the critical<br />

acceleration are less than 1 and 0.1g, respectively.<br />

Thus, even a small magnitude seismic event could<br />

generate rockfalls.<br />

Acknowledgements: The study of active faults and the<br />

evaluation of their potential have been financial supported<br />

by the General Secretariat for research and technology of<br />

Greece. The authors of this study would like to thank the<br />

anonymous reviewers for their constructive comments.<br />

References<br />

Ambraseys, N. Pantelopoulos, P. (1989). The Fokis<br />

earthquake of 1 August 1870. Journal European<br />

Earthquake Engineering, 3, 10-18.<br />

Angelier, J., Lyberis, N., Le-Pichon, X., Barrier, E., &<br />

Huchon, P. (1982). The tectonic development of the<br />

Hellenic Arc and the Sea of Crete: a synthesis.<br />

Tectonophysics, 86, 159 -196.<br />

De Boer J.Z., J. R. Hale (2000). The geological origins of<br />

the oracle at Delphi, Greece, in: McGuire, B. et al., (eds)<br />

The archaeology of geological catastrophes. Geological<br />

Society of London, Special Publication, 171, 399–412.<br />

Doutsos, T., Kontopoulos, N. & Poulimenos, G. (1988): The<br />

Corinth-Patras rift as the initial stage of continental<br />

fragmentation behind an active island arc (Greece). Basin<br />

Research, 1, 177-190.<br />

Jackson, J.A. McKenzie, D.P. (1988). The relationship<br />

between plate motions and seismic tensors and the rate<br />

of active deformation in the Mediterranean and Middle<br />

East. Geophysical Journal, 93, 45-73.<br />

Jibson, R.W, (1993). Predicting earthquake-induced<br />

landslide displacements using Newmark’s sliding block<br />

analysis. Transp Res Rec 1411:9–17<br />

Jibson, R.E, (2007). Regression models for estimating<br />

coseismic landslide displacement. Engineering Geology<br />

91:209–218<br />

Jibson, R.W., Keefer, D.K, (1993) Analysis of the seismic<br />

origin of landslides: examples from the New Madrid<br />

Seismic Zone. Geological Society America Bulletin<br />

105(4):521–536<br />

Jibson, R.E., Harp, E.L., & Michael, J.A, (1998) A method<br />

for producing digital probabilistic seismic landslide hazard<br />

maps: an example from the Los Angeles, California area.<br />

USGS open-file report, pp 98–113<br />

Keefer, D, (1984) Landslides caused by earthquakes.<br />

Geological Society America Bulletin 95:406–421<br />

Keefer, D.K., Wilson, R.C, (1989) Predicting earthquakeinduced<br />

landslides, with emphasis on arid and semi-arid<br />

environments. In: Landslides in a semi-arid environment.<br />

Inland Geological Survey Society, 2:118–149<br />

Khazai, B., Sitar, N, (2000). Landsliding in native ground: a<br />

GIS-based approach to regional seismic slope stability<br />

assessment,<br />

report.<br />

http://www.ce.berkeley.edu/*khazai/research<br />

Le-Pichon, X. Angelier, J. (1979). The Hellenic arc and<br />

trench system a key to the neotectonic evolution of the<br />

eastern Mediterranean area. Tectonophysics, 60, 1-42.<br />

Makris, J. (1976). A dynamic model of the Hellenic arc<br />

deduced from geophysical data. Tectonophysics, 36,<br />

339-346.<br />

Makris, J. (1978). The crust and upper mantle of the<br />

Aegean region from deep seismic soundings.<br />

Tectonophysics, 46, 269-284.<br />

McKenzie, D.P. (1972). Active tectonics of the<br />

Mediterranean region, Geophysical Journal Research<br />

Society, 30, 109-182.<br />

McKenzie, D. (1978). Active tectonics of the Alpine-<br />

Himalayan belt: the Aegean Sea and surrounding<br />

regions. Geophysical Journal Research Society, 55, 217-<br />

254.<br />

Miles, S.B., Keefer, D.K, (2000). Evaluation of seismic slope<br />

performance models using a regional case study.<br />

Environmental Engineering Geoscience 6(1):25–39<br />

Miles, S.B., Keefer, D.K, (2009). Evaluation of CAMEL–<br />

Comprehensive Areal Model of Earthquake-induced<br />

Landslides. Engineering Geology 104: 1–15<br />

Newmark, N.M, (1965). Effects of earthquake on dams and<br />

embankments. Geotechnique 15(2):139–160<br />

Papadopoulos, G.A. (ed.) (2000). Historical earthquakes<br />

and tsunamis in the Corinth rift, central Greece. National<br />

Observatory of Athens, Institute of Geodynamics,<br />

Publication No 12. 128p.<br />

Papadopoulos, G.A., Kondopoulou, D.P., Leventakis, G.A.<br />

& Pavlides, S.B. (1986). Seismotectonics of the Aegean<br />

region. Tectonophysics, 124, 67-84<br />

Papazachos, B.C. and Papadopoulos, GA. (1977). Deep<br />

tectonic and associated ore deposits in the Aegean area.<br />

Proc. 6th Colloq. Geology Aegean Region, 3, 1071-1080.<br />

Pavlides, S., Chatzipetros, A. & Valkaniotis, S. (2007).<br />

Seismically Capable Faults In: Aegean Broader Area:<br />

Criteria For Classification. Geological Society<br />

Bicentenary Conference: Earth Sciences in the Service of<br />

Society, 10-12 September 2007, London. Abstract Book,<br />

124-125.<br />

Pavlides, S., Chatzipetros, A. & Valkaniotis, S. (2008).<br />

Active faults of Greece and surroundings. In <strong>Proceedings</strong><br />

of the 33 rd<br />

International Geological Congress, Oslo,<br />

Norway. Abstract.<br />

Pèchoux, P. Y. (1977). Nouvelles remarques sur les<br />

versants Quaternaires du secteur de Delphes. Revue de<br />

Gèographie Physique et de Gèologie Dynamique,<br />

19:83–92<br />

Piccardi, L. (2000). Active faulting at Delphi, Greece:<br />

Seismotectonic remarks and a hypothesis for the<br />

geologic environment of a myth. Geology, 28: 651–654.<br />

Schmidt, J. (1879). Studien ueber Erdbeben, ed. Georgi, A.<br />

(Leipzig), 352p.<br />

Sebrier, M. (1977). Tectonique recente d’une tranversale a<br />

l’Arc Egeen . These de Docteur, Academie de Versailles,<br />

Universite de Paris XI, 100p.<br />

Skarlatoudis, A.A., Papazachos, C.B., Margaris, B.N.,<br />

Theodulidis, N., Papaioannou, Ch., Kalogeras, I.,<br />

Scordilis, E.M., & Karakostas, V, (2003). Empirical peak<br />

ground-motion predictive relations for shallow<br />

earthquakes in Greece. Bulletin of Seismological Society<br />

of America 93(6): 2591–2603<br />

Valkaniotis, S. (2009). Correlation Between Neotectonic<br />

Structures and Seismicity in the broader area of Gulf of<br />

Corinth (Central Greece). Unpublished PhD Thesis,<br />

Aristotle University of Thessaloniki, 2009, 247pp.<br />

258


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SEDIMENTARY BURIAL OF ANCIENT OLYMPIA (PELOPONNESE, GREECE)<br />

BY HIGH-ENERGY FLOOD DEPOSITS – THE OLYMPIA TSUNAMI HYPOTHESIS<br />

Vött, Andreas (1, Fischer, Peter (2), Hadler, Hanna (1), Handl, Mathias (3), Lang, Franziska (4), Ntageretzis, Konstantin (1),<br />

Willershäuser, Timo (1)<br />

(1) Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong>, 55099 <strong>Mainz</strong>, GERMANY. E-Mail: voett@uni-mainz.de<br />

(2) Institute for Geography, <strong>Universität</strong> zu Köln, Albertus-Magnus-Platz, 50923 Köln, GERMANY.<br />

E-mail: peter.fischer@uni-koeln.de<br />

(3) Philipps-<strong>Universität</strong> Marburg, Biegenstraße 10, 35037 Marburg, GERMANY<br />

(4) Department of Classical Archaeology, Technische <strong>Universität</strong> Darmstadt, El-Lissitzky-Str. 1, 64287 Darmstadt, GERMANY.<br />

E-mail: flang@klarch.tu-darmstadt.de<br />

Abstract(Sedimentary burial of ancient Olympia (Peloponnese, Greece) by high-energy flood deposits – The Olympia<br />

tsunami Hypothesis): Detailed geo-scientific studies were carried out in the Kladeos and lower Alpheios River valleys in order to<br />

clarify the mystery of the rapid burial of Olympia under 4-6 m of sediments after the 6 th cent. AD and subsequent erosion of the<br />

Kladeos River by 8-10 m down to the ancient flow level. Sedimentological, geophysical, geochemical and microfaunal analyses<br />

were conducted along the Olympia terrace by means of 22 vibracores and 70 resistivity tomography transects. Geomorphological<br />

studies revealed strong discrepancies between the present hydraulic potential of the Kladeos River and the dimension and structure<br />

of the Olympia terrace. Our results show that the Kladeos River valley and Olympia experienced at least four distinct phases<br />

of catastrophic high-energy flood events. Sedimentary, geochemical and faunal traces found in the adjacent Basin of Flokas-<br />

Pelopio clearly document multiple tsunami impact. Identical fingerprints and strong stratigraphical correlations were also detected<br />

along the Kladeos River beyond the Ridge of Flokas-Platanos. We thus set up and discuss the Olympia Tsunami Hypothesis<br />

saying that the shallow saddles of the ridge were repeatedly overflowed by tsunami waters and the cult site Olympia was rather<br />

destroyed by tsunami than by fluvial processes related to the Kladeos River.<br />

Key words: Olympia, high-energy deposits, tsunami, geoarchaeology<br />

INTRODUCTION<br />

Olympia, used as famous cult site for Panhellenic<br />

games between Archaic times and the 4 th cent. AD,<br />

is located at the confluence of the Kladeos and Alpheios<br />

Rivers in the western Peloponnese (Greece).<br />

The sedimentary burial of ancient Olympia is one of<br />

the most interesting geoarchaeological mysteries in<br />

the Mediterranean world. The sedimentological evolution<br />

since early medieval times shows two different<br />

steps. After the 6 th cent. AD, the site was covered by<br />

4-6 m of sediments; subsequently, the nearby<br />

Kladeos River eroded its bed by 8-10 m approximately<br />

reaching the level existing during antiquity.<br />

Previous studies presented different explanations for<br />

this setting. Büdel (1981), together with Dufaure &<br />

Fouache 1988, Fouache & Pavlopoulos 2011, are in<br />

favour of anthropogenically induced soil erosion as<br />

main factor for enhanced sediment accumulation at<br />

the mouth of the Kladeos River during phases of<br />

uncontrolled landuse, especially after the Slavic<br />

invasion in early medieval times. On the contrary,<br />

Fountoulis & Mavroulis (2008) hold distinct periods of<br />

wet climate responsible for accelerated sediment<br />

accumulation. However, both scenarios do not give<br />

explanations for the change from accumulation to<br />

subsequent erosion dynamics within the past 1500 or<br />

so years.<br />

259<br />

Before systematic excavation of the site by the German<br />

Archaeological Institute started in 1875, the<br />

archaeological remains of Olympia were integrated<br />

into a wide terrace, the so called Olympia terrace.<br />

From a geomorphological point of view, this terrace<br />

can be traced both downstream the Alpheios River<br />

towards the present coast of the Gulf of Kyparissia<br />

and upstream all along the lower and middle Kladeos<br />

River valley. In the Kladeos area (Fig. 1), the Olympia<br />

terrace is up to 300-500 m wide. It is present on<br />

both sides of the Kladeos River with a distance of up<br />

to 200 m between the opposite terrace faces. Considering<br />

that the Kladeos River is rather a creek than<br />

a river with a perennial runoff concentrated within a<br />

maximum 5-8 m-wide secondary river channel and<br />

maximum water flow depths of 2-3 m during winter<br />

and heavy rain events, there is a considerable discrepancy<br />

between the dimension and the hydraulic<br />

potential of the Kladeos River on the one hand, and<br />

the local geomorphology of the Olympia terrace on<br />

the other hand.<br />

The objectives of our investigations thus were (i) to<br />

establish a well-based stratigraphy of the Olympia<br />

terrace along the Kladeos River by detailed geomorphological<br />

and sedimentological studies, (ii) to compare<br />

these results with stratigraphies found along the<br />

Alpheios River, especially in the adjacent Basin of<br />

Flokas-Pelopio, and (iii) to find a geomorphological<br />

model which best explains the hydro-dynamic finger-


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

print and distribution pattern of the encountered<br />

sediments.<br />

METHODS<br />

We carried out geomorphological mapping of the<br />

Olympia terrace using topographical and geological<br />

maps and remote sensing data. Stratigraphical studies<br />

are based on 22 vibracores, 16 of which were<br />

drilled in the Kladeos River valley and 6 in the Alpheios<br />

River valley by means of a handheld Cobra<br />

mk1 vibracorer (Atlas Copco) and a Nordmeyer drill<br />

rig (type RS 0/2.3). We used core diameters of 6 and<br />

5 cm. Maximum coring depth was 17 m below ground<br />

surface (m b.s.). Earth resistivity tomography (ERT)<br />

was conducted to study subsurface structures and<br />

stratigraphies along 70 transects using a multielectrode<br />

geo-electrical Iris instrument (type Syscal<br />

Junior Switch 48). Selected sediment samples were<br />

analysed by their microfossil content. Key cores were<br />

additionally cored with an inliner system and analysed<br />

using the X-ray fluorescence technique (XRF).<br />

We also conducted grain size analyses for selected<br />

samples. Vibracoring sites and ERT transects were<br />

measured by means of a differential GPS (type Topcon<br />

HiPer Pro) with an accuracy of 2 cm or better. A<br />

local geochronostratigraphical framework was established<br />

using age estimations of diagnostic ceramic<br />

fragments, radiocarbon dating and Optically Stimulated<br />

Luminescence (OSL) approaches.<br />

RESULTS<br />

Vibracores ALP 3, 4, 5 and 8, drilled in the environs<br />

of ancient Olympia, especially on top of the Olympia<br />

terrace towards the south of the southeastern Roman<br />

baths and on top of the Olympia terrace to the west<br />

of the Kladeos, revealed characteristic sequences of<br />

light brown, silt- and clay-dominated alluvial or colluvial<br />

deposits with remains of freshwater fauna which<br />

are repeatedly interrupted by up to four sections out<br />

of light brown sand and gravel. These coarse-grained<br />

deposits were found in comparable stratigraphic<br />

positions in every core and thus indicate synchronous<br />

flooding of wide areas related to high flow velocities.<br />

Associated to these deposits, we found<br />

sedimentary structures such as basal erosional unconformities<br />

in underlying silt deposits, fining upward<br />

sequences with mud caps and abundant marine shell<br />

debris and marine microfauna. XRF analyses revealed<br />

clear maximum peaks of the Ca/Ti ratio for<br />

the coarse-grained sections. The calcium content<br />

documents the input of calcium carbonate from biogenic<br />

and bedrock sources, the titan content reflects<br />

terrigenous input by subaerial weathering into the<br />

sedimentary system. Results were also tested for<br />

masking and matrix effects which can be excluded as<br />

major sources of bias. Apart from a distinct and ca. 1<br />

m-thick palaeosol found on top of a coarse-grained<br />

section at ca. 4 m b.s. in core ALP 8 (Fig. 2), which<br />

includes Roman sherds, palaeosols are missing. A<br />

charcoal fragment from a fining upward sequence out<br />

of sand and gravel deposited under high-energy<br />

conditions at site ALP 5 was 14 C AMS radiocarbon<br />

dated to 585-647 cal AD (2 interval, 3.74 m b.s.).<br />

Vibracores ALP 12-15 and 19 were drilled in the<br />

middle Kladeos River valley around the villages of<br />

Mageiras and Kladeos on top of the Olympia terrace.<br />

Here, we also found predominating clayey to silty<br />

deposits accumulated under quiescent to moderate<br />

flow conditions. These deposits are grey in colour,<br />

partly include lots of plant material and freshwater<br />

shell and thus document a permanent water body of<br />

fluvio-limnic nature. Similar to the situation at Olympia,<br />

we found up to four intersecting layers of sand<br />

and gravel, partly grey (base and mid-section), partly<br />

light brown in colour (top), associated to high-energy<br />

sediment type structures such as basal unconformities,<br />

muddy intraclasts, fining upward sequences and<br />

including abundant faunal remains of marine origin.<br />

Based on XRF measurements, the Ca/Ti ratio again<br />

shows clear maximum peaks stratigraphically corresponding<br />

to the intersecting coarse-grained layers.<br />

Thus, all along the Kladeos River valley between<br />

Kladeos, Mageiras and Olympia, the Olympia terrace<br />

shows a similar inner structure with the individual<br />

coarse-grained high-energy layers lying in stratigraphically<br />

consistent positions. The same is true for<br />

both distal and proximal parts of each specific terrace<br />

section as documented by ERT transects.<br />

We thus conclude, in a first step, that the Olympia<br />

terrace between Kladeos and Olympia documents at<br />

least four phases of high-energy flood events that<br />

obviously affected the whole valley bottom to an<br />

extent far beyond the dimensions of the present<br />

Kladeos river channel.<br />

Vibracores 9-11 were drilled at the eastern fringe of<br />

the Basin of Flokas-Pelopio at a distance of 1 to 2 km<br />

to the west of the Kladeos River valley across the<br />

Flokas-Platanos Ridge. Vibracore ALP 10 lies around<br />

2.2 km distant from the Alpheios River at a right<br />

angle opposite to its seaward flow direction. Despite<br />

the vicinity to the Alpheios River, vibracore ALP 9<br />

does not include any pieces of gravel; it consists of<br />

homogeneously light brown (top) to grey (base),<br />

clayey to silty deposits accumulated in a low-energy<br />

freshwater lake environment. Associated to basal<br />

unconformities, we found several intersecting layers<br />

of sand with fining upward sequences and mud caps<br />

and marine faunal remains which clearly document<br />

episodic high-energy interruption of the allochthonous<br />

environment. Vibracore ALP 10, drilled at the<br />

western hill slope of the Flokas-Platanos Ridge,<br />

revealed a similar stratigraphic pattern with distinct<br />

interruptions of limnic (base and mid-section) and<br />

colluvial (top) deposits by predominantly sandy to<br />

gravelly high-energy deposits reaching up to 20 m<br />

above present sea level (m a.s.l.). In each case, at<br />

least the upper part of the intersecting layer consists<br />

of brown sediments of mostly terrestrial origin. The<br />

Ca/Ti ratios for both cores show distinct maximum<br />

peaks for the intersecting coarse-grained allochthonous<br />

material (Fig. 3).<br />

260


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 1: View of the middle Kladeos River valley (foreground) towards the west. The valley is separated from the adjacent<br />

Basin of Flokas-Pelopio (middleground) by the homonymous ridge. The Alpheios River valley connects the basin and the Gulf<br />

of Kyparissia (left background) in a direct line. Photo taken by A. Vött, 2011.<br />

