vrancea_99_hauser.pdf

Upload: madalinaviorica

Post on 04-Jun-2018

218 views

Category:

Documents


0 download

TRANSCRIPT

  • 8/14/2019 Vrancea_99_Hauser.pdf

    1/24

    VRANCEA99the crustal structure beneath the southeastern

    Carpathians and the Moesian Platform from a seismic

    refraction profile in Romania

    F. Hausera,*, V. Raileanu b, W. Fielitz c, A. Balab, C. Prodehl a,G. Polonic d, A. Schulze e

    aGeophysical Institute, University of Karlsruhe, Hertzstr. 16, D-76187 Karlsruhe, GermanybNational Institute for Earth Physics, P.O. Box MG-2, RO-76900 Bucuresti-Magurele, Romania

    cGeological Institute, University of Karlsruhe, Kaiserstr. 12, D-76131 Karlsruhe, GermanydThe Institute of Geodynamics, 19-21 J.L. Calderon St., RO-70201 Bucuresti-32, Romania

    eGeoForschungsZentrum, Telegrafenberg, D-14473 Potsdam, Germany

    Received 2 May 2001; accepted 3 September 2001

    Abstract

    The VRANCEA99 seismic refraction experiment is part of an international and multidisciplinary project to study the

    intermediate depth earthquakes of the Eastern Carpathians in Romania. As part of the seismic experiment, a 300-km-longrefraction profile was recorded between the cities of Bacau and Bucharest, traversing the Vrancea epicentral region in NNE

    SSW direction. The results deduced using forward and inverse ray trace modelling indicate a multi-layered crust. The

    sedimentary succession comprises two to four seismic layers of variable thickness and with velocities ranging from 2.0 to 5.8

    km/s. The seismic basement coincides with a velocity step up to 5.9 km/s. Velocities in the upper crystalline crust are 5.96.2

    km/s. An intra-crustal discontinuity at 1831 km divides the crust into an upper and a lower layer. Velocities within the lower

    crust are 6.77.0 km/s. Strong wide-angle PmP reflections indicate the existence of a first-order Moho at a depth of 30 km near

    the southern end of the line and 41 km near the centre. Constraints on upper mantle seismic velocities (7.9 km/s) are provided

    by Pn arrival times from two shot points only. Within the upper mantle a low velocity zone is interpreted. Travel times of a PLP

    reflection define the bottom of this low velocity layer at a depth of 55 km. The velocity beneath this interface must be at least

    8.5 km/s. Geologic interpretation of the seismic data suggests that the Neogene tectonic convergence of the Eastern Carpathians

    resulted in thin-skinned shortening of the sedimentary cover and in thick-skinned shortening in the crystalline crust. On the

    autochthonous cover of the Moesian platform several blocks can be recognised which are characterised by different lithologicalcompositions.This could indicate a pre-structuring of the platform at Mesozoic and/or Palaeozoic times with a probable active

    involvement of the Intramoesian and the CapidavaOvidiu faults. Especially the Intramoesian fault is clearly recognisable on

    the refraction line. No clear indications of the important Trotus fault in the north of the profile could be found. In the central part

    0040-1951/01/$ - see front matterD 2001 Elsevier Science B.V. All rights reserved.P I I : S 0 0 4 0 - 1 9 5 1 ( 0 1 ) 0 0 1 9 5 - 0

    * Corresponding author. Tel.: +49-721-608-4592; fax: +49-721-71173.

    E-mail address:[email protected] (F. Hauser).

    www.elsevier.com/locate/tecto

    Tectonophysics 340 (2001) 233256

  • 8/14/2019 Vrancea_99_Hauser.pdf

    2/24

    of the seismic line a thinned lower crust and the low velocity zone in the uppermost mantle point to the possibility of crustal

    delamination and partial melting in the upper mantle. D 2001 Elsevier Science B.V. All rights reserved.

    Keywords: Vrancea earthquakes; Crustal structure; Refraction seismic; Eastern Carpathians; Dobrogea

    1. Introduction

    The Carpathian Orogen in Romania (Fig. 1) is the

    result of Cretaceous and Neogene convergence and the

    resulting mountain range is the site of still ongoing

    neotectonic activity. This activity consists of near-sur-

    face crustal deformation (compressional, extensional

    and strike slip) and of strong seismicity at intermediate

    depth (60180 km), which is concentrated in a small

    area of the Eastern Carpathians in Romania called the

    Vrancea zone (Fig. 2, Oncescu et al., 1998). This

    localised seismicity is one of only three known places

    worldwide of such a concentration of intermediate

    depth earthquakes. The other two being the Bucara-

    manga region in the Andes of Columbia (e.g. Taboada

    et al., 2000) and a place in the HinduKushPamir

    area of central Asia (e.g. Mellors et al., 1995). In all

    these areas the geodynamic setting is not yet clear.

    Subduction related processes have been suggested, in

    particular, for the Vrancea area (Roman, 1970; Radu-

    lescu and Sandulescu, 1973; Airinei, 1977; Fuchs et al.,1979; Constantinescu and Enescu, 1984; Constanti-

    nescu et al., 1973; Oncescu, 1984; Linzer, 1996; Kovac

    et al., 2000; Csontos, 1995; Girbacea and Frisch, 1998;

    Nemcok et al., 1998; Seghedi et al., 1998).

    The Vrancea zone is overlapping with the south-

    eastern outer area of the Eastern Carpathian bend (Fig.

    2). While the shallow seismic activity scatters widely

    and has moderate magnitudes (Mw 5.6), the epicen-

    tral region of the intermediate depth seismicity is

    confined to an area of only about 40 km 80 km

    (Oncescu et al., 1998). These earthquakes occurbetween 60 and 180 km depth (Figs. 3 and 4) within

    an almost vertical elongated narrow zone and fre-

    quently have large magnitudes (Mw 7.4) that have

    caused a high toll of casualties and extensive damage

    over the last centuries. Deeper events have also been

    recorded, but show only small magnitudes. The depth

    interval of the strong events is separated from the

    crustal events by a zone of weak seismicity, which is

    located between about 40 and 60 km depth (Fuchs et

    al., 1979; Oncescu et al., 1998).

    Because of the unknown relationship of this zone of

    high seismicity to the geologic structures recognised at

    the surface, a joint GermanRomanian research pro-

    gram was set up. It comprises the Collaborative

    Research Centre 461 (CRC 461) Strong Earth-

    quakesa Challenge for Geosciences and Civil Engi-

    neering at the University of Karlsruhe (Germany) and

    the Romanian Group for Vrancea Strong Earthquakes

    (RGVE) at the Romanian Academy in Bucharest

    (Wenzel, 1997; Wenzel et al., 1998a).

    The VRANCEA99 seismic refraction project (Fig.

    1) presented in this paper is a contribution to this

    research program. It was designed to study the crustal

    and uppermost mantle structure to a depth of about 70

    km underneath the Vrancea epicentral region. The

    project was jointly performed by the Geophysical and

    Geological Institutes of the University of Karlsruhe

    (Germany), the GeoForschungsZentrum in Potsdam

    (Germany) and the National Institute for Earth Physics

    in Bucharest (Romania).

    The crustal structure and seismic velocities obtainedby this project will further be used to refine the geo-

    dynamic model for the Carpathian Orogen and to

    calibrate the relative velocities obtained by a subse-

    quently performed teleseismic tomography project

    (Wenzel et al., 1998b). The detailed knowledge of the

    velocity-depth structure and of the physical properties

    of the crust and the upper mantle will help to under-

    stand the propagation of seismic waves in the region. It

    will, therefore, also contribute to the seismic risk

    assessment for the metropolitan area of Bucharest,

    due to its proximity of 100160 km to the Vranceaepicentral region with its preferably NNE SSW-direc-

    ted wave propagation (Bonjer et al., 1998; Oncescu et

    al., 1998; Wenzel and Lungu, 2000; Wenzel et al.,

    1998a,c).

    2. Geological and tectonic setting

    The Eastern Carpathians are part of the Alpine

    Carpathian orogenic belt and the result of the collision

    F. Hauser et al. / Tectonophysics 340 (2001) 233256234

  • 8/14/2019 Vrancea_99_Hauser.pdf

    3/24

    of several microplates with the European margin,

    while closing the former Tethys Ocean (Dercourt et

    al., 1993; Csontos, 1995; Channell and Kozur, 1997;

    Stampfli et al., 1998; Linzer et al., 1998; Neugebauer

    et al., 2001). The result of two main compressional

    events, which took place during the Cretaceous and

    the Neogene, are several tectonic units, which have

    been accreted in the Carpathian area. During the

    Cretaceous the Inner Carpathians or Dacides (after

    the definition of Sandulescu, 1988, see Fig. 2 for

    Fig. 1. Topographic map of Romania showing the geographical location of the VRANCEA99 seismic refraction lines. The contour lines indicate

    the depth to Moho (after Radulescu, 1988). The dashed lines labelled with roman numbers are the early International Refraction Profiles of the

    1970s.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 235

  • 8/14/2019 Vrancea_99_Hauser.pdf

    4/24

    Fig. 2. Geological overview of the Eastern Carpathian bend area and its foreland with the main crustal units, nappe structures, and faults. The lo

    refraction lines are shown with their shot points. The thick solid lines indicate the location of the geological cross sections for Figs. 3, 4 and 10. C

    in the text.

