Long memory of mantle lithosphere fabric — European LAB constrained from seismic anisotropy

13
Long memory of mantle lithosphere fabric European LAB constrained from seismic anisotropy Jaroslava Plomerová , Vladislav Babuška Geophysical Institute, Academy of Sciences of the Czech Republic, 141 31 Praha 4, Czech Republic abstract article info Article history: Received 15 May 2009 Accepted 11 January 2010 Available online 25 January 2010 Keywords: Lithosphereasthenosphere boundary Fossil anisotropy Travel-time residuals We present a uniform updated model of the European lithosphereasthenosphere boundary (LAB) recalculated from data collected during our regional studies of seismic anisotropy and other tomographic experiments, and show results of mapping of large-scale domains of mantle lithosphere characterized by uniform fossil fabrics. Exploiting the long memory of the fabric of the deep continental lithosphere, we dene the LAB as a boundary between a fossil anisotropy in the lithospheric mantle and an underlying seismic anisotropy related to present-day ow in the asthenosphere. Analysis of static terms of teleseismic P-wave travel time deviations shows that the LAB topography is more distinct beneath the Phanerozoic part of Europe than beneath its Precambrian part and deepens down to 220 km beneath the two Alpine roots, the South Carpathians and eastward of the Trans-European Suture Zone. Separation of the relative P residuals into static and directional-dependent terms, shear-wave splitting analysis and joint inversion of anisotropic parameters, reveals that the mantle lithosphere consists of domains with consistent olivine fabrics, which can be modelled by peridotite aggregates with plunging foliation or lineation. Changes of the fossil fabric allow us to map the domain boundaries within the mantle lithosphere. We propose processes which could create the observed fossil fabric as a consequence of successive subduction and accretion of micro-continent fragments outboard of continental cratons and a gradual stabilization of the LAB by a mantle ow after a detachment of the lower parts of subducting slabs. We advocate the importance of combining different methods and data in Earth structure studies and applying 3D approaches considering both seismic anisotropy and the general orientation of symmetry axes to construct more realistic models of large-scale lithospheric structure. © 2010 Elsevier B.V. All rights reserved. 1. Introduction The topography of the lithosphereasthenosphere boundary (LAB) and structure of the continental lithosphere record the geodynamic development of the outer parts of the Earth. Though driving mech- anisms of plate tectonics throughout the planet history are disput- able, delineation of the architecture of continental plates can help to answer questions about how and when the plates were assembled, and to what extent they could later be deformed. There are different denitions and techniques used in imaging the LAB (e.g., Martinec and Wolf, 2005; Eaton et al., 2009 for reviews) including a mechanical, compositional, thermal and seismic-velocity boundary layer, or the top of a layer of increased electrical conductivity. Various denitions can result in various estimates of lithosphere thickness and a dis- agreement among the models might reect either real differences in physical fundamentals of various LABs, or inaccuracy and limitations of individual methods. The seismic LAB is dened as a seismic velocity boundary layer between the high-velocity outer portion of the Earth and the low velocity zone, or, as a region where velocity does not increase, in general (Anderson, 1989). Regions possessing thick or thin lithosphere can be mapped by velocity (or velocity perturbation) images resulting from seismic body-wave tomography or surface wave studies. These approaches represent isotropic and integrating views due to long wave lengths in case of surface waves (e.g., Pedersen et al., 2006) or due to volume parametrizations in body wave tomography which limit their resolution. Single station methods, e.g., those based on conversion of shear (S) to longitudinal (P) phases at the LAB (e.g., Kumar et al., 2005) or on static terms of relative travel-time residuals of P waves from teleseismic events (Babuška and Plomerová 1992), allow us to map not only large-scale features of plate thicknesses, but also to trace lateral variations of LAB topography on smaller scales. Based on velocity variations and seismic anisotropy changes, we model the lithosphere thickness and anisotropic structure in tectonic provinces of different ages and settings. Lateral changes of either depth variations or mean values of seismic velocity characteristics can be converted into models of the LAB topography, a parameter important for understanding the geodynamic formation and evolu- tion of the Earth. Deciphering fossil anisotropy within individual Lithos 120 (2010) 131143 Corresponding author. Geophysical Institute, Czech Acad. Sci., Bocni II 1401/1a, 141 31 Praha 4, Czech Republic. Tel.: +420 267 103 391, +420 605 256 291; fax: +420 272 761 549. E-mail address: [email protected] (J. Plomerová). 0024-4937/$ see front matter © 2010 Elsevier B.V. All rights reserved. doi:10.1016/j.lithos.2010.01.008 Contents lists available at ScienceDirect Lithos journal homepage: www.elsevier.com/locate/lithos

Transcript of Long memory of mantle lithosphere fabric — European LAB constrained from seismic anisotropy

Lithos 120 (2010) 131–143

Contents lists available at ScienceDirect

Lithos

j ourna l homepage: www.e lsev ie r.com/ locate / l i thos

Long memory of mantle lithosphere fabric — European LAB constrained fromseismic anisotropy

Jaroslava Plomerová ⁎, Vladislav BabuškaGeophysical Institute, Academy of Sciences of the Czech Republic, 141 31 Praha 4, Czech Republic

⁎ Corresponding author. Geophysical Institute, Czech141 31 Praha 4, Czech Republic. Tel.: +420 267 103 391,272 761 549.

E-mail address: [email protected] (J. Plomerová).

0024-4937/$ – see front matter © 2010 Elsevier B.V. Aldoi:10.1016/j.lithos.2010.01.008

a b s t r a c t

a r t i c l e i n f o

Article history:Received 15 May 2009Accepted 11 January 2010Available online 25 January 2010

Keywords:Lithosphere–asthenosphere boundaryFossil anisotropyTravel-time residuals

We present a uniform updated model of the European lithosphere–asthenosphere boundary (LAB)recalculated from data collected during our regional studies of seismic anisotropy and other tomographicexperiments, and show results of mapping of large-scale domains of mantle lithosphere characterized byuniform fossil fabrics. Exploiting the long memory of the fabric of the deep continental lithosphere, we definethe LAB as a boundary between a fossil anisotropy in the lithospheric mantle and an underlying seismicanisotropy related to present-day flow in the asthenosphere. Analysis of static terms of teleseismic P-wavetravel time deviations shows that the LAB topography is more distinct beneath the Phanerozoic part of Europethan beneath its Precambrian part and deepens down to ∼220 km beneath the two Alpine roots, the SouthCarpathians and eastward of the Trans-European Suture Zone. Separation of the relative P residuals into staticand directional-dependent terms, shear-wave splitting analysis and joint inversion of anisotropic parameters,reveals that themantle lithosphere consists of domains with consistent olivine fabrics, which can bemodelledby peridotite aggregates with plunging foliation or lineation. Changes of the fossil fabric allow us to map thedomain boundaries within the mantle lithosphere. We propose processes which could create the observedfossil fabric as a consequence of successive subduction and accretion ofmicro-continent fragments outboard ofcontinental cratons and a gradual stabilization of the LAB by amantleflowafter a detachment of the lower partsof subducting slabs. We advocate the importance of combining different methods and data in Earth structurestudies and applying 3D approaches considering both seismic anisotropy and the general orientation ofsymmetry axes to construct more realistic models of large-scale lithospheric structure.

Acad. Sci., Bocni II 1401/1a,+420 605 256 291; fax: +420

l rights reserved.

© 2010 Elsevier B.V. All rights reserved.

1. Introduction

The topography of the lithosphere–asthenosphere boundary (LAB)and structure of the continental lithosphere record the geodynamicdevelopment of the outer parts of the Earth. Though driving mech-anisms of plate tectonics throughout the planet history are disput-able, delineation of the architecture of continental plates can help toanswer questions about how and when the plates were assembled,and to what extent they could later be deformed. There are differentdefinitions and techniques used in imaging the LAB (e.g., MartinecandWolf, 2005; Eaton et al., 2009 for reviews) including amechanical,compositional, thermal and seismic-velocity boundary layer, or thetop of a layer of increased electrical conductivity. Various definitionscan result in various estimates of lithosphere thickness and a dis-agreement among the models might reflect either real differences inphysical fundamentals of various LABs, or inaccuracy and limitationsof individual methods.

The seismic LAB is defined as a seismic velocity boundary layerbetween the high-velocity outer portion of the Earth and the lowvelocity zone, or, as a region where velocity does not increase, ingeneral (Anderson, 1989). Regions possessing thick or thin lithospherecan bemapped by velocity (or velocity perturbation) images resultingfrom seismic body-wave tomography or surface wave studies. Theseapproaches represent isotropic and integrating views — due to longwave lengths in case of surface waves (e.g., Pedersen et al., 2006) ordue to volume parametrizations in body wave tomography — whichlimit their resolution. Single station methods, e.g., those based onconversion of shear (S) to longitudinal (P) phases at the LAB (e.g.,Kumar et al., 2005) or on static terms of relative travel-time residualsof P waves from teleseismic events (Babuška and Plomerová 1992),allow us to map not only large-scale features of plate thicknesses, butalso to trace lateral variations of LAB topography on smaller scales.

Based on velocity variations and seismic anisotropy changes, wemodel the lithosphere thickness and anisotropic structure in tectonicprovinces of different ages and settings. Lateral changes of eitherdepth variations or mean values of seismic velocity characteristicscan be converted into models of the LAB topography, a parameterimportant for understanding the geodynamic formation and evolu-tion of the Earth. Deciphering fossil anisotropy within individual

132 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

domains of the mantle lithosphere (see, e.g., Babuška and Plomerová,2006 for a review) or searching for a difference between anisotropy ofthe lithospheric mantle and that of the asthenosphere, is an importantnew tool for mapping the first-order active boundary related to in theEarth's plate-tectonic framework.