In the more seaward vibracore ALP 7, drilled some 8<br />

km inland in the midst of the Alpheios River valley<br />

halfway between ALP 9 and the present coast, we<br />

found autochthonous marine sand at approx. 2.50 m<br />

below present sea level (m b.s.l.). Covered by a 10<br />

m-thick layer of sandy gravel, this unit documents<br />

that the Gulf of Kyparissia extended far into today’s<br />

Alpheios valley during the Holocene.<br />

DISCUSSION<br />

Sedimentary structures, geochemical fingerprints and<br />

faunal remains encountered at the western fringe of<br />

the Basin of Flokas-Platanos indicate episodic highenergy<br />

marine flooding from the sea side. Stratigraphies<br />

of cores ALP 9 and 10 clearly show tsunami-type<br />

marine incursions into a shallow lake and<br />

runup-backflow sequences at the adjacent hillside,<br />

respectively. Considering that at least parts of the<br />

lower Alpheios River valley were flooded by the sea<br />

during the Holocene and manifold traces of palaeotsunami<br />

are known from the present coast (Vött<br />

et al. 2011), these findings are plausible.<br />

core sections revealed freshwater ostracods in finegrained<br />

silt-dominated deposits but exclusively marine<br />

species in high-energy flood sediments. In case,<br />

these were transported by a mega-Kladeos River<br />

with a width of several hundreds of meters – dimensions<br />

which are necessary to explain the consistent<br />

lateral distribution of this facies across the entire<br />

Olympia terrace – one would have to expect admixed<br />

freshwater or even river-borne species. This was not<br />

the case in the samples which we analysed. However,<br />

further attempts are needed to fully understand<br />

the fossil record of the high-energy flood deposits in<br />

the environs of Olympia.<br />

Geomorphological studies carried out within our<br />

project revealed erosion as well as scouring features<br />

across the lower saddles of the Flokas-Platanos<br />

Ridge lying at about 60 m a.s.l. which indicate possible<br />

flow paths across the ridge.<br />

Considering, however, that both the geochemical and<br />

the overall stratigraphic patterns of the cores from the<br />

west of the ridge and from the Kladeos River valley<br />

itself are principally identical – documenting episodic<br />

high-energy input of coarse-grained marine sediments<br />

into prevailing quiescent environments – we<br />

hypothesize that marine flooding also affected the<br />

Kladeos River valley and Olympia. Our main arguments<br />

thus are of sedimentological and geochemical<br />

nature and based on consistent stratigraphies on<br />

either side of the Flokas-Platanos Ridge. We call this<br />

scenario the “Olympia Tsunami Hypothesis”.<br />

As a major argument against our hypothesis one may<br />

bring forward the fact that Neogene bedrock units in<br />

the catchment area of the Kladeos River also include<br />

conglomerate and sand units, the latter being characterized<br />

by abundant remains of a Plio-Pleistocene<br />

marine fauna (IGME 1982). However, especially<br />

around Olympia and in most parts of the middle<br />

Kladeos River valley, Plio-Pleistocene marl and not<br />

sand is the predominant bedrock material provoking<br />

many landslides (IGME 1982, Christaras et al.<br />

2002). Moreover, microfaunal analyses of selected<br />

261<br />

Fig. 2: Facies distribution pattern of vibracore ALP 8 (38.37<br />

m a.s.l.) drilled on top of the Olympia terrace to the west of<br />

the Kladeos River some 250 m to the west of the Kronos<br />

hill. Several sandy to gravelly high-energy flood type (heft)<br />

deposits can be seen alternating with fine-grained colluvial<br />

or (fluvio-)limnic deposits. The mid-core palaeosol dates to<br />

Roman times. Please note the three fining upward sequences<br />

encountered in the uppermost heft unit. Photo<br />

taken by T. Willershäuser, 2010.<br />

Concerning the elevation which, at a first glance,<br />

seems to be too high for tsunami flooding, one has to<br />

consider (i) channelling and accelerating effects of<br />

the tube-like and ca. 8 km-long lower Alpheios River<br />

valley during inflow, (ii) backwatering and boosting of<br />

inundating marine water masses because of blocked<br />

backflow in the Alpheios River valley itself and at the


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

breakthrough through Drouva Ridge, where the modern<br />

Alpheios dam is located, (iii) the arrival of subsequent<br />

waves of a longer wave train before backflow<br />

was completely accomplished, as well as (iv) potential<br />

interim changes in the topography due to landslides<br />

of the predominating Tertiary marls typical for<br />

the ridge. From the nearby coast, there is evidence of<br />

multiple tsunami landfall since the mid-Holocene and<br />

for tsunami run-up up to 18 m a.s.l. around the ancient<br />

site Pheia, one of the harbours of Olympia (Vött<br />

et al. 2011).<br />

By the Olympia Tsunami Hypothesis, we suggest that<br />

tsunami waters repeatedly overflowed the lower<br />

saddles between Flokas and Platanos and then<br />

partly flowed upstream and partly downstream the<br />

Kladeos River valley, hereby creating a vast terrace<br />

structure way above the flow level of the Kladeos at<br />

that time. Subsequently, tsunami backflow concentrated<br />

along the Kladeos creek eroding an up to 200<br />

m-wide gap into the terrace. Tsunami backflow<br />

through the breakthrough of the Alpheios across the<br />

Drouva Ridge was hindered and possibly blocked, at<br />

least during tsunami inundation of the Basin of Flokas-Pelopio.<br />

One of the high-energy flood deposits<br />

encountered near Olympia was dated to 585-647 cal<br />

AD which fits well with the earthquake in 551 AD<br />

during which Olympia is reported to have been destroyed.<br />

However, there are no historic accounts on<br />

catastrophic flooding of Olympia.<br />

finally covered completely by sediments. (ii) Similar<br />

high-energy flood deposits were also encountered at<br />

the eastern fringe of the adjacent Basin of Flokas-<br />

Pelopio; due to their fossil content, their geochemical<br />

fingerprint, their geomorphological position and their<br />

stratigraphical pattern, they are interpreted as of<br />

tsunamigenic origin. (iii) In both areas, the highenergy<br />

facies is associated to sedimentary structures<br />

known from historic to recent tsunami events (basal<br />

unconformity, fining upward sequences, mud caps,<br />

ripped up intraclasts etc.) and is characterized by<br />

abundant fragments of a marine fauna. (iv) Geomorphological<br />

features such as marks of scouring and<br />

undercutting let us assume that tsunami overflow<br />

occurred across the comparatively shallow saddles<br />

between the villages of Flokas and Platanos wheras<br />

backflow was accomplished along the river valley as<br />

soon as blocking of the entrance of the Alpheios<br />

River into the Basin of Flokas-Pelopio by tsunami<br />

waters had ceased. We call this scenario “The Olympia<br />

Tsunami Hypothesis”. (v) Our results, together<br />

with manifold geoarchaeological destruction patterns<br />

at Olympia, rather indicate catastrophic event-related<br />

flooding by tsunami than by the River Kladeos itself.<br />

Acknowledgements: Sincere thanks are due to our cooperating<br />

partners G. Chatzi-Spiliopoulou (Olympia), I. Fountoulis<br />

(Athens), H.-J. Gehrke (Freiburg), A. Hoppe, R. Lehné<br />

(Darmstadt), K. Reicherter (Aachen), D. Sakellariou (Athens)<br />

and R. Senff (Athens/Olympia). Work permits were<br />

kindly issued by the Greek Institute for Geology and Mineral<br />

Exploration (IGME). We gratefully acknowledge funding by<br />

the German research Foundation (DFG, VO 938/3-1).<br />

References<br />

Fig. 3: Ca/Ti ratios for vibracore ALP 9A drilled at the western<br />

fringe of the Basin of Flokas-Pelopio at the foot of the<br />

homonymous ridge nearby Olympia. Allochthonous siltdominated<br />

deposits of a quiescent limnic environment are<br />

characterized by a Ca/Ti base level around 50. Episodic<br />

interferences from the sea side by tsunami waves left behind<br />

marine sand deposits with Ca/Ti ratios up to 200.<br />

Similar Ca/Ti profiles and stratigraphies were found for<br />

vibracores in the Kladeos River valley and nearby Olympia.<br />

CONCLUSIONS<br />

Detailed geomorphological, sedimentological, geophysical,<br />

geochemical and microfaunal studies of the<br />

Olympia terrace in the Kladeos and the lower Alpheios<br />

River valleys allow to draw the following conclusions.<br />

(i) The Kladeos valley and the environs of<br />

Olympia were affected by at least four distinct phases<br />

of high-energy flood events by which the site was<br />

Büdel, J. (1981). Klima-Geomorphologie. 2 nd<br />

edition.<br />

Bornträger. Berlin, Stuttgart. 304 p.<br />

Christaras, B., Mariolakos, I., Dimitriou, A., Moraiti, E. & D.<br />

Mariolakos (2002). Slope instability at Olympia archaeological<br />

site in southern Greece. Intern. Symp. “Landslides<br />

Risk Mitigation and Protection of Cultural and Natural<br />

Heritage”, Abstract Volume. Kyoto. 339-342.<br />

Dufaure, J.-J. & E. Fouache (1988). Variabilité des crises<br />

d’âge historique le long des vallées d’Elide (Ouest du Peloponnèse).<br />

Cahier Inter-universitaire d’Etudes Méditerranéennes,<br />

12: 259-278. Poitiers.<br />

Fouache, E. & K. Pavlopoulos (2011). The interplay between<br />

environment and people from Neolithic to Classical<br />

times in Greece and Albania. In: Landscapes and societies.<br />

Selected cases (Martini, I.P., Chesworth, W. eds.).<br />

Springer. Dordrecht. 155-166.<br />

Fountoulis, I., Mariolakos, I., Mavroulis, S. & I. Ladas<br />

(2008). Flood periods during prehistoric and Roman<br />

times in the Kladeos torrent basin – ancient Olympia<br />

(Greece). In: 8 th<br />

Intern. Hydrogeol. Congr. Greece, 3 rd<br />

MEM Workshop Fissured Rocks Hydrology, Abstract<br />

Volume (Migiros, G., Stamatis, G., Stournaras, G. eds.).<br />

Athens. 809-818.<br />

Institute for Geology and Mineral Exploration (IGME, 1982).<br />

Geological map of Greece, 1:50,000, Olympia Sheet.<br />

Athens.<br />

Vött, A., Bareth, G., Brückner, H., Lang, F., Sakellariou, D.,<br />

Hadler, H., Ntageretzis, K. & T. Willershäuser (2011).<br />

Olympia’s harbour site Pheia (Elis, western Peloponnese,<br />

Greece) destroyed by tsunami impact. Die Erde (in<br />

press).<br />

262


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

LANDSLIDE MAPPING TO ANALYSE EARTHQUAKE ENVIRONMENTAL EFFECTS (EEE)<br />

IN CARMONA, SPAIN – RELATION TO THE 1504 EVENT?<br />

Vollmert, Andre (1, Klaus Reicherter (2), Pablo G. Silva (3), Tomas M. Fernandez-Steeger (4)<br />

(1) Lehrstuhl für Ingenieur- und Hydrogeologie, RWTH Aachen University, Lochnerstr. 4-20, 52056 Aachen. GERMANY. Email:<br />

andre.vollmert@rwth-aachen.de<br />

(2) Lehr- und Forschungsgebiet Neotektonik und Georisiken, RWTH Aachen University, Lochnerstr. 4-20, 52056 Aachen.<br />

GERMANY. Email: k.reicherter@nug.rwth-aachen.de<br />

(3) Dpto. Geología, Escuela Politécnica Superior de Ávila, Universidad de Salamanca. Avda. Hornos Caleros, 50. 05003-Ávila.<br />

SPAIN. Email: pgsilva@usal.es<br />

(4) Lehrstuhl für Ingenieur- und Hydrogeologie, RWTH Aachen University, Lochnerstr. 4-20, 52056 Aachen. GERMANY. Email:<br />

fernandez-steeger@lih.rwth-aachen.de<br />

Abstract (Landslide mapping to analyse earthquake environmental effects in Carmona, Spain – relation to the 1504<br />

event?): The 1504 Carmona earthquake (intensity IX EMS) claimed the loss of human life and caused a number of Earthquake<br />

Environmental Effects. On the basis of historical data reported by George Bonsor (1918) this study is intended to estimate<br />

coseismic slope performance. The aim is to combine field investigations, geotechnical parameters and computerized models to<br />

generate digital probabilistic seismic landslide hazard maps on a local scale. GIS-based simulations of mass movements driven by<br />

hydrodynamical and gravitational processes are performed by means of the factor of safety, which is calculated for dry and fully<br />

water saturated conditions. Following Newmark´s sliding block model these approaches are extended to assess the potential of<br />

earthquake-triggered slope movements. Assuming a Peak Ground Acceleration of 0.3 g, representing the 1504 event, the most<br />

affected areas show a failure probability of 33.5 %.<br />

Key words: 1504 Carmona Earthquake, Seismic Landslide Hazard Assessment, South Spain<br />

INTRODUCTION<br />

In 1918 the archaeologist George Bonsor was the<br />

first scientist who published the effects of a strong<br />

earthquake near Carmona (South Spain) in 1504<br />

(Bonsor, 1918). Based on the ESI-2007 Intensity<br />

Scale, Silva et al. (2009) attract notice again on this<br />

event in order to update Bonsor´s data. They focus<br />

on ground cracks, liquefaction, anomalous waves,<br />

flooding in rivers, temporary turbidity changes in<br />

wells and, especially, on mass movements, since<br />

landslides and rock falls belong to the most relevant<br />

phenomena of all EEE being observed in Carmona.<br />

This study provides different approaches to calculate<br />

the site scaled slope instability in terms of the 1504<br />

earthquake (IX EMS) as a potential triggering factor<br />

for a number of observed landslides. Each of the<br />

methods combines geotechnical results and slope<br />

angles derived from a Digital Elevation Model (DEM).<br />

Figure 1 points out the sequential steps leading to<br />

the hazard-mapping procedure of the study. All<br />

simulations have been performed under dry and fully<br />

water saturated conditions.<br />

LANDSLIDES IN CARMONA<br />

Within the southern margin of the Guadalquivir river<br />

valley Carmona is founded on a small NE-SW<br />

trending ridge (Los Alcores). It consists of Miocene<br />

blue marls and grey clays coming from the southeast<br />

located Betic Cordillera front.<br />

Fig. 1: Flow chart showing the steps involved in producing a seismic landslide hazard map; white: input parameters; blue:<br />

results (modified after Jibson et al., 2000).<br />

263


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

This substratum is covered by a Late Neogene<br />

calcarenite unit, which outcrops in a steep cliff,<br />

surrounding large parts of the city. Both units can be<br />

subjected to massive landslides. In order to<br />

distinguish seismically triggered slope movements<br />

from others driven by hydrodynamical and<br />

gravitational processes, all landslide phenomena are<br />

classified according to type of movement, material<br />

and size of the displaced mass. Furthermore, all<br />

possible causes including geological, morphological,<br />

physical and human influences are determined as an<br />

important aim.<br />

All important investigation sites are indicated in<br />

figure 2 showing the studied area around Carmona.<br />

The map also includes joint diagrams from<br />

Fig. 2: Map of Carmona showing the sampling points for later geotechnical investigations, joint diagrams and the location of<br />

landslide phenomena illustrated in figure 3 (Gauss-Krüger coordinates).<br />

264


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

calcarenite strata and the location of observed rock<br />

falls, topples and slides as well as earth slides in the<br />

unit of blue marls and grey clays. Typical examples<br />

of these types are illustrated in Figure 3.<br />

66 %) are supposed to be essential for the<br />

occurrence of massive earth slides on steep slopes<br />

along the courses of streams.<br />

Apart from these invariant parameters, both, water<br />

saturated soils caused by intensive rainfalls and<br />

earthquake shaking can be seen to be the most<br />

relevant causal factors for landslides in Carmona.<br />

Therefore, these triggering factors are considered in<br />

the following simulations.<br />

Fig. 3: Huge rock fall underneath the Picacho (A);<br />

Toppling process of a column of calcarenite rock (B);<br />

Calcarenitic boulders are transported downslope on<br />

softer clayey materials (C); Earth slides due to water<br />

saturated conditions and steep slope angles (D).<br />

Landslides in Carmona are related to a number of<br />

preparative and triggering factors. Observed rock<br />

falls and topples in calcarenite strata can be mainly<br />

subjected to SSE- and ENE-striking tension cracks<br />

(fig. 2) and relatively low shearing parameters<br />

(c´ = 17 MPa; ´ = 41°). Laboratory results of<br />

analysed loose material samples indicate effective<br />

cohesions between 13.04 and 20.16 kPa and angles<br />

of internal friction between 17.35 and 26.23°. This<br />

data and the high contents of clay minerals (35 –<br />

265<br />

SIMULATION OF SLOPE STABILITY<br />

First slope stability has been simulated be means of<br />

the factor of safety, which is calculated by the ratio of<br />

the sum of the resisting forces that act to inhibit a<br />

slope failure to the sum of the driving forces that tend<br />

to cause a failure. The application of a Geographical<br />

Information System (ArcGIS 9.3) allowed a<br />

differentiated calculation for every grid cell (2 x 2 m),<br />

where input parameters vary due to different slope<br />

angles.<br />

Based on the factor of safety the site-specific critical<br />

acceleration was calculated in a second step.<br />

According to Newmark´s sliding block analogy<br />

(Newmark, 1965) the critical acceleration is defined<br />

as the minimum horizontal seismic acceleration that<br />

is necessary to overcome the shear resistance of a<br />

friction block, resting on an inclined plane. That<br />

means, the higher the degree of slope stability, the<br />

higher the critical acceleration, which is needed to<br />

cause a failure.<br />

To estimate the cumulative slope displacement<br />

during an earthquake, Wilson & Keefer (1983)<br />

developed a double integration approach based on<br />

numerically cumbersome calculations performed by<br />

Newmark (1965). Thereby, those sections of an<br />

earthquake accelerogram that exceed the critical<br />

acceleration of a slope are integrated two times to<br />

obtain the velocity and the cumulative displacement<br />

of the sliding block. Considering also the PGA of the<br />

Carmona earthquake (0.3 g) it was possible to<br />

determine the Newmark Displacement.<br />

Newmark displacement rates are not directly<br />

correlated to the potential of earthquake-triggered<br />

landslides. For this reason, Jibson et al. (2000)<br />

developed a probabilistic empirical model, which<br />

allows the estimation of the probability of a failure for<br />

every grid cell (Eq. 1):<br />

P<br />

<br />

<br />

<br />

1.565<br />

f 0.335<br />

1exp0.048D<br />

N [1]<br />

They have calibrated these parameters with data<br />

from Southern California and anticipate that the<br />

mapping procedure is applicable in any areas<br />

susceptible to seismic slope failure. Therefore the<br />

model was used to compile digital probabilistic<br />

landslide hazard maps for dry and fully water<br />

saturated conditions in the study area (fig. 4).


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 4: Probabilistic landslide hazard map showing the probability of failure in case of an intensity-IX (EMS) earthquake<br />

under fully water saturated conditions (Gauss-Krüger coordinates).<br />

The most affected areas show a failure probability of<br />

33.5%. They are generally related to the same slopes<br />

indicating a higher potential of landslides under nonseismic<br />

conditions, however, they are extended.<br />

CONCLUSION<br />

The performed assessment of earthquake-triggered<br />

landslides provides useful information to estimate<br />

potential damages during future earthquakes. In this<br />

sense, the designation of vulnerable areas can be<br />

used to predict interruptions of access roads, gas<br />

and water pipes or electrical lines in case of another<br />

strong earthquake in Carmona when landslides of<br />

large volumes will be triggered with high probability.<br />

Acknowledgements: This work would not have been<br />

possible without the support of many persons. It is a<br />

pleasure for me to thank those who always helped and<br />

inspired me during this project.<br />

References<br />

Bonsor, J. (1918): El terremoto de 1504 en Carmona y en<br />

Los Alcores. Boletín de la Real Sociedad Espan ola de<br />

Historia Natural, 18, 115-123.<br />

Jibson, R. W., Harp, E. L. & Michael, J. A. (2000): A method<br />

for producing digital probabilistic seismic landslide hazard<br />

maps. Engineering Geology, 58, 271-289.<br />

Newmark, N. M. (1965): Effects of earthquakes on dams<br />

and embankments. Géotechnique, 15, 139-160.<br />

Silva, P.G., M.A Rodríguez-Pascua, R. Pérez-López, J.L.<br />

Giner-Robles, J. Lario, T. Bardají, J.L. Goy & C. Zazo,<br />

(2009). Geological and archaeological record of the 1504<br />

AD Carmona earthquake (Guadalquivir Basin, South<br />

Spain): a review after Bonsor, 1918. In:<br />

Archaeoseismology and Palaeoseismology in the Alpine-<br />

Himalayan Collisional Zone (Pérez-López, R., Grützner,<br />

C., Lario, J., Reicherter, K., Silva, P.G. eds). Baelo<br />

Claudia, Spain, 139-142.<br />

Wilson, R. C. & Keefer, D. K. (1983): Dynamic analysis of a<br />

slope failure from the 6 August 1979 Coyote Lake,<br />

California, earthquake. Bulletin of the Seismological<br />

Society of America, 73, 863-877.<br />

266


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

STORM SURGE LAYERS WITHIN A CHANGEFUL HOLOCENE ENVIRONMENT OR<br />

SEDIMENTARY TRACES OF PALAEO-TSUNAMIGENIC EVENTS. PROS AND CONS OF<br />

ON-SITE FINDINGS, JADE BAY, SOUTHERN NORTH SEA, GERMANY<br />

Wartenberg, Wolfram (1), Andreas Vött (2), Hanna Hadler (2), Timo Willershäuser (2), Holger Freund (1), Stefanie Schnaidt (1)<br />

(1) Institute for Chemistry and Biology of the Marine environment (ICBM), Carl von Ossietzky <strong>Universität</strong> Oldenburg, 26382<br />

Wilhelmshaven. Germany. Email: wolfram.wartenberg@uni-oldenburg.de<br />

(2) Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong>, 55099 <strong>Mainz</strong>. Germany. Email: voett@uni-mainz.de<br />

Abstract (Storm surge layers within a changeful Holocene environment or sedimentary traces of palaeo-tsunamigenic<br />

events. Pros and cons of on-site findings, Jade Bay, Southern North Sea, Germany): Tsunamigenic events are well known to<br />

occur within the Mediterranean and along many seismo-tectonically active coasts all over the world. Their incidence is almost<br />

excluded for the southern North Sea, a shore line well known to be prone to storm surges that in parts occur at a regional scale.<br />