  • 8/14/2019 Vrancea_99_Hauser.pdf

    5/24

    location) converged and in the Neogene the accreted

    complex was transported to the present position (San-

    dulescu, 1988; Royden, 1988; Csontos, 1995; Ma-

    tenco and Bertotti, 2000). During the Neogene the

    Outer Carpathians or Moldavides (after the definition

    of Sandulescu, 1988, see Fig. 2 for location) got theirpresent-day geometry (Sandulescu, 1988; Ellouz et

    al., 1994; Zweigel et al., 1998; Matenco, 1997).

    The present-day Eastern Carpathians are made up

    of several major tectonic units derived from this

    activity (Fig. 2, Sandulescu, 1984; Badescu, 1998).

    To the west the nappes of the Median Dacides re-

    present a part of the Tisza Dacia microplate (after

    Csontos, 1995). Depending on the geodynamic inter-

    pre tation, the nappes of the Outer Dacides are

    explained in two different ways: (1) as a Jurassic to

    Cretaceous rift along the European margin (Sandu-

    lescu, 1988) or (2) as a Neogene oceanic domain

    between the Tisza Dacia microplate and the Euro-

    pean margin (Csontos, 1995; Nemcok et al., 1998;

    Stampfli et al., 1998; Linzer et al., 1998; Neugebauer

    et al., 2001). They mark the transition to the Molda-vides, which are part of the European margin and

    which grade into the foredeep. In the centre the

    Dacidian nappes are partially covered by the Paleo-

    gene to Neogene Transylvanian Basin, which was also

    affected by some Neogene to Quaternary compres-

    sional deformation (Ciulavu et al., 2000). Together

    with an area of Late Pliocene to Quaternary intra-

    mountain graben structures in the centre of the Carpa-

    thian bend area, the foredeep is affected by the

    youngest tectonic deformation. The front of the com-

    Fig. 3. Pre-1999 geological section along the main NNE SSW VRANCEA99 seismic-refraction line with alternative interpretations of the

    expected geological structures. Inside the Eastern Carpathians the section follows mostly the trend of main geological structures, whereas to the

    south it becomes transverse to the main geological structures. For location see Fig. 2. Circles represent the foci of the intermediate depth

    earthquakes of the Vrancea zone after Oncescu et al. (1998) projected onto this cross-section. Compiled from various sources given in the text.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 237

  • 8/14/2019 Vrancea_99_Hauser.pdf

    6/24

    pressional deformation is nowadays located inside the

    foredeep and is represented by the Pericarpathian

    Front (Fig. 2). These outermost nappes already thrust

    over the foreland platform and are covered by youngerand undeformed sedimentary layers. This cover

    extends further to the east and the south onto the

    Moesian and Scythian platforms.

    The Moldavidian zone is characterized by several

    nappes with a fold-and-thrust belt geometry (Sandu-

    lescu, 1984, 1988; Ellouz et al., 1994; Morley, 1996;

    Badescu, 1998; Zweigel et al., 1998; Matenco and

    Bertotti, 2000). The main nappes from west to east

    are: the Inner Moldavide nappes, the Tarcau nappes, the

    Marginal Folds and the Subcarpathian nappes (Fig. 2).

    The Tarcau and Marginal Folds nappes are made up

    mainly of Cretaceous marine basin sediments and

    Paleogene to Neogene flysch and other clastic sedimen-

    tary deposits, whereas the Subcarpathian nappe consistsmainly of molasse deposits. The lithology of the nappes

    consists of shales and sandstones with subordinate

    marls, limestones, tuff and conglomerates. The Mar-

    ginal Folds and Subcarpathian nappes contain also

    Neogene evaporitic formations like salt and/or gypsum

    (Sandulescu, 1988; Matenco and Bertotti, 2000).

    The thickness of the Moldavidian nappes is con-

    strained by reflection seismic data, mainly from oil

    exploration in the Subcarpathian nappe, but is only

    interpolated through balanced cross-sections for the

    Fig. 4. Pre-1999 geological section along the short W E VRANCEA99 seismic refraction line along the Putna valley with alternative

    interpretations of the expected geological structures. The section is transverse to the main structures. For location see Fig. 2. Circles represent

    the foci of the intermediate depth earthquakes of the Vrancea zone after Oncescu et al. (1998) projected onto this cross-section. Compiled from

    various sources given in the text.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256238

  • 8/14/2019 Vrancea_99_Hauser.pdf

    7/24

    Marginal Folds and Tarcau nappes (Stefanescu and

    Working Group, 1988; Ellouz et al., 1994; Morley,

    1996; Matenco and Bertotti, 2000). These cross-sec-

    tions are constrained by surface geology and boreholedata, but the structures of the deeper nappes are fairly

    unknown and their balancing therefore less accurate.

    The International Geotraverse XI (Fig. 1) which

    crossed the Carpathians between Focsani and Targu

    Secuiesc gives only some vague estimations on nappe

    thickness (Radulescu et al., 1976). Magnetotelluric

    data from Stanica and Stanica (1998) indicate a depth

    of 8 km for the base of the Moldavidian nappes and of

    about 1416 km for the basement in the area around

    shot point D (Figs. 13).

    The Carpathian foreland consists of older consoli-

    dated crustal blocks with their pre-orogenic autochtho-

    nous cover (Sandulescu, 1984; Visarion et al., 1988;

    Ellouz et al., 1994; Zugravescu and Polonic, 1997;

    Seghedi, 1998; Matenco and Bertotti, 2000). It is part

    of the Moesian Platform to the south and the Scythian

    Platform to the northeast (part of the Eastern European

    platform), which are distinguished by geophysical

    (mainly magnetic) anomalies in the crystalline base-

    ment and lithologic differences in the sedimentary

    cover (Seghedi, 1998 and references therein) and

    which are separated by the Trotus fault (Fig. 2). The

    crystalline basement of the two platforms is made up ofmetamorphic and intrusive magmatic rocks, which

    sometimes provide a weak acoustic contrast to their

    older sedimentary cover (Raileanu et al., 1994). The

    sedimentary rocks, separated by several unconform-

    ities, consist of Paleozoic and Mesozoic detritic and

    carbonaceous deposits and a very thin undeformed

    Neogen cover, which shows a slight dip towards the

    central part of the foredeep (Raileanu et al., 1994). Near

    the orogenic front of the Carpathians the platform

    sediments are partly covered by the foredeep sedi-

    ments. The sedimentary succession above the crystal-line basement has been especially well studied for oil

    exploration by seismic reflection techniques and by

    wells drilled in the southern part of the Moesian plat-

    form.

    In the autochthonous and overthrusted areas of the

    Moesian platform the seismic refraction line is

    expected to cross two major crustal faults, the Cap-

    idavaOvidiu Fault (COF) and the Intramoesian Fault

    (IMF). Whereas the COF is outcropping in the

    Dobrogea area near the Black Sea, sediments and

    nappes cover the supposed NW-prolongation of the

    fault, as well as the IMF (Figs. 2 and 3; Visarion et al.,

    1988; Polonic, 1996; Ellouz et al., 1994; Seghedi,

    1998). The COF separates a greenschist basement tothe north from a higher-grade metamorphic basement

    to the south (Seghedi, 1998 and references therein).

    This crystalline basement extends to the south as far

    as the IMF, which again separates two different parts

    of the Moesian platform. They are characterized by

    different structural orientations and lithologies, with

    magmatic intrusions into the western block (Seghedi,

    1998 and references therein). The IMF is an active

    fault along which many low magnitude earthquakes

    have been recorded (Radulescu et al., 1976; Cornea

    and Polonic, 1979; Zugravescu and Polonic, 1997).

    Mainly from seismic data all these crustal faults are

    thought to prolongate down to respectively cross the

    Moho discontinuity (e.g. Visarion et al., 1988; Radu-

    lescu and Diaconescu, 1998).