The accumulation of digital waveforms during the last decades anda search for directional variations of seismic velocities (e.g., inversionof both Rayleigh and Love waves, shear-wave splitting analysis) havedemonstrated the ubiquitous presence of seismic anisotropy in theEarth upper mantle, with peridotite as the main constituent (see, e.g.,Fouch and Rondenay, 2006 for a recent review of seismic anisot-ropy). The depth and lateral variations of anisotropy, its strength andorientation derived from (1) radial ξ and azimuthal G anisotropyfrom surface waves, (2) time delay δt and polarization φ of splitshear waves, and (3) directional variations of body-wave travel-timedeviations, indicate that the major part of the variable anisotropyis located within the mantle lithosphere (Plomerová et al., 1996;Babuška et al., 1998) and led to a concept of its domain-like structure(e.g., Babuška et al., 2002; Babuška and Plomerová, 2006). Each of thedomains retains its frozen anisotropy generated by lattice-preferredorientation of olivine (LPO), which we interpret as anisotropy relatedto the strain field at the time of domain formation (Babuška andPlomerová, 1989, 2006). The long memory of mantle–lithospherefabrics with their characteristic 3D general orientation in individualdomains, differs from the very large-scale anisotropy in the sub-lithospheric mantle formed by present-day mantle flow.

The fundamental difference in origin and orientation of seismicanisotropy in themantle lithosphere and in the sub-lithosphericmantlehas lead us to developing approaches in which the LAB is modelled as aboundary between the fossil anisotropy in the mantle lithosphere andthe anisotropy in the underlying mantle. On a global scale (Plomerováet al., 2002b), the resolution of the LAB model is comparable to that ofsurface waves and does not allow us to map an often expressive short-wavelength relief of the LAB. The LAB topography shows variations inamplitudeof almost oneorder ofmagnitude larger than variations of thesurface topography.

Since the early 1980's, we have been building step-by-stepregional models of the European LAB, by analyzing static terms ofrelative P-wave residuals, carefully evaluated from data recorded atdifferent arrays of both permanent and temporary seismic stations.The aim of this paper is to present a uniform updated model of theEuropean LAB recalculated from data collected for our regional studies(Babuška et al, 1984; Babuška and Plomerová, 2006; Plomerová et al.,2008, and references therein) and from tomographic experiments ofother authors (Granet et al., 1995, Arlitt, 1999; Ritter et al., 2001;Sandoval et al., 2004; Eken et al., 2008), as well as to map large-scaledomains of mantle lithosphere characterized by uniform fossil fabrics(Babuška and Plomerová, 1992).

2. Method and data

Our models of the LAB are derived from lateral variations of thestatic terms of travel time deviations of P waves calculated relative toa reference level beneath a region. Effects from the source regions(their structure and foci mislocations), or effects of heterogeneitiesalong deep mantle paths, are minimized by normalization. On thereceiver side — relative P residuals in each station contain correctionsfor crustal effects. Data processing includes the grouping of eventsaccording to source regions to minimize onset-time picking errorsand effects due to uneven distribution of foci, testing stability ofthe chosen reference level as well as relative residual at each station,etc. The relative residuals are separated at each station into the staticterm, a directional mean and directional (azimuth-incidence angledependent) terms (Babuška and Plomerová, 1992). We associatelateral variations of the static term, calculated from the steepest raysin all back-azimuth intervals, with variations of the lithosphere thick-

ness. The directional mean at a station, calculated from all rays aftersuppressing uneven foci distribution, characterizes a mean velocitybeneath the station. The directional terms exhibit dependence ofvelocities on azimuth and angle of propagation within the mantlelithosphere relative to the directional mean (station mean velocity)and are associated with anisotropy in the mantle lithosphere. 3Dimaging of the velocity dependencies in P-spheres reflects the fabric ofthe mantle lithosphere in 3D. We model domain fabrics as peridotiteaggregateswith generally oriented axes of hexagonal or orthorhombicsymmetry. Such orientations of plunging axes also result from jointinversion of body-wave anisotropic parameters: the P-travel timedeviations and shear-wave splitting (Šílený and Plomerová, 1996;Vecsey et al., 2007).

Primary association of the residuals (the static terms) with thelithosphere thickness was done through linking the thinnest litho-sphere with the most delayed relative arrivals in the Belgo-DutchPlatform and the thickest lithosphere with the early arrivals in thecentral part of theWesternAlps (Babuška et al., 1987). The static termswere evaluated from the residuals calculated relative to a referencelevel formed by 15 well distributed European stations with stable andhighest quality observations. Thicknesses of 50 km (Panza et al., 1980)and 220 km (Baer, 1980; Mueller, 1982) were assigned to static termsof 0.8 s and −1.0 s calculated from the original European dataset(Babuška et al., 1987), which resulted in the empirical relation withgradient of 9.4 km/0.1 s. This empirical relationship between thethickness of the lithosphere and the static terms of relative residualsdid not require any assumption on velocity–depth distribution exceptthat vlith>vasth expressing the fact that the lithosphere can beconsidered as a high velocity heterogeneity above the asthenosphere.Because the lithosphere represents the first order high-velocity layerin the uppermost mantle, the relation reflects either a lack or anabundance of high-velocity material in regions with a thin or a thicklithosphere, respectively. The high velocity contrast across the LAB(δvP∼0.6 km/s) required by the empirical gradient cannot beaccomplished by using isotropic velocities of materials forming theupper mantle (Babuška and Cara, 1991). However, consideringinclined fast directions in the mantle lithosphere and mostly sub-horizontal fast directions (lineation) below it, it is possible to re-produce the required velocity contrast. In such a configuration,teleseismic P waves steeply crossing the LAB propagate at about7.7 km/s below the LAB and at ∼8.25 km above it, assuming thecomposition and fabric of typical mantle xenoliths (Ben Ismail andMainprice, 1998). Taking into account an additional 0.1 km/s reduc-tion due to possible partial melting in the asthenosphere (Anderson,1989), the resulting velocity contrast across the LAB is δvP∼0.65 km/s,which gives an estimate of the gradient at about 9.6 km/0.1 s. Thisestimated gradient is almost identicalwith the gradient of 9.4 km/0.1 sin the empirical relation. Jousselin and Mainprice (1998) showed forOman asthenospheric samples that the melt could also enhancethe horizontal anisotropy component as it tends to be aligned inpockets in the flow plane. We keep the relationship constant, thoughinclination of the mantle lithosphere fabrics differ in provinces ofdifferent tectonic setting and age (Babuška et al., 1998).

Adopting the lithosphere as a high-velocity, in general subhor-izontal, layer of variable thickness requires comments on potentialinherent trade-offs. In regions without large active subductions, as incentral and northern parts of Europe, such trade off vertical resolution(by reducing the lithosphere to a single layer with uniform anisotropyand velocity) for lateral one is plausible. Due to vertical smearing inteleseismic body-wave tomography, the higher velocities in the uppermantle associated with the lithosphere are mapped down to depthsexceeding up to 50% of lithosphere thickness as deduced from surface-wave upper mantle tomography (Frederiksen et al., 2007; Priestleyand Tilmann, 2009). In our modelling we combine characteristicsof methods that favour vertical resolution (especially the surface-wave tomography) by fixing the end members of observed relative

133J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

residuals with the thinnest and thickest lithosphere in the region,with the high lateral resolution of teleseismic body waves (particu-larly with relative changes in P-wave propagation). By this particularmethodology, we maximise lateral resolution in mapping the LAB instable continental regions.

Standard deviations of the static terms are usually ±0.1 s or betterand therefore, the accuracy of lateral changes of the LAB determi-nation is estimated at ±10 km in most provinces. However, theassociation of the differences between the static terms of relativeresiduals with lithosphere thickness, carries the uncertainty betweenabsolute and relative values. Therefore, the application of appropriateresidual corrections for crustal effects at each station is a crucial task.In our calculations, each residual is corrected for deviations in crustalthickness and velocity relative to the 2-layer crust of the IASP91Earth model. External regional sources of information on sedimentthicknesses and velocities are exploited. To correct for the crust andsediments in case of newly incorporated P arrival times from the Eifelexperiment (Ritter et al., 2001), we consider the values according toBarth (2002), Wawerzinek (2008), Budweg et al. (2006), Dézes et al.(2004), Scheck-Wenderoth and Lamarche (2005). Delays of about0.07 s and 0.2 s per 1 km due to consolidated Mesozoic andunconsolidated Cenozoic sediments (Wawerzinek, 2008) must beconsidered. Neglecting such corrections would result in unrealisticLAB shallowing beneath regions with thick sediments.

At some permanent stations we have evaluated P residuals duringfour decades of their recordings. This has fortunately allowed us torecognize, that there are stations, where residuals changed over time.Such temporal changes are usually correlated with changes in in-strumentation. Some of new digital recording systems suffer withuncompensated internal delays. Such instrumental errors mask reallateral variations of static terms of relative residuals and could remainundetected in case of stations providing data only during relativelyshort-time experiments.

We have calculated the revised map of the LAB beneath centraland northern of Europe from data (Babuška and Plomerová, 2006;Plomerová et al., 2008, and references therein) collected for bothpermanent stations and temporary arrays (Fig. 1). Different datasetsthus cover different periods (Table 1) and have their own referencelevels. We link individual datasets through static terms of stationslocated in overlapping areas, which calibrates neighbouring datasetsto the same reference level. The levelled terms are used in theresidual-depth relation. Our current model of the LAB covers most ofthe European territory between 0–30°E and 41–70°N, except forthe East European Craton. Recently processed data of the SwedishNational Seismological Network (Eken et al., 2009) and data frompassive seismic experiments in the Bohemian Massif (Babuška et al.,2005, Plomerová et al., 2007) and the Rhenish Massif (Ritter et al.,2001) refined the LAB below the Swedish part of Fennoscandia, theEger Rift and the Rhenish Massif.

In several parts of Europe, the density of stations is sufficient tomap undulating LAB topography or a step-like changes in thelithospheric thickness over short distances, especially thanks to fieldmeasurements within projects like SVEKALAPKO (Hjelt et al., 2006)and TOR (Gregersen et al., 2002) in northern EUROPE, or, BOHEMA I–III (2001–2006). The recently dismounted temporary array of thePASSEQ experiment (2006–2008; Wilde-Piórko et al., 2008) will pro-vide data for studying the structure of mantle lithosphere in provinceswithout sufficient station coverage, e.g., around the central part ofthe Trans-European Suture Zone and in part of the East EuropeanCraton. Similarly, data from the most recent LAPNET (2007–2009)array will provide data for mapping the LAB beneath the northern-most Europe in greater detail (Kozlovskaya et al., 2006). Some com-plementary information on refining relief of the LAB in central Europeis expected from temporary arrays ALPASS (2005–2006, Brückl et al.,2007) and CBP (2005–2008) in the Pannonian Basin (Hausmann et al.,2008) in the near future.