Accordant chronicles describe floods of sheer enormity, for instance affecting the Dutch and the German coastal sections at the<br />

same time (e.g. Cosmas and Damian flood, 1509).<br />

Key words: tsunami, storm, North Sea, Germany<br />

INTRODUCTION<br />

The Jade Bay in Lower Saxony, northwest Germany,<br />

is the largest tidal inlet of the German North Sea<br />

coast. The modern embayed tidal flat system shows<br />

a changeful Holocene sedimentary record from<br />

terrestrial-driven to seaward-influenced environments<br />

(Streif, 2004). The Jade area has been investigated<br />

interpreting sedimentary markers, pollen and macro<br />

remains taken from 45 cores. Probing was carried<br />

out between 2009 and 2011. The sedentary<br />

chronology is based on 36 radiocarbon and pollen<br />

datings. Direct age determination of clastic sediments<br />

will be complemented in late 2011 using optical<br />

dating.<br />

The palaeo-geomorphology of the Jade area is<br />

influenced by a major north-south trending channel<br />

(Sindowski, 1972), representing the structural<br />

rudiment for an early stage of the advancing sea.<br />

From approximately 4500 cal BC onwards, the<br />

palaeo-coastline must have been close, starting to<br />

increase the groundwater level. Alder carr to<br />

Cyperaceae fen peat started to develop extensively<br />

before marine conditions became dominant from<br />

~3000 to ~2800 cal BC (Wartenberg & Freund, in<br />

press). Two different palaeo-environments are<br />

related to the present-day Jade Bay, each identifying<br />

a distinctive local depositional development<br />

(Wartenberg & Freund, in press). From the west to<br />

the centre, the equivalent early Holocene landscape<br />

morphology is drainage-driven, feeding the<br />

associated pronounced basal peat with minerogenic<br />

water but being autonomous from isochronic relative<br />

sea-level. To the east, basal peat is absent within the<br />

sedimentary succession. Here, the facies zone is<br />

dominated by tidal flat to brackish-lagoonal<br />

267<br />

sediments, in places intercalated by fen peat layers<br />

dating back to minimum 4490 cal BC.<br />

Comparing the western and the eastern palaeoenvironments<br />

reveals different sedimentary signals<br />

identifying coincident event-stratigraphic markers of<br />

early to late Holocene age (Fig 1). Their eventstratigraphic<br />

signal may be linked to distinct<br />

sedimentary horizons at a regional scale. The clastic<br />

material deposited in parts shows rhythmic layers of<br />

coarser material within one event horizon (Fig. 2).<br />

Fig. 1(above): Description of core PR 321, tidal flat Arngast<br />

Sand, central Jade Bay. Drawing: Stefanie Schnaidt.<br />

Fig. 2 (below): Event layer of core PR 321 at 1.77 to 1.64 m.<br />

The poster presents detailed<br />

discussion on the origin of these<br />

high-energy event layers with<br />

respect to major storm or tsunami<br />

influence. Feasible Tsunami<br />

triggers may have been landslides<br />

in the northern North Sea area<br />

(alike the Storegga event at ~8000<br />

BP) or earthquakes along active<br />

faults at the coastal sections of


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

southern Spain and Portugal.<br />

Acknowledgements: The geological field work was<br />

accomplished as part of an interdisciplinary study of the<br />

Jade Bay. We gratefully acknowledge project funding by the<br />

Ministry for Science and Culture of Lower Saxony<br />

(Sponsorship program: “Niedersächsisches Vorab der<br />

Volkswagen Stiftung”).<br />

Wartenberg, W., Freund, H. (2011): Late Pleistocene<br />

and Holocene sedimentary record within the Jade<br />

Bay, Lower Saxony, Northwest Germany – new<br />

aspects for the palaeo-ecological record.<br />

Quaternary International (in press).<br />

References<br />

Sindowski, K.-H. (1972): Zur Geologie des<br />

Jadebusen-Gebietes. Oldenburger Jahrbuch 72,<br />

175-181.<br />

Streif, H. (2004): Sedimentary record of Pleistocene<br />

and Holocene marine inundations along the North<br />

Sea coast of Lower Saxony, Germany. Quaternary<br />

International 112, 3-28.<br />

268


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

TESTING EARTHQUAKE RECURRENCE MODELS WITH 3D TRENCHING ALONG THE<br />

DEAD-SEA TRANSFORM<br />

Wechsler, Neta (1, 2), Thomas K. Rockwell (2), Yann Klinger (1), Amotz Agnon (3), Shmulik Marco (4)<br />

(1) Equipe de Tectonique, Institut de Physique du Globe de Paris, CNRS, Paris, France. Email: wechsler@ipgp.fr<br />

(2) Department of Geological Sciences, San Diego State University, San Diego, CA, United States.<br />

(3) Institute of Earth Sciences, Hebrew University of Jerusalem, Jerusalem, Israel.<br />

(4) The Department of Geophysics and Planetary Sciences, Tel Aviv University, Tel Aviv, Israel<br />

Abstract (Testing earthquake recurrence models with 3D trenching along the Dead-Sea Transform): We propose to test<br />

earthquake recurrence and slip models via high-resolution three-dimensional trenching of the Betieha site on the Dead Sea<br />

Transform (DST) in northern Israel. We extend the earthquake history of this simple plate boundary fault to establish how often<br />

earthquakes have occurred in the Holocene (past 50-100 centuries), to determine the amount of slip per event, and to test<br />

competing rupture models (characteristic, slip-patch, slip-loading, and <strong>Gutenberg</strong> Richter type distribution). We do this by 3D<br />

trenching and documentation of offset buried streams of various ages across the DST. This information is critical for improving<br />

seismic hazard analysis and earthquake forecast models in general, and for establishing the current earthquake risk in Israel and<br />

the surrounding Middle East.<br />

Key words: Dead Sea Transform, 3D Paleoseismology, slip per event, historical earthquakes<br />

INTRODUCTION<br />

Understanding earthquake production along major<br />

plate-boundary faults is critical for improving seismic<br />

hazard assessment and earthquake forecast models.<br />

Models used to forecast future seismicity make<br />

fundamental assumptions about fault behavior,<br />

whether it ruptures in a random, quasi-periodic, or<br />

clustered pattern. Those models are based on limited<br />

observations of recurrent slip at a point along a fault,<br />

or variations in recurrence times at multiple<br />

paleoseismic sites along individual faults. Models<br />

such as the “characteristic earthquake” model<br />

(Schwartz and Coppersmith, 1984) or the “slip-patch”<br />

model (Sieh, 1996) rely on assessments of fault<br />

segmentation and assume that large ruptures<br />

terminating at invariant segment boundaries<br />

(Wesnousky, 2008). However the self-similarity of<br />

large ruptures, both in terms of magnitude and of slip<br />

distribution, is not clearly established thus far, as<br />

there is yet to be a recorded repeated large event<br />

since the advent of modern instrumental<br />

measurements. In order to better understand longterm<br />

earthquake recurrence there is need for<br />

comprehensive event records that includes<br />

magnitude, location and displacement data. The<br />

Beteiha (Bet-Zayda) site, located on the Dead-Sea<br />

Transform fault (DST), provides us with an<br />

opportunity for constructing long term record for an<br />

active plate-boundary via high-resolution threedimensional<br />

trenching. The data can be used to test<br />

earthquake recurrence and slip models, as well as to<br />

address other key issues such as slip-rate variation<br />

with time, or GPS-geological slip-rate disparity.<br />

Geological Background<br />

The DST is a major plate boundary and a source of<br />

significant hazard in the Middle East, accommodating<br />

the relative motion between the African and Arabian<br />

269<br />

Fig. 1: a) Generalized tectonic framework of the Middle<br />

East. b) The DST (on land) from the Red Sea northward to<br />

the East Anatolian fault (EAF) in Turkey. Slip rate data are<br />

shown in black (Daeron et al., 2004, Ferry et al., 2007, Le<br />

Beon et al., 2010, Meghraoui et al., 2003).


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

plates. It transfers slip northward from the oblique<br />

opening at the Red Sea to the Taurus-Zagros<br />

collision zone, and a consequence of this northward<br />

motion of Arabia and collision with Asia is the<br />

westward extrusion of Anatolia along the North<br />

Anatolian Fault system (Fig. 1a). The cumulative<br />

offset of the DST is ~105 km, representing the total<br />

motion between the Arabian plate and Sinai subplate<br />

since the mid Miocene (e.g., Freund et al. 1968;<br />

Garfunkel 1981). The rate of sinistral motion<br />

measured across the fault is estimated to be between<br />

3 and 7 mm/year in northern Israel (Le Beon et al.<br />

2008 and references therein, Figure 1b). Despite the<br />

rich record of historical seismicity in the Middle-East,<br />

not enough data exists to constrain surface rupture<br />

extent for historical earthquakes. Previous work<br />

(Ellenblum et al., 1998; Marco et al., 2005) on this<br />

segment of the DST obtained slip per event for the<br />

last 2 historical earthquakes (1759 and 1202 CE) and<br />

established that they indeed ruptured this far south.<br />

Results<br />

We present results from a 2 year trenching campaign<br />

in the Beteiha valley, on the northern shore of Lake<br />

Tiberias. Previous trenches in the same locale by<br />

Marco et al. (2005) exposed 5 buried paleo-channels<br />

which provided slip-per-event data for the last two<br />

earthquakes, as well as slip-rate estimate for the last<br />

5k years. In the first year, we established the<br />

feasibility for an extended 3D trenching project in that<br />

locale by excavating a 300m long fault parallel trench<br />

(Figure 2) and exposing at least 7 additional buried<br />

channels that can be used as offset markers, in<br />

addition to the 5 channels that were originally<br />

mapped by Marco et al. (2005). In order to constrain<br />

each paleo-channel’s age, 30 14 C samples were<br />

dated. The obtained dates span the range 915 years<br />

BP to more than 4000 years BP. A re-dating of<br />

charcoal from the channel (CH1) previously dated to<br />

5kyr using bulk soil 14 C, found it to be younger by at<br />

least 1kyr (4k and not 5k). Based on the new dates,<br />

the slip-rate estimate of 3 mm/yr based on 15m offset<br />

measured for CH1 is revised to 3.7 mm/yr. The ages<br />

of the buried channels cover the period between the<br />

last large event on this segment (1202 CE) and over<br />

4000 years ago.<br />

The second year’s trenching season will commence<br />

in May 2011, with the main goal to follow the paleochannels<br />

across the fault and measure the amount of<br />

offset for each, thus refining the slip-rate and slip-perevent<br />

history. We plan to have several of the<br />

channels excavated and mapped by July, and we will<br />

present our results in Corinth, 2011.<br />

References<br />

Ellenblum, R., Marco, S., Agnon, A., Rockwell, T.K., and<br />

Boas, A., 1998, Crusader castle torn apart by earthquake<br />

at dawn, 20 May 1202, Geology 26, 303–306.<br />

Daeron, M., Benedetti, L., Tapponnier, P., Sursock, A., and<br />

Finkel, R.C., 2004, Constraints on the post 25-ka slip rate<br />

of the Yammouneh fault (Lebanon) using in situ<br />

cosmogenic 36 Cl dating of offset limestone-clast fans,<br />

Earth Planet. Sci. Lett., 227, 105–119.<br />

Ferry, M., M. Meghraoui, N. Abou Karaki, M. Al-Taj, H.<br />

Amoush, S. Al-Dhaisat, and M. O. Barjous, 2007, A 48-<br />

kyr-long slip rate history for the Jordan Valley segment of<br />

the Dead Sea Fault, Earth Planet. Sci. Lett., 260, 394–<br />

406.<br />

Freund, R., Zak, I., and Garfunkel, Z., 1968, Age and rate of<br />

the sinistral movement along the Dead Sea Rift, Nature,<br />

220, 253– 255.<br />

Garfunkel, Z., Zak, I., and Freund, R., 1981, Active faulting<br />

in the Dead Sea rift, Tectonophysics, 80, 1-26.<br />

Le Béon, M., Klinger, Y., Amrat, A. Q., Agnon, A., Dorbath,<br />

L., Baer, G., Ruegg, J. C., Charade, O., and Mayyas, O.,<br />

2008, Slip rate and locking depth from GPS profiles<br />

across the southern Dead Sea Transform, J. Geophys.<br />

Res., 113, B11403, doi:10.1029/2007JB005280.<br />

Le Beon, M., Klinger, Y., Al Qaryouti, M., Mériaux, A-S,<br />

Finkel, R. C., Elias, A., Mayyas, O., Ryerson, F. J., and<br />

Tapponnier, P., 2010, Early Holocene and Late<br />

Pleistocene slip rates of the southern Dead Sea Fault<br />

determined from 10 Be cosmogenic dating of offset alluvial<br />

deposits, J. Geophys. Res., 115, B11414,<br />

doi:10.1029/2009JB007198.<br />

Marco, S., Rockwell, T.K., Heimann, A., Frieslander, U.,<br />

Agnon, A., 2005, Late Holocene activity of the Dead Sea<br />

Transform revealed in 3D paleoseismic trenches on the<br />

Jordan Gorge segment. Earth Planet. Sci. Lett., 234,<br />

189–205.<br />

Meghraoui, M., et al., 2003, Evidence for 830 years of<br />

seismic quiescence from palaeoseismology,<br />

archaeoseismology and historical seismicity along the<br />

Dead Sea Fault in Syria, Earth Planet. Sci. Lett., 210, 35–<br />

52.<br />

Schwartz, D. P., and K. J. Coppersmith, 1984, Fault<br />

behavior and characteristic earthquakes: Examples from<br />

the Wasatch and San Andreas fault zones, J. Geophys.<br />

Res., 89, 5681–5698.<br />

Sieh, K., 1996, The repetition of large-earthquake ruptures,<br />

Proc. Natl. Acad. Sci. U.S.A. 93, 3764-3771.<br />

Wesnousky, S.G., 2008, Displacement and geometrical<br />

characteristics of earthquake surface ruptures: Issues<br />

and implications for seismic-hazard analysis and the<br />

process of earthquake rupture, Bull. Seismol. Soc. Am.<br />

98(4), 1609-1632.<br />

Acknowledgements: Thanks are due to all the trenching<br />

assistants, especially to K. Farrington, J. B. Salisbury and<br />

E. Bowles-Martinez. We thank the village of Almagor for<br />

letting us trench in their fields, with special thanks to Avshi<br />

Herzog for his assistance. This project is funded by NSF<br />

grant EAR-1019871 to T. K. Rockwell and by a grant from<br />

the city of Paris to N. Wechsler and Y. Klinger.<br />

270


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 2: Location of first year trenches, on an aerial photograph of the field from summer 2009. Approximate location of previous<br />

trenches is marked with a blue rectangle. The buried paleo-channel locations are marked in the trench by letters, and the youngest<br />

14C calibrated date is given for each. Marco et al. (2005) followed sub-units of channel E to constrain slip for the 1202 and 1759<br />

events. Their CH1, used to estimate the last 5kyr slip-rate, was exposed in T1, and its approximate location is marked on the<br />

photo, just south of our paleo-channel F.<br />

271


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

A TERRESTRIAL CLOSE RANGE VIEW OF THE NORMAL FAULT ZONE NEAR<br />

ARCHANES (EAST YIOUCHTAS MT., HERAKLION BASIN, CRETE)<br />

Wiatr, Thomas (1), Ioannis D. Papanikolau (2, 4, Klaus Reicherter (1), Tomás Fernández-Steeger (3)<br />

(1) RWTH Aachen University, Neotectonics and Natural Hazards Group, Lochnerstr. 4-20, 52056 Aachen, Germany<br />

(2) Laboratory of Mineralogy & Geology, Department of Sciences, Agricultural University of Athens, 75 Iera Odos Str., 11855<br />

Athens, Greece<br />

(3) RWTH Aachen University, Department of Engineering Geology and Hydrogeology, Lochnerstr. 4-20, 52064 Aachen,<br />

Germany<br />

(4) AON Benfield UCL Hazard Research Centre, Department of Earth Sciences, University College London, WC 1E 6BT,<br />

London UK<br />

Abstract (A terrestrial close range remote sensing view of the normal fault zone near Archanes (East Yiouchtas Mt.,<br />

Heraklion basin, Crete): The focus of investigation in this paper is the reconstruction of different fault plane conditions and the<br />

primary interpretation of them, based on terrestrial close range LiDAR (Light Detection and Ranging) data. For reconstruction the<br />

slip per event along the bedrock scarp and where possible (e.g. postglacial scarp) an estimate regarding the slip rate of the<br />

individual faults we used the backscattered signal of the laser beam, the geomorphological geometry and the fault plane<br />

conditions. In this paper we discuss the N-S striking normal fault near Archanes in the Heraklion basin. To the north of this fault<br />

zone the ancient Minoan temple Anemospillia (caves of the wind) is located, which was destroyed by an earthquake around 1700<br />