    Using published (Dumitrescu and Sandulescu,

    1970) as well as unpublished (Fielitz, personal com-

    munication) geological data, crustal cross-sections for

    both refraction lines were constructed and are shown

    in Figs. 3 and 4. They form the base for the inter-

    pretation of the velocity-depth model obtained from

    the seismic experiment (Figs. 9 and 10). Concerning

    the nappe and the basement structures, two alternativeinterpretations are possible. In the interpretation of

    Ellouz et al. (1994) the basement and the autochtho-

    nous cover of the Moesian platform is affected by the

    stacking of the Moldavidian nappes (Fig. 3, top and

    Fig. 4, left). In the interpretation of Sandulescu (1984)

    and Stefanescu and Working Group (1988) shortening

    affects only the Moldavidian nappes, i.e. the cover

    (Fig. 3, bottom and Fig. 4, right). Both cross-sections

    also show the foci of the intermediate-depth earth-

    quakes of the Vrancea zone for the last 10 years

    (Oncescu et al., 1998) projected onto the profiles.Based on seismic reflection and geological data,

    Matenco and Bertotti (2000) tried to estimate the

    depth of the different nappes. Using their results,

    the predictions for the area containing the seismic

    refraction profile can be summarized as follows.

    Between shot points B and C the Subcarpathian

    nappe is about 6 km thick, while the thickness of

    the Marginal Folds nappe around shot point D should

    be 7 km. The thickness of the Tarcau nappe (between

    shot points E and H) is expected to be about 67 km.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 239

  • 8/14/2019 Vrancea_99_Hauser.pdf

    8/24

    Near shot point K the Subcarpathian nappe should be

    6 km thick, while the top of the Mesozoic platform

    cover is expected to be about 5 km deep (Figs. 24).

    3. Earlier geophysical investigations

    The area crossed by the seismic refraction line

    VRANCEA99 (Fig. 1) has already been investigated

    by earlier geophysical work. While the flysch area

    is only poorly explored, the Moesian platform has

    been investigated in more detail by both seismic

    reflection and refraction methods. The seismic

    reflection investigations have revealed the structure

    of the Neogene and Mesozoic sediments and to a

    lesser extent the Paleozoic sediments and the crys-

    talline basement.

    The Neogene cover shows many reflections,

    which can be correlated over long distances. One

    such regional reflector, which can be observed over

    the whole platform, is the Cretaceous erosional relief

    unconformably overlain by Neogene sediments (Rai-

    leanu, 1998; Raileanu and Diacunescu, 1998; Rai-

    leanu et al., 1994). Across this interface a velocity

    jump of more than 1 km/s generates strong reflec-

    tions. On a short reflection line, about 20 km north-

    east of shot point M, the P wave velocities in theNeogene sediments range from 2.0 to 3.5 km/s,

    while the Mesozoic and underlying layers have Vp

    values between 3.5 km/s and over 6.0 km/s (Raileanu

    et al., 1993). For the Mesozoic and Paleozoic layers

    with clastic sedimentary rocks the seismic P wave

    velocities are less than 4 km/s, while they are higher

    for the limestone and dolomite (Raileanu et al.,

    1993).

    In the Moesian area the reflectivity on seismic

    reflection data decreases with depth. The contact of

    the sedimentary cover with the crystalline basementdoes not generate reflected waves for near-vertical

    angles of incident (Raileanu et al., 1994). In addition,

    the crystalline crust is only weakly reflective all the

    way down to the Moho. Some multiple waves,

    originating in the Neogene or Mesozoic successions,

    have been observed and may hide the deeper and

    weaker reflected waves (Raileanu et al., 1994). The

    base of the crust is marked by the complete disap-

    pearance of reflectivity and lies at a depth of 35 38

    km according to Raileanu et al. (1993).

    Seismic refraction data collected on the eastern

    part of the Moesian platform yielded P wave veloc-

    ities of 5.9 6.2 km/s for the top of the basement

    (Radulescu et al., 1976; Cornea et al., 1981). The baseof the crust was mainly detected by reflected waves at

    wide-angle distances and to a lesser extent by head

    waves, due to the short length of the seismic spreads.

    The depth to Moho is given between 30 and 40 km.

    A short (only 20 km long) seismic reflection line

    in the Subcarpathians, just west of shot point B (Fig.

    2), shows a weakly reflective upper and lower crust.

    The main reflectors according to Raileanu et al.

    (1994) are the seismic basement at 12 km depth, a

    mid-crustal discontinuity at 23 km and the Moho at

    43 km depth.

    Results from the International Geotraverse XI (Fig.

    1) on the Eastern Carpathians give the depth to base-

    ment with 11 km. The depth to the middle crust is

    given as 28 km, while the Moho discontinuity lies at

    45 km depth (Radulescu et al., 1976).

    Magnetotelluric data along a profile almost parallel

    to the transverse line (shot points R and S in Figs. 1

    and 2) have provided a crustal pattern of four layers

    with alternating electric conductivity. For the area of

    the refraction seismic line the first layer is assigned to

    the Marginal Folds nappe with a thickness of 8 km.

    The second one to the older sedimentary cover of theunderlying platform at 1416 km depth and the third

    one to the base of the crystalline crust at 50 km depth.

    The deepest layer was interpreted as upper mantle

    (Stanica and Stanica, 1998).

    The Bouguer anomaly map of Visarion (1998)

    shows maximum negative values for the foredeep

    region and the neighbouring flysch nappes. This is

    caused by the thick folded and thrusted nappes, the fill

    of the foredeep and the thickened crust. The Bouguer

    anomaly is negative along the entire refraction line

    with lowest values within the internal sector of theforedeep (i.e. from south of shot point G to shot point

    L in Figs. 1 and 2). Towards both ends of the

    refraction line the Bouguer anomaly increases to 0

    mgal in the south and 50 mgal around Bacau in the

    north.

    Using a large amount of geological and geophys-

    ical data Polonic (1996, 1998) compiled a map of the

    crystalline basement for Romania. Along the VRAN-

    CEA99 refraction line the depth to basement increases

    from 6 km near Bucharest to 15 km under shot point

    F. Hauser et al. / Tectonophysics 340 (2001) 233256240

  • 8/14/2019 Vrancea_99_Hauser.pdf

    9/24

    G. Further north it decreases again to 7 km around

    Bacau.

    Based on seismic refraction and reflection data, as

    well as on gravimetric and other geophysical data,Radulescu (1988) and Enescu et al. (1992) produced

    regional maps for the Conrad and Moho disconti-

    nuities in Romania. The Conrad discontinuity varies

    in depth from 14 km near the southern end of the

    VRANCEA99 line to 28 km in the Focsani depres-

    sion and 22 km near Bacau. The Moho contours

    show about 30 km depth near Bucharest, about 50

    km for the Carpathian bend area and about 40 km

    near Bacau (Fig. 1). In addition to the above-men-

    tioned maps, Enescu et al. (1992) also used travel

    time information of some shallow seismic events (0

    10 km depth) to derive empirical velocity-depth

    functions (P and S waves) for the crust and for the

    upper mantle down to 70 km depth. The reported P

    wave velocity at the base of the crust is about 6.95

    km/s. The mean P wave velocity for the entire crust

    including the sediments is 6.18 km/s for the orogenic

    area and 6.21 km/s for the Moesian and East Euro-

    pean platforms.

    4. The VRANCEA99 seismic experiment

    In order to achieve the objectives discussed above,

    the VRANCEA99 seismic refraction survey was car-

    ried out in May 1999. The project consisted of two

    lines intersecting each other at the northern edge of

    the Vrancea epicentral region (Fig. 1).

    The main line is 300 km long and runs from the city

    of Bacau through Bucharest and onto the Danube

    River, crossing the Vrancea region in a NNESSW

    direction. It crosses the bending area of the Eastern

    Carpathian Orogen in the Moldavidian zone, the fore-

    deep and its eastern and southern foreland (Scythianand Moesian platforms). The line (Figs. 13) starts in

    the Subcarpathian nappe, which outcrops between shot

    points A and D. Along this segment of the seismic line

    the Subcarpathian nappe overlies the Scythian plat-

    form to the north of the Trotus Fault (TF) and the

    Moesian platform to the south. Until about 5 km north

    of shot point E the seismic line cuts across the

    Marginal Folds nappe. The Tarcau nappe is crossed

    until about 10 km south of shot point G. From shot

    sites A to G the direction of the seismic line is almost

    parallel to the direction of the folded structures, there-

    after it is oblique to the geological structures. The next

    segment of the seismic line, between shot points H and

    K crosses again the Subcarpathian nappe. The fore-deep segment comprises shot points L and M, where

    the basement is already part of the Moesian platform.

    The southernmost shot point N south of Bucharest is

    fully located on the Moesian Platform outside the

    foredeep.

    An additional, short refraction line transverse to the

    geological structures passes along the Putna Valley,

    north of the bend area intersecting with the main

    refraction line at shot point D (Figs. 14). Two more

    shot points are located on this line, one in the Focsani

    foredeep to the east (shot point S) and one in the

    Tarcau nappe east of Targu Secuiesc to the west (shot

    point R).