3. European LAB

The LAB beneath Europe (Fig. 2) is distinctly undulating. Thelithosphere thins (∼ up to 60 km) around the NW limit of the Trans-European Suture Zone — beneath the south-western edge of Baltica,the Danish and North-German Basins and the Belgo-Dutch Platform.Incorporating data from the Eifel experiment (Ritter et al., 2001)resulted in refining the depth to the LAB beneath the Rhenish Massif.Previously recognised lithosphere thinning beneath most of themassif (Babuška and Plomerová, 2006) was based on few stationswith less well resolved crustal structure, now concentrates into thecentral part, above the small Eifel upper mantle plume (Mathar et al.,2006) in the revised model. The LAB shallows beneath the easternrim of the Rhenish Massif and the lithosphere thinning continuessouthward to the Rhine Graben and northward to the North GermanBasins. Though most of available information on sediments was col-lected and appropriate corrections were applied, the role of appro-priate corrections for sediments in the LAB relief remains open(J. Ritter, personal communication 2009). A significant feature ofthe LAB topography is a broad lithospheric thinning beneath thePannonian Basin. The LAB also shallows beneath the southern partof the French Massif Central, particularly beneath volcanic centres(Babuška et al., 2002), the Po Plain and towards the Tyrrhenian Sea.Elongated ridges of relatively shallow (thinned) lithosphere tend toappear along the Rhine Graben, Leine Gragen, Eger Rift and partlyLimagne Graben. The distinct local thinning of the lithosphere be-neath the Quaternary Depression in central Bulgaria (Babuška et al.,1986) may not be reliable due to poorly defined crustal structure,especially the thickness and velocity of sediments, in this region.

The LAB deepens to ∼ 220 km marking the two roots of theWestern and Eastern Alps (Babuška et al., 1990), beneath centralFennoscandia (Plomerová et al., 2008) and towards the East EuropeanCraton (Babuška et al., 1987; Babuška and Plomerová, 2006 andreferences therein). The South Carpathian arc is also marked by alithosphere thickening. The LAB depth changes dramatically at theTrans-European Suture Zone, the most prominent tectonic suture inEurope. On average, the LAB is shallower beneath the Phanerozoicpart of Europe than beneath its Precambrian part, east of the Trans-European Suture Zone. Lateral variations of lithosphere thickness aremore distinct and occur on shorter distances in Phanerozoic Europethan beneath the Precambrian provinces. Though the PrecambrianLAB topography is smoother on average, the lithospheric thinning atthe SW margin of Fennoscandia is as significant as those beneath thebasins of Phanerozoic Europe. Detailed modelling of the LAB acrossthe Trans-European Suture Zone bordering Fennoscandia reveals astep-like structure at the boundary marking bottoms of individualmantle lithosphere blocks with differently oriented anisotropy(Plomerová et al., 2002a, Babuška and Plomerová, 2004). Additionalconstraints on depth of the LAB beneath the Proterozoic Fennoscan-dia, obtained after incorporating travel times measured on theSwedish National Seismic Network (Eken et al., 2009) did not revealany significant small-scale variation of the LAB topography. TheFennoscandian lithosphere thins from the thick core of the continen-tal province towards its well resolved N–SW edge.

4. Fossil fabrics of the mantle lithosphere domains

Joint inversion and interpretation of directional variations ofP-wave velocities along with shear-wave splitting parametersresults in models of the lithospheric mantle fabric (e.g., Plomerováet al., 2001, Babuška and Plomerová, 2006; Kozlovskaya et al., 2007;Vecsey et al., 2007; Babuška et al., 2008). The lithospheric fabricis commonly consistent over large provinces extending severalhundreds of kilometres. To satisfy back-azimuth variations of thetwo sets of independent anisotropic parameters of body-waves,the mantle lithosphere fabric is modelled by anisotropic peridotite

Fig. 1. Seismic stations used in this study in a simplified tectonic sketch map of Europe along with location of four profiles (dashed lines) showed in Fig. 4. BM — Bohemian Massif,EEC — East European Craton, MC — French Massif Central, MoP — Moesian Platform, PB Pannonian Basin, NA — Northern Apennines, Po — Po Plain, PZ — Protogine Zone, RM —

Rhenish Massif, TESZ — Trans-European Suture Zone: TS — Thor Suture, STZ — Sörgenfrei–Tornquist Zone, TTZ — Teysseire–Tornquist Zone, TIB — Trans-Scandinavian Igneous Belt,CS, NS, NS — Central, Northern, and Southern Svecofennian.

134 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

aggregates with dipping symmetry axes (Fig. 3). The anisotropypattern commonly changes at prominent boundaries of adjacentlarge-scale tectonic units and reflects the different fossil fabrics ofindividual domains of European lithospheric mantle formed by acollage of independent micro-plates (Babuška and Plomerová,2006). The Moldanubian pattern, with high velocities dipping tothe south, extends towards the Alpine Foredeep and follows theAlpine chain (Babuška and Plomerová, 1992). The Saxothuringianpattern with NW–N dipping high velocities begins at the Saxothur-ingian/Moldanubian contact zone (Fig. 3) and extends further to theN–NW, including Sudetes in the northern part of the BohemianMassif. A detailed study of the western part of the Bohemian Massifdelimited a relatively small domain of the mantle lithospherebeneath the Teplá-Brrandian unit whose fabric wasmodelled by thehigh-velocity lineation dipping to the E (Plomerová et al., 2007;Babuška et al., 2008). Similarly to the Bohemian Massif, the litho-sphere of the French Massif Central consists of three domains, withfabrics characterized in the western part (Limousin) by lineations

dipping to the W and in two eastern domains by high-velocityfoliations dipping to the E (Babuška et al., 2002).

Distinct changes in the lithospheric fabric are associated withdomains around the Trans-European Suture Zone. The Danish block,situated between the two branches of the Trans-European Suture Zone,most probably represents an allochtonous fragment of the lithosphere, amicro-plate with its own pre-existing fabric formed before the blockdocked between Avalonia and Baltica, both of them with their ownunique fabrics (Babuška and Plomerová, 2004). Similarly, the Precam-brian part of Europe, north-east of the Trans-European Suture Zone,does not exhibit a uniform fabric. Models of different symmetries andaxis orientations fit body-wave anisotropic data around the contact oftheProterozoic andArcheanmantle lithosphere in central Fennoscandia(Plomerová et al., 2006a, Vecsey et al., 2007). The central Svecofennianis characterised by high-velocity foliations dipping to the NE, similar tothe mantle lithosphere fabric next the Sorgenfrei–Tornquist SutureZone. The lithosphere beneath the Trans-Scandinavian Igneous Belt andSouthern Svecofennian appears to consist of an assemblage of several

Table 1Data source.

Region Period Source Reference

Central Europe 1973–1979 ISC Bulletins Babuška et al. (1987)W. Alps 1975–1979 Swiss National Network Babuška and Plomerová (1988)GRF 1979–1984 GRF network Faber et al. (1986), Plomerová and Babuška (1988),

Babuška and Plomerová (1992)Balkan (Bulgaria) 1981–1984 National bulletins Babuška et al. (1986)Pannonian Basin-Carpathians, Mo.Pl 1978–1980 Romanian regional network Pajdušák et al. (1989), Praus et al. (1990)

1976–1979 ISC BulletinsItaly 1972–1985 Regional bulletins Babuška and Plomerová (1990, 1993)E. Alps 1979–1985 Nat. Seism. Net of Austria Aric et al. (1989), Babuška, et al. (1990)East Alpine network, ISC Babuška and Plomerová (1993)French MC, Europe 1985–1986 ISC Babuška and Plomerová (1992)

1991–1992 MC experiment Babuška et al., 2002Saxoth./Mold.Contact 1992 Exp'92 Plomerová et al. (1998)In Central Europe

Bohemian Massif 2001–2003 BOHEMA experiment Plomerová et al. (2007)Rhennish Massif 1997–1998 Eifel exp. Ritter et al. (2001)NW Europe (TESZ) 1997–1997 TOR experiment Plomerová et al. (2002a)Fennoscandia 1979–1982 ISC Babuška et al. (1988), Plomerová et al. (2008)South-central Sweden 1991 Varmland'91 Plomerová et al. (2001, 2008)Finland 1997–1999 Finnish National Network Plomerová et al. (2008)Central Fennoscandia 1998–1999 SVEKALAPKO Plomerová et al. (2008)

FinlandCentral Fennoscandia Sweden 2002–2006 Swedish Nat. Seism. Net. Eken et al. (2009)

ISC Bulletin data and locations to link several regional datasets with the European reference level.

135J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

small blocks (Plomerová et al., 2001; Korja et al., 2006) which com-plicate the modelling. The high velocity lineation dips to the SW in thenorthern Svecofennian (Eken et al., 2009).

In southern Europe, we interpret sharply bounded lateral varia-tions of anisotropic parameters along the Northern Apennines as dueto a fossil fabric in the mantle lithosphere of the Adriatic micro-plate(Plomerová et al., 2006b), and document that anisotropy in the sub-lithospheric mantle, generated by the present-day flow in the as-thenosphere around the retreating subducting plate, is not the onlysource of the observed anisotropy in the mantle.

Fig. 4 schematically shows fossil mantle lithosphere fabrics, aswell as the topography of the LAB, in cross-sections along four profiles(see Fig. 1 for their location) sampling different tectonic provinces ofthe European plate. The cross-sections integrate structures in bands of100 km broad. The longest SW–NE cross-section (AA′) runs throughthe Variscan orogenic belt and massifs in Phanerozoic Europe fromthe Pyrenees, through the eastern part of the French Massif Central,the Rhenish Massif, the Belgo-Dutch and North German Platforms.Further to the NE, it crosses the two north-western branches ofthe Trans-European Suture Zone — the Thor Suture and Sorgenfrei–Tornquist Suture Zone, and samples the mantle lithosphere inPrecambrian Fennoscandia, touching its Archean core at the very NEend of the profile. Changes in orientation of the mantle lithospherefabric correlate with prominent tectonic boundaries mapped on thesurface and track their continuation through the whole lithosphere(Babuška and Plomerová, 2006). Boundaries of anisotropic domainsof the mantle lithosphere are commonly linked with changes in lith-osphere thickness. Step-like lithosphere thickening across the Trans-European Suture Zone (Plomerová et al., 2002a) supports the ideaof different origin of the domains retaining their fossil anisotropicstructure since their origin (e.g., Babuška and Plomerová, 2004).