BC and forms the nearest seismic source to the site.<br />

Key words: close range LiDAR, normal fault, fault scarp morphology, Crete<br />

INTRODUCTION<br />

Crete is the largest Greek island with an area of ~<br />

8300 km² and approx. 900 km coastline. The<br />

development of a multidirectional tectonic regime on<br />

Crete is interpreted as a result of the Hellenic<br />

subduction zone in the south and the westward<br />

extrusion of the Anatolian plate in the north. The<br />

island is forms a horst structure in the Hellenic fore<br />

arc zone, which is also influenced by the roll back of<br />

the African plate. Rapid uplift of ~ 1.2 mm/yr can be<br />

observed on the entire island (Meulenkamp et al.,<br />

1994). Crete has been uplifted since the Middle<br />

Miocene from 1 up to 2 km depending on the<br />

influence of different tectonic blocks. The island of<br />

Crete lies on top of the active subduction zone for<br />

about 30 ma years, implying that it experiences high<br />

strain rates and constant deformation processes<br />

(Papanikolaou, 1993). Crete is characterised by a<br />

complex geological and tectonic structure that results<br />

from: i) the successive thrusting of the alpine<br />

geotectonic units on top of each other (Bonneau,<br />

1984), ii) the activity of major detachment faults<br />

(Fasoulas et al., 1994, Papanikolaou and Vassilakis,<br />

2010, Zachariasse et al., 2011), iii) by the intense<br />

neotectonic and active faulting (Monaco and<br />

Tortorici, 2004, Peterek and Schwarze, 2004, Caputo<br />

et al., 2010). Crete is located in a high seismicity<br />

area. Over the last 40 years the active Hellenic<br />

subduction zone produced earthquakes in a depth<br />

range from 18 to 162 km with magnitudes of M = 4.9<br />

up to 6.1 (Benetatos et al., 2004). Hypocentral<br />

depths of earthquakes showed that the north dipping<br />

Wadati-Benioff seismic zone close to the low angle<br />

subduction along the convex side of the Hellenic arc<br />

trench is located in a depth of around 60 to 90 km<br />

near Crete (Papazachos et al., 2000). But this region<br />

has also experienced strong thrusting<br />

paleoearthquakes with magnitudes up to M > 7.5 -<br />

8.0 and hence one of the most intense seismic<br />

activity area of the Aegean region (Papazachos &<br />

Papazachou, 1997).<br />

Decoding paleoearthquakes in fault bedrock scarps<br />

is important for seismic hazard assessment. Shallow<br />

earthquakes greater than M S 6 can produce an<br />

imprint in the landscape named fault scarps (Stewart<br />

& Hancock, 1990). Bedrock fault scarps are<br />

indicators of large surface faulting events and may<br />

provide not only slip rates, but also information on<br />

slip per events when they are analysed with<br />

cosmogenic isotope dating (Benedetti et al., 2003).<br />

Fault scarps are preserved in the landscape when<br />

the slip rate is greater than the erosion rate.<br />

Therefore, these are regarded as postglacial scarps<br />

that were formed since the last glaciation (Benedetti<br />

et al., 2002). The Neogene fault plane solutions and<br />

the seismic activities indicate large earthquakes and<br />

a rapid uplift with complex tectonic settings of Crete<br />

(Dewey & Sengör, 1979; Papazachos et al., 1987).<br />

This paper is focused on the N-S striking fault zone in<br />

the Heraklion basin to the south of Knossos (Fig.1).<br />

This basin is characterized by block tectonics and the<br />

Yiouchtas Mt. represent a neotectonic horst structure<br />

which is subdivides the Heraklion Basin in a western<br />

and an eastern subbasin (Papanikolaou and<br />

Nomikou 1998). Major aims of the investigation were<br />

to find quantitative and qualitative data for the<br />

reconstruction of surfaces and to analyze the tectonic<br />

geomorphology and paleoseismicity of active faults<br />

with terrestrial laser scanning (TLS) to reconstruct<br />

fault history and activity along surface rupturing<br />

scarps.<br />

The varying scale of structural heterogeneity and<br />

discontinuous geometry of the exhumed foot wall slip<br />

plane along a fault zone and the complexity of the<br />

surface features like the subslip-plane breccia sheet,<br />

brecciated colluvium or frictional water-wear<br />

striations on the rupture plane, makes it difficult to<br />

recognize the paleoevents on the fresh fault scarp<br />

272


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

above the level of exhumation (Stewart & Hancock,<br />

1991; Roberts, 1996). Even, more the detail<br />

geometrical characterisation of the slip surface<br />

depends on the view direction to the strike-slip<br />

direction (parallel or perpendicular), the calculation<br />

approaches and the scaling size of the analysis,<br />

because the anisotropy properties and fractal<br />

dimensions of fault morphology are decisive<br />

(Mandelbrot, 1985; Fardin et al., 2001, 2004;<br />

Rahman et al., 2006; Renard et al., 2006; Sagy et al.,<br />

2007, 2009; Candela et al., 2009).<br />

Additionally, several time-dependent and overlapping<br />

processes influence the condition of the free face<br />

fault plane that become degraded. These processes<br />

involve weathering, pedogenesis of the unbrecciated<br />

colluvium, vegetation, karstification and erosion of<br />

the colluviums and the fault outcrop.<br />

PRIMARY RESULTS OF THE LiDAR<br />

INVESTIGATION ON THE NORMAL FAULT ZONE<br />

NEAR ANEMOSPILIA "CAVES OF THE WIND"<br />

(1900-1700 BC MM II/III) IN THE HERAKLION<br />

BASIN<br />

In the northern part of the Mount Juktas about 7 km<br />

south of Knossos, the legendary area and tomb of<br />

Cretan Zeus is situated, which is considered to be<br />

one of the earliest Minoan temples: Anemospilia<br />

(Fig.1).<br />

The three small rooms, each of them opened into a<br />

corridor, were described and discovered by<br />

Sakellarakis during the 1979 expedition. They<br />

assumed that the temple was destroyed by a<br />

sequence of large earthquakes around 1700 BC and<br />

based this conclusion on pottery and artefacts (Nur,<br />

2008). Furthermore, they found a skeleton with<br />

broken legs under an ash layer implying that the<br />

earthquake was strong enough to damage the<br />

massive temple and was followed by a fire. Five<br />

different close range LiDAR scans were made in the<br />

middle of the N-S striking normal fault 2 km south of<br />

the Anemospilia temple (Fig.2). The free natural fault<br />

plane in the interesting area is around 6 m high and<br />

the scanned area is around 30 m wide. The whole<br />

outcrop is by this location around 70 m continuous<br />

wide. Our primary goal in this study was to use the<br />

TLS for fault tectonomorphology reconstruction. The<br />

TLS data have a point to point range between 3 and<br />

7 mm. The examples in this paper (Fig.3/4) had<br />

around 2.8 and 3.8 million numbers of shots (points)<br />

and the average range between LiDAR and defined<br />

scan window was 10 m. This allowed a spatial<br />

reconstruction of the scan sequence without gaps<br />

and ensured a good data quality and spatial<br />

resolution for the interpolation between the points<br />

and for the analysis of the plane morphology. The<br />

detailed structural analysis of rock surfaces has<br />

shown that the surface conditions are changing from<br />

base to top (Fig.4). The hillshades in figure 3b and 4b<br />

illustrates the plane morphology with different karstic<br />

features and degradation in the upper part of the<br />

scarp plane (variance of rougher surface conditions<br />

in section III, IV, V, VI in Fig.3).<br />

Fig. 1: Investigation area near Archanes in the Heraklion<br />

basin of Crete including earthquakes recorded since<br />

1979 (a) (SRTM, USGS, www.usgs.gov) and the main<br />

faults, the LiDAR position, location and looking direction<br />

of photos (see Fig.2), dip direction of the scanned fault<br />

plane, recent stress field and ancient sites (b) (modified<br />

from Fassoulas, 2001; ten Veen & Meijer, 1998).<br />

273<br />

Fig. 2: Photo of the east dipping continuous normal fault zone<br />

western Archanes and eastern Mount Yiouchtas within the<br />

ancient Minoan site Anemospilia in the northern part of<br />

Mount Juktas (a). b) Zoom in photo of the fault plane.<br />

Parts I and II of this section have no significant<br />

karstification. The sections are dominated by<br />

striation, slickensides and small fractures (compare<br />

Fig.4). The combination of the plane morphology with<br />

the detected backscattered signal of the LiDAR is<br />

shown in figure 3d/3e and 4d. By using this<br />

technique, a different point of view allowed that the<br />

detected near infrared laser signal can be used for<br />

the classification of different functions of weathering,<br />

morphological and erosion features on the fault plane


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

and also reveals the exhumation history. The signal<br />

intensity can be used to identify vegetation (including<br />

lichen) and its influence on the fault plane. The<br />

results shown the influence of the colluvium on the<br />

base of the scarp (I in Fig.3e), which can be detected<br />

by a change of the backscattered intensity.<br />

Furthermore it is possible to distinguish between the<br />

different parts of the fault plane, which are<br />

characterised and dominated by 1) degradation; 2)<br />

karstification; 3) slickensides; and 4) the influence of<br />

the colluvium. Following Giaccio (2002) and the<br />

model of natural free normal bedrock scarp, we<br />

identified different weathering mircomorphologies<br />

depending on the scarp height (Fig.4e). The different<br />

sections with determinates features on the fault plane<br />

represent the time-dependent fault scarp evolution<br />

(Stewart, 1996).<br />

Fig.4: a) photo of the fault plane includes the LiDAR scan<br />

window, b) hillshade of the fault plane for geometrical and<br />

morphological analysis, c) dip direction of the plane, d)<br />

distribution of backscattered signal with primary surface<br />

interpretation, e) brown, red, green and yellow boxes are<br />

demonstrate the different surface conditions on the fault<br />

plane in depending on the scarp height.<br />

Fig. 3: Primary results of the LiDAR investigation based<br />

on the model in Fig.1. a) photo of the fault plane includes<br />

the LiDAR scan window, b) hillshade of the fault plane for<br />

geometrical and morphological analysis, c) dip direction of<br />

the plane, d) distribution of the backscattered signal with<br />

primary surface interpretation, e) primary interpretation of<br />

all fault plane conditions.<br />

The time- and height-dependent features of bedrock<br />

fault scarps are shown figure 4e. The boxes illustrate<br />

examples of the fault plane morphology from bottom<br />

to top (young to old; brown, red, green, yellow).<br />

Conspicuous is the increasing roughness from young<br />

to old (brown to yellow) and the specific surface<br />

features in different heights. The reason could be the<br />

different bio-karstic, bio-erosional, physical and<br />

biochemical processes which depend on time<br />

(Giaccio et al., 2002). The brown box shows the<br />

striation in the lower part of the fault. The red box<br />

illustrates small fractures and a rougher surface than<br />

in the brown box. The green box the gradation of the<br />

karstification and the yellow box show the rillen karst<br />

in an advanced stage. Close range LiDAR<br />

investigation on postglacial natural normal fault<br />

scarps has shown that reconstruction of the spatial<br />

distribution of different plane evolution indicators is<br />

possible.<br />

These fundamental phenomena can be realized by<br />

and imaged with a high resolution digital elevation<br />

model (HRDEM) in combination with the<br />

backscattered laser impulse. The primary<br />

interpretations of the fault surface conditions and<br />

their interaction are described in figure 3. Based on<br />

the LiDAR results and the field survey we created a<br />

principal model of fault scarp alteration for the East<br />

Yiouchtas fault, following the general model of<br />

Giaccio et al. 2002 (Fig. 5).<br />

Fig. 5: The model of a normal fault with fault plane<br />

evolution indicators (modified from Giaccio et al.<br />

2002).<br />

274


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The investigation with LiDAR and the field survey has<br />

shown that the eastern Yiouchtas fault is active with<br />

several events in the post-glacial time period.<br />

References<br />

Benedetti, L., Finkel, R., King, G., Armijo, R.,<br />

Papanastassiou, D., Ryerson, F., Flerit, F., Farber, D. &<br />

Stavrakakis, G., 2003: Motion on the Kaparelli fault<br />

(Greece) prior to the 1981 earthquake sequence<br />

determined from 36Cl cosmogenic dating. Terra Nova,<br />

15:118-124.<br />

Benedetti, L., Finkel, R., Papanastassiou, D., King, G.,<br />

Armijo, R., Ryerson, F., Farber, D. & Flerit, F., 2002:<br />

Post-glacial slip history of the Sparta Fault (Greece)<br />

determined by 36 Cl cosmogenic dating: evidence for nonperiodic<br />

earthquakes. Geophys. Res. Lett., 29: 87-1–87-<br />

4.<br />

Bonneau, M., 1984: Correlation of the Hellinides nappes in<br />

the south-eastern Aegean and their tectonic<br />

reconstruction. Special Publication of the Geological<br />

Society of London 17, 517-527<br />

Benetatos, C., Kiratzi, A., Papazachos, C., & Karakaisis G.,<br />

2004: Focal mechanisms of shallow and intermediate<br />

depth earthquakes along the Hellenic Arc. J. Geodyn.,<br />

37, 253-296, doi:10.1016/j.jog.2004.02.002.<br />

Candela, T., Renard, F., Bouchon, M., Brouste, A., Marsan,<br />

D., Schmittbuhl, J. & Voisin, C., 2009: Characterization of<br />

fault roughness at various scales: implications of threedimensional<br />

high resolution topography measurements,<br />

Pure appl. geophys. doi: 10.1007/s00024-009-0521-2.<br />

Caputo, R., Catalano, S., Monaco, C., Romagnoli, G.,<br />

Tortorici G. & Tortorici, L., 2010: Active faulting on the<br />

island of Crete (Greece). Geophys. J. Int., 183, 111–126,<br />

doi: 10.1111/j.1365-246X.2010.04749.x.<br />

Dewey, J.F. & Sengör, A.M.C., 1979: Aegean and<br />

surrounding regions: Complex multiplate and continuum<br />

tectonics in a convergent zone. Bull. geol. Soc. Am., 90,<br />

84-92.<br />

Fardin, N., Stephansson, O. & Jing, L., 2001: The scale<br />

dependence of rock joint surface roughness. Int. J. Rock<br />

Mech. & Min. Sci., 38, p. 659-669.<br />

Fardin, N., Feng, Q. & Stephansson, O., 2004: Application<br />

of a new in situ 3D laser scanner to study the scale effect<br />

on the rock joint surface roughness. Int. J. Rock Mech. &<br />

Min. Sci., 41, p. 329-335.<br />

Fassoulas, C., 2001: The tectonic development of a<br />

Neogene graben at the leading edge of the active<br />

European margin, The Heraklion basin, Crete, Greece, J.<br />

Geodyn., 31, 49–70, doi:10.1016/S0264-3707(00)00017-<br />

X.<br />

Giaccio, B., Galadini, F., Spasato, A., Messina, P., Moro,<br />

M., Zreda, M., Cittadini, A., Salvi, S. & Todero, A., 2002:<br />

Image processing and roughness analysis of exposed<br />

bedrock fault planes as a tool for paleoseismological<br />

analysis: results from the Campo Felice fault (central<br />

Apennines, Italy). Geomorphology, 49, 281-301.<br />

Mandelbrot, B.B., 1985: Self-affine fractals and fractal<br />

dimension, Phys. Scripta, v. 32, p. 257-260.<br />

Meulenkamp, J.E., van der Zwaan, G.J., van Wamel, W.A.,<br />

1994. On late Miocene to recent vertical motions in the<br />

Cretan segment of the Hellenic arc. Tectonophysics, 234,<br />

53-72.<br />

Monaco, C. & Tortorici, L., 2004: Faulting and effects of<br />

earthquakes on Minoan archaeological sites in Crete<br />

(Greece). Tectonophysics, 382, 103-116.<br />

Nur, A., Burgess, D., 2008: Apocalypse: Earthquakes,<br />

Archaeology and the Wrath of God. Princeton University<br />

Press. Princeton and Oxford. 309 p.<br />

Papanikolaou, D. (1993). Geotectonic evolution of the<br />

Aegean. Bull. Geol. Soc. Greece, 28/1, 33-48.<br />

Papnikolaou, D.J. & Nomikou, P.V., 1998: Neotectonic<br />

blocks and planation surfaces in Iraklion basin, Crete,<br />

Greece. Bull. Geol. Soc. Greece, XXXII/1, 231-239.<br />

Papanikolaou, D. and Vassilakis, E. (2010). Thrust faults<br />

and extensional detachment faults in Cretan tectonostratigraphy:<br />

Implications for Middle Miocene extension.<br />

Tectonophysics 488, 233-247<br />

Papazachos, B.C., Papadimitriou, E.E., Kiratzi, A.A.,<br />

Papaioannou, C.A. & Karakaisis, G.F., 1987: Probabilities<br />

of Occurrence of Large Earthquakes in the Aegean and<br />

Surrounding Area During the Period 1986-2006.<br />

Pageoph., 125, 4, 597-612.<br />

Papazachos, B. & Papazachou, C., 1997: The Earthquakes<br />

of Greece. Zitis, Thessaloniki.<br />

Papazachos, B.C., Karakostas, V.G., Papazachos, C.B. &<br />

Scordilis, E.M., 2000: The geometry of the Wadati–<br />

Benioff zone and lithospheric kinematics in the Hellenic<br />

arc. Tectonophysics, 319, 275-300.<br />

Peterek, A. & Schwarze, J., 2004: Architecture and Late<br />

Pliocene to recent evolution of outer-arc basins of<br />

Hellenic subduction zone (south-central Crete, Greece).<br />

J. Geodyn., 38, 19-55.<br />

Rahman, Z., Slob, S. & Hack, R., 2006: Deriving roughness<br />

characteristics of rock mass discontinuities from<br />

terrestrial laser scan data. <strong>Proceedings</strong> of the 10th IAEG<br />

Congress, Engineering geology for tomorrow's cities,<br />

Nottingham, United Kingdom, 6-10 September 2006.<br />

Engineering geology for tomorrow's cities.<br />

Renard, F., Voisin, C., Marsan, D. & Schmittbuhl, J., 2006:<br />

High resolution 3D laser scanner measurements of a<br />

strike-slip fault quantify its morphological anisotropy at all<br />

scales, Geophys. Res. Lett. 33, L04305, doi:<br />

10.1029/2005GL025038.<br />

Roberts, G.P., 1996: Variation in fault-slip directions along<br />

active and segmented normal fault systems. J. Struct.<br />

Geol., Vol. 18, No. 6, 835-845.<br />

Sagy, A., Brodsky, E.E. & Axen, G.J., 2007: Evolution of<br />

fault-surface roughness with slip. Geology, 35: 283-286.<br />

Sagy, A. & Brodsky, E.E., 2009: Geometric and rheological<br />

asperities in an exposed fault zone. J. Geophys. Res.,<br />

114, B02301, doi: 10.1029/2008JB005701.<br />

Sakellarakis, Y. & Sapouna-Sakellaraki, E., 1981: Drama of<br />

death in a Minoan temple. National Geographic, February<br />

1981.<br />

Stewart, I.S. & Hancock, P.L., 1990: Brecciation and<br />

fracturing within neotectonic normal fault zones in the<br />

Aegean region. In: Knipe, R.J. and Rutter, E.H., (eds.)<br />

Deformation Mechanisms, Rheology and Tectonics.<br />

Geol. Soc. Spec. Publ., 54: 105-112.<br />

Stewart, I.S. & Hancock, P.L., 1991: Scales of structural<br />

heterogeneity within neotectonic normal fault zones in the<br />

Aegean region. J. Struct. Geol., Vol. 13, No. 2, 191-204.<br />

Stewart, I., 1996: A rough guide to limestone fault scarps. J.<br />

Struct. Geol., 18,1259-1264.<br />

ten Veen, J.H. & Meijer, P.T., 1998: Late Miocene to Recent<br />

tectonic evolution of Crete (Greece): geological<br />

observations and model analysis. Tectonophysics, 298,<br />

191-208.<br />

Zachariasse, W.J., van Hinsbergen, D.J.J., & Fortuin, A.R.,<br />

2011: Formation and fragmentation of a late Miocene<br />

supradetachment basin in central Crete: implications for<br />

exhumation mechanisms of high-pressure rocks in the<br />

Aegean forearc. Basin Research, doi: 10.1111/j.1365-<br />

2117.2011.00507.x.<br />

275


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE DISCONTINUITY OF A CONTINUOUS FAULT: DELPHI (GREECE)<br />

Wiatr, Thomas (1), Klaus Reicherter (1), Ioannis D. Papanikolau (2, 3) & Tomás Fernández-Steeger (4)<br />

(1) RWTH Aachen University, Neotectonics and Natural Hazards Group, Lochnerstr. 4-20, 52056 Aachen, Germany<br />

(2) Laboratory of Mineralogy & Geology, Department of Sciences, Agricultural University of Athens, 75 Iera Odos Str., 11855<br />

Athens, Greece. papanikolaou@ucl.ac.uk<br />

(3) AON Benfield UCL Hazard Research Centre, Department of Earth Sciences, University College London, WC 1E 6BT,<br />

London UK. papanikolaou@ucl.ac.uk<br />

(4) RWTH Aachen University, Department of Engineering Geology and Hydrogeology, Lochnerstr. 4-20, 52064 Aachen,<br />

Germany. fernandez-steeger@lih.rwth-aachen.de<br />

Abstract (The discontinuity of a continuous fault: Delphi (Greece): We used a terrestrial laser scanning system for the<br />

reconstruction and analysis of the morphotectonic features on a fault segment near the ancient Delphi, Greece. Delphi is located<br />

on the northern part of the Gulf of Corinth and embedded in a seismic landscape and is the major onshore fault north of the<br />

Corinth Gulf. This paper concentrates on the LiDAR long range investigation of the fault for generating a digital elevation model.<br />

The model was used for estimating the natural fault plane height by using several vertical profiles. The results show a high<br />