    Along the two segments a total of 140 recording

    sites were occupied with an average station spacing of

    2 km. Seismic recording equipment for the experiment

    was provided by the GeoForschungsZentrum Potsdam

    (Germany), by Leicester University (UK) and by the

    NERC geophysical equipment pool (UK).

    The spacing of shot points ranged from 12 to 14

    km between shot points H K and E G to 29 30 km

    between L M and G H, with an average of 22 km.

    The charge sizes varied from 300 to 900 kg with thelarger shots being A and M (900 kg each) and B and L

    (600 kg each), close to the end points of the main line

    (Figs. 1 and 2). All shots were recorded simultane-

    ously along the main line and the transverse line. For

    more technical details on the project see Prodehl et al.

    (2000) and Hauser et al. (2000).

    As the 70 km long transverse line running perpen-

    dicular to the main line along the Putna Valley (Fig. 2)

    did not resolve the deeper levels of the crust, it is not

    discussed further in this paper.

    5. The seismic sections

    The seismic record sections were compiled and

    plotted using the SeismicHandler program package of

    Stammler (1994). All record sections (Figs. 5 8)

    presented in this paper are vertical component data

    plotted with a reduction velocity of 6 km/s. Seismo-

    grams are trace normalised, which means that the

    amplitudes are scaled with respect to the maximum

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 241

  • 8/14/2019 Vrancea_99_Hauser.pdf

    10/24

    F. Hauser et al. / Tectonophysics 340 (2001) 233256242

  • 8/14/2019 Vrancea_99_Hauser.pdf

    11/24

    amplitude per trace. A general band pass filter

    between 3 and 14 Hz has been applied to the data

    in order to improve the signal-to-noise ratio.

    For this paper, the term travel time refers tothe reduced travel time of the seismic sections, the

    term offset to the distance between source and

    receiver, and the term distance to a location

    along the refraction model with respect to shot

    point A.

    The data quality for the experiment is very

    variable and seems to depend strongly on local

    conditions, like structure and physical properties of

    the corresponding shot sites and local receiver

    conditions. The maximum offset to which first or

    later arrivals can clearly be seen on the seismic

    sections may be considered as a measure of data

    quality. The northern segment of the refraction

    profile generally has a low signal-to-noise ratio.

    This may be due to the fact, that most of the shot

    points (A K and R S) are located within the

    Carpathians where nappes and faults cause strong

    scattering and absorption of seismic energy and

    therefore reduce the signal-to-noise ratio. Only for

    shot point A (Fig. 5) can arrivals be picked over an

    offset range of 200 km. Shot points L (Figs. 6 and

    7) and M (Fig. 8) on the other hand, which are

    located in the Carpathian foreland, have the highestsignal-to-noise ratios.

    Not all phases have been identified on all record

    sections, sometimes due to the limited offset, but

    sometimes also because of the low signal-to-noise

    ratio. We have only picked those phases, which are

    coherent over several traces. A summary of the

    phases observed on each section is given in Table 1.

    rUp to four separate first-arrival refractions with appa-

    rent velocities < 6 km/s could be identified on the

    seismic sections on the basis of apparent velocity and

    offset distribution. They have been named P1 to P4.Most of the data also show a clear Pg phase (a diving

    wave through the upper crust) as first arrivals between

    40 and 100 km offset. This phase is characterised by

    strong undulations and sometimes very small ampli-

    tudes, making picking generally difficult beyond 80 km

    offset.

    Within the deeper crust we could identify PcP

    (reflection from the top of the lower crust) as asecondary arrival behind Pg up to 200 km offset.

    The phase shows an apparent velocity of 6.1 6.3

    km/s at furthest offsets. It varies from being prom-

    inent and laterally coherent over 100 km offset to

    being weak and only visible in the sub-critical

    offset range. This could suggest that the mid-crustal

    boundary is laterally not continuous along the

    profile.

    The reflection from the Moho (PmP) is observed

    on several record sections, especially from shots at

    both ends of the profile. The phase is characterised by

    high lateral coherency, very strong amplitudes and has

    an apparent velocity of 6.8 km/s at furthest offset.

    This would suggest that the Moho is a laterally

    continuous, sharp discontinuity or a thin transition

    zone.

    A diving wave through the upper mantle (Pn) can

    only be seen for the southern shot points L (Figs. 6

    and 7) and M (Fig. 8). It shows an apparent velocity of

    8 km/s and can be picked out to the maximum offset

    of over 200 km. For the same two shot points (L and

    M) a mantle phase (PLP) can be seen in the data as

    well. It is characterised by a high apparent velocity(8.6 km/s) and strong amplitudes (sometimes as large

    as or larger than PmP). This phase is coherent

    between 100 km offset and the end of the seismic

    sections.

    6. Interpretation techniques and the velocity model

    There are several major steps in the modelling

    procedure:

    (1) Travel times and associated errors were pickedfor each of the seismic phases described above. The

    integrity of the picked travel times and the consistency

    of the phase identification were checked by compar-

    ing reciprocal travel times where possible.

    Fig. 5. (a) Trace normalised P wave record section and (b) synthetic seismograms for shot point A, reduced with 6 km/s. The calculated travel

    times from the model in Fig. 9 are plotted on top of the data. Travel times are labelled as follows: P1P4, first-arrival phases refracted within the

    sedimentary cover; Pg, diving wave through the upper crust; PcP, reflection from the top of the lower crust; PmP, reflection from the Moho; Pn,

    diving wave through the upper mantle; PLP, reflection from the upper mantle.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 243

  • 8/14/2019 Vrancea_99_Hauser.pdf

    12/24

    Fig. 6. (a) Trace normalised P wave record section and (b) synthetic seismograms for shot point L observed to the north. For further explanations

    see Fig. 5.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256244

  • 8/14/2019 Vrancea_99_Hauser.pdf

    13/24

    Fig. 7. (a) Trace normalised P wave record section and (b) synthetic seismograms for shot point L observed to the south. For further explanations

    see Fig. 5.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 245

  • 8/14/2019 Vrancea_99_Hauser.pdf

    14/24

    F. Hauser et al. / Tectonophysics 340 (2001) 233256246

  • 8/14/2019 Vrancea_99_Hauser.pdf

    15/24

    (2) One-dimensional velocity-depth functions were

    calculated for each shot point.

    (3) The resulting 1-D models were merged into a 2-

    D cross-section along the main profile, making due

    allowance for the offsets in the different phases. Next

    the data were interpreted by 2-D forward modelling,

    initially using the structure deduced from the 1-D

    models. The ray-tracing program SEIS83 of Cerveny

    and Psencik (1983) was used for this step.

    (4) Finally, when the velocity model obtained by

    forward modelling was good enough to identify theobserved arrivals confidently, a travel time inversion

    was carried out, using the method of Zelt and Smith

    (1992). The misfit between the picked and calculated

    travel times was minimised using a damped least-

    squares inversion, which also allows the resolution for

    individual model parameters to be quantified.

    (5) The relative amplitudes of individual phases

    were estimated qualitatively where possible and triedto match by varying the velocity gradient within

    individual layers of the model. However, a detailed

    trace-by-trace amplitude modelling was not attemp-

    ted. Synthetic seismograms were calculated again

    using the program SEIS83 (Cerveny and Psencik,

    1983). The resulting synthetic seismograms are shown

    in Figs. 58 together with the observed data.

    6.1. The velocity structure of the crust and upper

    mantle

    The final 2-D velocity model derived using the

    methods described is shown in Fig. 9a. It has a

    multi-layered character with a mean velocity for the

    whole crust of 6.2 km/s. The main structures crossed

    by the seismic line can be separated into three

    groups. (1) The sedimentary cover showing veloc-

    ities < 6 km/s. (2) The crystalline crust down to

    Moho. (3) An upper mantle structure at sub-Moho

    level.

    The sedimentary succession along the main line

    consists of two to four layers with velocities ranging

    from 2.0 to 5.8 km/s. The acoustic basement coincideswith a velocity step up to 5.9 km/s. The crustal

    velocity is fairly constant lateral direction and in-

    creases gradually to 6.2 km/s at the base of the upper

    crust. An intra-crustal discontinuity, defined by sec-

    ondary PcP reflections is present and divides the crust

    into an upper and a lower layer. Velocities within the

    lower crust again seem fairly constant in a horizontal

    direction and increases vertically from 6.7 to 7.0 km/s

    at the Moho.

    Strong wide-angle Moho reflections (PmP) indicate

    the existence of a first-order crustmantle boundary.The depth to Moho increases from 38 km at the north-

    ern end of the profile to 41 km between shot points F

    and L. South of L it decreases to 30 km under shot point

    N (Fig. 9a). A constraint on upper mantle seismic

    Fig. 8. (a) Trace normalised P wave record section and (b) synthetic seismograms for shot point M observed to the north. For further

    explanations see Fig. 5.