Lithosphere thinning beneath the Eifel region and the PannonianBasin, and a thickening beneath the South Carpathians and theMoesian Platform dominate along the NW–SE profile (BB′). Theshallowing of the LAB beneath the Eger Rift in the Bohemain Massif ismuch less developed in comparison with those beneath the RhenishMassif or FrenchMassif Central (cf. AA′ and CC′). The profile intersectsseveral domains with different fabrics as well. The two separatedroots of theWestern and Eastern Alps (Babuška et al., 1990) dominatethe WSW–ENE profile (CC′). The distinct LAB shallowing beneath the

Pannonian Basin is disturbed by a local thickening beneath the BukkMts. (Babuška et al., 1987).

An approximately N–S profile (DD′) runs from the Tyrrhenian,through the Alps, cuts the easternmost part of the Eger Rift in thewestern part of the Bohemian Massif, crosses the Teysseire–Tornquistsegment of the Trans-European Suture Zone and follows the axisof the Swedish National network. Anisotropy studies in this part ofFennoscandia (Plomerová et al., 2001; 2002a; Eken et al., 2009) haverevealed sharply bounded Proterozoic lithospheric domains accretedto the Archean core, characterized by different fabric orientations.The coincidence of the sharply bounded domains of the PrecambrianFennoscandian lithosphere exhibiting the relatively flat LAB (seealso profile AA′) with images of the velocity perturbations indepen-dently retrieved in 3D tomography (Eken et al., 2009) supports theidea of fossil fabrics and demonstrates that the mantle lithospheredomains can differ both in average (isotropic) velocities and in seis-mic anisotropy.

5. Discussion

Only regional P-velocity tomography with the use of dense tem-porary networks of stations results in images of P-velocity perturba-tions with sufficient lateral resolution to withstand a comparisonwithsingle-method based analyses, e.g., P-wave travel times (Babuška andPlomerová, 1992) or, shear-to-longitudinal (Sp) converted phases(Kind and Vinnik, 1988; Kumar et al., 2005). This was demonstratedby distinguishing two separated roots in the Western and EasternAlps (Babuška et al., 1990), and was also later found in regionalteleseismic tomography by Lippitsch et al. (2003) and Kissling et al.(2006). The step-like thickening around the Trans-European SutureZone (Plomerová et al., 2002a) derived from the static terms of Presiduals was confirmed by images from P and P+S tomography(Arlitt, 1999 and Shomali et al., 2002) and surprisingly also detectedin a model from surface waves (Cotte et al., 2002). The lithospherethickness estimated from seismic tomography beneath the Protero-zoic part of central Fennoscandia (Plomerová et al., 2001; Eken et al.,2007, 2008) as well as the lithospheric root beneath the Proterozoic–Archean contact in south-central Finland (Sandoval, 2004) iscomparable with our model (Fig. 2), though the isotropic teleseismictomography generally overestimates depths of the LAB beneath the

Fig. 2. Model of the LAB depths beneath Europe and selected seismic stations.

136 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

cratons due to the vertical smearing of velocity perturbations alongteleseismic raypaths (Babuška et al., 1990; Mathar et al., 2006).

The structure of the mantle lithosphere beneath the Variscanmassifs deserves a special attention. Their common feature is thatthey consist of several lithospheric domains, each of them possessingits own frozen fabric (e.g., Babuška et al., 2002, 2008; Plomerová et al.,2007). The crust of the three massifs— FrenchMassif Central, RhenishMassif, Bohemian Massif — is dissected by several rifts belonging tothe European Cenozoic Rift System along which Neogene–Quaternaryvolcanic centres are aligned (Ziegler, 1992). Some of the rifts de-veloped above the boundaries of the lithospheric mantle domains,which can be considered as a continuation of crustal terrane bound-aries, locally with an offset of ∼20 km between their crustal andmantle parts (Babuška et al., 2002; Plomerová et al., 2007). Therevised model (Fig. 2) locates the shallow lithosphere at about 60 kmbeneath the central and eastern parts of the Rhenish Massif, whilein its western part the lithosphere thickness slightly exceeds thattypical for the Variscan domains (∼ 120 km). Distinct lithosphereshallowing resulting in a LAB depth ∼70 km is shown beneath thesouth-eastern domain of the Massif Central (Babuška et al., 2002). TheLAB beneath the western and north-eastern domains of the French

Massif Central lies at depths ∼120 and 140 km, respectively. No suchdistinct and localized changes of the LAB depths similar to those in theFrench Massif Central or in the Rhenish Massif were found beneaththe Bohemian Massif, where the LAB seems to be quite flat withmoderate thickening to the south and a slight thinning along the EgerRift (see also Figs. 2 and 4).

An additional remarkable difference between structures of theVariscan massifs concerns the existence of mantle plumes. BeneaththeMassif Central and the RhenishMassif— such narrow low-velocityheterogeneities were detected within the upper 300–400 km of themantle in regional P-velocity tomography (Granet et al., 1995, Ritteret al., 2001). An effect of themantle plume beneath the RhenishMassifwas recognized also in the shear-wave splitting analysis (Walkeret al., 2005). Rayleigh- and Love-wave upper mantle models (Matharet al., 2006) indicate an upper boundary of a low-velocity zone at50 km, which the authors interpret as the top of the small Eifelplume. On the other hand, no steep narrow cylindrical low-velocityheterogeneity, similar to those beneath the Massif Central or RhenishMassif, has been imaged beneath the Bohemain Massif for P-velocitytomography down to 250 km (Plomerová et al., 2007). An asthenos-peric diapir at depths reaching the Moho has been suggested from

Fig. 3. Anisotropic models of several mantle lithosphere domains approximated by hexagonal symmetrywith dipping high-velocity foliation (a,c) or lineation a. Fabrics are stable withindomains and change at domain boundaries (yellow dashed).

Fig. 4. LAB depths and schematic representation of variable fabrics of the mantle lithosphere along four profiles (for their location see Fig. 1). Wemodel domain fabrics by peridotiteaggregates with generally oriented axes of hexagonal or orthorhombic symmetry. The diagrams are polar projections of fabrics approximated by peridotite aggregates.

137J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

138 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

integrated gravity-thermal modelling of the southern Scandes (Pascaland Olesen, 2009) to explain why the topography inferred fromtheir study is not compensated by amantle thermal anomaly at depth.Such structure could contribute to the observed distinct lithospherethinning in our model of the LAB (Fig. 2).

Either P-to-S or S-to-P phases, converted at the crust and uppermantle discontinuities, became a powerful tool to study the Earthstructure. Geissler et al. (2009) analyzed Sp receiver functions (shear-to-longitudinal converted phases) at selected European stations andwith the use of an empirical P velocity value of 8.94 km/s transformedthe differential time between the direct shear waves (S) and theirprecursors (Sp) converted at the LAB into the interface depth. Fig. 5shows the singular estimates of the LAB depths from the Sp phasesplotted over our LAB topography derived from the static terms ofrelative P residuals (Fig. 2). In general, the LAB depths from the twoindependent datasets match well. Because the stacked radial Spreceiver functions do not account for inclination of the boundary andback-azimuth distribution of rays, and often contain several dominantpeaks of similar intensity, we re-interpreted some of the picks atseven stations located mostly at sites above the steep LAB (labelledstations in Fig. 5), carefully keeping coherence with peaks related tothe Moho depth. The ambiguity in determining proper differentialtime between several peaks tends interpreter to choose values leadingto smoother LAB topography (Fig. 5). The S-receiver function methodis sensitive to the sharpness of the LAB, which may vary from diffuse(>50 km thick) to a relatively very sharp (<20 km thick) transition(Eaton et al., 2009). Moreover, the presence of any regional sharpdiscontinuity, as well as ignoring the LAB inclinations and velocityanisotropy, can create multiple peaks in times of expected LAB con-versions. Weaker conversions at the LAB can be masked by suchadditional peaks in the stacked radial receiver functions. An in-dependent knowledge of velocity, by the time difference tS− tSp

Fig. 5. Comparison of the LAB depths derived from the P-residual study (Fig. 2) and fromreinterpreted the Sp receiver functions for the LAB depth estimates, are labelled.

transformed into the LAB depth, is needed. However, commonly usedempirical values, based on a reference global Earth model (e.g., IASP91, Kennett and Engdahl, 1991), can also affect accuracy of LAB depthestimates.

Intensive studies of both radial and transversal components ofreceiver functions, especially their azimuthal dependence, allow us toinfer dipping interfaces and/or anisotropy (e.g., Bostock, 1998;Savage, 1998; Levin and Park, 1998). For instance Bostock (1998)derived a multi-layer structure of the upper mantle beneath the SlaveProvince (Canada) from Ps (longitudinal-to-shear wave) convertedphases. Anisotropy is confined in their synthetic model to only three5–10 km thin layers within the isotropic upper mantle. On the otherhand, ourmodel of the uppermantle associates the seismic anisotropywith the large-scale fabric of the mantle material, namely due tolattice preferred orientation of olivine.

Significant lithospheric thickening eastward of the Trans-EuropeanSuture Zone appears also in LAB models based on heat-flow mea-surements (Majorowicz, 2004). Babuška and Plomerová (1993) foundgood correlation between the ‘seismic’ and ‘heat-flow’ LAB beneaththe Bohemian Massif and the Pannonian Basin, in spite of the factthat unlike the inferred seismic anisotropy, estimates of the LAB depthcan suffer, especially beneath basins, from insufficient knowledge ofvelocities and thicknesses of sediments. But the distinct lithosphericthinning beneath the North German Basin (Fig. 2) is also modelledfrom heat-flow data (Norden et al., 2008) at approximately thesame depth (∼75 km) and also fits results from Sp receiver functions(Geissler et al., 2009).

Magnetotelluric estimates of the LAB depth have become a widelyused method in several regions (e.g., Korja et al., 2002; Jones et al.,2003; Eaton et al., 2009). We found a quite good coherence betweenthe magnetotelluric (e_LAB) and seismic LAB (s_LAB) in severalEuropean regions (Praus et al., 1990; Babuška and Plomerová 1993;

the Sp receiver functions (diamonds) after Geissler et al. (2009). Stations, where we

139J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

Plomerová et al., 2006a; Vecsey et al., 2007), especially consideringthe lower accuracy and resolution of the e_LAB, as well as the largervolume the estimate represents. For a detailed comparison of thes_LAB and e_LAB we refer to Jones et al. (2010-this issue).