horizontal and vertical variability on a 120 m long fault scarp.<br />

Key words: long range LiDAR, discontinuous fault, Corinth Gulf, Delphi<br />

INTRODUCTION<br />

The ancient Delphi with its oracle dedicated to the<br />

god Apollo, was the most popular place of worship<br />

around 700 B.C. to 400 A.D. in ancient Greece.<br />

Delphi is located at the southern flank of Mount<br />

Parnassus that is compiled by thick-bedded neritic<br />

Mesozoic limestones with significant bauxite<br />

deposits. The mountain range is situated on the<br />

northern coast of the Gulf of Corinth and bounded by<br />

an active fault zone that dips southwards (Fig.1).<br />

Ambraseys & Jackson, 1998; Papazachos &<br />

Papazachou, 2003 and Pavlides & Caputo, 2004).<br />

Piccardi (2000) described the 373 B.C. earthquake,<br />

which partly destroyed the ancient Delphi showing<br />

that there is a post-earthquake reconstruction phase<br />

at the shrine of Athena (located around 500 m east of<br />

the Temple of Apollo) (Fig.2).<br />

Fig. 2: Setting of the archaeological site of the shrine<br />

of Athena Pronaia (modified from Piccardi, 2000).<br />

Fig. 1: Tectonic and topographic overview map of the<br />

Delphi area.<br />

The Gulf of Corinth is a graben like tectonic structure<br />

with E-W trending normal faults and characterized<br />

one of the fastest extending regions worldwide with<br />

up to 20 mm/yr rate (Billiris et al., 1991; Briole et al.,<br />

2000). Destructive historical earthquakes in the area<br />

are reported for 373 B.C. (Piccardi, 2000), 515 A.D.<br />

(Ambraseys & Jackson, 1998; Papazachos &<br />

Papazachou, 2003) and 1870 (1 st of August, Ms=6.7,<br />

Furthermore, he postulates that the Temple of<br />

Athena was relocated from its original position.<br />

Aim of our study was the reconstruction of the<br />

morphotectonic features of the Delphi fault scarp,<br />

located 2 km west of the Temple of Apollo by using a<br />

terrestrial remote sensing technique. We scanned a<br />

120 m long fault scarp with long range LiDAR. We<br />

used a terrestrial LiDAR (Light Detection and<br />

Ranging) system from Optech Inc.<br />

METHODS<br />

The ground-based LiDAR (Light Detection And<br />

Ranging) or TLS (terrestrial laser scanning) remote<br />

sensing method has been established as a versatile<br />

data acquisition tool in photogrammetry, engineering<br />

technologies, atmospheric studies and as a good<br />

data acquisition tool in geosciences and geological<br />

engineering in difficult accessible areas.<br />

276


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

As the TLS has a high spatial and temporal<br />

resolution it is an effective remote sensing<br />

technology for reconstruction, monitoring and<br />

observation of geosciences phenomena.<br />

The fundamental principle of ground-based LiDAR is<br />

to generate coherent a laser beam with little<br />

divergence by stimulated emission. LiDAR is a<br />

contact- and destructionless, non-penetrative active<br />

recording system which is stationary during the<br />

recording. The electromagnetic waves are reflected<br />

by surfaces and the receiver detects portions of the<br />

backscattered signal. All scan sequences were<br />

mapped with first pulse detection mode. The laser<br />

ranging system is based on measuring the time-offlight<br />

(two-way travel time) of the short wave laser<br />

signal. Advantages of the terrestrial method are the<br />

flexible handling, a relatively quick availability of an<br />

actual dataset, and a very high spatial resolution of<br />

the object with information about intensity, x-y-zcoordinates<br />

and range. The combination with a digital<br />

camera allows combining the point cloud with<br />

panchromatic information in order to achieve<br />

additionally the RGB colour-coding.<br />

For the morphological analysis twelve vertical profiles<br />

with 10 m distance to each other were generated<br />

from the LiDAR data. All profiles start on the same<br />

height level on top of the fault zone on the foot wall<br />

and are perpendicular to the fault plane towards the<br />

hanging wall.<br />

Results show a variation of the free face fault height<br />

with a difference of 5.4 m. The height data range of<br />

the natural bedrock fault plane is between 4 and 9.4<br />

m. Furthermore, horizontal variations of 3 m were<br />

detected (Fig.5).<br />

Laser scanning allows 3D surface data acquisition,<br />

which is specifically characterized by a digital data<br />

record and a computerized data analysis.<br />

Furthermore, an implementation of the dataset in a<br />

geographical information system (GIS) is<br />

uncomplicated with accurate digital elevation models<br />

(DEM) or digital terrain models (DTM) sourced<br />

directly from the raw dataset.<br />

The infrared laser scanner detected the<br />

monochromatic information of the backscattered<br />

intensity in 256 grey values. The information of the<br />

monochromatic wavelength, the detected<br />

backscattered intensity, reflexes the surface<br />

properties in the near infrared range. This<br />

wavelength is invisible for human eyes. Hence, the<br />

results show a different kind of view of the surface<br />

conditions. The quality of the reflection depends on<br />

the inclination angle of the laser beam, the range<br />

between the object and scanner, the material, the<br />

colour, the surface condition (weathering/roughness),<br />

and the spatial resolution.<br />

DISCONTINUOUS FAULT SEGMENT ON A<br />

CONTINUOUS FAULT NEAR DELPHI<br />

The scan position for the long range investigation<br />

was 250 m south of the fault plane (Fig.3). The raw<br />

data point cloud includes around 6.2 million points for<br />

an 130 m long and 60 m wide scan window. For this<br />

study a point resolution of 2 cm was chosen.<br />

Moreover, we recorded 9 close range scans with<br />

4 mm point resolution (Fig.4).<br />

After data validation and cleaning, the point cloud<br />

has been geo-referenced and imported into a GIS. In<br />

the GIS, the scans have been converted in a<br />

triangulated irregular network (TIN) and in a raster<br />

format. With the grid format it is possible to calculate<br />

the basic applications for morphologic specifications.<br />

Fig. 3: Panchromatic images of the investigation area.<br />

a) The geoeye satellite image shows the continuous<br />

fault by Delphi and includes the long range LiDAR<br />

study area (modified from Google earth). b) Photo of<br />

the south dipping continuous fault zone 2 km westward<br />

from the ancient shrine of Apollo by Delphi.<br />

We found that the Delphi fault has an oscillation in<br />

three dimensions (horizontal and vertical<br />

displacement), across the 120 m long scan<br />

sequence. The absolute elevation range of all profiles<br />

on the natural free face fault surface (vertical range<br />

of the fault plane) is between 574.5 m and 588.7 m<br />

above sea level (see Fig.5 right site number 1) and<br />

the horizontal variation is between 11.2 m and 21 m<br />

(see Fig.5 number 1 under the profiles). The<br />

horizontal and vertical value range (number 1)<br />

includes the variation of the knick point on top of the<br />

free face to the foot wall (number 2) and the knick<br />

point of the bottom of the free face to the hanging<br />

wall (number 3). Furthermore, the results had shown<br />

a variability of the alluvial deposits (vertical and<br />

horizontal displacement) of the hanging wall (number<br />

3).<br />

It turns out that the long term slip rate for postglacial<br />

scarps (last 19 ka ± 3 ka) in this case ranges<br />

between 0.21 ± 0.04 mm/yr and 0.49 ± 0.09 mm/yr<br />

277


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

depending on the fault segment analysed, and<br />

between 0.3 and 0.72 mm/yr for the last 13 ka (The<br />

13 ka pertains to Benedetti et al., 2002; 2003).<br />

Hence, the long term slip rate had a range between<br />

0.25 ± 0.05 mm/yr and 0.58 ± 0.09 mm/yr for the last<br />

16 ka ± 3ka. This implies for the long term slip rate<br />

estimation a variation of 0.28 mm/yr (19 ka ± 3 ka<br />

(0.4 mm/yr worst case scenario)) and 0.42 mm/yr (13<br />

ka), which is enormous.<br />

CONCLUSION<br />

Post-glacial throw variation along strike is evident,<br />

even over short distances, including locations that<br />

are apparently undisturbed by incision or deposition<br />

processes (Papanikolaou et al., 2005). This is the<br />

natural variation associated with coseismic surface<br />

slip and constitutes a major source of uncertainty that<br />

in central Apennines was measured at ± 20%<br />

(Papanikolaou et al., 2005). A fairly irregular surface<br />

slip distribution has been documented from several<br />

normal faulting events such as the 1981 Alkyonides<br />

earthquake sequence (Jackson et al., 1982). Herein,<br />

in Delphi we measured a higher variability of around<br />

± 28 %. Processes like erosion, deposits and debris<br />

as well as catchments analysis are not including in<br />

this research until now.<br />

But the LiDAR investigation on a fault scarp segment<br />

in Delphi has shown a massive variation in the<br />

vertical and horizontal displacement. The data<br />

collection for calculation of the long term slip rate with<br />

the profile method, it is necessary to produce a lot of<br />

profiles in the field, to get an impression of the fault<br />

variation. Due to the digital elevation model of the<br />

long range LiDAR data with 2cm spatial resolution<br />

and without gaps it is possible to get a high data<br />

quality for morphological analysis and ensured a<br />

documentation in a tectonic environment.<br />

Fig. 4: Close and long range terrestrial LiDAR<br />

investigation for morphological analysis on the<br />

southwards dipping continuous bedrock fault scarp<br />

2 km west of the archaeological site of Delphi (a). b)<br />

Illustrates the calculation of the largest postglacial<br />

vertical displacement in the scan window on one<br />

profile. c) Fault plane dip direction derived from<br />

structural mapping in the scan area. d) Fault plane<br />

reconstruction based on LiDAR data was realised by<br />

a hillshade in a GIS environment. e) Photo of the fault<br />

plane including the LiDAR scan window. f) Hillshade<br />

of the close range fault plane and g) distribution of<br />

backscattered signal with primary surface<br />

interpretations.<br />

Fig. 5: Twelve perpendicular profiles from the 120 m<br />

long Delphi scarp segment show the horizontal and<br />

vertical discontinuity of the fault plane. 1) Variations of<br />

the free face fault plane. 2) Top knick point of the free<br />

face to the foot wall. 3) Bottom knick point of the free<br />

face to the hanging wall.<br />

References<br />

Ambraseys, N.N. & Jackson, J.A., 1998: Faulting<br />

associated with historical and recent earthquakes in the<br />

Eastern Mediterranean region. Geophysical Journal<br />

International, 133, 390–406.<br />

278


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Benedetti, L., Finkel, R., Papanastassiou, D., King, G.,<br />

Armijo, R., Ryerson, F., Farber, D. & Flerit, F., 2002:<br />

Post-glacial slip history of the Sparta Fault (Greece)<br />

determined by 36 Cl cosmogenic dating: evidence for nonperiodic<br />

earthquakes. Geophys. Res. Lett., 29: 87-1–87-<br />

4.<br />

Benedetti, L., Finkel, R., King, G., Armijo, R.,<br />

Papanastassiou, D., Ryerson, F., Flerit, F., Farber, D. &<br />

Stavrakakis, G., 2003: Motion on the Kaparelli fault<br />

(Greece) prior to the 1981 earthquake sequence<br />

determined from 36Cl cosmogenic dating. Terra Nova,<br />

15:118-124.<br />

Billiris, H., Paradissis, D., Veis, G., England, P.,<br />

Featherstone, W., Parsons, B., Cross, P., Rands, P.,<br />

Rayson, M., Sellers, P., Ashkenazi, V., Davison, M.,<br />

Jackson, J. & Ambraseys, N., 1991: Geodetic<br />

determination of tectonic deformation in central Greece<br />

from 1900 to 1988. Nature, 350, 124– 129,<br />

doi:10.1038/350124a0.<br />

Briole, P., A. Rigo, H. Lyon-Caen, J. C. Ruegg, K.<br />

Papazissi, C. Mitsakaki, A. Balodimou, G. Veis, D.<br />

Hatzfeld, & Deschamps, A., 2000: Active deformation of<br />

the Corinth rift, Greece: Results from repeated Global<br />

Positioning System surveys between 1990 and 1995, J.<br />

Geophys. Res., Solid Earth, 21, 25, 605–625.<br />

Jackson, J., Gagnepain, A., Houseman, J., King, G.,<br />

Papadimitriou, G.C.P., Soufleris, P. & Virieux, C., 1982:<br />

Seismicity, normal faulting, and the geomorphological<br />

development of the Gulf of Corinth (Greece): the Corinth<br />

earthquakes of February and March 1981. Earth and<br />

Planetary Science Letters 57, 377– 397.<br />

Papanikolaou, I.D., Roberts, G.P. & Michetti, A.M., 2005:<br />

Fault scarps and deformation rates in Lazio-Abruzzo,<br />

Central Italy: comparison between geological fault sliprate<br />

and GPS data. Tectonophysics 408, 147–176.<br />

Papazachos, C. & Papazachou, C., 2003: The Earthquakes<br />

of Greece. Ziti, Thesaloniki [in Greek].<br />

Pavlides, S. & Caputo, R. 2004. Magnitude versus faults’<br />

surface parameters: quantitative relationships from the<br />

Aegean Region. Tectonophysics, 380, 159–188.<br />

Piccardi, L., 2000: Active faulting at Delphi: seismotectonic<br />

remarks and a hypothesis for the geological environment<br />

of a myth. Geology, 28, 651–654.<br />

279


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

POSTSEISMIC DEFORMATION OF THE 2009 L’AQUILA EARTHQUAKE (M6.3) SURFACE<br />

RUPTURE MEASURED USING REPEAT TERRESTRIAL LASER SCANNING<br />

Wilkinson, Maxwell (1, Ken McCaffrey (1), Gerald Roberts (2), Patience Cowie (3), Richard Phillips (4)<br />

(1) Department of Earth Sciences, Durham Univerity, United Kingdom. Email: maxwell.wilkinson@durham.ac.uk<br />

(2) School of Earth Sciences, Birkbeck College, University of London, United Kingdom.<br />

(3) School of Geosciences, University of Edinburgh, United Kingdom.<br />

(4) Institute of Geophysics and Tectonics, University of Leeds, United Kingdom<br />

Abstract (Postseismic deformation of the 2009 L’ Aquila Earthquake (M6.3) surface rupture measured using repeat<br />

terrestrial laser scanning): We conducted an innovative survey using repeat terrestrial laser scan (TLS) technology at four sites<br />

on the surface rupture of the 2009 L’Aquila earthquake (M6.3), Central Italy. Between 8 – 126 days after the earthquake we<br />

repeatedly laser scanned four road sections cross-cut and vertically offset by the surface rupture. A method was developed to<br />

quantify the postseismic deformation. We modelled rupture afterslip and associated near-field postseismic deformation in the<br />

hangingwall at each site with millimetre to sub-centimetre precision. The observed postseismic deformation coincides with the<br />

coseismic slip deficit within the fault zone, which we suggest is the driving mechanism for afterslip and near-field postseismic<br />

deformation. Repeat TLS survey of actively deforming surface ruptures provide a new method to monitor and quantify postseismic<br />

deformation.<br />

Key words: Postseismic deformation, Laser scanning.<br />

INTRODUCTION<br />

We report the use of Terrestrial Laser Scan (TLS)<br />

technology to monitor near field postseismic<br />

deformation at four sites along the surface rupture of<br />

the 6th April 2009 L’Aquila earthquake, in the<br />

Abruzzo region, Apennines, central Italy. The main<br />

shock (Mw 6.3) occurred at 03:32 local time. The city<br />

of L’Aquila and its surrounding suburbs were<br />

subjected to the largest intensity shaking, resulting in<br />

308 deaths, 1,500 injuries and over 50,000 people<br />

made homeless.<br />

discontinuous surface rupture, ~ 12 km in length<br />

(Falcucci et al. 2009) along the Paganica fault with<br />

normal sense displacement, down thrown to the SE,<br />

with a maximum throw of 0.1 m as defined by the<br />

EMERGEO working group (2009) and Vittori et al. (in<br />

press).<br />

Fig. 1: Riegl LMS-z420i laser scanner set up at site PAG,<br />

one of the four surface rupture study sites.<br />

The source of the L’Aquila seismicity was identified<br />

as the Paganica fault, to the NE of the city of<br />

L’Aquila, defined by focal mechanisms, aftershock<br />

distribution (Chiarabba et al. 2009) and differential<br />

interferometry. The earthquake created a<br />

280<br />

Fig. 2: Modelled postseismic deformation and rupture<br />

afterslip at PAG for the various time intervals. (From<br />

Wilkinson et al. 2010)<br />

Between 8 – 124 days after the earthquake we<br />

repeatedly laser scanned four road sections crosscut<br />

and vertically offset by the surface rupture (Fig.<br />

1). By comparing each subsequently acquired<br />

dataset to the first at each site, we were able to


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

model the resultant near-field postseismic<br />

deformation and rupture afterslip over various time<br />

intervals with millimetre to decimetre precision. We<br />

used reflective targets to measure the horizontal<br />

component of deformation.<br />

DISCUSSION<br />

We observe progressive surface deformation in our<br />

modelled datasets whose rate decreases over time.<br />

We present the modelled data from site PAG in figure<br />

2 as this is site with the greatest magnitude of<br />

deformation. The three equations shown in figure 3<br />

are used to describe afterslip and postseismic<br />

deformation for three different studies of the<br />

Guatemala (1) and Superstition Hills (2 & 3)<br />

earthquake surface ruptures. The three equations<br />

differ slightly as they were derived using a mix of<br />

theoretical and empirical approaches, as well as<br />

datasets from different earthquakes or study sites. In<br />

figure 4 we compare these three equations with their<br />

earthquake specific parameters to our measurements<br />

for the L’Aquila earthquake derived from figure 2. We<br />

find that our data, with a decaying rate over time is<br />

consistent with previously published theoretical and<br />

empirical laws derived to explain afterslip<br />

phenomenon.<br />

At the study site PAG, we observe surface<br />

deformation of significant magnitude in the form of a<br />

developing syncline in the hangingwall, up to 30 m<br />

from the surface rupture. We interpret this<br />

deformation as the signal of shallow afterslip in the<br />

fault zone. Unpublished data from the three<br />

remaining study sites, plus a supplementary dataset<br />

of total station line of sight measurements from a fifth<br />

site show similar results, confirming postseismic<br />

deformation along the surface rupture is attributable<br />

to afterslip within the fault zone.<br />

We note our study sites experienced significant<br />

postseismic deformation, and are located above a<br />

zone of coseismic slip deficit within the fault zone<br />

(Fig. 5, after Cheloni et al. 2010). The correlation<br />

between a coseismic slip deficit in the fault zone and<br />

significant postseismic deformation at the surface<br />

suggests the coseismic slip deficit as the driving<br />

mechanism for afterslip and near-field deformation.<br />

Fig. 3: Theoretical and empirical afterslip models with<br />

parameters obtained from afterslip datasets of previous<br />

earthquakes.<br />

Fig. 4: Surface motion measurements of rupture throw<br />

(purple diamonds), hangingwall syncline subsidence (pink<br />

squares) and extension between reflective targets (green<br />

triangles), derived from the modelled data in Fig 2. Our<br />

measurements are compared to data for the three<br />

theoretical and empirical models for afterslip shown in figure<br />

3. (Figure from Wilkinson et al. 2010)<br />

Fig. 5: Coseismic fault slip map derived from coseismic<br />

GPS data. The contours represent coseismic slip (m). Note<br />

the location of study site PAG at the surface and the<br />

coseismic slip deficit (white patch) within the fault zone<br />

beneath it. (Figure adapted from Cheloni et al, 2010)<br />

CONCLUSIONS<br />

Repeat TLS survey of actively deforming surface<br />

ruptures provide a new method to monitor and<br />

quantify postseismic deformation. We have<br />

measured near-field postseismic deformation at four<br />

sites along the surface rupture of the L’Aquila<br />

earthquake. We observe rupture afterslip and<br />

hangingwall deformation whose decay in rate over<br />

time is consistent with field observations from other<br />

earthquakes, as well as theoretical and empirical<br />

laws derived to explain such phenomenon. We<br />

interpret the near-field postseismic deformation as<br />

the signal of afterslip within the fault zone. We note<br />

the correlation between the location of our sites<br />

experiencing significant postseismic deformation and<br />

a coseismic slip deficit within the fault zone beneath<br />

them. We suggest that the coseismic slip deficit is the<br />

driving mechanism for near-field postseismic<br />

deformation and afterslip within the fault zone.<br />

281


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Acknowledgements: Funded by NERC grants<br />

NE/H003266/1 and NE/E016545/1 and Durham University<br />

Doctoral Fellowship (M. Wilkinson). Thanks to A. Michetti,<br />

E. Vittori, L. Guerrieri, A. M. Blumetti, N. De Paola, A.<br />

Yates, A. Bubeck and G. Sileo for assistance in the field.<br />

References<br />

Buckham, R.C., G. Plafker, and R.V. Sharp (1978), Fault<br />

movement (afterslip) following the Guatemala earthquake<br />

of February 4, 1976. Geology, 6, 170-173.<br />

Cheloni, D., et al. (2010), Coseismic and initial post-seismic<br />

slip of the 2009 M-w 6.3 L'Aquila earthquake, Italy, from<br />

GPS measurements. Geophys J Int, 181 (3), 1539-1546,<br />

doi: 10.1111/j.1365-246X.2010.04584.x<br />

Chiarabba, C. et al. (2009), The 2009 L’Aquila (central Italy)<br />

Mw6.3 earthquake: Main shock and aftershocks.<br />

Geophys Res Lett, 36, L18308,<br />

doi:10.1029/2009GL039627<br />

Emergeo Working Group (2010), Evidence for surface<br />

rupture associated with the Mw 6.3 L’Aquila earthquake<br />

sequence of April 2009 (central Italy). Terra Nova, 22, 43-<br />

51.<br />

Falcucci, E., et al. (2009), The Paganica Fault and Surface<br />

Coseismic Ruptures Caused by the 6 April 2009<br />

Earthquake (L’Aquila, Central Italy), Seismol. Res. Lett.,<br />

80, 940–950, doi:10.1785/gssrl.80.6.940.<br />

Marone, C. J., S. H. Scholtz, and R. Bilham (1991), On the<br />

mechanics of Earthquake Afterslip, J. Geophys. Res., 96,<br />

8441–8452, doi:10.1029/91JB00275<br />

Scholz, C. H. (1990), The Mechanics of Earthquakes and<br />

Faulting, Cambridge Univ. Press, New York.<br />

Sharp, R.V. et al. (1989), Surface faulting along the<br />

Superstition Hills fault zone and nearby faults associated<br />

with the earthquakes of 24 November 1987, Bull.<br />

Seismol. Soc. Am., 79, 252-281.<br />

Vittori, E. et al. (2010), Surface faulting of the April 6th 2009<br />

Mw 6.3 L’Aquila earthquake in central Italy. Bulletin of the<br />

Seismological Society of America, in press.<br />

Wilkinson, M., et al. (2010), Partitioned postseismic<br />

deformation associated with the 2009 Mw 6.3 L’Aquila<br />

earthquake surface rupture measured using a terrestrial<br />

laser scanner, Geophys Res Lett, 37, L10309, doi:<br />

10.1029/2010GL043099<br />

Williams, P.L. and H.W. Magistrale (1989), Slip along the<br />

Superstition Hills fault associated with the 24 November<br />

1987 Superstition Hills, California, Earthquake. Bull.<br />

Seismol. Soc. Am., 79, 390-410.<br />

282


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SEDIMENTARY EVIDENCE OF HOLOCENE TSUNAMI IMPACTS AT THE GIALOVA<br />

LAGOON (SOUTHWESTERN PELOPONNESE, GREECE)<br />

Willershäuser, Timo (1), Andreas Vött (1), Georg Bareth (2), Helmut Brückner (2), Hanna Hadler (1), Konstantin Ntageretzis (1)<br />

(1) Institute for Geography, <strong>Johannes</strong> <strong>Gutenberg</strong>-<strong>Universität</strong> <strong>Mainz</strong>, Johann-Joachim-Becher-Weg 21, 55099 <strong>Mainz</strong>. GERMANY.<br />

Email: Timo.Willershaeuser@uni-mainz.de<br />

(2) Institute for Geography, <strong>Universität</strong> zu Köln, Albertus-Magnus-Platz, 50923 Köln. GERMANY<br />

Abstract (Sedimentary evidence of Holocene tsunami impacts at the Gialova Lagoon (Southwestern Peloponnese,<br />