    Table 1

    Summary of the phase correlation

    P1 P2 P3 P4 Pg PcP PmP Pn

    A S XXX XXX XXX XX X XB N XXX XXX XX

    B S XXX XX XX XX X

    C N XXX XXX XX XX

    C S XXX XXX XX X

    D N XX XX XX X

    D S XXX XXX XX

    E N XXX XXX XX X

    E S XX XX XX

    F N XX XX XX X

    F S XX XX

    G N XX XX XX X

    G S XXX XXX XX X

    H N XXX XXX XXX XX XX X X

    H S XXX XXX XX XX X X XK N XXX XXX XX XX XX X X

    K S XXX XXX XX XX X X X

    L N XXX XXX XXX XXX XX XX XX XX

    L S XXX XXX XX XX XX XX XXX

    M N XXX XXX XXX XX XX XXX XX

    M S XXX XX XX XX XXX

    N N XXX X X

    N S XXX X X

    XXX indicates an easy correlation, XX a less clear correlation,

    while X a very difficult correlation. A blank field means that this

    phase was either not present or could not be correlated.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 247

  • 8/14/2019 Vrancea_99_Hauser.pdf

    16/24

    velocities (7.98.0 km/s) is provided by Pn arrival

    times picked from shot points L (Fig. 6) and M (Fig. 8).

    Based on PLP reflections from the same two shot

    points, a low-velocity zone with a velocity of 7.6 km/s

    was modelled within the upper mantle. The base of

    this velocity inversion lies at a depth of 55 km. The

    velocity beneath this interface must be at least 8.5 km/

    s in order to match the amplitudes and critical dis-

    F. Hauser et al. / Tectonophysics 340 (2001) 233256248

  • 8/14/2019 Vrancea_99_Hauser.pdf

    17/24

    tances of the PLP reflections as seen on shot points L

    (Fig. 6) and M (Fig. 8).

    6.2. Model resolution and uncertainties

    In a qualitative sense it is obvious that velocities

    derived from refracted phases are certainly more reli-

    able than those derived from reflected phases. It is also

    intuitive that the reliability to the depth of a reflector

    increases with the number of rays reflected from aninterface. The advantages of using an inversion algo-

    rithm like the one described by Zelt and Smith (1992)

    are the ability to construct models that satisfy a large

    number of shots and receivers and to assess the model

    resolution, uncertainty and non-uniqueness of the

    derived model in a more objective way.

    The reliability of the model was quantified during

    the inversion procedure, where estimates of the reso-

    lution and the absolute parameter uncertainties were

    calculated. Travel time fits are assessed using the

    normalised form of the misfit parameter v2 (Beving-

    ton, 1969). In general a value of v2, as close to 1 as

    possible, is sought as this indicates that the computed

    travel times fit the data to within their assigned

    uncertainties. Values much less than 1, indicate that

    the data have been overfit. As a result the model will

    contain structure not required by the data. Final values

    of v2 much greater than 1 generally indicate the data

    have sampled small-scale heterogeneities they cannot

    resolve (Zelt and Smith, 1992). In practice, final v2

    values greater 1 are acceptable if the parameter

    resolution is high (see below) and if it is possible to

    trace rays to all stations. The root mean square travel

    time residual for the whole model is 0.12 s, and the

    normalised v2 value is 1.339. The values for individ-

    ual phases are shown in Table 2. While the final

    model obtained through travel time and amplitude

    modelling is shown in Fig. 9a, ray coverage and travel

    time fits for all shots and all phases are shown in Fig.

    9b and c, respectively.

    Fig. 9. (a) Final 2-D velocity-depth model along the main VRANCEA99 line, starting from the city of Bacau in the north, passing through

    Bucharest and ending near the river Danube. The profile is traversing the Vrancea epicentral region in a NNE SSW direction. Labelled dots at

    the top of the model indicate the shot points, while numbers indicate the P wave velocities in km/s. Thick solid lines indicate areas which are

    constrained by reflections and/or refractions. (b) Ray coverage through the final model connecting all shot and receiver pairs. (c) Comparison of

    observed and calculated travel times for all shots and all phases. Vertical bars indicate observed data with height representing pick uncertainties.

    Solid lines indicate calculated travel times.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 249

  • 8/14/2019 Vrancea_99_Hauser.pdf

    18/24

    The diagonal elements of the resolution matrix

    range between zero and one. While a value close to

    1 indicates a well resolved parameter, a low value (i.e.

  • 8/14/2019 Vrancea_99_Hauser.pdf

    19/24

    It shows some topography, and two steps are obvious

    in this layer. The first one occurs between shot points

    B and C and has a vertical offset of about 1 km. It is

    located just a few kilometres south of the Trotusvalley and can be interpreted as either being the

    Trotus Fault or a local fault striking transverse to the

    nappe structures and separating two blocks of the

    same nappe as a transfer fault. The second step is

    located south of shot point G and has a vertical offset

    of about 4 km. This coincides with a slight increase in

    thickness, but a decrease in velocity. Furthermore,

    halfway between shot points G and H this layer

    plunges beneath layer 1 and pinches out near shot

    point M, south of the Pericarpathian Front. This

    second step corresponds to the onset of the foredeep

    cover and the pinching out of the layer further south

    coincides with the location of the Intramoesian Fault.

    Near the fault, layer 2 corresponds to the undeformed

    pre-Sarmation beds of the foredeep. These same beds

    have been folded and thrusted further north until the

    front of the Subcarpathian nappe between shot points

    K and L (the Pericarpathian Front in Fig. 2). Velocities

    in this layer range from 3.43.9 km/s for the Molda-

    vian nappes to 3.33.5 km/s for the subsided part of

    the Subcarpathian nappes.

    The third seismic layer starts from the northern end

    of the model, but pinches out south of shot point M(Fig. 9a). It consists of the Subcarpathian, the Mar-

    ginal Folds, and the Tarcau nappes as well as of

    foredeep rocks. The P wave velocities in this layer

    range from 4.0 to 4.8 km/s. The marked decrease in

    velocity under shot point G could indicate a change in

    petrological composition towards less consolidated

    rocks. The layer is at a constant depth of 4 km at

    the northern end of the line, but deepens to 8 km

    between shot points G and H. South of shot point H

    its depth decreases again, while the velocity increases

    until it pinches out.Geological cross-sections for the area (Matenco

    and Bertotti, 2000) show the occurrence of Miocene

    salt deposits between shot points A and F at 4 km

    depth, which could explain the observed velocity of

    4.7 km/s at the base of this layer. Therefore, it is

    very likely that between those shot points the

    seismic energy was reflected from the top of the

    salt beds. South of shot point F, where according to

    Matenco and Bertotti (2000) the salt beds disappear

    due to facies changes in the more inner parts of the

    Moldavidian nappes, the reflector deepens and the

    velocities decrease. In this part of the seismic

    profile the base of the third seismic layer would

    correspond to the base of the Moldavidian nappes.The velocity increase south of shot point H can be

    interpreted with the reappearance of the salt beds

    facies in the again more outer parts of the Molda-

    vidian nappes.

    The fourth seismic layer is the only sedimentary

    layer that can be observed along the entire seismic

    refraction profile (Fig. 9a). It probably represents the

    autochthonous Palaeozoic, Mesozoic, and maybe, the

    very thin Cenozoic sedimentary cover rocks of the

    Moesian platform. The velocities within this layer

    vary from 4.7 to 5.8 km/s. The greatest thickness of

    about 9 km is reached between shot points F and G,

    but it decreases towards both ends to 5 km.

    The marked lateral velocity changes within this

    layer allow a sub-division into several blocks. (1) A

    southern block with highest velocities (5.35.8 km/

    s). (2) A central block with lowest velocities (4.7

    5.2 km/s) between shot points L and G, which also

    coincides with the position of a minimum in the

    Bouguer anomaly. (3) This is followed by a smaller

    high-velocity block (5.35.7 km/s) and (4) a north-

    ern block with intermediate velocities of 5.0 5.5

    km/s.These observations led us to the conclusion that the

    Moesian platform is made up of several blocks with

    different lithological compositions. These blocks

    could correspond to some kind of Mesozoic and/or

    Palaeozoic pre-structuring of the platform. The Intra-

    moesian (IMF) and the Capidava Ovidiu (COF)

    faults could have played an important role, because

    of their proximity to the described block boundaries.

    It is possible that the observed lateral velocity

    changes across the COF correlate also with known

    differences in the basement, which is made up near itstop of greenschist to the north and crystalline rocks to

    the south of the fault. This would point to different

    geological evolutions of the two blocks. No velocity

    changes have been observed for this seismic layer

    across the Trotus Fault (TF).