Our idea of the LAB definition as a boundary between the frozenseismic anisotropy in the lithosphere and the anisotropy related tothe present-day flow in the sub-lithospheric mantle (Babuška andPlomerová, 1992; Plomerová et al., 2002b; see also Gung et al., 2003)is based on self-consistent 3D anisotropic models of the mantlelithosphere with plunging symmetry axes derived from joint inver-sions/interpretations of anisotropic parameters of body waves(Babuška et al., 1993; Plomerová et al., 1996) and on age-dependenttypes of surface-wave anisotropy-depth distribution (Babuška et al.,1998). Geographical variations of depth dependences of radial(polarization) anisotropy document changes in orientation of anisot-ropy at shallow depth (down to ∼ 250 km— beneath the Precambrianshields). Provinces with radial shear velocities larger than tangentialshear velocities (vSV≥vSH) represent regions with steep fabric(Babuška et al., 1998). Gung et al. (2003) also locate vSV>vSH belowthe Precambrian provinces at depths down to ∼200 km in models oftransverse isotropy from inversion of surface and body waveforms.Below the 300-km depth the authors found that the shields are nolonger prominent in vSV. Therefore, this depth can be accepted as themaximum depth of the LAB, though isotropic body-wave tomogra-phies often trace the high velocities associated with the thicklithosphere even deeper (Fishwick et al., 2005). Gung et al. (2003)explain the incompatibility among different tomographic models inthe depth range 250–400 km by the different sensitivity of differenttypes of data to anisotropy and not by a lack of resolution. In principle,neglecting seismic anisotropy in the isotropic velocity images tends tomap high-velocity heterogeneities into larger depths (Babuška et al.,1990, Sobolev et al., 1997; Plomerová et al., 2002a, Vecsey et al., 2007).

Due to the tendency toward vertical smearing of large anomaliesin teleseismic traveltime tomography, Frederiksen et al. (2007) admitthat their lithosphere thickness of at least 400 km beneath the SuperiorProvince may be exaggerated by 50% or even more. The coherencebetween boundaries of the high and low-velocity heterogeneities intheir models and provinces of consistent azimuthal anisotropymappedfrom split SKS waves provides a strong indication that the observedboundaries of lithospheric domains represent changes in both isotropicvelocity and anisotropic fabric, though the authors did not constrainplunging anisotropic symmetry axes from their splitting results. Wefound similar coincidence between boundaries of velocity heterogene-ities and changes of generally oriented fabrics of the Fennoscandianlithospheric mantle domains retrieved from joint inversion of body-wave anisotropic parameters (Eken et al., 2009). Sometimes a weakazimuthal anisotropy of Rayleigh waves (Bruneton et al., 2004) mightreflect a complex and variable structure of the lithosphere (e.g., Vecseyet al., 2007), which otherwise remains invisible due to low lateralresolution of surface waves (Eaton et al., 2009).

Different techniques applied to various geophysical observables(magnetotellurics, mantle creep modelling, xenoliths thermobaro-metry, seismic anisotropy, receiver functions) converge to values∼200–230 km for depths of the LAB beneath regions with the thickestlithosphere. Only isotropic velocity tomography and some thermalmodels tend to locate the LAB deeper beneath cratons (e.g., Artemievaand Mooney, 2001; Fishwick et al., 2005; see also Martinec and Wolf,2005, and Eaton et al., 2009, for reviews).

We interpret our anisotropic models, which consist of sharplybounded domains of the continentalmantle lithospherewith plungingsymmetry axes oriented generally in 3D, as fragments of micro-platesretaining fossil anisotropy formed during their origin (Babuška andPlomerová, 1989; 2006). The 3D approach of our modelling of generalorientation of symmetry axes of anisotropy allows us to detect abruptchanges of stable fabric at provinces separated by prominent tec-tonic sutures even in those cases, when azimuthal anisotropy from

shear-wave splitting evaluated in 2D exhibits uniform average fast Spolarization azimuths over the regions (e.g., Plomerová et al., 1998;Babuška et al., 2008). The stability of domain fabrics supports theirfrozen origin as opposed to fabrics being produced by a progressivepresent-day deformation, e.g., in the asthenosphere, or those pro-duced during a last tectonic event in the past. Moreover, Vauchez et al.(2005) document, using ultramafic xenoliths, that it is difficult toreorient mantle lithosphere fabric once it has been created.

To explain the observed dipping anisotropic fabrics in the con-tinental mantle lithosphere, Babuška and Plomerová (1989) proposeda scenario of successive subduction cycles of oceanic lithosphere andaccretions of micro-plates with pre-existing fabrics to a continentalcore (craton) (Fig. 6). Anisotropy of originally thin, but later thickenedoceanic lithosphere is characterized by sub-horizontal high-velocities(a-axis) oriented in direction of its movement away from a mid-oceanic ridge. Orthorhombic or hexagonal models with vertical ‘slow’

symmetry axis b approximate the oceanicmantle lithosphere. Stackingof subducting oceanic lithospheres at a craton margin may result ina block of lithosphere with a dipping fabric. The anisotropy of thisaccretion can be modelled by a hexagonal olivine (peridotite) ag-gregate with plunging symmetry axis b (Babuška et al., 1993). Iforientations and dips of fabrics of individual accreted blocks differ, weare able to decipher their boundaries and map a mosaic of continentalfragments forming large continental plates.

The model of a system of dipping palaeo-subduction zones issupported by other seismic or geological evidence for lateral andvertical growth of continents (e.g., Cooper et al., 2006; Brown, 2008).Cook (1986) published a model of continental “shingling” based onseismic reflection data indicating dipping structures in the deep crustof accreted terranes. Musacchio et al. (2004) found relics of oceaniclithosphere accreted at the base of the crust during the final stagesof lithospheric assembly in the western Superior Province at about2.7 Ga. Results of the CD-ROM Experiment, integrating seismicreflection, refraction, teleseismic and geologic data, suggested thatthemantle lithosphere in the southern RockyMountains preserves oldsubduction structures (Levander et al., 2005). Inclined reflectorswhichpenetrate into the mantle (Snyder, 2002) may represent mega-thrustzones at a contact of two subsequent subductions. These seismologicalmodels for generation of the subcontinental lithospheric root byrepeated stacking of subducted oceanic slabs are supported by somepetrological observations. Helmstaedt and Gurney (1995) and Hartet al. (1997) suggest that the Archean continental lithosphere consistsof alternating layers of basalt and peridotite derived from subductedand obducted Archean oceanic lithosphere. On the other hand, Griffinand O'Reilly (2007) argue that cratonic peridotites of exposed massifsand xenoliths differ markedly from oceanic and ophiolitic peridotitesin their Fe–Cr–Al relationships and abundances of trace elements(Li and B) diagnostic of subduction, and conclude that there is nounambiguous evidence that mantle lithosphere peridotites andeclogites represent subducted ocean-floor material. Mantle xenolithscarry fundamental information on evolution of the continental lith-osphere (Sleep, 2005), from which the author concludes that cratonallithosphere about 200 km thick was formed in the Archean byprocesses analogous to modern plate tectonics and has been stablebeneath large cratons since that time (Condie and Benn, 2006). Sim-ilarly, Bostock (1998) suggests the proto-Slave craton was assembledthrough processes of shallow subduction.

The European lithsophere includes terranes from Archean toPhanerozoic ages with different thickness of the lithospheric mantle.Griffin et al. (2009) point out that interpreting seismic velocityvariation retrieved in mantle tomography only in terms of temper-ature differences is inadequate. The authors stress and document theimportance of considering links between thickness, composition andgeotherm. On the other hand, tomographic images of the anisotropicEarth are affected by ignoring anisotropy in their isotropic images.This effect exceeds velocity variations due to compositional changes

Fig. 6. Model of four stages of a possible growth of continental mantle lithosphere resulting in inclined fossil fabric (modified from Babuška and Plomerová, 1989). (I) Oceaniclithosphere (B) subducting beneath an old continental craton (A) retains its preferred olivine orientation (LPO) in its mantle part; (II) the ongoing subduction process ceases and anew one starts (B) outboard of the enlarged continent; (III) the continent–continent plate fragment collision, each fragment with its ready-made fabric, is followed by a newsubduction of an oceanic plate (D) creating a new pattern of inclined anisotropic structure outboard of the accreted continental fragments (B,C); (IV) lower parts of the successivelysubducting slabs detach and the LAB is being gradually stabilized by a mantle flow. The LPO of olivine in peridotite aggregates is approximated by hexagonal or orthorhombicsymmetry with inclined lineation a, or foliation (a,c).

140 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

accounting for ∼20% at depth of 100–175 km beneath the continents.Moreover, sites of kimberlites, which tend to concentrate at marginsof high velocity volumes, rather than within them (Griffin et al.,2009), might be indicative of ancient and possibly partly healedsutures between blocks of mantle lithosphere (e.g., Proterozoic/Archean in Fennoscandia, Vecsey et al., 2007). Babuška et al. (1998)found in their global study of the age dependence of large-scale fabricof the upper mantle that an extent of structures bearing the old(Archean) signature is broader than if it were only beneath the

Archean/Proterozoic crust. This finding is in agreement with infer-ences of O'Reilly et al. (2001) and Griffin et al. (2009) pointing out thatmost of lithospheric roots beneath Pre-Cambrian crust were gener-ated in Archean time and recommending to allow for a higher rate ofcrustal growth in the early Archean, which has significant impact oncurrently accepted mechanisms of crustal growth, its recycling andlithospheric evolution.