Greece): The coastal area around the Gialova Lagoon (southwestern Peloponnese, Greece), directly exposed to the tectonically<br />

highly active Hellenic Trench, was repeatedly affected by tsunamigenic impacts as known from historical sources. Detailed geoscientific<br />

studies were carried out in coastal environments in search of corresponding tsunami deposits using terrestrial<br />

vibracorings. Geomorphological, sedimentological and geochemical methods were applied to reconstruct the sedimentary<br />

fingerprints of Holocene tsunami events and the palaeogeographical evolution. Our results show that the palaeogeographical<br />

setting was strongly affected by high-energy tsunami impacts. Coarse-grained allochthonous sediments of marine origin were<br />

found intersecting muddy deposits of low-energy lagoonal and limnic environments. The Voidokilia washover fan and beachrock<br />

structures along the coastline also seem to be of tsunamigenic origin. Both sedimentological and geochronological criteria suggest<br />

multiple tsunami landfall since the mid-Holocene.<br />

Key words: Palaeotsunami, beachrock-type tsunami deposits, Holocene stratigraphy, washover fan.<br />

INTRODUCTION AND AIMS<br />

The eastern Mediterranean is a tectonically active<br />

region with a high tsunami risk (Papazachos &<br />

Dimitriou, 1991). The plate boundary of the Hellenic<br />

Arc, where the African Plate is being subducted<br />

beneath the Aegean microplate, is a hot spot for<br />

earthquakes and therefore highly capable for<br />

triggering tsunamis. Numerous historical accounts<br />

show that the surrounding coastal areas and their<br />

geomorphology were affected by multiple<br />

tsunamigenic impacts (Soloviev, 1990). Thus,<br />

palaeotsunami research in the eastern<br />

Mediterranean has been distinctly intensified in the<br />

last 20 years. Sedimentary characteristics of recent<br />

and subrecent tsunami deposits comprise e.g. (a)<br />

shell debris layers, (b) mixture of littoral and<br />

sublittoral material, (c) multi-modal grain size<br />

distribution, (d) rip up-clasts, (e) basal erosional<br />

unconformities, (f) fining upward and thinning<br />

landward tendencies, (g) lithified beachrock-type<br />

tsunamites and (h) washover deposits (Dominey-<br />

Howes et al., 2006; May, 2010; Vött et al., 2009a,<br />

2009b, 2010a, 2010b).<br />

The main objectives of this study are (i) to detect<br />

allochthonous high-energy deposits in the local<br />

stratigraphical record and (ii) to reconstruct<br />

palaeotsunami events against the background of the<br />

palaeogeographical evolution of the Gialova coastal<br />

area during the Holocene.<br />

REGIONAL SETTING AND METHODS<br />

The coastal area of Pylos and the Gialova Lagoon<br />

are located in the southwestern Peloponnese. The<br />

study area is directly exposed to the subduction zone<br />

283<br />

of the Hellenic Trench holding a high tsunami risk<br />

(Hollenstein et al., 2008, Sachpazi et al., 2000,<br />

Papazachos & Dimitriou, 1991). The Gialova Lagoon<br />

is located in the northern fringe of the Navarino Bay,<br />

a tectonic depression. The shallow lagoon is<br />

separated from the Bay of Navarino by a beach<br />

barrier system to the south and the semi-circular Bay<br />

of Voidokilia to the west (Fig. 1).<br />

Fig. 1: (a) Topographic and geomorphological<br />

overview of the study area and selected coring sites.<br />

General map based on Google Earth images (2003).<br />

(b) Bird’s eye view of Gialova Lagoon, view to the<br />

east. Photo taken by. T. Willershäuser, 2009.<br />

Vibracores in the environs of the Gialova Lagoon<br />

were retrieved by an Atlas Copco mk1 corer. In the


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

field, vibracores were analyzed by sedimentological<br />

and pedological methods. Laboratory studies<br />

comprised analyses of organic content (loss on<br />

ignition), concentration of calcium carbonate, pHvalue<br />

and electrical conductivity. All sediment<br />

samples were analysed for contents of Ca, Mn, Fe<br />

and more than 20 other elements using the XRF<br />

technique.<br />

SEDIMENTARY RECORD IN QUIESCENT NEAR-<br />

SHORE ENVIRONMENTS<br />

The Gialova geo-archive is dominated by quiescent<br />

low-energy conditions. For the Gialova Lagoon<br />

vibracoreprofiles PYL 2 and PYL 3 are considered to<br />

be most representative for the local coastal evolution<br />

and to detect potential high-energy signatures in the<br />

stratigraphical record.<br />

stratum of lagoonal mud documents that the preexisting<br />

quiescent conditions were again reestablished.<br />

Finally, the lagoonal mud is covered by<br />

sandy sediments of the recent dunes and marshy<br />

deposits.<br />

In summary, both Vibracores show distinct signatures<br />

of abruptly environmental changes in the<br />

sedimentary record. We detected a minimum of two<br />

distinct allochthonous layers of high-energetic<br />

conditions compared to the autochthonous,<br />

predominantly limnic pre-existing environments.<br />

Vibracoring site PYL 2 (N 36°57’46.9’’,<br />

E 21°41’29.3’’, ground surface at 0.47 m above sea<br />

level) and PYL 3 (N 36°57’51.6’’, E 21°39’51.3’’,<br />

ground surface at 0.22 m above sea level) are<br />

located at the eastern and western shores of the<br />

Gialova Lagoon, respectively (Fig. 1).<br />

At its base, the stratigraphy of PYL 2 (based on<br />

sedimentological and geochemical parameters),<br />

consists of homogenous silty sediments showing<br />

quiescent, most probably limnic conditions. This<br />

basal stratum is intersected by high-energy deposits<br />

of unsorted coarse-grained material of marine origin.<br />

This event layer is overlain by homogeneous and<br />

well sorted silty clay indicating an immediate reestablishment<br />

of pre-existing quiescent conditions.<br />

The limnic environment was again influenced by a<br />

second input of unsorted allochthonous grus, gravel<br />

and limestone fragments embedded in a loamy<br />

matrix. The associated sharp basal erosional contact<br />

again documents that sediment input occurred<br />

abruptly. Towards the top, the high-energy sediments<br />

are covered by a sequence of peat, organic mud,<br />

limnic mud and finally by recent marshy sediments.<br />

The base of PYL 3 is made out of well sorted fine<br />

sandy silt of a quiescent (fluvio-)limnic environment.<br />

This facies is separated by a sharp erosional<br />

unconformity from following allochthonous coarsegrained<br />

and unsorted sediments with distinct fining<br />

upward sequences and sublayers including wellrounded<br />

gravel. Subsequently, autochthonous limnic<br />

conditions were re-established. A second sharp<br />

erosional unconformity indicates another abrupt<br />

environmental change and the input of allochthonous<br />

marine material (Fig. 2). The fairly unsorted<br />

sediments consist of a mixture out of gravel, grus,<br />

sand and loam. This part of the profile is again<br />

characterized by distinct fining upward sequences<br />

and rip up-clasts of eroded underlying limnic<br />

sediments. After the event, quiescent limnic<br />

conditions were quickly re-established, subsequently<br />

turning into lagoonal conditions. Towards the top of<br />

vibracore PYL 3, the lagoon was influenced by<br />

another third distinct input of allochthonous marine<br />

sand. The intersecting event layer is characterised by<br />

rip up-clasts and several fining upward sequences. A<br />

284<br />

Fig. 2: Vibracore PYL 3 and facies interpretation.<br />

Details (a + b) show intersecting allochthonous<br />

marine sediments in between predominantly limnic<br />

deposits.<br />

BEACHROCK AS LITHIFIED TSUNAMI DEPOSIT<br />

First studies on the occurrence of beachrock along<br />

the coastline of Pylos were made by Kraft et al.<br />

(1980). These authors described that the beachrock<br />

includes sherds of probably Roman age, but no<br />

further information on the internal structure,<br />

geomorphological and sedimentary context was<br />

given. However, recent studies on beachrock-type<br />

calcarenitic deposits at adjacent coastal areas<br />

revealed that post-depositional pedogenetic<br />

decalcification and cementation of tsunami deposits<br />

must be assumed for their evolution (Vött et al.,<br />

2010a & Scheffers et al., 2008).<br />

In the investigation area, several locations along the<br />

coastline show onshore and offshore occurrences of<br />

eroded and fragmented beachrock. The internal<br />

structure is characterized by typical sedimentological<br />

features of tsunami deposits. At Vromoneri (10 km to<br />

the north of Gialova Lagoon), we found beachrock<br />

sequences injected in between bedrock units and<br />

characterized by basal erosional discontinuities,<br />

partly well-laminated structure, embedded intra clasts<br />

and features of a distinct landward flow direction are<br />

detectable.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

At Romanou (Fig. 3), the basal section of the<br />

beachrock is dominated by gravel, followed by<br />

coarse, medium and fine sand with distinct fining<br />

upward sequences. The sedimentary features are<br />

untypical of littoral environments and rather<br />

document high-energy flow dynamics. Thus,<br />

beachrock-type lithified sediments along the Pylos<br />

coastline are interpreted as deposits of high-energy<br />

impact.<br />

Fig. 3: Beachrock at Romanou (about 5 km to the<br />

north of the Gialova Lagoon - see Fig. 1). The present<br />

position of the fragmented beachrock is offshore (I)<br />

and onshore (II).<br />

DISCUSSION AND CONCLUSION<br />

Our case-study, based on sediment cores and<br />

geomorphological findings, show that tsunami<br />

signatures in coastal sedimentary environments are<br />

highly variable on large scales.<br />

The sediment profiles recovered from near-shore<br />

environments are characterised by allochthonous<br />

gravelly to sandy high-energy deposits intersecting<br />

fine-grained autochthonous sediments of a lower<br />

energetic potential. As the energetic potential as well<br />

as the landward extent of storm influence in the<br />

Mediterranean are known to be restricted to the<br />

immediate littoral zone (Vött et al., 2010b),<br />

allochthonous coarse-grained deposits found in the<br />

environs of Gialova Lagoon are interpreted as of<br />

tsunamigenic origin. Moreover, numerous findings of<br />

beachrock along the coastline must not be<br />

considered as consolidated littoral sediments. We<br />

assume that the origin of the beachrock in this case<br />

is attributed to sediment deposition by high-energy<br />

events followed by post sedimentary cementation<br />

and calcification (Vött et al., 2010a). Associated<br />

sedimentary characteristics, such as laminated<br />

structures, erosional contact at the base, fining<br />

upward sequences and embedded stones and<br />

ceramic fragments are characteristic for tsunami<br />

influence but cannot be described by recent littoral<br />

processes. The paleogeographical evolution of the<br />

Pylos area is characterised by predominant limnic<br />

and lagoonal conditions. The shifting and destruction<br />

of the palaeo-environment is mainly controlled by<br />

high-energy tsunami impacts induced by earthquake<br />

focal mechanisms. Gradual changes seem to be<br />

merely responsible for the re-arrangement of<br />

sediments after high-energy impacts.<br />

Acknowledgements: We acknowledge funding of the<br />

project by the German Research Foundation, DFG (Bonn,<br />

Gz. VO 938/3-1). Work permits were kindly issued by the<br />

Greek Institute of Geology and Mineral Exploration (IGME,<br />

Athens).<br />

References<br />

Dominey-Howes, D.T.M., G.S. Humphreys & P.P. Hesse<br />

(2006). Tsunami and palaeotsunami depositional<br />

signatures and their potential value in understanding the<br />

late-Holocene tsunami record. The Holocene 16 (8),<br />

1095-1107.<br />

Hollenstein, C., M.D. Müller, A. Geiger & H.G. Kahle (2008).<br />

Crustal motion and deformation in Greece from a decade<br />

of GPS measurements, 1993–2003. Tectonophysics<br />

449, 17-40.<br />

Kraft, J.C., G.R. Rapp, Jr. & S.E. Aschenbrenner (1980).<br />

Late Holocene Palaeogeographic Reconstructions in the<br />

Aerea of the Bay of Navarino: Sandy Pylos. Journal of<br />

Archaeological Science 7, 187-210.<br />

May, S.M. (2010). Sedimentological, geomorphological and<br />

geochronological studies on Holocene tsunamis in the<br />

Lefkada – Preveza area (NW Greece) and their<br />

implications for coastal evolution. PhD Thesis 159 pp.,<br />

Cologne, Germany.<br />

Papazachos, B.C. & P.P. Dimitriu (1991). Tsunamis In and<br />

Near Greece and Their Relation to the Earthquake Focal<br />

Mechanisms. Natural Hazards 4, 161-170.<br />

Sachpazi, M., A. Hirn, C. Clément, F. Haslinger, M. Laigle,<br />

E. Kissling, P. Charvis, Y. Hello, J. C Lépine, M. Sapin&<br />

J. Ansorge (2000). Western Hellenic subduction and<br />

Cephalonia Transform: local earthquakes and plate<br />

transport and strain. Tectonophysics 319(4), 301–319.<br />

Scheffers, A., D. Kelletat, A. Vött, S.M. May & S. Scheffers<br />

(2008). Late Holocene tsunami traces on the western<br />

and southern coastlines of the Peloponnesus (Greece).<br />

Earth and Planetary Science Letters 269, 271-279.<br />

Soloviev, S.L. (1990). Tsunamigenic Zones in the<br />

Mediterranean Sea. Natural Hazards 3, 183-202.<br />

Vött, A., H. Brückner, S. Brockmüller, M. Handl, S.M. May,<br />

K. Gaki-Papanastassiou, R. Herd, F. Lang, H.<br />

Maroukian, O. Nelle & D. Papanastassiou (2009a).<br />

Traces of Holocene tsunamis across the Sound of<br />

Lefkada, NW Greece. Global and Planetary Change 66<br />

(1-2), 112-128.<br />

Vött, A., H. Brückner, S.M. May, D. Sakellariou, O. Nelle, F.<br />

Lang, V. Kapsimalis, S. Jahns, R. Herd, M. Handl & I.<br />

Fountoulis (2009b). The Lake Voulkaria (Akarnania, NW<br />

Greece) palaeoenvironmental archive - a sediment trap<br />

for multiple tsunami impact since the mid-Holocene.<br />

Zeitschrift für Geomorphologie N.F., Suppl. Issue 53 (1),<br />

1-37.<br />

Vött, A., G. Bareth, H. Brückner, C. Curdt, I. Fountoulis, R.<br />

Grapmayer, H. Hadler, D. Hoffmeister, N. Klasen, F.<br />

Lang, P. Masberg, S.M. May, K. Ntageretzis, D.<br />

Sakellariou & T. Willershäuser (2010a). Beachrock-type<br />

calcarenitic tsunamites along the shores of the eastern<br />

Ionian Sea (western Greece) – case studies from<br />

Akarnania, the Ionian Islands and the western<br />

Peloponnese. Zeitschrift für Geomorphologie N.F.,<br />

Suppl- Issue 54 (3), 1-50.<br />

Vött, A., et al. (2010b), Sedimentological and<br />

geoarchaeological evidence of multiple tsunamigenic<br />

imprint on the Bay of Palairos-Pogonia (Akarnania, NW<br />

Greece), Quaternary International (2010),doi:10.1016/<br />

j.quaint.2010.11.002<br />

285


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

IS THE RURRAND FAULT (LOWER RHINE GRABEN, GERMANY) RESPONSIBLE FOR<br />

THE 1756 DÜREN EARTHQUAKE SERIES?<br />

Winandy, Jonas (1, Christoph Grützner (1), Klaus Reicherter (1), Thomas Wiatr (1), Peter Fischer (2), Thomas Ibeling (3)<br />

(1) RWTH Aachen, Neotectonics and Natural Hazards Group, Aachen, Germany (jonas.winandy@rwth-aachen.de, +49 241<br />

8096358),<br />

(2) Köln University, Geographisches Institut, Köln, Germany<br />

(3) Archäologische Grabungen und Sondagen GBR, Köln, Germany<br />

Abstract (Is the Rurrand Fault (Lower Rhine Graben, Germany) responsible for the 1756 Düren earthquake series?): In<br />

1756, several strong earthquakes (M5-6.1) occurred close to Düren (Lower Rhine Graben, LRG) in Germany. The Rurrand Fault in<br />

the LRG located in the middle between Aachen and Cologne possibly indicates the Düren earthquake sequence. This fault is one<br />

of the most prominent NW-SE trending normal faults with a morphological expression in the area within the Lower Rhine Graben.<br />

Holocene sediments with significant offsets covered by thin colluvial sediments were found and a complex fault geometry was<br />

observed during archaeological excavations. DC geoelectrics and georadar were applied in order to image the deeper parts of the<br />

fault zone. Radiocarbon and luminescence dating of sediment samples are in progress, but the morphological expression of the<br />

fault, the shallow depths of the offset sediments, and geophysical data allow concluding on recent seismicity along this active fault<br />

with at least four surface-rupturing events.<br />

Key words: earthquake, geophysics, Rhine Graben, Rurrand Fault<br />

INTRODUCTION: THE DÜREN EARTHQUAKES<br />

1755/1756<br />

The area between Aachen and Cologne in western<br />

Germany was hit by a series of earthquakes in<br />

1755/1756. On 18 th February, 1756, the strongest<br />

event took place most likely west of the city of Düren,<br />

leaving two people dead (some reports claim three<br />

fatalities) and causing significant damages also in<br />

Aachen, Cologne, and nearby villages. Chimneys<br />

were destroyed in up to 70 km distance (Liège,<br />

Belgium), light damages were recorded in Brussels,<br />

Gießen and Osnabrück (200 km distance). A<br />

landslide was triggered 15 km SW from Düren. The<br />

shaking was felt as far as 400 km from the epicentre<br />

in London, Magdeburg, Halle, Paris, and Strasbourg<br />

(Meidow, 1995). Epicentral intensities of VIII are<br />

reported by Skupin et al. (2008) for the Eschweiler<br />

area (15 km W of Düren). A magnitude of 6.3 is<br />

assumed for the main event by Skupin et al. (2008),<br />

Meidow (1995) assumes M L =6.1. Our study shows<br />

that the Rurrand Fault was possibly activated during<br />

the Düren earthquake sequence.<br />

The Lower Rhine Embayment and especially its<br />

western part is one of the tectonically most active<br />

areas in Germany and dominated by the Lower<br />

Rhine Graben. NW-SE trending normal faults form a<br />

horst and graben structure with a number of single<br />

blocks (Krefeld-, Köln-, Venlo-, Erft-, and Rur blocks,<br />

from NE to SW). The faults show offsets of more than<br />

50 m in the Quaternary. Düren is situated in a NW-<br />

SE striking graben, which is flanked by the Rurrand<br />

Fault in the NE and the Stockheimer Sprung in the<br />

GEOLOGY AND TECTONIC SETTING<br />

The Lower Rhine Embayment (LRE) underwent<br />

subsidence from Miocene to recent, accompanied by<br />

uplift of the Rhenish Massif to the southwest and east<br />

of the study area. Tertiary and Quaternary sediments<br />

of more than 1 km thickness were deposited and also<br />

include lignites which are extracted in open pit mines.<br />

Fluvial and aeolian Pleistocene-Holocene sediments<br />

as well as loess cover wide areas.<br />

286<br />

Fig. 1: Neotectonics and historical earthquakes in the<br />

study area, the Lower Rhine Embayment. RRF:<br />

Rurrand Fault; SSF: Stockheimer Sprung


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

SW (Fig. 1). The Rurrand Fault is a NW-SE striking<br />

normal fault dipping to the SW and expressed by an<br />

escarpment.<br />

A large number of damaging earthquakes since<br />

Karolingian times has been reported for this area.<br />

The most recent one was the 1992 Roermond<br />

earthquake which reached M L =5.9. Recent studies<br />

reported that active faults in the study area are<br />

characterized by recurrence periods in the order of<br />

tens of ka (Skupin et al., 2008; Camelbeeck et al.,<br />

2007), and that present day aseismic slip is assumed<br />

for the Rurrand Fault, resulting form the lowering of<br />

the groundwater level due to the nearby mining<br />

activities (Vanneste & Verbeeck, 2001). Active faults<br />

in Germany are often not visible in the field due to<br />

relatively high erosion rates. Therefore, the seismic<br />

hazard might be under-estimated. The Rurrand Fault<br />

was trenched already only approx. 2-3 km away, and<br />

only Pleistocene faulting evidence was found (Skupin<br />

et al., 2008).<br />

measurements. Data processing was done with<br />

ReflexW by Sandmeier Scientific Software. The<br />

geoelectrics data were gathered with the 4-point-light<br />

system (Lippmann), and 80 electrodes with 1.5 m<br />

spacing. Schlumberger, Dipole-dipole and Wenner<br />

configurations were applied. Soil samples were taken<br />

for radiocarbon and luminescence dating, which is<br />

currently in progress.<br />

RESULTS<br />

METHODS<br />

Fig. 2: Location of trenches, outcrops and geophysical<br />

profiles at the study area.<br />

Due to construction works for a new highway,<br />

extensive archeological excavations have proven<br />

findings from Roman times until recent. At this<br />

occasion, the Rurrand Fault was trenched in several<br />

places, where we mapped layer offsets, sediment<br />

deformation, and structural data (Fig. 2). The trench<br />

walls were sketched and photographed. Ground<br />

penetrating radar (GPR) and electric resistivity<br />

measurements have been applied in order to image<br />

deeper sediment structures and to map the fault<br />

trace. We used the GSSI 100, 270, and 400 MHz<br />

antennas with the SIR 3000 controller, a survey<br />

wheel and a GPS data logger for georadar<br />

Fig. 3: 270 MHz GPR profile (64) crossing the fault<br />

with an angle of 45°.<br />

The GPR data revealed sediment layers inclined<br />

towards the fault (Fig. 3). The fault cropped out 2 m<br />

away from the GPR profile, allowing a direct<br />

comparison. Down to a depth of 4 m several<br />

reflectors dip towards the NW.<br />

287


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Fig. 4: Photograph of the fault, outcropped in an archeological trench (upper image), and reconstruction of the faulting<br />

history (lower image). S = sedimentation and erosional stages; E W - E Z = last earthquake events.<br />