    7.2. The crustal structure

    The next two seismic layers make up the crystalline

    crust, which consists of an upper (layer 5) and a lower

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 251

  • 8/14/2019 Vrancea_99_Hauser.pdf

    20/24

    (layer 6) crustal layer. The crystalline crust shows some

    thickness variations, while the lateral velocity structure

    along the entire seismic line remains constant. The

    upper crustal velocity of 5.9 km/s is low and may beinterpreted as a low-grade metamorphic sedimentary

    composition.

    The total thickness of the crustexcluding sedi-

    mentsis 2830 km for the northern and central part

    of the model. Further to the south it decreases to 25 km

    under shot point N. The marked decrease in thickness

    corresponds to the location of the Intramoesian Fault.

    An intra-crustal boundary was recognised from wide-

    angle reflections. It separates an upper crust with

    velocities of 5.96.2 km/s from a lower crust with

    velocities of 6.7 7 km/s.

    The upper crust reaches its greatest thickness of 20

    km in the centre of the model. Towards both ends it

    decreases to 13 km. On the other hand the thickness of

    the lower crust gradually increases from 10 km near the

    centre of the seismic model towards both ends where it

    reaches 12 kmin the south and about 16 km in the north.The fact that the velocity structure of the crust shows

    no lateral variations while the thickness does so could

    suggest the existence of an originally homogeneous

    crust of approximately constant thickness north of the

    Intramoesian Fault (IMF) and a thinner crust of com-

    parable composition south of it. Later on the crustal

    segment north of the IMF was deformed at mid-crustal

    level by shortening near the centre of the seismic line.

    The overthrusted body must have had the same phys-

    ical properties and/or lithological composition as the

    undeformed crust, because the present crust has the

    same velocity. As the thickened upper crust coincides

    with the area of emplacement of the Moldavidian

    Fig. 10. Interpreted geological section along the main VRANCEA99 seismic refraction line between Bacau and Bucharest from the 2-D seismic

    model of Fig. 9a. Inside the Eastern Carpathians the section follows mostly the trend of the main geological structures, whereas to the south it

    becomes transverse to the main geological structures due to the bending of the Carpathians. For location see Fig. 2. Circles represent the foci of

    the intermediate depth earthquakes of the Vrancea zone after Oncescu et al. (1998) projected onto this cross-section.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256252

  • 8/14/2019 Vrancea_99_Hauser.pdf

    21/24

    nappes, it is suggested that this thickening is due to the

    same shortening process, with internal thrusting of the

    crystalline crust, which could have been favoured by

    the lower rheologic strength of the middle crust. TheCapidavaOvidiu and the Intramoesian faults could

    bound this thrust sheet, respectively.

    The thinned lower crust between the Capidava

    Ovidiu and the Intramoesian faults correlates well

    with a low velocity layer at the top of the upper

    mantle (Figs. 9a and 10). An explanation for this

    could be delamination of the lower crust, which

    allowed partial melting of the uppermost mantle.

    7.3. Moho and upper mantle structure

    The wide-angle Moho reflections indicate the exis-

    tence of a first-order discontinuity. The depth to Moho

    increases from 38 km under shot point A to 41 km

    between shot points F and L. Further to the south it

    decreases again to 30 km under shot point N. These

    values are shallower than shown in previous studies

    (Enescu et al., 1992), with the maximum depth shifted

    somewhat to the south (Fig. 1). It is also interesting to

    note that the deepest part of the Moho lies between the

    Intramoesian and the CapidavaOvidiu faults, while

    the hypocenters of the intermediate depth earthquakes

    project onto the profile slightly to the north of the laterfault zone (Fig. 10).

    The uppermost mantle has a velocity of 7.9 km/s,

    which suggests a homogeneous lithological composi-

    tion. As has been pointed out above we modelled a 4-

    km-thick layer below the Moho with a small velocity

    gradient, which is underlain by a low velocity zone.

    The travel times of the PLP reflections define an inter-

    face in the mantle that lies at a depth of 55 km. The

    velocity beneath this interface must be at least 8.5 km/s.

    The depth interval of the low velocity zone coincides

    well with the seismic gap between crustal and inter-mediate depth earthquakes in the Vrancea zone (Fig.

    10, Oncescu et al., 1998). Fuchs et al. (1979) proposed

    a zone of low strength in the upper mantle in order to

    explain this seismic gap. Such a zone of low strength

    could show up as a sub-crustal low velocity zone,

    which was also observed for the area by other authors

    (Lazarescu et al., 1983; Fan et al., 1998).

    The low velocity of 7.6 km/s could suggest a

    mixture of crustal and upper mantle rocks or a process

    of eclogitisation, i.e. a transition from crustal to mantle

    rocks. Another interpretation could be a zone of partial

    melting in the uppermost part of the mantle. This could

    coincide with a delamination of the lower crust as

    described above. A sub-crustal low-velocity zone alsoplays a crucial role in the deep lithospheric model

    proposed by Chalot-Prat and Girbacea (2000) for the

    Eastern Carpathians. These authors suggest that slab

    rollback and break-off induced delamination of the

    European mantle lithosphere and upwelling of the

    asthenosphere into the newly created space.

    8. Conclusions

    A 300-km-long NNE SSW trending seismic

    refraction line was carried out in Romania in order

    to study the lithosphere underneath the Vrancea epi-

    central region within the SE Carpathians. The inter-

    pretation of the data by forward and inverse modelling

    gave the following results.

    The sedimentary succession is up to 13 km thick

    and can be sub-divided into two to four layers. It

    comprises the Moldavidian nappes, the Neogene infill

    of the foredeep and cover of the Moesian and Scy-

    thian platforms, as well as the autochthonous Meso-

    zoic and Palaeozoic sedimentary rocks of the Moesian

    and Scythian platforms.The underlying crystalline crust shows thickness

    variations, but at the same time the lateral velocity

    structure along the seismic line remains constant. An

    intra-crustal boundary separates an upper crust with

    velocities of 5.96.2 km/s from the lower crust with

    velocities of 6.77 km/s. Within the upper mantle we

    observe a low velocity zone down to a depth of 55 km.

    Using the VRANCEA99 seismic refraction experi-

    ment in addition to other geophysical data, a crustal

    cross-section can be derived (Fig. 10). In the centre of

    the model the crystalline crust is thickened andcovered by the Moldavidian flysch nappes and the

    Carpathian foredeep up to the Pericarpathian Oro-

    genic Front. The deepest segment is situated approx-

    imately between shot points L and G under the

    Carpathian foreland and is also confirmed by both

    Bouguer and isostatic gravity anomalies (Visarion,

    1998; Rosca, 1998). It seems that the Neogene tec-

    tonic convergence resulted in thin-skinned shortening

    of the sedimentary cover and in thick-skinned short-

    ening in the deeper part of crust. This would correlate

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 253

  • 8/14/2019 Vrancea_99_Hauser.pdf

    22/24

    with the first alternative interpretation of the geo-

    logical cross-sections from Figs. 3 and 4. On the

    Moesian platform itself several blocks with different

    lithological compositions can be recognised, whichpoints to a Mesozoic and/or Palaeozoic pre-structur-

    ing of the platform, where the Intramoesian and the

    CapidavaOvidiu faults probably played an important

    role, because of their proximity to two of the

    described block boundaries. Especially the Intramoe-

    sian fault is evident from interpretations of the seismic

    refraction data. A reactivation of these crustal faults

    during the Neogene shortening is probable. However,

    no clear indications of the important Trotus fault in the

    north have been found.

    A thinned lower crust between the Capidava

    Ovidiu and the Intramoesian faults correlates well

    with a low velocity layer at the top of the upper

    mantle. This could be explained by delamination of

    the lower crust, which may have allowed partial

    melting of the uppermost mantle.

    Acknowledgements

    This investigation was only enabled by the

    continuous effort of many volunteers. In particular

    we thank all additional participants in the field work:G. Danci, M. Diaconescu, A. Hlevca, D. Mateciuc, L.

    Munteanu, V. Nacu (NIEP, Bucharest), V. Dumitrescu

    and M. Georgescu (Geotec, Bucharest), T. Orban

    (University, Bucharest), J. Bribach (Potsdam), P.

    Denton and A. Myers (Leicester), G. OBrien (Dublin),

    H. Raue (Wurzburg), S. Bourguignon, A. Goertz, P.

    Heidinger, C. Jaeger, I. Koglin, H-M. Rumpel and C.

    Weidle (Karlsruhe). The National Institute for Earth

    Physics (NIEP) provided the logistics for the fieldwork

    in Romania. We express our sincere thanks to

    Professor D. Enescu and Dr. G. Marmureanu, GeneralDirectors. Dr. Mihaela Rizescu and Mihaela Popa

    ensured the recordings of the NIEP network. We are

    extremely grateful to Professor C. Dinu and Dr. V.