Successive subductions of oceanic lithosphere is a mechanismwhich can work in modern-style plate tectonics as we know it now,

141J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

being considered as widespread since 2.7 Ga, and which can be, fromsome plate-tectonic indicators, tracked back locally by 3.0 Ga, or evenearlier (Condie and Kröner, 2008 and references therein). Ouranisotropy studies show that even the mantle lithosphere of theFennoscandian Archean craton is composed of several domainsexhibiting its own frozen fabrics (e.g., Plomerová et al., 2006a; Vecseyet al., 2007). Though modern plate tectonics is the most distincttectonic style acting up to now, we have to consider a mechanismcreating oriented structures (fabrics) in the pre-plate-tectonic style.According to Stern (2008), earlier tectonic modes in a hotter moredynamic Earth might be similar in some respects to those of themodern-plate tectonics. Basaltic “rockbergs” on convecting magmaocean in the Hadean Earth are supposed to turn to either proto-platetectonics with platelets and supercratonal, or, to unstable stagnantlithospheric lid models in the Archean (∼4.0 Ga), both evolving intothe modern-style plate tectonics (Ernst 2007). The stage of plateletsor supercratonal lasted during the deep mantle and plume-drivencirculations, when plate motions were controlled by asthenosphericconvection dragging buoyant lithosphere along and down. The earlylithosphere formed in dynamic conditions, far from simple coolingwhich could give rise a sub-horizontal layered structure of the lith-osphere. Differential motions between pairs of Precambrian cratonsindicated in paleomagnetic records suggest supercontinental amalgam-ation and dispersal involving Archean cratons (Evans and Pisarevsky,2008).

6. Conclusions

Modelling seismic anisotropy of the mantle lithosphere in 3Drepresents apowerful tool to studyboth its thickness and fabric. Becauseof the long memory of the fabric of the continental mantle lithosphere,we define the LAB as a boundary between fossil anisotropy in thelithosphere and seismic anisotropy related to present-day flow inthe sub-lithospheric mantle. Based on the analysis of static terms ofteleseismic P-wave travel time deviations, we present a model of theEuropean LAB, in which

• LAB topography is more distinct beneath the Phanerozoic part ofEurope compared with its Precambrian part;

• the LAB shallows up to about 60 km beneath major European basinsand deepens down to ∼220 km beneath the two Alpine roots, theSouth Carpathians and eastward of the Trans-European Suture Zone.

Separation of relative residuals into static terms and directional-dependent terms of P residuals, shear-wave splitting analysis andjoint inversion of these anisotropic parameters reveal that the mantlelithosphere is formed by domains with consistent fabrics, which

• extend up to several hundreds of km;• due to LPO of olivine, can be modelled by peridotite aggregates withplunging foliation or lineation;

• show abrupt changes, allowing the mapping of their boundarieswithin the mantle lithosphere.

We propose processes acting during the modern-style platetectonics, which could create the observed fossil fabric as a con-sequence of

• cycles of successive subductions and accretions of micro-continentfragments outboard of continental cratons

• gradual stabilization of the LAB by a mantle flow after a detachmentof lower parts of subducting slabs.

In general, there are similarities between the LAB models derivedfrom various geophysical parameters and techniques, as well as di-versities, which might reflect differences in resolution and accuracyof individualmethods. On the other hand, different physical parameterscan ‘see’ different LABs and diverse lithosphere structures. Therefore,we advocate the necessity of combining differentmethods and datasets,

and especially 3D approaches that allow us to consider seismic an-isotropy with general orientation (plunging symmetry axes).

Acknowledgements

We thank J. Ritter for providing us with P arrival times collectedduring the Eifel seismic experiment, as well as A. Barth andM. Scheck-Wenderoth for data on the crust and sediments. Special thanks goto W. Geissler who has provided results of the Sp receiver functionsfor the comparison with our LAB depth estimate. The manuscriptbenefits substantially from reviews of S. O'Reilly, D. Mainprice and ananonymous reviewer. The research was partly supported by grants ofthe Grant Agency of the Czech Academy of Sci. No. IAA300120709 andof the Czech Grant Agency No. 205/07/1088. We used GMT software(Wessel and Smith, 1995) to prepare most of the figures.

References

Anderson, D.L., 1989. Theory of the Earth. Blackwell Science Publications. 366 pp.Aric, K., Gutdeutsch, R., Leichter, B., Lenhardt, W., Plomerová, J., Babuška, V., Pajdušák,

P., Nixdorf, U., 1989. Structure of the lithosphere in the Eastern Alps derived fromP-residual analysis. Arbeiten aus Zentralanstalt für Meteorologie und Geodyna-mik Nr. 370, H.73. Wien. 26 pp.

Arlitt, R., 1999. Teleseismic body wave tomography across the Trans-European SutureZone between Sweden and Denmark. PhD. Thesis of ETH Zurich No. 13501, 110 pp.

Artemieva, I.M., Mooney, W.D., 2001. Thermal thickness and evolution of Precambrianlithosphere: a global study. Journal of Geophysical Research 106 (B8), 16387–16414.

Babuška, V., Plomerová, J., 1989. Seismic anisotropy of the subcrustal lithosphere inEurope: another clue to recognition of accreted terranes? In: Hillhouse, J.W. (Ed.),Deep structure and past kinematics of accreted terranes. Geophysical Monograph50, vol. 5. IUGG, Washington DC, pp. 209–217.

Babuška, V., Cara, M., 1991. Seismic anisotropy in the Earth. Kluwer Acad. Publishers,Dordrecht. 217 pp.

Babuška, V., Plomerová, J., 1988. Subcrustal continental lithosphere: a model of itsthickness and anisotropic structure. Physics of the Earth and Planetary Interiors 51,130–132.

Babuška, V., Plomerová, J., 1990. Tomographic studies of the upper mantle beneathItalian region. Terra Nova 2, 569–576.

Babuška, V., Plomerová, J., 1992. The lithosphere in central Europe — seismological andpetrological aspects. Tectonophysics 207, 141–163.

Babuška, V., Plomerová, J., 1993. The lithosphere thickness and velocity anisotropy —

seismological and geothermal aspects. Tectonophysics 225, 78–89.Babuška, V., Plomerová, J., 2004. Sorgenfrei–Tornquist Zone as themantle edge of Baltica

lithosphere: new evidence from three-dimensional seismic anisotropy. Terra Nova16, 243–249.

Babuška, V., Plomerová, J., 2006. European mantle lithosphere assembled from rigidmicroplates with inherited seismic anisotropy. Physics of the Earth and PlanetaryInteriors 158, 264–280. doi:10.1016/j.pepi.2006.01.010.

Babuška, V., Plomerová, J., Šílený, J., 1984. Spatial variations of P residuals and deepstructure of the European lithospere. Geophysical Journal of the Royal astronomicalSociety 79, 363–383.

Babuška, V., Plomerová, J., Spasov, E., 1986. Lithosphere thickness beneath the territoryof Bulgaria — a model derived from teleseismic P residuals. Geologica Balcanica 16,51–54.

Babuška, V., Plomerová, J., Šílený, J., 1987. Structural model of the subcrustal lithospherein central Europe. In: Froidevaux, C., Fuchs, K. (Eds.), The composition, structureand dynamics of the lithosphere– asthenosphere system. AGU Geophysical Series,vol. 16. Washington DC, pp. 239–251.

Babuška, V., Plomerová, J., Pajdušák, P., 1988. Seismologically determined deep lithospherestructure in Fenoscandia. GFF Meeting Proceedings, Geologiska Foreningens i Stock-holm Forhandlingar 110, 380–382.

Babuška, V., Plomerová, J., Granet, M., 1990. The deep lithosphere in the Alps: a modelinferred from P residuals. Tectonophysics 176, 137–165.

Babuška, V., Plomerová, J., Šílený, J., 1993. Models of seismic anisotropy in deepcontinental lithosphere. Physics of the Earth and Planetary Interiors 78, 167–191.

Babuška, V., Montagner, J.-P., Plomerová, J., Girardin, N., 1998. Age-dependent large-scale fabric of the mantle lithosphere as derived from surface-wave velocityanisotropy. Pure and Applied Geophysics 151, 257–280.

Babuška, V., Plomerová, J., Vecsey, L., Granet, M., Achauer, U., 2002. Seismic anisotropyof the French Massif Central and predisposition of Cenozoic rifting and volcanismby Variscan suture hidden in the mantle lithosphere. Tectonics 21, 11–1–20.doi:10.1029/2001TC901035,2002.

Babuška, V., Plomerová, J., Vecsey, L., Jedlička, P., Růžek, B., 2005. Ongoing passiveseismic experiments unravel deep lithosphere structure of the Bohemian Massif.Studia Geophysica and Geodaetica 49, 423–430.

Babuška, V., Plomerová, J., Vecsey, L., 2008. Mantle fabric of western Bohemian Massif(central Europe) constrained by 3D seismic P and S anisotropy. Tectonophysics 462,149–163. doi:10.1016/j.tecto.2008.01.020.

Baer, M., 1980. Relative travel-time residuals for teleseismic events at the new Swissstation network. Annals of Geophysics 36, 119–126.

142 J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

Barth, A., 2002. P-Wellen-Tomographie des oberen Mantels und der Übergangszoneunter Eifel und Zentralmassiv. Master thesis (in german), University of Göttingen.

Ben Ismail, W., Mainprice, D., 1998. An olivine fabric database: an overview of uppermantle fabrics and seismic anisotropy. Tectonophysics 296, 145–157.

Bostock, M.G., 1998. Mantle stratigraphy and evolution of the Slave province. Jornal ofGeophysical Research 103 (B9), 21183–21200.

Brown, M., 2008. Characteristic thermal regimes of plate tectonics and theirmetamorphic imprint throughout Earth history: when did Earth first adopt aplate tectonic mode of behaviour? In: Condie, K., Pease, V. (Eds.), When did PlateTectonics Begin? Geological Society of America, Special Paper, vol. 440, pp. 97–128.

Brückl, E., Behm, M., Chvatal, W., Mitterbauer, U., CELEBRATION 2000, ALP 2002,ALPASS Working Groups, 2007. Erkenntnisse über den Tiefbau der Ostalpen ausneuen seismischen Großexperimenten. Beiträge zur Geologie Oberösterreichs,Arbeitstagung 2007 GBA. ISBN: 978-3-85316-035-0, pp. 143–150.

Bruneton, M., Pedersen, H.A., Farra, V., Arndt, N.T., Vacher, P., SVEKALAPKO SeismicTomography W.G., 2004. Complex lithospheric structure under the central BalticShield from surface wave tomography. Journal of Geophysical Research 109, B10303.doi:10.1029/203JB002947.

Budweg, M., Bock, G., Weber, M., 2006. The Eifel Plume — imaged with convertedseismic waves. Geophysical Journal International 166, 579–589.