A sharp contrast in reflection amplitudes marks the<br />

fault itself. On the footwall, only few layers appear to<br />

dip towards the Rurrand Fault. Similar observations<br />

were made at other GPR profiles that crossed the<br />

fault. The high-frequency antennas allowed<br />

identifying tilted sediments and the fault itself at<br />

various locations. The resolution of the 100 MHz<br />

288<br />

antenna was found too low for imaging fault features<br />

in this case. Geoelectrics data revealed low-resistivity<br />

anomalies (higher conductivities) at the fault zone,<br />

which are most likely related to an increase in water<br />

content at the fault zone.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

Holocene surface-near sediments with significant<br />

offsets covered by thin colluvial sediments were<br />

found and a complex fault geometry was observed<br />

during the archaeological excavations (Figs. 4 - 6).<br />

Some deformation structures seem to be related to<br />

liquefaction. The offset of surface-near sediments is<br />

in the order of 5 cm (Fig. 5), deeper layers show<br />

greater offsets (Fig. 6). Growing displacement of the<br />

major fault downsection suggest more than one<br />

major, surface-rupturing earthquakes along the<br />

Rurrand Fault in the Holocene/Late Pleistocene. We<br />

developed a deformation model for the fault,<br />

assuming at least four surface-faulting events (E W -<br />

E Z ) that led to the present day geometry (Fig. 4) and<br />

seven stages of seismic quiescence (S 1 -S 7 ). Roots<br />

penetrated the soil at stage four, and liquefaction of<br />

fine grained material is likely to have occurred during<br />

the last earthquake event. Despite the results from<br />

dating are yet to come, we can assume four events<br />

since Late Pleistocene, which would result in slightly<br />

shorter recurrence periods than estimated by<br />

previous studies. The fault may also be responsible<br />

for the Düren 1756 events, as the evidence for<br />

surface-faulting earthquakes proves that it is capable<br />

for destructive events with magnitudes > 5.5.<br />

However, only the dating will allow associating the<br />

Düren earthquakes with the Rurrand Fault.<br />

We found evidence for four surface-faulting events at<br />

the active Rurrand Fault close to Düren, which might<br />

be responsible for the 1755/1756 earthquake series.<br />

Offset sediments clearly point to seismic activity<br />

since Late Pleistocene and at least four earthquake<br />

events. Geophysical data allowed mapping the fault<br />

trace where there were no outcrops (Geoelectrics<br />

and GPR) and revealed the fault geometry in depths<br />

of up to 6 m (GPR).<br />

Fig. 6: Offset surface-near layers have been found<br />

during the archeological excavations.<br />

References<br />

Fig. 5: A) Step-like offset clearly points to seismic<br />

deformation instead of soil creep. Displacement is<br />

about 5 cm. The offset reddish layer is made up of<br />

clayey-silty, loess material, 50 cm below the surface.<br />

B) The fault zone is clearly visible in the trenches, not<br />

only at the walls, but also on the floor. This enabled a<br />

very good correlation with the geophysical data and a<br />

precise analysis of subsurface features.<br />

CONCLUSION<br />

Camelbeeck, T., Vanneste, K., Alexandre, P., Verbeeck, K.,<br />

Petermans, T., Rosset, P., Everaerts, M., Warnant, R. &<br />

Van Camp, M. (2007). Relevance of Active Faulting And<br />

Seismicity Studies To Assessments Of Long-Term<br />

Earthquake Activity and Maximum Magnitude In<br />

Intraplate Northwest Europe, between the Lower Rhine<br />

Embayment and the North Sea. In: Geological Society of<br />

America, Special Paper, vol. 425, pp. 193-224.<br />

Meidow, H. (1995). Rekonstruktion und Reinterpretation von<br />

historischen Erdbeben in den nördlichen Rheinlanden<br />

unter Berücksichtigung der Erfahrungen bei dem<br />

Erdbeben von Roermond am 13. April 1992. Dissertation,<br />

<strong>Universität</strong> Köln, Leverkusen, 305 p.<br />

Skupin, K., Buschhüter, K., Hopp, H., Lehmann, K., Pelzing,<br />

R., Prüfert, J., Salamon, M., Schollmayer, G., Techmer,<br />

A. & Wrede, V. (2008). Paläoseismische Untersuchungen<br />

im Bereich der Niederrheinischen Bucht. Scriptum 17,<br />

Krefeld (Geol. Dienst Nordrh.-Westf.), 72 p.<br />

Vanneste, K. & Verbeeck, K. (2001). Paleoseismological<br />

analysis of the Rurrand fault near Jülich, Roer Valley<br />

graben, Germany: Coseismic or aseismic faulting<br />

history? Netherlands Journal of Geosciences / Geologie<br />

en Mijnbouw, 80 (3-4): 155-169.<br />

289


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

FIND AND PRIMARY SEARCH OF AN ACTIVE FAULT AT THE GAIXIA SITE, GUZHEN<br />

COUNTY, ANHUI PROVINCE, P.R. CHINA<br />

YAO Da-quan (1), SHUO Zhi (2),TANG Ji-ping (3), WANG Zhi (2),SHEN Xiao-qi (1),CHEN An-guo (1) , LI Lin-li (1)<br />

(1)Seismological Administration of Anhui Province,Hefei,Anhui P.R.China,230031. email: daquany@aheq.gov.cn<br />

(2)Cultural Relic Archaeology Research Institute of Anhui Provine, Hefei,Anhui P.R. China,230001<br />

(3)Institute of Cultural Relic Archaeology of Tongling City,Anhui Province,Tongling, Anhui P.R. China,244001<br />

Abstract (Find and primary search of an active fault at the Gaixia Site, Guzhen County, Anhui Province, P.R. China):<br />

Recently a large number of ancient sites were excavated due to construction works in eastern China, which makes it possible to<br />

identify and trace thousands of years of natural deformation history. With this opportunity, one can benefit from the precise<br />

archaeological layer technology. In a collaboration of earthquake research institutions and archaeological departments, we<br />

analysed Quaternary tectonic deformation, in particular the deformation phenomena hosted in the prehistoric cultural layer. This<br />

paper reports the working progress of a special excavation at the Gaixia archaeological site in Haocheng town (Guzhen county,<br />

Anhui province, P.R. China).<br />

Key words: Quaternary deformation, prehistoric culture, Gaixia site, Anhui Province<br />

THE BASIC FEATURES OF THE VESTIGE<br />

The Gaixia site is located east of the Tancheng-<br />

Lujiang fault (TanLu Fault on Fig. 1). In the west<br />

there is the Guzhen-Fengtai fault, where the Ms 6.2<br />

earthquake event of 1831 took place in FengTai. The<br />

southern border is the NW-trending GuoHe fault,<br />

where a magnitude Ms 6 earthquake occurred in<br />

1481 in Guoyang (Fig.1).<br />

The Gaixia site lies in Bawangcheng village,<br />

Haocheng town, 24 kilometers eastward of Guzhen<br />

county, which is the centre battlefield of the<br />

competition between Chu and Han and the final<br />

Gaixia battle, where the ‘Gaixia Battle’ took place in<br />

202 B.C. The entire archaeological area is about<br />

150,000 km 2 , with an earth wall around the main site,<br />

which is 2-3 m higher than the outer surface level.<br />

The inner terrain is higher at the sides and lower than<br />

the surroundings in the middle. West and north of the<br />

archaeological site there is the Tuohe River.<br />

In order to clarify the specific age, the floor width and<br />

construction methods of the ancient city wall,<br />

archaeologists excavated a trench since March 2008<br />

(serial number: 2008 GGTG3, TG3 for short) with a<br />

length of 40 m and a width of 3 m, about 70 m from a<br />

former excavation in 2007. The trench extends from<br />

the city wall to the inner moat, which completely<br />

reveals the whole section of the city wall and<br />

excavates some of the cultural layers in the inner<br />

wall. Just in the Dawenkou Cultural layer in both<br />

sides and at the bottom of TG3, seismologists and<br />

archaeologists discovered an active fault (Fig.2).<br />

DISCOVERY AND ANALYSIS OF THE TENSION-<br />

SHEAR FAULT IN THE CULTURAL LAYER<br />

On May 12 th , 2008, when excavations reached greyblack<br />

ash layers, seismologists and archaeologists<br />

Fig. 1: A sketch of regional geology seismology of Gaixia ruins<br />

Fig. 2: TG3 prospecting trenchfault of Gaixia ruins<br />

290


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

found fault dislocation evidence. According to this<br />

clue, they carefully scraped the pace of the layer<br />

again for further identification, and confirmed the<br />

range of fault dislocation.<br />

point on the south wall is 1.68 meters below ground<br />

surface, and also extends to the raw soil layer. From<br />

the 18th layer downwards, 12 layers of the north wall<br />

had dislocation, not including the raw soil layer: the<br />

fault penetrated the wall with the layers 18, 33, 36,<br />

37, 39, 40, 41, 47, 54, 61, 68, 70, and also<br />

penetrated the narrow channel G6, which is S-N<br />

trending. The southern part of the prospecting trench<br />

also showed normal faulting (Fig.3). From the 18th<br />

layer downwards to the raw soil layer there are 10<br />

layers that were dislocated, not including the raw soil<br />

layer, and also showing normal faults (Figs.4,5).<br />

Fig.3: Tensile-shear fault on the northern wall. Left:<br />

Photo of the northern wall; Right: sketch of the<br />

northern wall.<br />

The dislocation was located in TG3, and the west<br />

segment is about 15.4 m to the west wall. According<br />

to the change of quality and color of the soil and the<br />

analysis of samples taken at the dislocation position,<br />

it could be confirmed that the dislocation is beneath<br />

the 17th layer of the city wall (serial number: “the<br />

17th layer wall”, following are the same), and the<br />

dislocation extends from the 18th layer, including raw<br />

soil layer (undisturbed soil). A crack with a width of<br />

Fig.5: Tensile-shear fault on the trough bottom. Left:<br />

photo towards the southern overlook; Right: photo<br />

towards the eastern overlook.<br />

DISCUSSION AND CONCLUSIONS<br />

In conclusion, active faulting was found in the culture<br />

layers of the vestige belong to the Late Dawenkou<br />

Dynasty (about 4300a BP), and the observation of<br />

the fault plane shows the characteristic of high speed<br />

deformation (He Yong-nian et al.,1985; Yao Daquan,2004).<br />

Fig.4: Tensile-shear fault on the southern wall. Left:<br />

photo of the southern wall; Right: sketch of the<br />

southern wall.<br />

2 mm - 4 mm can be seen at both the plane and the<br />

profile of the dislocation, which was very clear to see<br />

on the bottom of the soil layer, gradually becoming<br />

thinner upwards in the 17th layer. The crack was<br />

filled with grey clay and had a different colour than<br />

the surrounding soil. The soil’s thickness was<br />

symmetric and completely anastomosing on the two<br />

sides of the fault.<br />

The western part of the fractured plane was 3.8 cm<br />

higher than the eastern (= 3.8 cm vertical offset). The<br />

trend of the fracture plane is 353°, and the section<br />

dipped towards the E with an angle of ~60°. The fault<br />

plane is curved, with its highest point 1.79 meters<br />

below the ground surface at the western wall of the<br />

trench, extending to the raw soil layer. The highest<br />

The archaeological site is located right on the NNW<br />

trending Tancheng-Lujiang fault. According to the<br />

history records, several earthquakes of Ms~6<br />

occurred near this area. Our discovery of this<br />

earthquake in the ruins enriches the seismotectonic<br />

knowledge in this area and completes the earthquake<br />

catalogue for future hazard studies.<br />

Acknowledgements: This paper is a contribution to the<br />

Scientific Research Special Project of the Earthquake<br />

Calling200808064and Science and Technology Tackle<br />

Key Problem Plan Project of Anhui Province <br />

08010302204<br />

References<br />

He Yong-nian, Yang Zhu-en. 1985. Research and the<br />

significance of microscopic marks of ancient earthquakes<br />

[J]. Earthquake Research in China,1(3):76-81(in<br />

Chinese).<br />

Yao Da-quan 2004. Macroscopic and Microscopic Evidence<br />

of Periodical Stick-Slip Deformation in an Active Fault [J].<br />

Recent Developments in World Seismology, (4):6-10 (in<br />

Chinese).<br />

291


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

THE MOVRI MOUNTAIN EARTHQUAKE: UNDERSTANDING ACTIVE DEFORMATION OF<br />

THE NW PELOPONNESE<br />

Zygouri, Vasiliki (1, Sotiris Kokkalas (1), Paris Xypolias (1), Ioannis Koukouvelas (1), Gerassimos Papadopoulos (2)<br />

(1). Department of Geology, Division of General, Marine Geology and Geodynamics, University of Patras, Patras 26500, Greece.<br />

Email: iannis@upatras.gr<br />

(2). Institute of Geodynamics, National Observatory of Athens, Lofos Nymfon, Thissio, 11810, Athens, Greece. Email:<br />

papadop@noa.gr<br />

Abstract (The Movri Mt earthquake: understanding active deformation of the NW Peloponnese): The Mw 6.4 June 8 th<br />

2008 Movri Mountain earthquake struck NW – Peloponnesus, Greece, caused widespread deformation and<br />

damage in buildings, as well as extensive ground hazards. Three surface ruptures were triggered by the<br />

earthquake, with the most promising for paleoseismology analysis lying near the epicenter of the event, attaining a<br />

maximum offset of 25 cm. In this surface rupture a paleoseismological trench was excavated. Based on seven 14 C<br />

samples, we identify two surface – rupturing earthquakes in the last 1Kyr prior the recent event. Thus, observations<br />

from paleoseismology suggest that the Nisi fault appear to be related to surface ruptures and events. In addition,<br />

our 14 C data support the view that the Nisi fault displays a slip rate in the order of 1.5 mm/yr during the last 1Kyr.<br />

Key words: paleoseismology, surface ruptures, colluvial wedge.<br />

INTRODUCTION<br />

On June 8 th 2008, a Mw 6.4 strike – slip earthquake<br />

(hypocentral depth of almost 20 km) occurred in the<br />

northwestern Peloponnese (Greece) without obvious<br />

relation to any mapped fault. Although strong<br />

earthquakes are common in western Greece, this<br />

event took place in a region previously considered as<br />

seismically quiet. This earthquake ruptured along an<br />

unknown dextral strike slip fault segment striking NE-<br />

SW, resulted in new geological conclusions for the<br />

area. During the Movri Mt earthquake, three main<br />

fault ruptures emerged on the epicentral area, the 4.5<br />

km long Vithoulkas surface rupture, the 5.0 km long<br />

Michoi surface rupture and the 6.0 km long Nisi<br />

surface rupture (Fig. 1).<br />

Fig. 1: Geological map of the study area showing lithology, active faults and surface ruptures during the Movri Mt<br />

earthquake.<br />

292


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

In order to investigate whether these ruptures and<br />

associated fractures record an individual or unique<br />

tectonic offset or a response to strong ground motion<br />

we performed a paleoseismological study across the<br />

Nisi surface rupture. The Nisi rupture trends NNW-<br />

SSE and has an almost straight, segmented trace<br />

and attained a maximum offset of 25 cm during the<br />

event. The trench has a length of 10 m, a width of 1.5<br />

m and a depth of 2.5 m.<br />

DESCRIPTION OF THE NISI TRENCH<br />

The sedimentary environment observed inside the<br />

trench suggests fine – grained coastal and fluvial to<br />

lagoonal sediments (fig. 2). It comprises sandy to<br />

clay horizons with generally yellow, brownish yellow<br />

and grey to dark brown sediments. Shallow coastal<br />

sediments are mostly found on the uplifted part of the<br />

trench. Shallow coastal sediments are overlain in the<br />

upthrown block of the fault either directly by the<br />

modern plowed soil, or by a succession of almost 0.5<br />

m thick yellow sandy and grey clay fluvial sediments<br />

in the downthrown block. The fault zone is almost 1<br />

m wide with open voids formed also during the latest<br />

event. The deeper excavated part of the fault zone<br />

includes rotated coastal sediments. Open voids allow<br />

the precipitation of meteoric water, and disturb the<br />

deposition of new material after every possible new<br />

event. The material inside the fault zone is<br />

characterized as an unconsolidated, unsorted and<br />

mixed assemblage of coastal sediments derived<br />

mainly from the fine grained footwall’s coastal<br />

sediments. Two small colluvial wedges (fig. 3) with a<br />

maximum thickness of 30-40 cm, though obscure,<br />

can be detected. These colluvials are characterized<br />

by wedge geometry, thickest at the fault surface and<br />

thinning away from it, and colour differences.<br />

Fig. 3: Detailed view of the upper colluvial wedge<br />

formed in the Nisi fault zone (The black box shows the<br />

Ps3 sample).<br />

Fig. 2: Trench stratigraphy of the Nisi Fault. he<br />

sampling sites within the trench are mentioned<br />

with Ps abbreviation .<br />

293<br />

DISCUSSION AND CONCLUSIONS<br />

Despite the flat and rather featureless landscape,<br />

certain evidences support the view that Nisi rupture is<br />

a newly emerged fault with recent seismic history.<br />

The colluvial wedges in the trench fulfil most of the<br />

parameters defining deposits in fault zones such as<br />

their shape and chaotic sedimentation (sensu<br />

Pavlides et al., 2004). The colluvial deposits are<br />

supplied by fine grained material of the upthrown<br />

block, controlled by the sedimentation environment<br />

and the ruptured sediments and therefore is bound to<br />

be formed by the finer deposits accumulating on the<br />

downthrown block’s free face. Open voids, filled<br />

cracks, the formation of at least one datable colluvial<br />

wedge and the increased deformation in upthrown<br />

horizons attests to possible episodic activity during<br />

the past 1000 years.