    Mocanu (University of Bucharest) for providing the

    facilities at the Geological Institute of the University of

    Bucharest. We thank Dr. M. Melinte (Geological

    Institute of Romania), Dr. D. Badescu, Dr. L. Matenco,

    and Dr. D. Ciulavu (University of Bucharest), who

    contributed much with their discussions about the

    Romanian geology. Dr. Mihail Ianas, President of the

    National Agency for Mineral Resources issued the

    general permit for the fieldwork. Col.eng. Neculai

    Cioanca, Deputy chief of the Military Topographic

    Department, provided permission to obtain maps in thescale 1:100,000, revised and re-published in 1996. The

    governmental forestry offices at Casin, Tulnici, Nereju

    and Gura Teghii as well as many other institutions and

    individuals provided logistical support for suitable

    sites. The Romanian exploration company PROSPEC-

    TIUNI, Bucharest, carried out the drilling and shooting

    operations. Dr. V. Varodin and Eng. M. Milea,

    Technical Directors, co-ordinated the whole operation,

    from permitting through property access formalities to

    the drilling and shooting works. Data were collected

    using the seismic equipment of the GeoForschungs-

    Zentrum Potsdam (120 units) as well as of the

    Department of Geology, Leicester University, UK

    (12 units) and of the NERC Geophysical Equipment

    Pool, UK (8 units). Dr. J. Bribach (GFZ Potsdam)

    trained the German participants before the experiment

    proper. Some figures were created using the Generic

    Mapping Tools (GMT) software of Wessel and Smith

    (1995). The project was funded through the Deutsche

    Forschungsgemeinschaft (German Research Society)

    by providing the funding for the Sonderforschungsber-

    eich 461 (CRC 461) at the University of Karlsruhe,

    Germany: Strong Earthquakes a Challenge forGeosciences and Civil Engineering. The NATO

    Science Collaborative Research Linkage Grant no.

    EST.CLG 974792 assisted the project by additional

    funding of travel and living expenses for data

    interpretation, and the NATO Computer Network

    supplement EST.CNS 976375 supported the labora-

    tory of NIEP. We also appreciate helpful comments by

    G. Fan and L. Ratschbacher.

    References

    Airinei, St., 1977. Lithosphere microplates on the Romanian terri-

    tory reflected in regional gravimetric anomalies (in Romanian).

    St. Cerc., Geol., Geofiz., Geogr., Geofiz. 15 (1), 19 30.

    Badescu, D., 1998. Geology of the East Carpathiansan overview.

    CERGOP South Carpathians Monogr. 7 (37), 49 69,

    Warszawa.

    Bevington, P.R., 1969. Data Reduction and Error Analysis for the

    Physical Sciences. McGraw-Hill, New York.

    Bonjer, K.-P., Oncescu, M.C., Rizescu, M., Driad, L., 1998. A note

    on empirical site responses in Bucharest areas. In: Wenzel, F.,

    Lungu, D., Novak, O. (Eds.), Vrancea Earthquakes: Tectonics,

    F. Hauser et al. / Tectonophysics 340 (2001) 233256254

  • 8/14/2019 Vrancea_99_Hauser.pdf

    23/24

    Hazard and Risk Mitigation. Kluwer Academic Publishing, Dor-

    drecht, Netherlands, pp. 149162.

    Cerveny, V., Psencik, I., 1983. 2-D seismic ray tracing program

    package. Charles University, Prague.

    Chalot-Prat, F., Girbacea, R., 2000. Partial delamination of conti-nental mantle lithosphere, uplift-related crust-mantle decou-

    pling, volcanism and basin formation: a new model for the

    Pliocene-Quaternary evolution of the southern East-Carpathians,

    Romania. Tectonophysics 327, 83 107.

    Channell, J.E.T., Kozur, H.W., 1997. How many oceans? Meliata,

    Vardar, and Pindos oceans in Mesozoic Alpine paleogeography.

    Geology 25 (2), 183 186.

    Ciulavu, D., Dinu, C., Szakacs, A., Dordea, D., 2000. Neogene

    kinematics of the Transylvanian basin (Romania). AAPG Bull.

    84 (10), 1589 1615.

    Constantinescu, L., Enescu, D., 1984. A tentative approach to pos-

    sibly explaining the occurrence of Vrancea earthquakes. Rev.

    Roum. Geol., Geophys., Geogr., Geophys. 28.

    Constantinescu, L., Cornea, S., Lazarescu, D., 1973. An approach to

    the seismotectonics of the Romanian Eastern Carpathians. Rev.

    Roum. Geol., Geophys., Geogr., Geophys. 17 (2), 133143.

    Cornea, S., Polonic, G., 1979. Data concerning the seismicity and

    seismotectonics of the Eastern Moesian Platform (in Romanian).

    St. Cerc., Geol., Geofiz., Geogr., Geofiz. 17 (2), 167 176.

    Cornea, I., Radulescu, F., Pompilian, A., Sova, A., 1981. Deep

    seismic soundings in Romania. PAGEOPH 119, 1144 1156.

    Csontos, L., 1995. Tertiary tectonic evolution of the Intra-Carpathi-

    an area: a review. Acta Vulcanol. 7, 113.

    Dercourt, J., Ricou, L.E., Vrielinck, B., 1993. Atlas Tethys, Paleo-

    environmental Maps. Gauthier-Villars, Paris, 307 pp.

    Dumitrescu, I., Sandulescu, M., 1970. Tectonic map of Romania at

    a Scale 1:1000,000. Inst. Geol. Rom., Bucharest.Ellouz, N., Roure, F., Sandulescu, M., Badescu, D., 1994. Balanced

    cross-sections in the Eastern Carpathians (Romania): a tool to

    quantify Neogene dynamics. In: Roure, F., Ellouz, N., Shein,

    V.S., Skvortsov, I. (Eds.), Geodynamic Evolution of Sedimen-

    tary Basins. Proceedings Volume, International Symposium in

    Moscow, pp. 305325.

    Enescu, D., Danchiv, D., Bala, A., 1992. Lithosphere structure in

    Romania II. Thickness of the Earth crust. Depth-dependent

    propagation velocity curves for the P and S waves. Stud. Cercet.

    Geol. Geofiz. Geogr., Geofiz. 30, 319.

    Fan, G., Wallace, T.C., Zhao, D., 1998. Tomographic imaging of

    deep velocity structure beneath the Eastern and Southern Carpa-

    thians, Romania: implications for continental collision. J. Geo-

    phys. Res. 103 (B2), 2705 2723.Fuchs, K., Bonjer, K., Bock, G., Cornea, I., Radu, C., Enescu, D.,

    Jianu, D., Nourescu, A., Merkler, G., Moldoveanu, T., Tudor-

    ache, G., 1979. The Romanian earthquake of March 4, 1977: II.

    Aftershock and migration of seismic activity. Tectonophysics 53

    (3 4), 225 247.

    Girbacea, R., Frisch, W., 1998. Slab in the wrong place: lower

    lithospheric mantle delamination in the last stage of the Eastern

    Carpathians subduction retreat. Geology 26, 611614.

    Hauser, F., Raileanu, V., Prodehl, C., Bala, A., Schulze, A., Denton,

    P., 2000. The Seismic-Refraction Project VRANCEA-99, Open-

    File Report. Geophysical Institute, University of Karlsruhe.

    Kovac, M., Marton, E., Sefara, J., Konecny, V., Lexa, J., 2000.

    Miocene development of the Carpathian chain and the Panno-

    nian Basin: movement trajectory of lithospheric fragments, sub-

    duction and diapiric uprise of asthenospheric mantle. Slovak

    Geol. Mag. 6 (23), 7784.Lazarescu, V., Cornea, I., Radulescu, F., Popescu, M., 1983. Moho

    surface and recent crustal movements in Romania; geodynamic

    connections. Annu. Inst. Geol. Geofiz., 8391.

    Linzer, H.G., 1996. Kinematics of retreating subduction along the

    Carpathian arc, Romania. Geology 24, 167 170.

    Linzer, H.-G., Frisch, W., Zweigel, P., Girbacea, R., Hann, H.-P.,

    Moser, F., 1998. Kinematic evolution of the Romanian Carpa-

    thians. Tectonophysics 297, 133 156.

    Matenco, L., 1997. Tectonic evolution of the outer Romanian Car-

    pathians. PhD thesis, Vrije Universiteit, Amsterdam, 160 pp.

    Matenco, L., Bertotti, G., 2000. Tertiary tectonic evolution of the

    external East Carpathians (Romania). Tectonophysics 316,

    255286.