Bulletins of the International Seismological Centre (ISC) 1979–1982; 1988–1992.Condie, K.C., Benn, K., 2006. Archean geodynamics: similar to or different from modern

geodynamics? In: Benn, K., Mareschal, J.-C., Condie, K.C. (Eds.), Archean Geody-namics and Environments. AGU Geophysical Monograph, vol. 164. Washington, DC,pp. 47–60. 206.

Condie, K.C., Kröner, A., 2008. When did plate tectonics begin? Evidence from thegeological record. In: Condie, K.C., Peace, V. (Eds.), When did plate tectonics beginon Planet Earth? The Geological Society of Am., Special Paper, vol. 440, pp. 281–284.

Cook, F.A., 1986. Continental evolution by lithosphere shingling. In: Barazangi, M.,Brown, L. (Eds.), Reflection Seismology; The Continental Crust. AGU GeodynamicSeries, vol. 16. Washington D.C, pp. 13–16.

Cooper, C.M., Lenardic, A., Levander, A., 2006. Creation and preservation of cratoniclithosphere: seismic constraints and geodynamic models. In: Benn, K., Mareschal,J.-C., Condie, K.C. (Eds.), Archean Geodynamics and Environments. AGU Geophys-ical Monograph, vol. 164. Washington, DC, pp. 75–88. 206 pp.

Cotte, N., Pedersen, H.A., TOR Working Group, 2002. Sharp contrast in lithosphericstructure across the Sorgenfrei–Tornquist Zone as inferred by Rayleigh wave analysisof TOR1 project data. Tectonophysics 360, 75–88.

Dézes, P., Schmid, S.M., Ziegler, P.A., 2004. Evolution of the European Cenozoic RiftSystem: interaction of the Alpine and Pyrenean orogens with their forelandlithosphere. Tectonophysics 389, 1–33.

Eaton, D.W., Darbyshire, F., Evans, R.L., Grütter, H., Jones, A.G., Yuan, X., 2009. The elusivelithosphere–asthenosphere boundary (LAB) beneath cratons. Lithos 109, 1–22.

Eken, T., Shomali, H., Roberts, R., Bodvarsson, R., 2007. Upper mantle structure of theBaltic Shield below the Swedish National Seismological Networks (SNSN) resolvedby teleseismic tomography. Geophysical Journal International 169, 617–630.

Eken, T., Shomali, Z.H., Roberts, R., 2008. S and P velocity heterogeneities within theupper mantle below the Baltic Shield. Tectonophysics 462, 109–124.

Eken, T., Plomerová, J., Roberts, R., Vecsey, L., Babuška, V., Shomali, H., Bodvarsson, R.,2009. Seismic anisotropy of the mantle lithosphere beneath the Swedish NationalSeismological Network (SNSN). Tectonophysics 480, 241–258.

Ernst, W.G., 2007. Speculations on evolution of the terrestrial lithosphere–astheno-sphere system — plumes and plates. Gondwana Research 11, 38–49. doi:10.1016/j.gr.2006.02.007.

Evans, D.A.D., Pisarevsky, S.A., 2008. Plate tectonics on early Earth? Weighing thepaleomagnetic evidences. In: Condie, K.C., Peace, V. (Eds.), When did plate tectonicsbegin on Planet Earth? The Geological Society of America, Special Paper, vol. 440,pp. 249–263.

Faber, S., Plomerová, J., Babuška, V., 1986. Deep-seated lateral variations beneath theGRF array inferred frommislocation patterns and P residuals. Journal of Geophysics60, 139–148.

Fishwick, S., Kennett, B.L.N., Reading, A.M., 2005. Contrasts in lithospheric structurewithin the Australian Craton — insights from surface wave tomography. Earth andPlanetary Science Letters 231, 163–176.

Fouch, M.J., Rondenay, S., 2006. Seismic anisotropy beneath stable continental interiors.Physics of the Earth and Planetary Interiors 158, 292–320.

Frederiksen, A.W., Miong, S.K., Darbyshire, F.A., Eaton, D.W., Rondenay, S., Sol, S., 2007.Lithospheric variationsacross the Superior Province, Ontario, Canada: evidencefrom omography and sudar-wave splitting. Journal of Geophysical Research 112,B07318. doi:10.1029/2006JB004861.

Geissler W., H., Sodoudi, F., Kind, R., 2009. Thickness of the European lithosphere asseen by S receiver functions, Geophyscal Journal International, in review.

Granet, M., Wilson, M., Achauer, U., 1995. Imaging a mantle plume beneath the MassifCentral (France). Earth and Planetary Science Letters 17, 1109–1112.

Gregersen, S., Voss, P., TORWorking Group, 2002. Summary of project TOR: delineationof stepwise, sharp, deep lithosphere transition across Germany–Denmark–Sweden.Tectonophysics 360, 61–73.

Griffin, W.L., O'Reilly, S.Y., 2007. Cratonic lithospheric mantle: is anything subducted?Episodes 30, 43–53.

Griffin, W.L., O'Reilly, S.Y., Alfonso, J.C., BEGG, G.C, 2009. The composition and evolutionof lithospheric mantle: a re-evaluation and its tectonic implications. Journal ofPetrology 50, 1185–1204. doi:10.1093/petrology/egn033.

Gung, Yu., Panning, M., Romanowicz, B., 2003. Global anisotropy and the thickness ofcontinents. Nature 422, 707–711.

Hart, R.J., Tredoux,M., deWit,M.J., 1997. Refractory trace elements in diamond inclusions:further clues to the origins of the ancient cratons. Geology 25, 1143–1146.

Hausmann, G., Hoyer, S., Zeisz, K., Schurr, B., Brückl, E., CBP Team, ALPASS Team, 2008.Seismicity of the Vienna Basin from the temporary ALPASS and CBP deployments.Geophysical Research Abstracts 10, EGU2008-A-05285, 2008, SRef-ID: 1607–7962/gra/EGU2008-A-05285, EGU General Assembly 2008.

Helmstaedt, H., Gurney, J.J., 1995. Geotectonic controls of primary diamond deposits:implications for area selection. In: Griffin,W.L. (Ed.), Diamond Exploration: Into the21st Century: Journal of Geochemical Exploration, vol. 53, pp. 125–144.

Hjelt, S.-E., Korja, T., Kozlovskaya, E., Yliniemi, J., Lahti, I., BEAR and SVEKALAPKOWorking Groups, 2006. Electrical conductivity and seismic velocity structures ofthe lithosphere beneath the Fennoscandian Shield. In: Gee, D., Stephenson, R.(Eds.), European Lithosphere Dynamics. Geological Society, London, pp. 541–559.

Jones, A.G., Lezeata, P., Ferguson, I.A., Chave, A.D., Evans, R.L., Garcia, X., Spratt, J., 2003.The electrical structure of the Slave craton. Lithos 71, 505–527.

Jones, A.G., Plomerová, J., Korja, T., Sodoudi, F., Spakman, W., 2010. Europe from thebottom up: A statistical examination of the central and northern Europeanlithosphere-asthenosphere boundary from comparing seismological and electro-magnetic observations. Lithos 120, 14–29 (this issue).

Jousselin, D., Mainprice, D., 1998. Melt topology and seismic anisotropy in mantleperidotites of the Oman ophiolite. Earth Planetary Science Letters 167, 553–568.

Kennett, B.L.N., Engdahl, E.R., 1991. Travel times for global earthquake location andphase identification. Geophysical Journal International 105, 429–465.

Kind, R., Vinnik, L.P., 1988. The upper-mantle discontinuities underneath the GRF arrayfrom P-to-S converted phases. Journal of Geophysics 62, 138–147.

Kissling, E., Schmid, S.M., Lippitsch, R., Ansorge, J., Fügenschuh, B., 2006. Lithospherestructure and tectonic evolution of the Alpine arc: new evidence from high-resolution teleseismic tomography. In: Gee, D., Stephenson, R.A. (Eds.), EuropeanLithosphere Dynamics. : Memoirs, vol. 32. Geological Society, London, pp. 129–145.

Korja, T., Hjelt, S.-E., Kaikkonen, P., Kozlovskaya, E., Lahti, I., Pajunpää, K., Pulkkinen, A.,Viljanen, A., BEAR Working Group, 2002. Crustal and upper mantle beneathFennoscandia as imaged by the Baltic Electromagnetic Array Research (BEAR). In:Lahtinen, R., et al. (Ed.), Lithosphere: Second Symposium on the Structure,Composition and Evolution of the Lithosphere in Finland. Inst. of Seismology, Univ.of Helsinki, Report S- 42, pp. 41–48.

Korja, A.K., Lahtinen, R., Nironen, M., 2006. The Svecofennian orogen: a collage ofmicrocontinents and island arcs. In: Gee, D., Stephenson, R. (Eds.), EuropeanLithosphere Dynamics. Geological Soc. London, pp. 561–578. 662 pp.

Kozlovskaya, E., Poutanen, M., POLENET/FI colleagues, 2006. POLENET/FI — a multi-disciplionary geophysical experiment in northern Fennoscandia and Antarcticaduring the International Polar Year 2007–2008. Proceedings of the LITHOSPHERE2006 Symposium, Espoo, Geological Survey of Finland S-46, pp. 75–79.

Kozlovskaya, E., Vecsey, L., Plomerová, J., Raita, T., 2007. Joint inversion of multimethodgeophysical data with use of multi-objective optimization: problem formulationand application to the seismic anisotropy investigation. Geophysical JournalInternational 171, 761–779. doi:10.1111/j.1365-246X.2007.03540.x.

Kumar, P., Kind, R., Hanka, W., Wylegalla, K., Reigber, Ch., Yuan, X., Wölbern, I.,Schwintzer, P., Fleming, K., Dahl-Jensen, T., Larsen, T.B., Schweitzer, J., Priestley, K.,Gudmundsson, O., Wolf, D., 2005. The lithosphere–asthenosphere boundary inthe North West Atlantic region. Earth and Planetary Science Letters 236, 249–257.

Levander, A., Zelt, C., Magnani, M.B., 2005. Crust and upper mantle velocity structure ofthe southern Rocky Mountains from Jemez Lineament to the Chyenne belt. In:Karlstrom, K.E., Keller, G.R. (Eds.), The Rocky Mountain region: an evolvinglithosphere tectonics, geochemistry and geophysics: Geophysical MonographSeries, vol. 154, pp. 293–309.