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

EARTHQUAKE<br />

ARCHAEOLOGY<br />

INQUA PALEOSEISMOLOGY<br />

AND ACTIVE TECTONICS<br />

The seven samples (fig. 2) were collected from both<br />

the uplifted and subsided block of the fault as well as<br />

from the disturbed fault zone. The disturbed zone<br />

samples yielded two kind of ages, a younger one at<br />

the lowest part of the zone (sample Ps5) and an<br />

older one (sample Ps3) above the previous sample<br />

and at the base of the colluvial wedge (figs. 2, 3).<br />

This attests the effect of precipitation of meteoric<br />

water through the fault zone. After considering the full<br />

distribution of 14 C ages and taking into account the<br />

stratigraphy of the trench we interpret three events<br />

(with the more recent included) that can be<br />

evidenced.<br />

Our trenching study showed a recurrence interval<br />

between 300 to 600 years, we suggest that the Nisi<br />

fault shapes fine grained colluvial wedges without<br />

following a constant time pattern. Thus, we assume a<br />

Quaternary slip rate in the order of 1.5mm/yr and<br />

strong slip-rate similarities to the Egion and Schinos<br />

faults in Gulf of Corinth (Koukouvelas, 1998;<br />

Koukouvelas et al., 2005).<br />

Acknowledgements: We would like to express our<br />

appreciation to S. Pavlides, R. Caputo and A., Chatzipetros<br />

for fruitful discussion of the trench and S. Verroios and V.<br />

Chatzaras for their help during fieldwork.<br />

References<br />

Koukouvelas, I.K., (1998). The Egion fault, earthquakerelated<br />

and long term deformation, Gulf of Corinth,<br />

Greece. Journal of Geodynamics 26, 501-513.<br />

Koukouvelas, I.K., D. Katsonopoulou, S. Soter, & P.<br />

Xypolias, (2005). Slip rates on the Helike Fault, Gulf of<br />

Corinth, Greece: new evidence from geoarchaeology.<br />

Terra Nova 17, 158-164.<br />

Pavlides, S., I. Koukouvelas, S. Kokkalas, L.<br />

Stamatopoulos, D. Keramydas & I. Tsodoulos, (2004).<br />

Late Holocene evolution of the East Eliki fault, Gulf of<br />

Corinth (Central Greece), Quaternary International 115-<br />

116, 139-154.<br />

.<br />

294


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

Index<br />

Preface<br />

Baize, Stéphane, Laurence Audin, Thierry Winter, Alexandra Alvarado, Luis<br />

Pilatasig, Mercedes Taipe, Paul Kauffmann, Pedro Reyes<br />

First paleoseismic evidences in Ecuador: The Pallatanga Fault Record.............................................. 3<br />

Salvatore Barba and Debora Finocchio<br />

Some notes on earthquake and fault relationships for dip-slip events .............................................. 7<br />

Boulton, Sarah J. and I. S. Stewart<br />

Holocene coastal notches in the Mediterranean: palaeoseismic or palaeoclimatic<br />

indicators? .......................................................................................................................... 11<br />

Braun, Anika, Tomas M. Fernandez-Steeger, Hans-Balder Havenith, Almaz Torgoev, Romy<br />

Schlögel<br />

Analysing the landslide susceptibility with statistical methods in Maily-Say,<br />

Kyrgyzstan ......................................................................................................................... 14<br />

Burchfiel, B. C. and Royden, L. H.<br />

Tectonic interpretation of the 2008 Wenchuan Earthquake: Why it only<br />

propagated in one direction - the future? ................................................................................. 17<br />

Campos, Corina, Christian Beck, Christian Crouzet, Eduardo Carrillo<br />

C haracterization of Late Pleistocene-Holocene earthquake-induced<br />

“homogenites” in the Sea of Marmara through magnetic fabric. Implication for<br />

co-seismic offsets detection and measurements ........................................................................ 19<br />

Carmo, Rita, José Madeira, Ana Hipólito, Teresa Ferreira<br />

Paleoseismological evidence for historical surface rupture events in S. Miguel<br />

Island (Azores) .................................................................................................................... 22<br />

Čyžienė, Jolanta<br />

Fault tectonics regarding the Neotectonic period and influence of tectonic<br />

structures on glacial process in areas of thick Quaternary cover ................................................... 26<br />

Figueiredo, Paula M., João Cabral, Thomas K. Rockwell<br />

Plio–Pleistocene tectonic activity in the Southwest of Portugal...................................................... 30<br />

Foumelis Michael, Ioannis Fountoulis, Ioannis Papanikolaοu, Dimitrios Papanikolaou<br />

Geodetic evidence of the control of a major inactive tectonic boundary on the<br />

contemporary deformation field of Athens (Greece).................................................................... 34<br />

Fountoulis, Ioannis, Mavroulis Spyridon<br />

Neotectonics and comparison of the Environmental Seismic Intensity scale (ESI<br />

2007) and the traditional scales for earthquake intensities for the Kalamata<br />

(SW Greece) earthquake (Ms=6.2R, 13-09-1986) ...................................................................... 38<br />

Fountoulis, Ioannis D., Emmanuel Vassilakis, Mavroulis Spyridon, John Alexopoulos,<br />

Athanasia Erkeki<br />

Quantification of river valley major diversion impact at Kyllini coastal area (W.<br />

Peloponnesus, Greece) with remote sensing techniques .............................................................. 42<br />

Garduño-Monroy, Víctor Hugo, Raúl Pérez-López, Miguel Ángel Rodríguez-Pascua, Julián<br />

García Mayordomo, Isabel Israde-Alcántara and Jim Bischoff<br />

Could large palaeoearthquakes break giant stalactites in Cacahuamilpa? (Taxco,<br />

Central Mexico) ................................................................................................................... 46<br />

Gath Eldon and Tania Gonzalez<br />

Three-dimensional investigation of the AD 1621 Pedro Miguel fault rupture for<br />

design of the Panama Canal’s Boringquen dam.......................................................................... 50<br />

Georgiev, Ivan, Dimitar Dimitrov, Pierre Briole, Emil Botev<br />

Velocity field in Bulgaria and Northern Greece from GPS campaigns spanning<br />

1993-2008.......................................................................................................................... 54


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

Gielisch, Hartwig<br />

Acrocorinth - Geological history and the influence of paleoseismic events to<br />

recent archaeological research ............................................................................................... 57<br />

Goodman-Tchernov, Beverly N.<br />

Interpreting offshore submerged tsunami deposits: An incompletely complete<br />

record ................................................................................................................................ 60<br />

Guerrieri, Luca, Anna Maria Blumetti, Elisa Brustia, Eliana Esposito, Mauro Lucarini,<br />

Alessandro M. Michetti, Sabina Porfido, Leonello Serva & Eutizio Vittori, & the INQUA<br />

TERPRO Project #0811 Working Group<br />

Earthquake Environmental Effects, intensity and seismic hazard assessment:<br />

the EEE catalogue (INQUA Project #0418)................................................................................ 62<br />

Guzman, Oswaldo, Jean-Louis Mugnier, Rexhep Koçi, Riccardo Vassallo, Julien Carcaillet,<br />

Francois Jouanne, Eric Fouache<br />

Active tectonics of Albania inferred from fluvial terraces geometries .............................................. 66<br />

Hadler, Hanna, Andreas Vött, Benjamin Koster, Margret Mathes-Schmidt, Torsten Mattern,<br />

Konstantin Ntageretzis, Klaus Reicherter, Dimitris Sakellariou, Timo Willershäuser<br />

Lechaion, the ancient harbour of Corinth (Peloponnese, Greece) destroyed by<br />

tsunamigenic impact ................................................................................................................ 70<br />

Han, S.-R., M. Lee, J. Park, Y.-S., Kim<br />

Structural characteristics and evolution of the Yangsan-Ulsan Fault System, SE Korea ................................... 74<br />

Havenith, Hans-Balder<br />

Where landslides represent the most important earthquake-related hazards: the mountain<br />

areas of Central Asia ............................................................................................................... 77<br />

Hinzen, Klaus-G., Helen Kehmeier, Stephan Schreiber, Sharon K.Reamer<br />

A case study of earthquakes and rockfall - induced damage to a Roman mausoleum in<br />

Pinara, SW Turkey .................................................................................................................. 81<br />

Hipólito, Ana, José Madeira, Rita Carmo, João Luís Gaspar<br />

Neotectonics of Graciosa Island (Azores) – uncertainty in seismic hazard<br />

assessment in a volcanic area with variable slip-rates................................................................. 84<br />

Hoffmann, Gösta, Klaus Reicherter, Thomas Wiatr, Christoph Grützner<br />

Evidence for Holocene tsunami-impact along the shoreline of Oman ........................................................ 88<br />

Hoffmeister, Dirk, Konstantin Ntageretzis, Nora Tilly, Constanze Curdt, Georg Bareth,<br />

Helmut Brückner, Andreas Vött<br />

Monitoring coastal changes on the Ionian Islands (NW-Greece) by multi-temporal terrestrial<br />

laser scanning ....................................................................................................................... 92<br />

Jamšek, Petra, Lucilla Benedetti, Miloš Bavec, Jure Atanackov, Marko Vrabec, Andrej<br />

Gosar<br />

Preliminary report on the Vodice Fault activity and its potential for seismic hazard in the<br />

Ljubljana Basin, Slovenia........................................................................................................... 96<br />

Jankaew, Kruawun, Dominik Brill, Maria E. Martin, Yuki Sawai<br />

Distribution and sedimentary characteristics of tsunami deposits on Phra Thong<br />

Island, Thailand ................................................................................................................... 99<br />

Kázmér, M., Kamol Sanittham, Punya Charusiri, Santi Pailoplee<br />

Archaeoseismology of the AD 1545 Earthquake in Chiang Mai, Northern Thailand ...................................... 102<br />

Kinugasa, Yoshihiro<br />

Outline of the 3.11 Tohoku Earthquake in Japan.............................................................................. 106<br />

Koster, Benjamin, Klaus Reicherter, Andreas Vött, Christoph Grützner<br />

The evidence of tsunamigenic deposits in the Gulf of Corinth (Greece) w ith geophysical<br />

methods for spatial distribution .................................................................................................. 107<br />

Kübler, S., A. M. Friedrich, M. R. Strecker<br />

Coseismic surface rupturing in the epicentral area of Germany’s strongest historical<br />

earthquake ......................................................................................................................... 111


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

Lazauskiene, Jurga and Andrius Pacesa<br />

Seismotectonic and seismic hazard maps of Lithuania (Baltic region) – recent implications<br />

of intracratonic seismicity......................................................................................................... 114<br />

Lee, Minjung and Young-Seog Kim<br />

Preliminary study on damaged stone monuments in Gyeongju, SE Korea................................................. 118<br />

Malik, Javed N., Michio Morino, Mahendra S. Gadhavi, Khalid Ansari, Chiranjeeb Banerjee,<br />

B. K. Rastogi,Fumio Kaneko, Falguni Bhattacharjee, Ashok. K. Singhvi<br />

Earthquake geology and related hazard in Kachchh, Gujarat, Western India.............................................. 121<br />

Marco, Shmuel and G. Ian Alsop<br />

Seismogenic slumps in palaeo-dead sea sediments.......................................................................... 125<br />

McCalpin, James P.<br />

Mapping and measuring Holocene fault scarps in dense forests with LIDAR.............................................. 128<br />

Meilianda, Ella, Ben Maathuis, Marjolein Dohmen-Janssen<br />

Changes on the geomorphic settings of sand-poor environment coast of Banda Aceh,<br />

Indonesia subject to tectonic and tsunami events ............................................................................. 132<br />

Meskouris, Konstantin, Britta Holtschoppen, Christoph Butenweg, Julia Rosin<br />

Seismic analysis of liquid storage tanks ........................................................................................ 136<br />

Michetti, Alessandro M., Leonello Serva, Andrea Berlusconi, Livio Bonadeo, Fabio<br />

Brunamonte, Francesca FerrarioGianfranco Fioraso, Franz Livio, Giancanio Sileo, Eutizio<br />

Vittori<br />

Geological criteria for evaluating seismicity: Lessons learned from the Po Plain, Northern<br />

Italy ................................................................................................................................. 140<br />

Mishra, Anurag, D.C. Srivastava, Jyoti Shah<br />

Ancient seismites as geodynamical indicator: approach to construct a reactivation event on<br />

the main boundary thrust in the Himalayan region ............................................................................ 144<br />

Mouslopoulou Vasiliki, Andrew Nicol, John J. Walsh, John G. Begg, Dougal B. Townsend,<br />

Dionissios T. Hristopulos<br />

Sampling biases in the paleoseismological data .............................................................................. 148<br />

Niemi Tina M.<br />

Earthquakes in Aqaba, Jordan over the past 2,000 years: Evidence from historical,<br />

geological, and archaeological data............................................................................................. 152<br />

Nomikou, P., Alexandri M., Lykousis V., Sakellariou D., Ballas D.<br />

Sw ath bathymetry and morphological slope analysis of the Corinth Gulf................................................... 155<br />

Pantosti Daniela, Stefano Pucci, Paolo Marco De Martini, Alessandra Smedile<br />

Is the decadence of Leptis Magna (Lybia) the consequence of a destructive earthquake? .............................. 159<br />

Papageorgiou Elena and Paraskevi Nomikou<br />

On-shore prolongation of bathymetrically recognized fault zones based on geodetic GPS<br />

observations along Santorini Volcano .......................................................................................... 163<br />

Papaloizou Loizos and Petros Komodromos<br />

The dynamic analysis of multi-drum ancient structures under earthquake excitations.................................... 167<br />

Papanikolaοu Dimitrios, Royden Leigh, Vassilakis Emmanuel<br />

Neotectonic and active diverging rates of extension in the Northern and Southern<br />

Hellenides across the Central Hellenic Shear Zone........................................................................... 170<br />

Papanikolaοu Ioannis and Gerald Roberts<br />

Clustering and anticlustering in the Southern Apennines as evidenced from geological fault<br />

slip-rate seismic hazard maps and the historical record ...................................................................... 174<br />

Papanikolaοu Ioannis, Gerald Roberts, Georgios Deligiannakis, Athina Sakellariou,<br />

Emmanuel Vassilakis<br />

The Sparta Fault, Southern Greece: tectonic geomorphology, seismic hazard mapping and<br />

conditional probabilities........................................................................................................... 178


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

Papanikolaοu Ioannis, Maria Triantaphyllou, Aggelos Pallikarakis, Georgios Migiros<br />

Active faulting tow ards the Eastern tip of the Corinth Canal: Studied through surface<br />

observations, borehole data and paleoenvironmental interpretations....................................................... 182<br />

Passchier, Cees W., Gilbert Wiplinger, Gül Sürmelihindi, Paul Kessener, Talip Güngör<br />

Roman aqueducts as indicators of historically active faults in the Mediterranean Basin ................................. 186<br />

Pérez-López, R., J.L. Giner-Robles, M.A. Rodríguez-Pascua, F. Martín-González,J. García<br />

Mayordomo,J.A. Álvarez-Gómez, M.J. Rodríguez-Peces, J.M. Insua-Árévalo, J. J.<br />

Martínez-Díaz and P.G. Silva<br />

Testing archaeoseismological techniques w ith instrumental seismic data caused by the M<br />

5.1 Lorca Erthquake (5-11-2011, SE of Spain) ................................................................................ 190<br />

Reicherter, Klaus<br />

Frontiers of earthquake archaeology: the Olympia and Samicum cases (Peloponnese,<br />

Greece)............................................................................................................................. 194<br />

Roberts, Gerald, Joanna Faure Walker, Patience Cowie, Richard Phillips, Ken McCaffrey,<br />

Ioannis Papanikolaou, Max Wilkinson, Alessandro Michetti, Peter Sammonds<br />

Regional strain-rates on active normal faults and variability in the seismic cycle: an<br />

example from the Italian Apennines............................................................................................. 198<br />

Rockwell, Thomas K. and, Yann Klinger<br />

The variability of along-strike co-seismic slip: a new example from the Imperial Fault of<br />

Southern California................................................................................................................ 200<br />

Rodríguez-Pascua M.A., P.G. Silva, Perucha Atienza, M.A., J.L. Giner-Robles, R. Pérez-<br />

López<br />

Earthquake archaeological effects generated by the Lisbon Earthquake (first of November<br />

1755) in the Coria´s Cathedral (Cáceres, Western Spain) ................................................................... 204<br />

Rojas, Wilfredo, Nury Simfors-Morales, Luis Sanez, Åke Sivertun<br />

Neotectonic of the longitudinal fault system in Southern Costa Rica ....................................................... 207<br />

Ruano, Patricia, Antonio J. Gil, Jesús Galindo-Zaldívar, Gracia Rodríguez-Caderot, María<br />

Clara de Lacy, Antonio M. Ruiz, María Jesús Borque, Juan A. Armenteros, Antonio<br />

Herrera, Antonio Jabaloy, Angel C. López-Garrido, Antonio Pedrera, Carlos Sanz de<br />

Galdeano<br />

Geodetic studies in the Zafarraya Fault (Betic Cordilleras) .................................................................. 210<br />

Rudersdorf, Andreas, Jochen Hürtgen, Christoph Grützner, Klaus Reicherter<br />

Neotectonic activity of the Granada Basin – new evidence from the Padul-Nigüelas Fault<br />

Zone ................................................................................................................................ 214<br />

Sakellariou, Dimitris, Lykousis Vasilis, Rousakis Grigoris<br />

Holocene seafloor faulting in the Gulf of Corinth: the potential for underw ater<br />

paleoseismology................................................................................................................... 218<br />

Schreiber,Stephan and Klaus-G. Hinzen<br />

Damage assessment in archaeoseismology: methods and application to the archaeological<br />

zone Cologne, Germany ......................................................................................................... 222<br />

Scholz, Christopher H.<br />

Earthquake Triggering, Clustering, and the Synchronization of Faults ..................................................... 226<br />

Silva, Pablo G., Alex Ribó, Moises Martín Betancor, Pedro Huerta, M. Ángeles Perucha,<br />

Cari Zazo, Jose L. Goy, Cristino J. Dabrio, Teresa Bardají<br />

Relief production, uplift and active tectonics in the Gibraltar Arc (South Spain) from the Late<br />

Tortonian to the Present .......................................................................................................... 227<br />

Sintubin, Manuel, Simon Jusseret, Jan Driessen<br />

Reassessing ancient earthquakes on Minoan Crete getting rid of catastrophism ......................................... 231<br />

Smedile, Alessandra, Paolo Marco De Martini, Daniela Pantosti<br />

Paleotsunamis evidence from a combined inland and offshore study in the Augusta Bay<br />

area (Eastern Sicily, Italy) ........................................................................................................ 233<br />

Solakov D., L.Dimitrova, S.Nokolova, St. Stoyanov, S. Simeonova, L. Zimakov, L.Khaikin<br />

Bulgarian National Digital Seismological Netw ork............................................................................. 235


2 nd INQUA-IGCP-567 International Workshop on Active Tectonics, Earthquake Geology, Archaeology and Engineering, Corinth, Greece (2011)<br />

Štěpančíková Petra, Nývlt Daniel, Hók Jozef, Dohnal Jiří<br />

Paleoseismic study of the Sudetic marginal Fault at the locality Bílá Voda (Bohemian<br />

Massif) .............................................................................................................................. 239<br />

Tsang, Rebecca Y., Thomas K. Rockwell, Aron J. Meltzner, Paula M. Figueiredo<br />

Tow ard development of a long rupture history of the Imperial Fault in Mesquite Basin,<br />

Imperial Valley, Southern California............................................................................................. 243<br />

Vacchi, Matteo, Alessio Rovere, Nickolas Zouros, and Marco Firpo<br />

Mapping paleo shorelines in Lesvos Island: new contribution to the Late Quaternary relative<br />

sea level changes and to the neotectonics of the area ....................................................................... 247<br />

Vacchi, Matteo, Alessio Rovere, Marco Firpo and Nickolas Zouros<br />

Boulder deposits in Southern Lesvos: an evidence of the 1949’s Chios-Karaburum<br />

Tsunami?........................................................................................................................... 251<br />

Valkaniotis, Sotiris, George Papathanassiou, Spyros Pavlides<br />

Active faulting and earthquake-induced slope failures in archeological sites: case study of<br />

Delphi, Greece..................................................................................................................... 255<br />

Vött, Andreas, Peter Fischer, Hanna Hadler, Mathias Handl, Franziska Lang, Konstantin<br />

Ntageretzis,Timo Willershäuser<br />

Sedimentary burial of ancient Olympia (Peloponnese, Greece) by high-energy flood<br />

deposits – the Olympia tsunami hypothesis.................................................................................... 259<br />

Vollmert, Andre, Klaus Reicherter, Pablo G. Silva, Tomas M. Fernandez-Steeger<br />

Landslide mapping to analyse Earthquake Environmental Effects (EEE) in Carmona, Spain<br />

– relation to the 1504 event?..................................................................................................... 263<br />

Wartenberg, Wolfram, Andreas Vött, Hanna Hadler, Timo Willershäuser, Holger Freund,<br />

Stefanie Schnaidt<br />

Storm surge layers w ithin a changeful Holocene environment or sedimentary traces of<br />

palaeo-tsunamigenic events? Pros and Cons of on-site findings, Jade Bay, Southern North<br />

Sea, Germany ..................................................................................................................... 267<br />

Wechsler, Neta, Thomas K. Rockwell, Yann Klinger, Amotz Agnon, Shmulik Marco<br />

Testing earthquake recurrence models w ith 3D trenching along the Dead-Sea Transform .............................. 269<br />

Wiatr, Thomas, Ioannis D. Papanikolaοu, Klaus Reicherter, Tomás Fernández-Steeger<br />

A terrestrial close range view of the normal fault zone near Archanes (Heraklion Basin,<br />

Crete) ............................................................................................................................... 272<br />

Wiatr, Thomas, Klaus Reicherter, Ioannis D. Papanikolaοu & Tomás Fernández-Steeger<br />

The discontinuity of a continuous fault: Delphi (Greece) ..................................................................... 276<br />

Wilkinson, Maxwell, Ken McCaffre, Gerald Roberts, Patience Cowie, Richard Phillips<br />

Postseismic deformation of the 2009 L‘ Aquila earthquake surface rupture measured using<br />

repeat terrestrial laser scanning ................................................................................................ 280<br />

Willershäuser, Timo, Andreas Vött, Georg Bareth, Helmut Brückner, Hanna Hadler,<br />

Konstantin Ntageretzis<br />

Sedimentary evidence of Holocene tsunami impacts at the Gialova Lagoon (Southw estern<br />

Peloponnese, Greece) ........................................................................................................... 283<br />

Winandy, Jonas, Christoph Grützner, Klaus Reicherter, Thomas Wiatr, Peter Fischer,<br />

Thomas Ibeling<br />

Is the Rurrand Fault (Low er Rhine Graben, Germany) responsible for the 1756 Düren<br />

Earthquake Series? ............................................................................................................... 286<br />

YaoDa-quan, Shuo Zhi,Tang Ji-ping, Wang Zhi,Shen Xiao-qi,Chen An-guo, Li Lin-li<br />

Find and primary search of an active fault at the Gaixia Site, Guzhen County, Anhui<br />

Province, P.R. China.............................................................................................................. 290<br />

Zygouri, Vasiliki, Sotiris Kokkalas, Paris Xypolias, Ioannis Koukouvelas, Gerassimos<br />

Papadopoulos<br />

The Movri Mountain Earthquake: understanding active deformation of the NW Peloponnese........................... 292


M.A Rodríguez Pascua et al. (2011). Quaternary International, 242 (1), 20-30. Sp. Vol. on<br />

Earthquake Archaeology and Paleoseismology (P.G. Silva, M. Sintubin & K. Reicherter Eds.)<br />

INQUA<br />

TERPRO<br />

Paleoseismology<br />

and ActiveTectonics<br />

<strong>Proceedings</strong> INQUA - IGCP 567 CORINTH (GREECE), 2011

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!