    Mellors, R.J., Pavlis, G.L., Hamburger, M.W., Al-Shukri, H.J.,

    Lukk, A.A., 1995. Evidence for a high-velocity slab associated

    with the Hindu Kush seismic zone. J. Geophys. Res. 100,

    40674078.

    Morley, C.K., 1996. Models for relative motion of crustal blocks

    within the Carpathian region, based on restorations of the outer

    Carpathian thrust sheets. Tectonics 15, 885904.

    Nemcok, M., Pospisil, L., Lexa, J., Donelick, R.A., 1998. Tertiary

    subduction and slab break-off model of the Carpathian Panno-

    nian region. Tectonophysics 295, 307340.

    Neugebauer, J., Greiner, B., Appel, E., 2001. Kinematics of the

    Alpine-West Carpathian orogen and palaeogeographic implica-

    tions. J. Geol. Soc. Lond. 158, 97110.

    Oncescu, M.C., 1984. A three-dimensional distribution of the seis-mic wave velocities under the Carpathians region. PhD thesis,

    Institute of Atomic Physics, Bucharest.

    Oncescu, M.C., Bonjer, K.-P., Rizescu, M., 1998. Weak and strong

    ground motion of intermediate depth earthquakes from the Vran-

    cea region. In: Wenzel, F., Lungu, D., Novak, O. (Eds.), Vrancea

    Earthquakes: Tectonics, Hazard and Risk Mitigation. Kluwer

    Academic Publishing, Dordrecht, Netherlands, pp. 27 42.

    Polonic, G., 1996. Structure of the crystalline basement in Romania.

    Rev. Roum. Geophys. 40, 5770.

    Polonic, G., 1998. The structure and morphology of the crystalline

    basement in Romania. CERGOP South Carpathians Monogr.

    7 (37), 127131, Warszawa.

    Prodehl, C., Raileanu, V., Hauser, F., Bala, A., Rumpel, H.-M.,

    Schulze, A., 2000. EUROPROBE PanCarDi project: the seis-mic-refraction project VRANCEA-99. EUROPROBE News 13,

    1518.

    Radulescu, F., 1988. Seismic models of the crustal structure in

    Romania. Rev. Roum. Geol. Geophys. Geogr. Ser. Geophys.

    32, 13 17.

    Radulescu, F., Diaconescu, M., 1998. Deep seismic data in Roma-

    nia. CERGOP South Carpathians Monogr. 7 (37), 177 192,

    Warszawa.

    Radulescu, D.P., Sandulescu, M., 1973. The plate-tectonics concept

    and the geological structure of the Carpathians. Tectonophysics

    16, 155161.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256 255

  • 8/14/2019 Vrancea_99_Hauser.pdf

    24/24

    Radulescu, D.P., Cornea, I., Sandulescu, M., Constantinescu, P.,

    Radulescu, F., Pompilian, A., 1976. Structure de la croute ter-

    restre en Roumanie, Essai dinterpretation des etudes sismiques

    profondes. Ann. Inst. Geol. Geofiz., L 50, 5 36.

    Raileanu, V., 1998. Studiul unor parametri fizici ai litosferei in unelezone din Romania. PhD Thesis, Inst. Fizica Atomica, Bucharest,

    232 pp.

    Raileanu, V., Diaconescu, C., 1998. Seismic signature in Romanian

    crust. Tectonophysics 288, 127136.

    Raileanu, V., Talos, D., Vorodin, V., Stiopol, D., 1993. Crustal

    seismic reflection profiling in Romania on the UrziceniMizil

    line. Tectonophysics 223, 401409.

    Raileanu, V., Diaconescu, C., Radulescu, F., 1994. Characteristics of

    Romanian lithosphere from deep seismic reflection profiling.

    Tectonophysics 239, 165185.

    Roman, C., 1970. Seismicity in Romaniaevidence for the sin-

    kung lithosphere. Nature 2, 1176 1211.

    Rosca, V., 1998. The isostatic anomaly maps. CERGOP South

    Carpathians Monogr. 7 (37), 139155, Warszawa.

    Royden, L.H., 1988. Late Cenozoic tectonics of the Pannonian

    basin system. In: Royden, L.H., Horvath, F. (Eds.), The Panno-

    nian Basin, a Study in Basin Evolution. AAPG Memoir, vol. 45,

    pp. 27 48.

    Sandulescu, M., 1984. Geotectonics of Romania. Technical Publish-

    ing House, Bucharest, 336 pp. (in Romanian).

    Sandulescu, M., 1988. Cenozoic tectonic history of the Carpathi-

    ans. In: Royden, L.H., Horvath, F. (Eds.), The Pannonian

    Basin, a Study in Basin Evolution. AAPG Memoir, vol. 45,

    pp. 17 25.

    Seghedi, A., 1998. The Romanian Carpathian foreland. CERGOP

    South Carpathians Monogr. 7 (37), 2148, Warszawa.

    Seghedi, I., Balintoni, I., Szakacs, A., 1998. Interplay of tectonicsand neogene post-collisional magmatism in the intracarpathian

    region. Lithos 45, 483497.

    Stammler, K., 1994. SeismicHandler programmable multichannel

    data handler for interactive and automatic processing of seismo-

    logical data. Comput. Geosci. 19 (2), 135140.

    Stampfli, G.M., Mosar, J., Marquer, D., Marchant, R., Baudin, T.,

    Borel, G., 1998. Subduction and obduction processes in the

    Swiss Alps. Tectonophysics 296, 159 204.

    Stanica, D., Stanica, M., 1998. 2D modelling of the geoelectric

    structure in the area of the deep-focus Vrancea earthquakes.

    CERGOP South Carpathians Monogr. 7 (37), 193 203,

    Warszawa.

    Stefanescu, M., Working Group, 1988. Geological Cross-sections at

    Scale 1:200,000 A9 A14. Inst. Geol. Geofiz., Bucharest.

    Taboada, A., Rivera, L.A., Fuenzalida, A., Cisternas, A., Philip, H.,

    Bijwaard, H., Olaya, J., Rivera, C., 2000. Geodynamics of the

    northern Andes: subductions and intracontinental deformation(Colombia). Tectonics 19 (5), 787 813.

    Visarion, M., 1998. Gravity anomalies on the Romanian territory.

    CERGOP South Carpathians Monogr. 7 (37), 133 138,

    Warszawa.

    Visarion, M., Sandulescu, M., Stanica, D., Veliciu, S., 1988. Con-

    tributions a la connaissance de la structure profonde de la plat-

    forme Moesienne en Roumanie. St. Tehn. Econ., Ser. Geogiz., D

    15, 211222.

    Wenzel, F., 1997. Strong earthquakes: a challenge for geosciences

    and civil engineeringa new collaborative research center in

    Germany. Seismol. Res. Lett. 68, 438443.

    Wenzel, F., Lungu, D., 2000. Earthquake risk mitigation in Roma-

    nia. Proceedings Volume, 2nd EuroConference on Global

    Change and Catastrophe Risk Management, Luxembourg.

    Wenzel, F., Lungu, D., Novak, O. (Eds.), 1998a. Vrancea Earth-

    quakes: Tectonics, Hazard and Risk Mitigation. Selected papers

    of the First International Workshop on Vrancea Earthquakes,

    Bucharest, November 1 4, 1997. Kluwer Academic Publishing,

    Dordrecht, Netherlands, 374 pp.

    Wenzel, F., Achauer, U., Enescu, D., Kissling, E., Russo, R., Mo-

    canu, V. and Mussachio, G., 1998b. The final stage of plate

    detachment; International tomographic experiment in Romania

    aims to a high-resolution snapshot of this process. EOS, 79: 589,

    592594.

    Wenzel, F., Oncescu, M.C., Baur, M., Fiedrich, F., Ionescu, C.,

    1998c. 25 seconds for Bucharest. In: EWC98 Group (Eds.),

    Proceedings Volume of the International IDNDR Conferenceon Early Warning Systems for the Reduction of Natural Disas-

    ters, Potsdam, Sept. 7 11, GFZ Potsdam, Germany.

    Wessel, P., Smith, W.H.F., 1995. Free software helps map and display

    data. EOS Transactions, Am. Geophys. Union 72, 445446.

    Zelt, C.A., Smith, R.B., 1992. Seismic traveltime inversion for 2-D

    crustal velocity structure. Geophys. J. Int. 108, 1634.

    Zugravescu, D., Polonic, G., 1997. Geodynamic compartments and

    present-day stress on the Romanian territory. Rev. Roum. Geo-

    phys. 41, 3 24.

    Zweigel, P., Ratschbacher, L., Frisch, W., 1998. Kinematics of an

    arcuate fold-thrust belt: the southern East Carpathians (Roma-

    nia). Tectonophysics 297, 177 207.

    F. Hauser et al. / Tectonophysics 340 (2001) 233256256