Levin, V., Park, J., 1998. P-SH conversions in layered media with hexagonal symmetricanisotropy: a cookbook. Pure and Applied Geophysics 151, 669–697.

Lippitsch, R., Kissling, E., Ansorge, J., 2003. Upper mantle structure beneath the Alpineorogen fromhigh-resolution teleseismic tomography. Journal ofGeophysical Research108 (B8), 2376.

Majorowicz, J.A., 2004. Thermal lithosphere across the Trans-European Suture Zone inPoland. Geology Quarterly 48, 1–13.

Martinec, Z., Wolf, D., 2005. Inverting the Fennoscandian relaxation-time spectrum interms of an axisymmetric viscosity distribution with a lithospheric root. Journal ofGeodynamics 39, 143–163.

Mathar, J.P., Ritter, J.R.R., Friederich, W., 2006. Surface waves image the top of the Eifelplume. Geophysical Journal International 154, 377–382.

Mueller, S., 1982. Deep structure and recent dynamics in the Alps. In: Hsü, K. (Ed.),Mountain building processes. Acad. Press London, pp. 85–93.

Musacchio, G., White, D.J., Asudeh, I., Thomson, C.J., 2004. Lithospheric structure andcomposition of the Archean western Superior Province from seismic refraction/wide-angle reflection and gravity modelling. Journal of Geophysical Research 109,B03304. doi:10.1029/2003jb002427, 1-28.

Norden, B., Förster, A., Balling, N., 2008. Heat flow and lithospheric thermal regime inthe Northeast German Basin. Tectonophysics 460, 215–229.

O'Reilly, S.Y., Griffin, W.L., Poudjom Djomani, Y., Morgan, P., 2001. Are lithospheresforever? Tracking changes in subcontinental lithspheric mantle through time. GSAToday 11, 4–9.

Pajdušák, P., Plomerová, J., Babuška, V., 1989. A model of the lithosphere thickness inthe region of the Carpathians. Studia Geophysica et Geodaetica 33, 11–21.

Panza, G., Mueller, S., Calcagnile, G., 1980. The gross features of the lithosphere–asthenosphere system in Europe from seismic surface waves and body waves. Pureand Applied Geophysics 118, 1209–1213.

Pascal, C., Olesen, O., 2009. Are the Norwegian mountains compensated by a mantlethermal anomaly at depth? Tectonophysics 475, 160–168.

Pedersen, H.A., Bruneton, M., Maupin, V., STWG, S.V.E.K.A.L.A.P.K.O., 2006. Lithosphericand sublithospheric anisotropy beneath the Baltic shield from surface wave arrayanalysis. Earth and Planetary Science Letters 244, 590–605.

143J. Plomerová, V. Babuška / Lithos 120 (2010) 131–143

Plomerová, J., Babuška, V., 1988. Lithosphere thickness in the contact zone of theMoldanubicum and Saxothuringicum in Central Europe. Physics of the Earth andPlanetary Interiors 51, 159–165.

Plomerová, J., Šílený, J., Babuška, V., 1996. Joint interpretation of upper mantleanisotropy based on teleseismic P-travel time delays and inversion of shear-wavesplitting parameters. Physics of the Earth and Planetary Interiors 95, 293–309.

Plomerová, J., Babuška, V., Šílený, J., Horálek, J., 1998. Seismic anisotropy and velocityvariations in the mantle beneath the Saxothuringicum–Moldanubicum contact incentral Europe. Pure and Applied Geophysics 151, 365–394.

Plomerová, J., Arvidsson, R., Babuška, V., Granet, M., Kulhánek, O., Poupinet, G., Šílený, J.,2001. An array study of lithospheric structure across the Protogine zone, Varmland,south-central Sweden — signs of a paleocontinental collision. Tectonophysics 332,1–21.

Plomerová, J., Babuška, V., Vecsey, L., Kouba, D., TOR Working Group, 2002a. Seismicanisotropy of the lithosphere around the Trans-European Suture Zone (TESZ) basedon teleseismic body-wave data of the TOR experiment. Tectonophysics 360, 89–114.

Plomerová, J., Kouba, D., Babuška, V., 2002b. Mapping the lithosphere–asthenosphereboundary (LAB) through changes in surface-wave anisotropy. Tectonophysics 358,175–185.

Plomerová, J., Babuška, V., Vecsey, L., Kozlovskaya, E., Raita, T., SSTWG, 2006a.Proterozoic-Archean boundary in the upper mantle of eastern Fennoscandia asseen by seismic anisotropy. Journal of Geodynamics 41, 400–410. doi:10.1016/j.jog.2005.10.008.

Plomerová, J., Margheriti, L., Park, J., Babuška, V., Pondrelli, S., Vecsey, L., Piccinini, D.,Levin, V., Baccheschi, P., Salimbeni, S., 2006b. Seismic Anisotropy beneath theNorthern Apennines (Italy): mantle flow or lithosphere fabric? Earth and PlanetaryScience Letters 247, 157–170.

Plomerová, J., Achauer, U., Babuška, V., Vecsey, L., BOHEMA working group, 2007. Uppermantle beneath the Eger Rift (Central Europe): plume or asthenosphere upwelling?Geophysical Journal International 169, 675–682. doi:10.1111/j.1365-246X.2007.03361.x.

Plomerová, J., Babuška, V., Kozlovskaya, E., Vecsey, L., Hyvonen, L.T., 2008. Seismicanisotropy — a key to resolve fabrics of mantle lithosphere of Fennoscandia.Tectonophysics 462, 125–136. doi:10.1016/j.tecto.2008.03.018.

Praus, O., Pěčová, J., Petr, V., Babuška, V., Plomerová, J., 1990. Magnetotelluric andseismological determination of the lithosphere–asthenosphere transition in CentralEurope. Physics of the Earth and Planetary Interior 60, 212–228.

Priestley, K., Tilmann, F., 2009. Relationship between the upper mantle high velocityseismic lid and the continental lithosphere Submitted to Elsevier Lithos 109,112–124. doi:10.1016/j.lithos.2008.10.021.

Ritter, J.R.R., Jordan, M., Christensen, U.R., Achauer, U., 2001. A mantle plume below theEifel volcanic fields, Germany. Earth and Planetary Science Letters 186, 7–14.

Sandoval, S., Kissling, E., Ansorge, J., 2004. High-resolution body wave tomographybeneath the SVEKALAPKO array: II. Anomalous upper mantle structure beneath thecentral Baltic Shield. Geophysical Journal International 157, 200–214.

Savage, M.K., 1998. Lower crustal anisotropy or dipping boundaries? Effects on receiverfunctions and a case study in New Zealand. Journal of Geophysical Research 103(B7), 15069–15087.

Scheck-Wenderoth, M., Lamarche, J., 2005. Crustal memory and basin evolution in theCentral EuropeanBasin System—new insights from3Dstructuralmodel. Tectonohysics397, 143–165.

Shomali, Z.H., Roberts, R.G., TORWorking Group, 2002. Non-linear body wave teleseismictomography along the TOR array. Geophysical Journal International 148, 562–574.

Šílený, J., Plomerová, J., 1996. Inversion of shear-wave splitting parameters to retrievethree-dimensional orientation of anisotropy in continental lithosphere. Physics ofthe Earth and Planetary Interior 95, 277–292.

Sleep, N.H., 2005. Evolution of the continental lithosphere. Annual Review of Earth andPlanetary Sciences 33, 369–393.

Snyder, D.B., 2002. Lithospheric growth at margins of cratons. Tectonophysics 355, 7–22.Sobolev, S.V., Zeyen, H., Granet, M., Achauer, U., Bauer, Ch., Werling, F., Altherr, R., Fuchs,

K., 1997. Upper mantle temperatures and lithosphere–asthenosphere systembeneath the French Massif Central constrained by seismic, gravity, petrological andthermal observations. Tectonophysics 275, 143–164.

Stern, R.J., 2008. Modern-style plate tectonics began in Neoproterozoic time: analternative interpretation of Earth's tectonic history. In: Condie, K.C., Peace, V.(Eds.), When did plate tectonics begin on Planet Earth? The Geological Society ofAmerica, Special Paper, vol. 440, pp. 265–280.

Vecsey, L., Plomerová, J., Kozlovskaya, E., Babuška, V., 2007. Shear-wave splitting as adiagnostic of varying upper mantle structure beneath south-eastern Fennoscandia.Tectonophysics 438, 57–77. doi:org/10.1016/j.tecto.2007.02.017.

Vauchez, A., Dineur, F., Rudnick, R., 2005.Microstructure, texture and seismic anisotropyof the lithosphereic mantle above a mantle plume: insights from the Labait volcanoxenoliths (Tanzania). Earth and Planetary Science Letters 232, 295–314.

Walker, K.T., Bokelmann, G.H.R., Klemperer, Bock, G., 2005. Shear-wave splitting aroundthe Eifel hotspot: evidence for a mantle upwelling. Geophysical Journal International163, 962–980. doi:0.1111/j.1365-246X.2005.02636.x.

Wawerzinek, B., 2008. Bericht über Krustenlaufzeitkorrekturen an den TIMO Stationen.Geophysikalisches Institut, Universität Karlsruhe, Unpublished report. . 11pp.

Wessel, P., Smith,W.H.F., 1995. New version of the generic mapping tools released. EOS,Transactions, American Geophysical Union 76, 329.

Wilde-Piórko, M., Geissler, W.H., Plomerová, J., Grad, M., Babuška, V., Brückl, E., Cyziene, J.,Czuba,W., England, R., Gaczyński, E., Gazdova, R., Gregersen, S., Guterch, A., Hanka,W.,Hegedűs, E., Heuer, B., Jedlička, P., Lazauskiene, J., Keller, G.R., Kind, R., Klinge, K.,Kolinsky, P., Komminaho, K., Kozlovskaya, E., Krüger, F., Larsen, T., Majdański, M.,Málek, J., Motuza, G., Novotný, O., Pietrasiak, R., Plenefisch, T., Růžek, B., Sliaupa, S.,Środa, P., Świeczak,M., Tiira, T., Voss, P.,Wiejacz, P., 2008. PASSEQ2006-2008: passiveseismic experiment in Trans-European Suture Zone. Studia Geophysica et Geodaetica52, 439–448. doi:10.1007/s11200-008-0030-2.

Ziegler, P.A., 1992. European Cenozoic rift system. Tectonophysics 208, 91–111.